You are on page 1of 8

Energy and Buildings 62 (2013) 222–229

Contents lists available at SciVerse ScienceDirect

Energy and Buildings


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

On the improvement of natural ventilation models


Roberto Z. Freire a,∗ , Marc O. Abadie b , Nathan Mendes c
a
Pontifical Catholic University of Parana – PUCPR, Industrial and Systems Engineering Graduate Program – PPGEPS, Rua Imaculada Conceição, 1155, Prado
Velho, 80215-901 Curitiba, PR, Brazil
b
Laboratoire des Sciences de l’Ingénieur pour l’Environnement – LaSIE, University of La Rochelle, 17000 La Rochelle, France
c
Pontifical Catholic University of Parana – PUCPR, Mechanical Engineering Graduate Program, Thermal Systems Laboratory – LST, Rua Imaculada
Conceição, 1155, Prado Velho, 80215-901 Curitiba, PR, Brazil

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

Article history: This paper aims to validate and improve three cross and single-sided natural ventilation models imple-
Received 18 November 2012 mented in a whole-building hygrothermal and energy simulation program. The tested models are the
Received in revised form 17 February 2013 British Standard for cross ventilation, the de Gids and Phaff’s and Larsen’s for single-sided ventilation.
Accepted 20 February 2013
Airflow rates obtained by those models have been compared to the measurements performed in two
full-scale buildings: one single room house located in a wind tunnel facility and one real three-storey
Keywords:
building. Results show a large variation of airflow rates provided by the different models. The Larsen’s
Single-sided natural ventilation
model can be improved if coupled to the CPCALC algorithm, providing better results for both wind tunnel
Wind tunnel
Building simulation
and on-site experiments.
Airflow © 2013 Elsevier B.V. All rights reserved.

1. Introduction ventilation through a unique aperture (single-sided ventilation)


and through multiple ones (cross ventilation). If the modeling
Healthy indoor climate conditions and, at the same time, energy of cross ventilation is currently well defined [8], that of single-
efficient and environmentally friendly building is a clear chal- sided ventilation is widely discussed since early 80s because of the
lenge. This is valid for existing buildings as well as for early-stage complexity for existing models to reproduce the airflow through
design processes. Creating better indoor air conditions to the occu- just one aperture in a building zone. Differently from the cross-
pants is certainly the main aspect when health and productivity ventilation, the turbulence of the wind and variation in the pressure
are taken into account. Full air-conditioned systems were in the gradients induced by e.g. wind gusts strongly affect the airflow
past considered as the ultimate choice, but today a more bal- through an opening [9] in single-sided ventilation. Since those
anced view is found in many countries and among many people parameters are unsteady, the airflow in single-sided ventilation is
[1]. Nowadays, the combination of natural ventilation and air much more difficult to evaluate.
conditioning systems are essential to reduce building energy con- In 1982, the first most relevant model to calculate natural
sumption. Studies considering the occupants behavior related to single-sided ventilation in buildings was presented by de Gids and
window opening are now simulated and the effects are found Phaff [10]. They proposed an approach to calculate the airflow in
to be relevant when energy consumption and thermal comfort single-sided ventilation zones, where the air change resulting from
are taken into account [2]. In this way, natural ventilation has opening a window in a room was the subject of investigation. Mea-
been widely adopted in different ways as strategy to reduce surements have been carried out in different locations, all located
energy consumption and improve thermal comfort condition in on the first floor of buildings situated in an urban environment
buildings. Natural ventilation has already proved its efficiency as and surrounded by buildings up to 4 floors high. They have then
night time ventilative cooling [3,4], wind towers [5], atrium ven- proposed an empirical expression to calculate the airflow through
tilation [6] and to reduce roof solar heat gain [7] among other an opening based on wind velocity, air temperature variation and
applications. opening area.
Natural ventilation can be achieved by means of inlet grilles and Seventeen years after the presentation of the de Gids and Phaff
duct system or simply by large openings (façade or roof windows, model, the British Standards published a slighlty different mathe-
doors). For the latter case, two configurations can be considered: matical expression to calculate the airflow in a single-zone with just
one opening [8]. In 2005, the American Society of Heating, Refriger-
ating and Air-Conditioning Engineers proposed another expression
∗ Corresponding author. Tel.: +55 41 3271 1333; fax: +55 41 3271 2579. to calculate the airflow in a single-sided ventilation residential
E-mail addresses: roberto.freire@pucpr.br, rozafre@terra.com.br (R.Z. Freire), buildings [11]. That expression was the first to take into account
mabadie@univ-lr.fr (M.O. Abadie), nathan.mendes@pucpr.br (N. Mendes). the wind incidence angle.

0378-7788/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.enbuild.2013.02.055
R.Z. Freire et al. / Energy and Buildings 62 (2013) 222–229 223

Finally, in 2006, Larsen concluded that a more precise design


expression found from wind tunnel measurements can be used to
predict airflows from single-sided ventilation [9] by completing the
de Gids and Phaff’s model. From Larsen’s experimental work (2006),
the authors noticed that the wind prevails on the windward side
and that the temperature difference stands out on the leeward side
of the building.
Recently, more advanced analysis to calculate the airflow
through openings has been performed. Based on the evolution of
computational hardware and on advances of CFD (Computational
Fluid Dynamics) softwares, new methods to study the airflow in
single or more than one openings started to be developed [12–18].
However, in the case of natural ventilation, the CFD approach
requires the modeling of large computational domain and the use
of high-order turbulence models such as Reynolds Stress Models
[19] in order to represent correctly the airflow around the build-
ings. Consequently, the use of CFD is complex and time consuming
so that empirical expressions are still widely used when whole-
building hygrothermal analysis are necessary.
However, according to Zhai et al. [20], empirical natural venti-
lation models implemented in whole-building energy simulations
Fig. 1. Pressure difference around the neutral plane in a building ventilated only by
tools usually present problems as significant functional limitations. thermal buoyancy.
In this way, the main idea of this paper is to validate the integration
of the cross ventilation model proposed by the British Standard [8]
and to compare the prediction given by the single-sided ventilation
indoor temperature is lower than the outdoor temperature, then
models of de Gids and Phaff [10] and Larsen [9] to results obtained
everything is the opposite. If one opening is sectioned by the neu-
from wind tunnel and on-site experiments. The model descriptions
tral plane level, its area should be divided into the n and m opening
of the cross and single-sided ventilations are presented in the first
sum.
two sections. The two methodologies for evaluating the wind pres-
To calculate the airflow rate caused by wind in a cross-
sure coefficient, that is necessary as an input parameter of both
ventilation problem, the following equation should be used [8]:
ventilation models, are given in the third section. The fourth sec-
tion presents the wind tunnel and on-site configuration. Results are  
presented and discussed in the last section. QV = CD Aw U (z = BH) CP,1 − CP,2  (3)

2. Cross ventilation model


where Aw can be calculated by using Eq. (2) but, in these case, the
n-openings are the openings with positive CP (pressure coefficient)
According to the literature, cross-ventilation airflow rates are
values, otherwise they are m-openings. U(z = BH) is the wind veloc-
much simpler to calculate than single-sided ventilation rates. The
ity at the building’s height and CP,i=1,2 are the pressure coefficients
flows induced by either thermal buoyancy, wind or a combination
that can be evaluated according to the expressions given in third
of them presented by the British Standards [8] have been adopted
section. The wind velocity at the building’s height is calculated
to define the cross-ventilation airflow formulation employed here.
using Counihan’s expression [21]:
The airflows through more than one opening could be driven
either by thermal buoyancy (if there are openings at different U(z)
 z ˛
heights in the building) or by the wind. A general expression to = (4)
U(10) 10
calculate the pressure driven airflow (in m3 /s) through “m + n”
openings clustered on two heights due to thermal buoyancy is:
where ˛ depends on the terrain roughness (from 0.10 to 0.45).

Following the British Standards, if the inequality:
2TgH
QV = CD Ab (1)
TM 
U(z = BH) Ab H
where the discharge coefficient for window openings CD is often √ < 0.26 (5)
T Aw CP
within the range between 0.60 and 0.75 (an average value of 0.67
is used in the present study), T is the temperature difference
is satisfied, the airflow should be calculated through Eq. (1), other-
between inside and outside (K), g is the gravity acceleration (m/s2 ),
wise Expression (3) should be used.
H is the distance between the two levels of openings (m), TM is
the mean temperature between the indoor and outdoor air (K) and
Ab is the area of “m + n” openings (m2 ) that can be calculated as 3. Single-sided ventilation models
presented in Eq. (2).
1 1 1 In what follows, two models to calculate the airflow for
= n 2 + m 2 (2)
A2b single-sided ventilated environments are presented. The modeling
A
n=1 n
A
m=1 m procedure performed by de Gids and Phaff [10] used measure-
where n and m are the number of openings. If the indoor temper- ments performed in a building located in an urban environment,
ature is higher than the outdoor temperature, the n-openings will while the model proposed by Larsen [9] used data from both
be the ones over the neutral plane level (where the heated air flows wind tunnel and on-site experiments. The main difference is
outward) and the m-openings are the ones under the neutral plane that of Larsen’s model integrates the effect of wind incidence
level (where the cool air flows inward), see Fig. 1. In cases where the angle.
224 R.Z. Freire et al. / Energy and Buildings 62 (2013) 222–229

Table 1
Constants C1 , C2 and C3 [22].

Direction Incidence angle (ˇ) C1 C2 C3

Windward 0◦ –75◦ ; 285◦ –360◦ 0.0015 0.0009 −0.0005


Leeward 105◦ –255◦ 0.0050 0.0009 0.0160
Parallel 90◦ or 270◦ 0.0010 0.0005 0.0111

3.1. de Gids and Phaff (1982)

From measurements, de Gids and Phaff found an expression


that describes the airflow rate (in m3 /s) in a single-sided ventilated
building [10]:

QV = C1 U(10)2 + C2 hT + C3 (6)


2

where A is the opening area (in m2 ), h is the opening height (in m),
U(10) is the meteorological station reference velocity at 10 m high
(m/s) and T is the mean temperature difference between inside
and outside (K). Fig. 2. Variation of surface-averaged wall pressure coefficients for low-rise build-
The coefficients of this model account for wind (dimensionless ings [26].

coefficient C1 = 0.001), buoyancy constant (C2 = 0.0035 m/[s2 .K])


and wind turbulence constant (C3 = 0.01 m2 /s2 ). 4. Pressure coefficient calculation

3.2. Larsen (2006) The pressure created by the wind on the building is described
by Eq. (10). It is calculated by multiplying a dimensionless pressure
As a result of experiments performed in 2006, Larsen proposed a coefficient CP by the dynamic pressure.
new model [9] to describe the airflow for a single-sided ventilation 1
case, which is presented in Eq. (7). Pwind = CP e U(z)2 (10)
2
where U(z) is calculated by Eq. (4). The CP coefficient is a function of
  CP,opening T
QV = A C1 f (ˇ) CP  U(z = BH)2 + C2 hT + C3
2 the shape of the building, the wind direction and the surrounding
2
U(z = BH) terrain. In what follows, two models to calculate the distribution
(7) of CP are presented. The first one considers a unique value for the
whole surface whereas the second one calculates the CP value at
where CP is the pressure coefficient which can be calculated by any any location on the surface.
of the methods presented in the third section of this paper, U(z = BH)
is the wind velocity at the building height, h is the opening height 4.1. Mean CP calculation
(m). The dimensionless coefficient depending on the wind effect
C1 , the buoyancy constant C2 , and the turbulence constant C3 are According to the American Society of Heating, Refrigerating and
defined in Table 1. If U(z) tends to zero, C2 is the only term to be Air Conditioning Engineers [23], the distribution of CP on a low-rise
considered, assuring the consistency of the equation. building associated to the variation of the incidence angle can be
In Eq. (7) CP,opening and f(ˇ) are calculated from: estimated through the curve presented in Fig. 2. Deru and Burns
[24] showed that there are several correlations for the wind pres-
CP,opening = 9.1894−9 ˇ3 − 2.62610−6 ˇ2 − 0.0002354ˇ + 0.113
sure coefficient derived from wind tunnel experimental data in
(8) order of increasing complexity and accuracy, as those proposed
by Walton [25], Swami and Chandra [26] and the COMIS group
[27]. These correlations are potentially inaccurate in situations
that introduce high turbulence intensity to the airflow, e.g. high
f (ˇ) = −3.284312 × 10−9 ˇ4 + 2.363134 × 10−6 ˇ3 terrain roughness or local shielding, irregular shaped buildings
− 5.24549 × 10−4 ˇ2 +3.581077 × 10−2 ˇ+0.3017574 (9) (nonrectangular or rectangular with aspect ratios far from a cube)
or buildings with overhangs or fins. The model developed by Swami
Eqs. (8) and (9) are calculated using ˇ, which is the wind incidence and Chandra [26] was selected as the best fit for the needs of this
angle (◦ ). work:
It has also been seen that the values of the constants C1 , C2 and C3
ˇ
depend on the wind direction. This is due to the fact that the flows Cp = Cp (ˇ = 0) × ln 1.248 − 0.703 sin − 1.175 sin2 ˇ
in the three cases (windward, leeward and parallel) are very differ- 2
ent one from each other and therefore also have different weighting ˇ ˇ
of the terms including wind pressure, thermal forces and fluctuat- +0.131 sin3 (2ˇG) + 0.769 cos + 0.07G2 sin2
2 2
ing forces. Contrarily to what could be expected, C1 does not have
the largest weight factor at windward side, but it remains the most ˇ
+0.717 cos2 (11)
dominating factor in this case. In the case with an opening located 2
on the leeward side of the building, the fluctuating term (the third
one of Eq. (7)) prevails. This is also the case for parallel wind situa- This expression calculates the surface pressure coefficient nor-
tions; however, the difference is not as high as it is for the leeward malized to the pressure coefficient at zero incidence angle as a
case. function of the wind incidence angle (ˇ) within a [0◦ , 180◦ ] domain,
R.Z. Freire et al. / Energy and Buildings 62 (2013) 222–229 225

Fig. 3. Experiments designs performed into the PowerDomus [33] software: (a) and (b) wind tunnel for cross and single-sided cases; (c) NOA building (single-sided case).

where values higher than 180◦ are obtained by symmetry, and 5. Experiments
G, the natural logarithm of the side ratio (ratio of the lengths of
adjacent walls L and W). For vertical walls, Swami and Chandra In this section, the experiments used to analyze and compare
recommend using a value of 0.6 for the pressure coefficient at zero the cross and single-sided ventilation models shown in the previous
incidence angle [27]. section are presented. The wind tunnel experiment presents results
for both cross and single-sided ventilation whereas the on-site one
is dedicated to the validation of the single-sided ventilation model.
4.2. CPCALC model

The CPCALC algorithm has been developed within the frame of 5.1. Wind tunnel experiment
the European Research Programme PASCOOL (Passive Cooling of
Buildings) of the Commission of the European Communities, Direc- The wind tunnel experiment has been carried out in a full-scale
torate General for Energy [28,29]. In 1992, the algorithm started to wind tunnel at the Japanese Building Research institute (BRI) by
be developed at the Lawrence Berkeley National Laboratory [27,30] Larsen [9] to investigate the airflow through openings in cross
within the COMIS workshop on infiltration and ventilation, and and single-sided ventilation situations. The building dimensions
being upgraded within the IEA-ANNEX 23 on multizone airflow are 5.56 m × 5.56 m × 3.00 m, which means that scale effects were
modeling [31]. avoided. The openings width and height are 0.86 m and 0.15 m for
CPCALC has been developed in order to fulfill the requirements both windows in the cross-ventilation case, positioned 0.54 m away
of multizone airflow models, which need a detailed evaluation of from the right edge and 0.925 m from the top as shown in Fig. 3a.
the wind pressure distribution around buildings. Scientists and pro- For the single-sided case, the openings dimensions are 0.86 m and
fessionals using this program, and who do not have the possibility 1.40 m of width and height, respectively. In this case, they are pos-
to test a scale model of their building in a wind tunnel, do not need itioned at 0.54 m away from the right edge and 0.69 m away from
to extrapolate CP data from tables usually yielding wall-averaged the top of the building as illustrated in Fig. 3b. The internal room
CP values [32]. height is 2.4 m and the thickness of the walls is 0.10 m. The room
The CPCALC model adopted in this work allows for calculating volume is 68.95 m3 .
wind pressure coefficients on the envelope of a block-shaped build- The experiment consisted in varying the wind speed in the tun-
ing with flat roof and uses the following input variables: ˇ wind nel (1, 3 and 5 m/s) with a turbulence intensity less than 5% while
incidence angle (◦ ), ˛: wind velocity profile [21], sbh: surrounding imposing distinct temperature differences of 0, 5 and 10 K between
building height (m), pad: plan area density (%), building height (m), the internal and external air. The wind speed profile created in this
wall azimuth (m) the coordinates x and y of the middle of the open- wind tunnel was almost uniform, which resulted in a wind profile
ing related to the origin of the building (m) and the frontal and side that differs from outdoor conditions as it was not able to reproduce
aspect ratios of the building (m). the atmospheric boundary layer. The buildings was also rotated
Based on these input data, the CPCALC algorithm is able to cal- between 0◦ and 345◦ with either a 15◦ or a 30◦ increase to get mea-
culate the pressure coefficient value at any point on a building surements for different angles of the wind. A total of 159 different
surface. In the present study, the interest point is the center of the cases were studied. The air-change rate was measured with the
openings. tracer gas decay method.
226 R.Z. Freire et al. / Energy and Buildings 62 (2013) 222–229

Table 2
Opening configuration and mean climatic conditions for single-sided ventilation
experiments in the NOA building.

Experiment TI TO U(10) ˇ

A1 + A2 31.4 31.3 6.8 40


B1 + B2 31.8 32.6 3.0 70
C 32.1 30.6 5.0 30
A2 + B2 31.5 32.5 6.7 50
A1 + A2 + B1 + B2 31.5 30.5 1.7 50
B1 + B2 + C 29.2 28.8 1.6 45
All 31.0 30.2 3.6 12
A2 + C 31.7 31.2 5.4 30
B2 + C 31.8 30.7 4.9 70
A1 + A2 + C 31.0 30.8 4.2 50
A1 + B1 + C 28.8 27.6 2.0 35
A2 + B2 + C 31.6 30.1 5.0 20
A1 + A2 + B1 + C 31.0 29.6 3.1 35
A1 + A2 + B2 + C 31.0 28.2 3.4 37

6. Results

This section presents the comparisons between the results


obtained by using the PowerDomus software [35] and the
experimental data obtained for both wind tunnel and on-site exper-
iments.

6.1. Wind tunnel simulation

6.1.1. Cross ventilation


Fig. 5 presents the results of wind tunnel cross ventilation exper-
iments of Larsen [9] and the predictions of the British Standard [8]
implemented in the PowerDomus software.
Among the six different experiments (three different wind
velocities and three distinct temperature differences between the
Fig. 4. Window parts of the NOA building. indoor air and the outdoor one) performed by Larsen [9], only one
is presented here (wind velocity of 1 m/s, isothermal case). In fact,
as shown by Eq. (3), the airflow rate is a linear function of the wind
velocity so that results for other velocities can be easily obtained
from the presented ones. The experimental results actually clearly
5.2. On-site experiment showed this linear tendency. Moreover, because of the fact that
the openings are located at the same height, there is no thermal
The building selected for the on-site experiment is the Institute buoyancy effect. In this way the negligible air change rate vari-
of Meteorology and Physics of the Atmospheric Environment, ations caused by temperature differences has not been treated in
which is a three-storey, naturally ventilated, office building this case of dominant wind effect. Results obtained by the model are
referred as the NOA (National Observatory of Athens) building in the same for distinct temperature differences in the present config-
the sequence. Experimental data have been collected from previ- uration. The experimental results showed also that tendency with
ous studies [33,34] with the purpose of comparing the air flow rate a slight variation probably due to the precision of the experimen-
of a single-sided ventilated office room to simulation results. Each tal measurements. The experimental results presented in Fig. 5 are
floor of NOA building is about 4.50 m high and the dimensions are
10.20 m × 16.30 m of length and width respectively. Ventilation
experiments were held on the first floor [31]. The selected office
room (zone in red in Fig. 3c) was isolated from the rest of the
building. The room has a 13.59 m2 floor area, while its length is
equal to 3.00 m.
The only external window is on the west wall and is divided in
five parts, Al , A2 , B1 , B2 and C, which can open separately, providing
the possibility to vary the opening area (Fig. 4) by opening different
parts. The total window area is 2.50 m2 and its angle to the North
is 315◦ .
Dimensions and area of each part of the opening are presented
in [33]. Airflow rates across the openings have been performed
according to the single tracer gas decay technique. Fourteen
different experiments have been taken into account. The mean
climatic conditions for each experiment are given in Table 2, where
TI is the indoor temperature (◦ C); TO is the outdoor temperature
(◦ C), U10 is the wind velocity at 10 m high (m/s) and ˇ is the wind
Fig. 5. Comparisons between the experimental and simulation results performed
incidence angle (◦ ). for the wind tunnel – cross ventilation case.
R.Z. Freire et al. / Energy and Buildings 62 (2013) 222–229 227

Fig. 6. Comparisons between the experimental and simulation results performed into the PowerDomus software for the wind tunnel case with wind speeds of: (a) 1, (b) 3
and (c) 5 m/s – single-sided ventilation case.

the airflow rate averages obtained from three distinct temperature 6.1.2. Single-sided ventilation
differences results. In order to illustrate the behavior of each single-sided ven-
Results show that the British Standard (Mean CP ) model is not tilation model and to analyze the effect from different wind
capable of predicting the variation of the airflow according to the speeds (1, 3 and 5 m/s), temperature differences (0, 5 and 10 ◦ C)
wind direction and predict a nearly constant air change rate of and incidence angles (varying from 0 to 345◦ ) on the airflow,
2 h−1 . The British Standard (CPCALC) model tends to better follow 27 simulations using the PowerDomus software have been per-
this variation with air change rates varying between 0.5 and 3.5 h−1 . formed. Every simulation consists in varying the incidence angle
One particular drawback of the Mean CP -based model occurs when and obtaining the air change rate to each natural ventilation
the wind is parallel to the openings. In this case, this model predicts model.
no flow (the CP difference between the openings is null) whereas As simulation parameters, for the pressure coefficient calcula-
the CPCALC-based model is able to detect a small but noticeable tion through the Mean CP method an ˛ = 0.10 has been adopted,
airflow. which is the value when there are no obstructions affecting the
One small difference remains, i.e. the results between the 90◦ wind. For the CPCALC method the same ˛ = 0.10 has been used and
and 270◦ cases. This comes from the fact that CPCALC tables provide for the plan area density and surrounding building height the values
different measurement results for both angles according to the of pad = sbh = 0, have been adopted because there are no obstruc-
windward and leeward configurations. In fact, those two cases are tions inside the wind tunnel.
particularly sensitive as a very small deviation from 90◦ or 270◦ Fig. 6 presents the air-change rates as a function of the incidence
will change the airflow pattern around the building (and the CP angle and the temperature difference of 5 ◦ C. Each graphic repre-
values) and the windward/leeward status of the wall. In CPCALC, sents one of the selected wind speeds. Larsen’s model does present
two distinct correlations (one for windward, the other for leeward) the expected angular dependency. For the lower wind velocity,
exist and they don’t exactly match at those “boundary” angles. The wind and temperature gradient effects are about the same so that
authors have chosen to keep the original algorithm instead of cor- the third term of Eq. (6) really affects the results. In particular, the
recting them by averaged value or so, as they are not so different influence of the non-symmetrical term CP is visible for the inci-
in reality (0.1 on CP values). dence angle in 120◦ ≤ ˇ ≤ 240◦ . For higher wind velocity, the first
However, it can be seen that the two models present almost the term of Eq. (6), and f(ˇ) term, predominates so that the obtained air
same mean relative difference of about 30% considering the whole change rate becomes more symmetrical, at least when the Mean CP
set of data (Table 3). method is used.
It is can be noticed, from the results obtained by the de Gids
and Phaff model (Fig. 6), that there is no variation of the air change
rates with incidence angle. This happens because the model does
Table 3
Relative differences (%) for the wind tunnel experiment–cross ventilation case. not take into account the incidence angle in its formula. In this way,
a constant single value is obtained for each wind speed. However,
Model Relative differences (%)
the overall performance of de Gids and Phaff model should be taken
Mean CP (A) 32.46 into account especially in cases where orthogonal structures and
CPCALC (A) 31.11 low-rise buildings have to be simulated.
228 R.Z. Freire et al. / Energy and Buildings 62 (2013) 222–229

Table 4 Table 5
Relative differences (%) for the wind tunnel experiment – single-sided ventilation Relative differences for the NOA building experiment.
case.
Model Relative differences (%)
Model Windward Leeward Parallel
Larsen (mean CP ) 27.71
Larsen (mean CP ) 34.25 20.85 22.38 Larsen (CPCALC) 24.03
Larsen (CPCALC) 24.99 19.75 7.68 de Gids and Phaff 49.06
de Gids and Phaff 29.89 20.39 14.68

The differences noticed between the Larsen (Mean CP ) and


better prediction in the wind tunnel experiment. The slightly better
Larsen (CPCALC) models originate from the calculation of the pres-
results of Larsen (CPCALC) compared to Mean CP essentially occur
sure coefficient. While the Mean CP method calculates the wall
for 3 points (C, All, A2 + C) where the incidence angle is about 70◦
mean pressure coefficient, the CPCALC method estimates the CP
for which this model also showed better prediction on the wind
value for the geometric center of the opening. As a consequence,
tunnel experiment. In the other side, the de Gids and Phaff model,
the CP values calculated by the CPCALC method are higher when
which presented great results for the wind tunnel experiment, was
the wind incidents directly on the window (angle in the interval
not capable to provide good results in this case.
of 270◦ ≤ ˇ ≤ 360◦ ) than for angles between 0◦ ≤ ˇ ≤ 90◦ . The Mean
CP method is not able to represent this actual behavior. The rela-
tive differences for the windward, leeward and parallel incidence
7. Conclusions
angles have been presented in Table 4. It is noticed for the leeward
incidence angles, the Larsen’s model by using the CPCALC calcu-
Several cross and single-sided natural ventilation simulations
lation presents slightly better results than the two other models.
through openings were performed to test and compare three mod-
However, the CPCALC based model provides much better results
els. Those simulations have been contrasted to experimental results
for the two other cases, especially when the incidence angle goes
obtained from both a full-scale wind tunnel and an on-site three-
near 90◦ (parallel incidence).
storey office building.
All models have been implemented into PowerDomus tool and
6.2. On-site simulation calculations have been performed using the software interface.
However, it is true that for the present study, the whole build-
The comparisons of the on-site experiment and the simulations ing and energy calculation capabilities of the tool have not been
performed by PowerDomus are compared in this section. According used as all boundary conditions (temperatures) have been imposed.
to Dascalaki et al. [34], the building is located in an open urban Furthermore, the natural ventilation algorithm available into Pow-
environment on top of a hill across from the Acropolis of Athens, erDomus has been developed in order to recognize and to adopt the
consequently, the simulation parameter ˛ = 0.28 has been chosen correct natural ventilation model (cross or single-sided ventilation)
for the pressure coefficient calculation by the Mean CP and by the according to the n-openings available at each time step during sim-
CPCALC methods. For the last one, the pad = sbh = 0 have also been ulation. In this way, distinct geometries of block-shaped buildings
used. Fig. 7 shows the comparisons between the simulation and the can be simulated.
measured results. Results showed that the use of the CPCALC algorithm can sub-
According to the results presented in Fig. 7, it has been noticed stantially improve the predictions while avoiding poor prediction
that the Larsen model using the CPCALC method to calculate the for particular wind-to-wall angle, especially for parallel wind. The
pressure coefficient has provided the best results. When the relative benefits of using the CPCALC algorithm are greater for the case of
differences for all the obtained results are calculated, the graphical single-sided configurations.
analysis is confirmed (Table 5). Even more elaborate models, e.g. as Larsen using CPCALC
It should be noticed that the wind incidence angle stayed around calculation, are not as accurate to predict single-sided natural ven-
90◦ during the experiments for which Larsen model already showed tilation. It can be explained by the fact that single-sided ventilation
is a nonlinear complex configuration. However, the model devel-
oped by de Gids and Phaff should not be disused and its overall
performance should be taken into account especially in cases where
orthogonal structures and low-rise buildings have to be simulated.
In these particular circumstances, de Gids and Phaff model can be
adopted to reduce computational efforts and consequently simula-
tion time, even not considering wind incidence angle and openings
location.
It has also been found that the current state-of-the-art empir-
ical models for natural ventilation are capable of predicting the
actual trends. However, the predictions still present a high differ-
ence of about 30%. More accurate empirical models are still needed
to evaluate the air change rate by cross and single-sided natural
ventilation.

Acknowledgments

The authors thank the national financial supports from CNPq


(Conselho Nacional de Desenvolvimento Científico e Tecnológico),
Fig. 7. Comparisons between the experimental and simulation results performed CAPES (Coordenação de Aperfeiçoamento de Pessoal de Nível Supe-
into the PowerDomus software for the NOA Building case. rior) and Eletrobrás (Centrais Elétricas Brasileiras S.A.).
R.Z. Freire et al. / Energy and Buildings 62 (2013) 222–229 229

References [19] M.G. Emmel, M.O. Abadie, N. Mendes, New external convective heat transfer
coefficient correlations for isolated low-rise buildings, Energy and Buildings 39
[1] A. Van der Aa, P.O.’t Veld, Technical report, RESHYVENT – A EU Cluster Project on (2007) 335–342.
Demand Controlled Hybrid Ventilation for Residential Buildings, RESHYVENT [20] Z.J. Zhai, M. Johnson, M. Krarti, Assessment of natural and hybrid ventilation
Project, 2004. models in whole-building energy simulation, Energy and Buildings 43 (2011)
[2] I. Oropeza-Pereza, P.A. Østergaardb, A. Remmen, Model of natural ventilation 2251–2261.
by using a coupled thermal-airflow simulation program, Energy and Buildings [21] J. Counihan, Adiabatic atmospheric boundary layer: a review and analysis of
49 (2012) 388–393. data from the period 1880–1972, Atmospheric Environment 10 (9) (1975)
[3] N. Artmann, R.L. Jensen, H. Manz, P. Heiselberg, Experimental investigation of 871–905.
heat transfer during night-time ventilation, Energy and Buildings 42 (2010) [22] T.S. Larsen, P. Heiselberg, Single-sided natural ventilation driven by wind
366–374. pressure and temperature difference, Energy and Buildings 40 (2008)
[4] S.J. Emmerich, B. Polidoro, J.W. Axley, Impact of adaptive thermal comfort on cli- 1031–1040.
matic suitability of natural ventilation in office buildings, Energy and Buildings [23] ASHRAE, Handbook of Fundamentals, in: Airflow Around Build, American Soci-
43 (2011) 2101–2107. ety of Heating, Refrigerating and Air Conditioning Engineers, Atlanta, GA, 2005
[5] B.R. Hughes, J.K. Calautit, S.A. Ghani, The development of commercial wind (Chapter 16).
towers for natural ventilation: a review, Applied Energy 92 (2012) 606–627. [24] M. Deru and P. Burns, Technical Report NREL/CP-550-33698, Infiltration and
[6] G. Gan, Simulation of buoyancy-driven natural ventilation of buildings: impact natural ventilation model for whole-building energy simulation of residen-
of computational domain, Energy and Buildings 42 (2010) 1290–1300. tial buildings, National Renewable Energy Laboratory (NREL), Golden, CO, USA,
[7] L. Susanti, H. Homma, H. Matsumoto, Y. Suzuki, M. Shimizu, Numerical simu- 2003.
lation of natural ventilation of a factory roof cavity, Energy and Buildings 42 [25] G.N. Walton, Airflow and multiroom thermal analysis, ASHRAE Transactions 88
(2010) 1337–1343. (2) (1982) 78–91.
[8] British Standards, Code of Practice for Design of Buildings: Ventilation Princi- [26] M.V. Swami, S. Chandra, Correlations for pressure distribution of buildings and
ples and Designing for Natural Ventilation, 1st ed., British Standards Institution calculation of natural-ventilation airflow, ASHRAE Transactions 94 (4) (1988)
(BS5925), 1999. 244–266.
[9] T.S. Larsen, Ph.D. thesis. Aalborg University – Department of Civil Engineering, [27] H.E. Feustel, A. Rayner-Hooson, COMIS fundamentals, in: Technical Report,
Group of Architectural Engineering. ISSN 1901-7294, Denmark, 2006. Applied Science Division – Lawrence Berkeley National Laboratory (LBL-
[10] W. de Gids, H. Phaff, Ventilation rates and energy consumption due to open 28560), Berkeley, CA, USA, 1990.
windows, in: A Brief Overview of Research in the Netherlands, vol. 1, Proc. of [28] M. Grosso, Modelling wind pressure distribution on buildings for passive
the Third IEA Air Infiltration Center Conference, 1982, pp. 4–5. cooling, in: Proceedings of the International Conference Solar Energy in
[11] ASHRAE, Handbook of Fundamentals, in: Ventilation and Infiltration, American Architecture and Urban Planning, Commission of the European Communities,
Society of Heating, Refrigerating and Air Conditioning Engineers, Atlanta, GA, Florence, 1993, pp. 17–21.
2005 (Chapter 27). [29] M. Grosso, D. Mariano, E. Parisi, Wind pressure distribution on flat and tilted
[12] K.A. Papakonstantinou, C.T. Kiranoudis, N.C. Markatos, Numerical simulation roofs: a parametrical model, in: Proceedings of the European Conference on
of air flow field in single-sided ventilated buildings, Energy and Buildings 33 Energy Performance and Indoor Climate in Buildings, Lyon, France, 1994, pp.
(2000) 41–48. 24–26.
[13] Y. Jiang, D. Alexander, H. Jenkins, Q. ad, R.A. Chen, Natural ventilation in [30] M. Grosso, Wind pressure distribution around buildings: a parametrical model,
buildings: measurement in a wind tunnel and numerical simulation with large- Energy and Buildings 18 (2) (1992) 101–131.
eddy simulation, Journal of Wind Engineering and Industrial Aerodynamics 91 [31] International Energy Agency (IEA) – Annex 23: Multizone airflow modelling.
(2003) 331–353. Evaluation of COMIS Appendices, Technical Report, Swiss Federal Institute of
[14] H-M. Kao, T-J. Chang, Y-F. Hsieh, C-H. Wang, C-I. Hsieh, Comparison of airflow Technology, Lausanne. LESO.PB, Institute of Building Technology, Torino, Italy,
and particulate matter transport in multi-room buildings for different natural 1996.
ventilation patterns, Energy and Buildings 41 (2009) 966–974. [32] M.W. Liddament, Technical Report, Air infiltration calculation techniques – an
[15] K.-S. Nikas, N. Nikolopoulos, A. Nikolopoulos, Numerical study of a naturally applications guide, Air Infiltration and Ventilation Centre, Bracknell, UK, 1986.
cross-ventilated building, Energy and Buildings 42 (2010) 422–434. [33] E. Dascalaki, M. Santamouris, M. Bruant, C.A. Balaras, A. Bossaer, D. Ducarme, P.
[16] C. Walker, G. Tan, L. Glicksman, Reduced-scale building model and numerical Wouters, Modeling large openings with COMIS, Energy and Buildings 30 (1999)
investigations to buoyancy-driven natural ventilation, Energy and Buildings 43 105–115.
(2011) 2404–2413. [34] E. Dascalaki, M. Santamouris, A. Argiriou, C. Helmis, D.N. Asimakopoulos, K.
[17] M.Z.I. Bangalee, S.Y. Lin, J.J. Miau, Wind driven natural ventilation through mul- Papadopoulos, A. Soilemes, Predicting single sided natural ventilation rates in
tiple windows of a building: a computational approach, Energy and Buildings buldings, Solar Energy 55 (5) (1995) 327–341.
45 (2012) 317–325. [35] N. Mendes, R.C.L.F. Oliveira, G.H. Santos, Domus 2.0: A Whole-Building
[18] S. Hussain, P.H. Oosthuizen, Numerical investigations of buoyancy-driven natu- Hygrothermal Simulation Program, Proc. of the Eighth International Conference
ral ventilation in a simple atrium building and its effect on the thermal comfort on Building Performance Simulation (IBPSA’03), Eindhoven, The Netherlands,
conditions, Applied Thermal Engineering 40 (2012) 358–372. 2003.

You might also like