You are on page 1of 26

Article

Journal of Building Physics

Assessment of the Ó The Author(s) 2018


1–26

ventilation efficiency in Reprints and permissions:


sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/1744259118771314
the breathing zone during journals.sagepub.com/home/jen

sleep through
computational fluid
dynamics techniques

Susana Hormigos-Jimenez1,
Miguel Ángel Padilla-Marcos2, Alberto Meiss2,
Roberto Alonso Gonzalez-Lezcano1 and
Jesús Feijó-Muñoz2

Abstract
In developed countries, presence at home varies between 60% and 90% of the day,
sleeping supposes 30%. Therefore, it is essential to ensure good indoor air quality that
enhances health and benefits rest and recovery. In this context, it is necessary to
achieve a balance between energy efficiency and air distribution parameters; thus, the
influence exerted by the furniture of a bedroom on the air exchange efficiency, in the
breathing zone during sleep, is assessed in this study. Computational fluid dynamics
techniques, experimentally validated by the tracer gas (SF6) concentration decay
method, are used to analyze 52 case studies corresponding to the same space, but vary-
ing both the number and the arrangement of the furniture inside. It is concluded that,
in order to achieve a significant improvement in the air exchange efficiency, the number
of elements included in the bedroom is not relevant, but the position of them. The
highest increase in the ventilation efficiency in breathing zone is observed when the fur-
niture is located avoiding the airflow obstruction in the area near the inlet and creating
an unfilled volume of air in the area close to the outlet.

1
Escuela Politécnica Superior, Universidad CEU San Pablo, Madrid, Spain
2
G.I.R. Arquitectura & Energı́a, E.T.S. Arquitectura, Universidad de Valladolid, Valladolid, Spain

Corresponding author:
Susana Hormigos-Jimenez, Escuela Politécnica Superior, Universidad CEU San Pablo, Monteprı́ncipe Campus,
Boadilla del Monte, 28668 Madrid, Spain.
Email: sus.hormigos@ceindo.ceu.es
2 Journal of Building Physics 00(0)

Keyword
Age of air, ventilation efficiency, breathing zone, computational fluid dynamics, indoor air
quality

Practical application
The ventilation efficiency analysis in the breathing zone (BZ) during sleep in a
bedroom is essential to assess the air quality to which a person is exposed during
that time. ANSYS Fluent 17 (computational fluid dynamics (CFD) code) was
used to develop the numerical study, which was experimentally validated by the
tracer gas (SF6) concentration decay method. 52 different arrangements of furni-
ture were considered to assess the influence of these elements on the ventilation
efficiency in BZ. Improvements of up to 6.5% in the ventilation efficiency can be
achieved. Adding furniture to an enclosure does not result in improved ventila-
tion efficiency.

Introduction
Today, in developed countries, 80%–90% of the time is spent indoors (Klepeis
et al., 2001), especially in households, since presence at home varies between 60%
and 90% of the day and 30% of the time is spent sleeping (Wargocki, 2016; Wouter
Borsboom et al., 2016). For this reason, homes contain the air that is inhaled most
of the time, constituting the indoor spaces of greater exposure. Indoor environment
in housing should benefit rest and recovery (Wargocki, 2016); therefore, poor indoor
air quality (IAQ) prevents this purpose since it has harmful effects on health.
Since the Energy Crisis of the 1970s, buildings have become increasingly airtight,
leading to the onset of diseases related to IAQ such as sick building syndrome
(SBS) (Cao et al., 2014). In addition, the relationship between air movement in
buildings due to ventilation and the spread of infectious diseases has been demon-
strated (Li et al., 2007). Poor IAQ also results in a loss of productivity, leading to
significant economic losses (Seppänen, 2008). Therefore, from the 21st century,
efforts are being focused on finding a balance between energy efficiency and air dis-
tribution parameters, IAQ and thermal comfort (Awbi, 2003; Li et al., 2012;
Nielsen, 2000; Seppänen, 1999, 2008; Valkeapaa and Siren, 2010). In this context,
the benefits of indoor air exchange have been shown, although the quantitative
influence of ventilation on the spread of infectious diseases is not clear (Cao et al.,
2014). Consequently, since there is no quantitative influence of the airflow rates on
health, the ventilation rates specified in the regulations are usually set according to
comfort criteria (perceived conditions) (Wargocki, 2015).
However, a healthy indoor environment can be achieved by applying the neces-
sary IAQ improvement strategies which are, in addition to increasing fresh air sup-
ply, controlling the pollution of the emission sources, air cleaning and the
improvement of the ventilation efficiency (Liddament, 2000); the latter is within
the scope of this study.
Hormigos-Jimenez et al. 3

The ventilation efficiency shows the distribution of fresh air within a space, thus
showing a qualitative measure of the performance of the ventilation system; addi-
tionally, it can be used as an IAQ indicator assuming that the air supply has good
quality. The assessment of the ventilation efficiency should be determined by the
objective of the ventilation system such as removal of heat, removal of contamina-
tion, protection from cross-infection, or supply of fresh air to the breathing zone
(BZ) (Cao et al., 2014).
According to ISO 15202-1 (2012) and Cao et al. (2014), the BZ consists of the
volume of air comprised within a hemisphere of 0.3 m radius, which extends in
front of a person’s face. This hemisphere is centered in the mid-point of the imagin-
ary line that connects the ears, and its base is placed in the plane formed by that
line, the larynx, and the top of the head. According to the flow pattern and there-
fore, according to the distribution of the age of the air inside a space, a better or
worse IAQ will be perceived in BZ. It is necessary to assess the air quality in BZ,
mainly in bedrooms, due to the amount of time invested in its interior, which leads
to a high exposure.
Ventilation efficiency depends on the flow pattern (Chung and Hsu, 2001;
Meiss, 2009), since it shows the extent to which an occupied area is adequately ven-
tilated compared to the piston flow model (optimum theoretical model, with an
efficiency of 100%). In order to characterize the flow patterns, the previous assess-
ment of the age of the air should be developed. Specifically, the room mean age of
the air (RMA) and the local mean age of the air (LMA), which are parameters
directly related to the path followed by the air, should be obtained considering that
the age of the air at the inlet is equal to zero (100% fresh). The ventilation effi-
ciency is determined by comparing the existing flow model in the case under study
with the piston model (Meiss, 2009). Therefore, the efficiency is calculated by
means of the relationship between LMA at the outlet and twice the RMA, since it
is an arithmetic mean, according to equation (1)
te
ea = 2t  100 ð1Þ

where ea is ventilation efficiency (%), te is LMA at the outlet (s), and t is RMA (s).
LMA and RMA are calculated by experimental analysis, using the tracer gas
methods; and by numerical analysis, using computational fluid dynamics (CFD)
techniques, validated by the experimental study. The experimental analysis consists
on injecting a tracer gas into a space, to mark the indoor air, and subsequently
initiating a ventilation process with fresh (unmarked) air, while performing the
continuous recording of the gas concentration in order to differentiate the air pres-
ent in the room from the fresh air entering the space. The numerical method con-
sists of solving partial differential equations, defined by the user, to describe the
transport of the ‘‘air age’’ scalar, thus it consists of assessing the ventilation effi-
ciency using the steady-state method (Chang et al., 2003; Meiss, 2009; Noh et al.,
2008; Roos, 1998).
4 Journal of Building Physics 00(0)

The objective of this study was to assess the ventilation efficiency in the breath-
ing area of a person in a bedroom during sleep, considering the influence of the
location of the bed, and of the rest of the furniture, on the age of the air. For this,
two different types of bedrooms were considered, a single one and a double one,
both of them with a volume of 30 m3, in isothermal conditions, studying 52 differ-
ent arrangements of furniture. A numerical study of the age of the air, validated by
the tracer gas (SF6) concentration decay method, was developed using the steady-
state method and through the user-defined function (UDF) of the age of the air.

Methodology
Implementation and validation of the numerical model
ANSYS Fluent 17 (CFD code) was used to carry out the numerical study consist-
ing of the resolution of the transport equation of a user-defined scalar (Fi). When
considering steady-state conditions (steady-state numerical method), convection-
diffusion equation (2) is used to solve Fi (Chanteloup and Mirade, 2009; Gan and
Awbi, 1994)
SFi = r  ðr~
vFi  Gi rFi Þ ð2Þ

where r is the density of the fluid (kg m3); y is the velocity of the fluid (m s21); Fi is
the scalar or the age of the air (t i); GFi is the diffusion coefficient of the scalar,
which depends on the effective viscosity of the air (meff) according to equation (3);
and SFi is the scalar source term which is normally considered equal to 1 (Bartak
et al., 2001; Gan, 2000; Hu and Chuah, 2003)
meff
Gti = 2:883105 r + 0:7 ð3Þ

The numerical method was included in the Fluent CFD code through a prede-
fined function in programming language C, known as UDF. The accuracy of this
programming was validated through the tracer gas concentration decay method
(SF6). The experimental tests were developed in the test room of the Ventilation
Laboratory of the Valladolid School of Architecture (Camino Olea et al., 2005). To
perform the concentration decay tests, a bedroom is simulated in the chamber, with
a floor area of 3.00 m 3 4.00 m base and 2.50 m height. The geometry of this space
is also used to perform the numerical simulations. There is an inlet of 0.40 m 3 0.01
m (40 cm2) and an outlet of 0.70 m 3 0.01143 m (80 cm2), as shown in Figure 1.
According to previous studies (Bartak et al., 2001; Chen and Srebric, 2002;
Davidson and Olsson, 1987), a variation between 20% and 30% is accepted
between the values of LMA obtained experimentally, by tracer gas techniques,
compared to those exported from numerical analysis. According to Bartak et al.
(2001), these differences between results are mainly due to the procedure developed
during the experiments, such as inconsistency when determining the volume of the
room, experimental accuracy, and repeatability.
Hormigos-Jimenez et al. 5

Figure 1. Test chamber dimensions (in meters) and inlet and outlet arrangements.

Two case studies are chosen to perform the validation, one of them with the least
number of furniture elements (single bed) and the other with all the elements (com-
plete equipment), these models are case 1 and case 2, respectively. The furniture
introduced inside the test chamber was built with real-scale cardboard, the arrange-
ment of the different elements, according to case 1 and case 2, is shown in Figure 2.
Both the choice of the number of measurement points and the distribution of
them depend on the purpose of the measurement, since the sampling must be per-
formed in representative places of the enclosure under study (ISO 16000-8, 2007).
This research aims to obtain information on the local ventilation conditions, so
that samples are taken in nine locations distributed evenly throughout the cham-
ber. For each test, one position is chosen (1–9), where measurements are taken at a
height of 0.75 and 2.00 m and in the exhaust; therefore, a total of 19 points are
sampled (Figure 3).
For each of the nine established locations, samples are taken three times with the
purpose of performing the calculation of the LMA values obtained in the experi-
mental study; subsequently, the average of the three measurements is compared
with the corresponding results found in the numerical analysis. Thus, 54 tests are
performed, 27 for each of the two case studies.
The tests were performed during the months of January and early February
2017 for the case 1, and during the second half of February and the month of
March 2017 for the case 2, at the School of Architecture of Valladolid. To reach
three measurements for each sampling point, in order to establish an average
between the values, a first round (A) of data collection of the nine fixed points is
completed, then a second (B), and finally, a third (C). The days on which the
tests were performed were subject to the wind velocity, so that only the tests were
performed as long as the weather forecast indicated a wind velocity of less than
6 Journal of Building Physics 00(0)

Figure 2. Case studies considered for the validation: (a) case 1 and (b) case 2.

Figure 3. Spatial distribution (m) of the nine points sampled at a height of 0.75 and 2.00 m.

10 km h21; in this way, a significant influence of the outdoor conditions during the
sampling process is avoided.
Hormigos-Jimenez et al. 7

The equipment used for the dosing and analysis of tracer gas consists of the fol-
lowing elements: a gas doser and analyzer, tubes of non-absorbent material (nylon)
by means of which the suction and injection of the gas is performed, SF6 bottle,
mixing fans that promote the mixture of the tracer gas with the air contained in
the enclosure, and anemometer for the measurement of the exhaust velocity and
extractor fan controlled by a frequency shifter.
The test procedure is performed according to the following successive steps:

 The test chamber is sealed and the mixing fans are left in operation.
 The analyzer pump is turned on and a sample of ambient air is taken to
record the initial concentration of SF6.
 The tracer gas is released into the room and the increase in the concentration
of SF6 is recorded continuously.
 The air is mixed with fans until a uniform concentration set to 125 mg m23 is
reached. These fans are switched off when the injection is stopped.
 The air exhaust is performed and the tracer gas concentration decay is
recorded continuously.

Based on the concentration readings given as a function of time, LMA is calcu-


lated according to the initial concentration (the concentration at the beginning of
the concentration decay) and the area under the concentration decay curve. For the
calculation of LMA, equation (4) is used
Б
Cdt
t0
t = Ct = t 0 ð4Þ

where t is the duration of the test (h), C is the concentration decay (m3 m23), and
Ct = t0 is the initial concentration of tracer gas (m3 m23) at the beginning of the con-
centration decay
During the process of validation of the numerical analysis, the meteorological
data are collected by means of a local meteorological station nearby, and in no case
the wind velocity reaches a speed of 13 km h21; additionally, this velocity occurred
at specific moments, so that it does not significantly alter the measurements, with
the velocity being less than 8 km h21 during most of the measurement time in all
tests. The indoor temperature was close to 20° C for all tests, with a relative humid-
ity of 40%, approximately.

Case studies
In total, 52 case studies were assessed considering the same space of 30 m3, but
varying the type of bedroom (single or double), as well as the number and arrange-
ment of the furniture elements. The assessment procedure of the ventilation effi-
ciency started by including the bed (configuration A) in the individual bedroom
8 Journal of Building Physics 00(0)

Figure 4. Geometric definition of the pieces of furniture and their configurations.

and considering different arrangements for it (1–8) in the room. Subsequently, a


bedside table (configuration B) was added and placed according to the previous
locations considered for the bed (1–8), in this way, a successive process was fol-
lowed with the cabinet (C), the shelving (D), the desk (E) and the chair (F). Finally,
a double bedroom was assessed in the same way, initially considering the bed (AA
configuration) and varying its layout in the room (1–2) and then incorporating two
bedside tables on each side of the bed (BB configuration) and a cabinet (CC
configuration).
Figure 4 shows the geometrical characteristics of the furniture and the configura-
tions established. Figure 5 shows each of the case studies considered, according to
the variations in the arrangement of the furniture. The items included in this study
had standard dimensions and simple geometries, thus avoiding gaps since it was
considered that the closet doors and the table drawers would be closed and that the
shelves holes would be filled; therefore, solid prisms were studied.

Ventilation efficiency in BZ
This study focuses on assessing the ventilation efficiency in BZ during sleep hours.
For this purpose, and taking into account that a person adopts different positions
while sleeping, the volume of air considered as representative of BZ is located in the
area of the pillow, on the bed (0.50 m high). BZ had a surface of 1.2 m 3 0.6 m and
0.3 m height when the bed was located in a corner, and another surface of 1.5 m
3 0.6 m and 0.3 m height if the bed was centered against a wall. Figure 6 shows the
locations of the established BZ according to the seven possible arrangements that
have been set for the bed in a single bedroom and according to the other two possi-
ble ones for the double bedroom.
To calculate the ventilation efficiency in BZ, equation (1) will take the form of
equation (5), where RMA is replaced by the mean age of the air in BZ (BZMA), as
follows
Hormigos-Jimenez et al. 9

Figure 5. Configurations and arrangements of the pieces of furniture.

Figure 6. Breathing zone, in floor plan, according to different arrangements of the bed.

te
ea = 2t BZ  100 ð5Þ

where ea is the ventilation efficiency (%), te is LMA at the outlet (s), and tBZ is
BZMA (s).
10 Journal of Building Physics 00(0)

Figure 7. Case F8 mesh.

Preprocessing
ANSYS 17 software was used for the construction and meshing of the geometry
through ANSYS DesignModeler and ANSYS Meshing, respectively. All meshes were
hexahedral (Figure 7), since it was considered the most appropriate according to the
simple geometries used. For the realization of the mesh, the criterion of saving compu-
tational cost was followed so that, in any case, 500,000 elements were not exceeded
(Figure 8), giving a higher mesh density to the singular elements of each case. In addi-
tion, in all cases the meshing was optimal, according to the criterion of Orthogonal
Quality (OQ) (ANSYS, 2013), since the average was above 0.95, as shown in Figure 9.
Reynolds-averaged Navier–Stokes (RANS), specifically the two-equation mod-
els (k-e), were used to develop the numerical analysis, since they provide high accu-
racy in the results (ANSYS, 2013; Chen, 1996). In this investigation, a simulation
process was followed, in which the calculation was developed by successive
approaches, through three turbulence models: standard k-e, Re-Normalisation
Group (RNG) k-e, and Realizable (RKE) k-e, setting 10,000 iterations for each one
(Meiss, 2009; Wilcox, 1998).
In this research, isothermal conditions were assumed; the air density was estab-
lished at 1.225 kg/m3 and its dynamic viscosity at 1.7894 3 1025 kg/(m s).
Regarding the boundary conditions, a value of 5% was chosen for the turbulence
intensity (default value) and a scale length of 0.01 (normalized) was set. However,
the exhaust air velocity differed for a single bedroom and a double bedroom since,
according to the Spanish regulation (RD314/2006) (Código Técnico de la
Edificación, 2009), a single bedroom, with a single occupant, requires a supply
Hormigos-Jimenez et al. 11

Figure 8. Number of elements and nodes of each of the 52 case studies.

Figure 9. Mesh quality of each of the 52 case studies.

airflow of 5 L/s; while a double bedroom, with two occupants, requires 10 L/s.
Therefore, in the case studies corresponding to a single bedroom (configurations
A–F), a constant velocity of exhaust air at the outlet (door) of 0.625 m/s was set.
Likewise, this value for the velocity was 1.25 m/s in the double bedroom models
(configurations AA–CC).
In short, and according to the established methodology, this research intends to
perform an assessment of the ventilation efficiency, in the BZ of a bedroom, during
the hours of sleep. For this purpose, a CFD numerical analysis, validated through
the tracer gas concentration decay method, is used in order to assess the distribu-
tion of the age of the air and to find both LMA at the outlet and BZMA values,
which are essential for the calculation of the ventilation efficiency.

Results
Experimental study
Tables 1 and 2 show the LMA values obtained in the tests, for each of the test
points (1–9) and each of the three sample-takings (A–C), of the corresponding
sample at a height of 2.00 m (M1), 0.75 m (M2), and exhaust (M3). The results of
12 Journal of Building Physics 00(0)

Table 1. LMA results for case 1 obtained through both the numerical and the experimental
analysis.

Code Age of air (s): Age of air (s): Difference between values
Experimental analysis Numerical analysis
M1 M2 M3 N1 N2 N3 N1/M1 N2/M2 N3/M3

1-01A 4779.14 4497.03 4780.40 6076.12 5984.57 5840.58 1.27 1.31 1.22
1-01B 4818.49 4776.67 5162.46 1.26 1.25 1.13
1-01 C 4688.36 4717.73 4672.96 1.30 1.27 1.25
1-02A 5371.18 5478.34 5453.81 6031.66 5939.63 1.12 1.08 1.07
1-02B 4500.34 4590.40 4907.09 1.34 1.29 1.19
1-02 C 5051.09 4921.25 5097.67 1.19 1.21 1.15
1-03A 4799.84 4526.98 4677.63 5973.16 5826.34 1.24 1.29 1.25
1-03B 5014.54 4986.54 5037.86 1.19 1.17 1.16
1-03 C 4957.37 4668.41 4838.16 1.20 1.25 1.21
1-04A 4897.44 5038.45 5142.91 5972.57 6096.90 1.22 1.21 1.14
1-04B 4974.73 4594.91 4756.03 1.20 1.31 1.23
1-04 C 4903.55 4888.29 5008.33 1.22 1.25 1.17
1-05A 4901.61 5156.67 4987.72 5827.69 5820.35 1.19 1.13 1.17
1-05B 4859.90 5278.12 5204.18 1.20 1.10 1.12
1-05 C 5012.82 4767.03 4712.89 1.16 1.22 1.24
1-06A 5053.40 5100.50 5267.34 5885.01 5907.54 1.16 1.16 1.11
1-06B 4833.54 4571.09 4848.83 1.22 1.29 1.20
1-06 C 4841.31 4833.77 4922.29 1.22 1.22 1.19
1-07A 5205.17 5537.23 5074.18 5870.54 6073.29 1.13 1.10 1.15
1-07B 5004.37 5021.10 4984.31 1.17 1.21 1.17
1-07 C 4870.32 4791.75 4789.21 1.21 1.27 1.22
1-08A 5183.36 5116.07 5367.21 6057.27 5812.77 1.17 1.14 1.09
1-08B 4778.54 4585.18 4713.49 1.27 1.27 1.24
1-08 C 4756.24 4889.27 4998.81 1.27 1.19 1.17
1-09A 4503.62 4559.14 4869.47 6021.63 5889.77 1.34 1.29 1.20
1-09B 4759.98 5027.05 4981.25 1.27 1.17 1.17
1-09 C 4827.88 4713.97 4833.58 1.25 1.25 1.21

LMA: local mean age of the air.

the numerical study are also shown, so that the corresponding value is presented
for each test location for a height of 2.00 m (N1), 0.75 m (N2), and exhaust (N3).
In addition, there is a single LMA value in the exhaust for each of the case studies.
As shown in these Tables 1 and 2, the greatest difference found in 34%; however,
in the majority of the cases, 30% is not exceeded, being close to 20%, in several
cases less than this percentage. Therefore, the results obtained are within the accep-
tance range.
Tables 3 and 4 show the data obtained in the experimental study once applied
the correction factor (Z) obtained from the relationship between the LMA values
in the exhaust of both the numerical study and the experimental study (N3/M3). In
addition, these values are compared with the result obtained from the numerical
Hormigos-Jimenez et al. 13

Table 2. LMA results for case 2 obtained through both the numerical and the experimental
analysis.

Code Age of air (s): Age of air (s): Difference between values
Experimental analysis Numerical analysis
M1 M2 M3 N1 M1 M2 M3 N2/M2 M1

2-01A 4763.94 4512.05 4728.97 5864.48 5764.99 5474.17 1.23 1.28 1.16
2-01B 4803.26 4844.49 4749.03 1.22 1.19 1.15
2-01 C 4629.82 4716.34 4658.68 1.27 1.22 1.18
2-02A 4574.29 4975.80 4765.59 5930.84 5702.33 1.30 1.15 1.15
2-02B 4460.13 4517.63 4617.40 1.33 1.26 1.19
2-02 C 4727.31 4536.99 4583.55 1.25 1.26 1.19
2-03A 4549.95 4644.03 4740.53 5954.52 5820.63 1.31 1.25 1.15
2-03B 4655.53 4587.22 4699.42 1.28 1.27 1.16
2-03 C 4745.17 4599.47 4802.37 1.25 1.27 1.14
2-04A 4642.67 4524.88 4702.48 5497.48 5630.81 1.18 1.24 1.16
2-04B 4213.67 4247.53 4531.07 1.30 1.31 1.21
2-04 C 4506.17 4597.01 4799.07 1.22 1.22 1.14
2-05A 4639.29 4619.84 4857.50 5330.28 5588.01 1.15 1.21 1.13
2-05B 4126.81 4544.74 4636.99 1.29 1.23 1.18
2-05 C 4644.71 4587.32 4849.01 1.15 1.22 1.13
2-06A 4598.58 4796.24 4892.56 5693.93 5891.19 1.24 1.23 1.12
2-06B 4624.75 4557.39 4630.55 1.23 1.29 1.18
2-06 C 4882.47 4595.28 4825.35 1.17 1.28 1.13
2-07A 4710.27 4713.13 4766.84 5727.55 5661.19 1.22 1.20 1.15
2-07B 4553.64 4345.93 4633.69 1.26 1.30 1.18
2-07 C 4320.91 4579.67 4639.24 1.33 1.24 1.18
2-08A 4579.61 4516.02 4679.03 5942.93 5656.96 1.30 1.25 1.17
2-08B 4693.44 4535.33 4667.47 1.27 1.25 1.17
2-08 C 4643.09 4703.74 4814.04 1.28 1.20 1.14
2-09A 4410.04 4436.15 4597.63 5786.82 5677.39 1.31 1.28 1.19
2-09B 4527.77 4507.36 4694.07 1.28 1.26 1.17
2-09 C 4658.12 4691.18 4821.10 1.24 1.21 1.14

LMA: local mean age of the air.

analysis for the locations at 2.00 m in height (N1) and at 0.75 m (N2) calculating
the difference between the values of both analyzes. In this study, the aim is to
achieve an average difference of less than 10% in all cases in order to consider the
model as validated.
A difference of less than 10% was found between the numerical and the experi-
mental analysis (Tables 3 and 4). Thus, this was an appropriate correlation, which
also demonstrated that the validation process and the tracer gas technique
employed were adequate. Most of the results obtained from the CFD analysis,
except for point 7 (at 2.00 m) and f (at 0.75 m) are higher than those found in the
laboratory tests, a fact that also occurs in other studies (Bartak et al., 2001; Chen
and Srebric, 2002; Davidson and Olsson, 1987).
Table 3. Difference in the LMA results for case 1 obtained both the experimental study and the numerical study.
14
Pa Zb Sampling points at 2.00 m (M1) Average Sampling points at 0.75 m (M2) Average
difference difference
M1 Z*M1 N1 Difference M2 Z*M2 N2 Difference

1A 1.22 4779.14 5839.04 3.90 4497.03 5494.37 8.19


1B 1.13 4818.49 5451.42 6076.12 10.28 5.91 4776.67 5404.11 5598.34 9.70 6.45
1C 1.25 4688.36 5859.83 3.56 4717.73 5896.54 1.47
2A 1.07 5371.18 5752.09 4.64 5478.34 5866.85 1.23
2B 1.19 4500.34 5356.45 6031.66 11.19 6.63 4590.40 5463.65 5656.32 8.01 4.77
2C 1.15 5051.09 5787.21 4.05 4921.25 5638.45 5.07
3A 1.25 4799.84 5993.18 20.34 4526.98 5652.48 2.98
3B 1.16 5014.54 5813.54 5973.16 2.67 0.72 4986.54 5781.09 5689.74 0.78 2.34
3C 1.21 4957.37 5984.49 20.19 4668.41 5635.65 3.27
4A 1.14 4897.44 5561.81 6.88 5038.45 5721.96 6.15
4B 1.23 4974.73 6109.15 5972.57 22.29 2.95 4594.91 5642.71 5688.42 7.45 6.70
4C 1.17 4903.55 5718.39 4.26 4888.29 5700.60 6.50
5A 1.17 4901.61 5739.75 1.51 5156.67 6038.42 23.75
5B 1.12 4859.90 5454.20 5827.69 6.41 0.44 5278.12 5923.56 5956.55 21.77 22.34
5C 1.24 5012.82 6212.28 26.60 4767.03 5907.68 21.50
6A 1.11 5053.40 5603.35 4.79 5100.50 5655.58 4.27
6B 1.20 4833.54 5822.16 5885.01 1.07 2.75 4571.09 5506.03 5632.39 6.80 4.66
6C 1.19 4841.31 5744.49 2.39 4833.77 5735.55 2.91
7A 1.15 5205.17 5991.36 22.06 5537.23 6373.57 24.94
7B 1.17 5004.37 5864.09 5870.54 0.11 21.04 5021.10 5883.69 6033.64 3.12 0.65
7C 1.22 4870.32 5939.49 21.17 4791.75 5843.67 3.78
8A 1.09 5183.36 5640.51 688 5116.07 5567.28 4.22
8B 1.24 4778.54 5921.18 6057.27 2.25 5.79 4585.18 5681.58 5653.82 2.26 2.73
8C 1.17 4756.24 5557.16 8.26 4889.27 5712.60 1.72
9A 1.20 4503.62 5401.76 10.29 4559.14 5468.36 7.15
9B 1.17 4759.98 5581.15 6021.63 7.32 6.91 5027.05 5894.28 5686.23 20.08 3.46
9C 1.21 4827.88 5833.70 3.12 4713.97 5696.05 3.29

LMA: local mean age of the air.


a
P is the simple location and the first, second, or third simple-taking (A, B, and C, respectively).
b
Z is the correction factor (Z = N3/M3).
Journal of Building Physics 00(0)

The bold values shows the mean values (obtained experimentally) with the greatest difference compared to the ones obtained in the numerical analysis.
Table 4. Difference in the LMA results for case 2 obtained both the experimental study and the numerical study.

Pa Zb Sampling points at 2.00 m (M1) Average Sampling points at 0.75 m (M2) Average
difference difference
M1 Z*M1 N1 Difference M2 Z*M2 N2 Difference

1A 1.16 4763.94 5514.65 5.97 4512.05 5223.07 9.40


1B 1.15 4803.26 5536.69 5864.48 5.59 6.26 4844.49 5584.22 5764.99 3.14 5.47
1C 1.18 4629.82 5440.27 7.23 4716.34 5541.93 3.87
2A 1.15 4574.29 5254.43 11.40 4975.80 5715.64 20.23
Hormigos-Jimenez et al.

2B 1.19 4460.13 5287.73 5930.84 10.84 9.02 4517.63 5355.90 5702.33 6.08 3.61
2C 1.19 4727.31 5645.87 4.80 4536.99 5418.57 4.98
3A 1.15 4549.95 5254.10 11.66 4644.03 5362.74 7.75
3B 1.16 4655.53 5423.05 5947.62 8.82 9.85 4587.22 5343.48 5813.42 8.08 8.55
3C 1.14 4745.17 5408.97 9.06 4599.47 5242.89 9.81
4A 1.16 4642.67 5404.55 1.69 4524.88 5267.43 6.45
4B 1.21 4213.67 5090.71 5497.48 7.40 5.20 4247.53 5131.62 5630.81 8.87 7.40
4C 1.14 4506.17 5140.07 6.50 4597.01 5243.68 6.88
5A 1.13 4639.29 5228.25 1.91 4619.84 5206.34 6.83
5B 1.18 4126.81 4871.88 5330.28 8.60 4.05 4544.74 5365.26 5588.01 3.99 6.05
5C 1.13 4644.71 5243.53 1.63 4587.32 5178.74 7.32
6A 1.12 4598.58 5145.24 9.64 4796.24 5366.40 8.91
6B 1.18 4624.75 5467.32 5693.93 3.98 5.45 4557.39 5387.68 5891.19 8.55 9.65
6C 1.13 4882.47 5538.97 2.72 4595.28 5213.16 11.51
7A 1.15 4710.27 5409.21 5.56 4713.13 5412.50 4.39
7B 1.18 4553.64 5379.60 5727.55 6.07 7.54 4345.93 5134.22 5661.19 9.31 6.08
7C 1.18 4320.91 5098.55 10.98 4579.67 5403.88 4.55
8A 1.17 4579.61 5357.85 9.84 4516.02 5283.47 6.60
8B 1.17 4693.44 5504.64 5942.93 7.37 9.46 4535.33 5319.19 5656.96 5.97 6.01
8C 1.14 4643.09 5279.79 11.16 4703.74 5348.75 5.45
9A 1.19 4410.04 5250.81 9.26 4436.15 5281.90 6.97
9B 1.17 4527.77 5280.23 5786.82 8.75 8.87 4507.36 5256.43 5677.39 7.41 6.85
9C 1.14 4658.12 5289.11 8.60 4691.18 5326.66 6.18

LMA: local mean age of the air.


a
P is the simple location and the first, second, or third simple-taking (A, B, and C, respectively).
b
15

Z is the correction factor (Z = N3/M3).


The bold values shows the mean values (obtained experimentally) with the greatest difference compared to the ones obtained in the numerical analysis.
16 Journal of Building Physics 00(0)

In the previous Tables 3 and 4, the values of greatest average difference between
the numerical and the experimental study are highlighted. For case 1 (Table 3), the
maximum difference between results is 6.91%, which occurs at location 9 at the
height of 2.00 m. There is also a difference greater than 6% in point 1 to 0.75 m, in
point 2 to 2.00 m, and in point 3 to 0.75 m. All other locations have a difference of
less than 6%, being point 5, located in the center of the room, which shows minor
differences with 0.44% (at 2.00 m) and point 3 with a 0.72% (at 2.00 m). Point 7
also has differences of less than 1%, both in the measure taken at 2.00 and 0.75 m.
For the case 2 (Table 4), the maximum difference between results is 9.85%, which
occurs in point 3 at the height of 2.00 m; there is also a difference greater than 9%
in points 2 (at 2.00 m), 6 (at 0.75 m), and 8 (at 2.00 m). Point 3 (at 0.75 m) shows
the smallest difference with 3.61%.

Numerical study
Figure 10(a) and (b) shows the distribution of the age of the air in the 52 case stud-
ies, through a plane that occupies the whole room, although in this study, just the
area corresponding to BZ was assessed. The horizontal plane displayed in these fig-
ures corresponds to a height of 0.8 m, which is the upper limit of the established
BZ, since it was observed that it was the one that shows the highest age of the air in
BZ in comparison with the other horizontal planes forming the ‘‘breathable’’ vol-
ume during the hours of sleep. This was because, due to the Coanda effect, the air
adhered to the surfaces it encountered in its path; thus, in most cases, the age of the
air was lower in the plane near the bed and it was higher as the distance from the
bed increased. Figure 10(a) shows that, in cases B4, B5, and D2, there were areas of
stagnation in BZ, since they did not follow the clear and logical tendency of the
other cases toward a lower age of the air, as the volume of the fluid decreased,
through the addition of furniture.
Figure 11 shows both BZMA and LMA values at the outlet for all the case stud-
ies considered. The relationship between these data, according to equation (4), con-
stitutes the ventilation efficiency values in BZ (Figure 12). As shown in Figures 11
and 12, the closer the ages of the air in BZMA and LMA are, in the same case
study, the ventilation efficiency will be higher, and vice versa; however, if LMA at
the outlet is higher than BZMA, the efficiency will exceed 50% (perfect mixing
model). Thus, the case with the highest efficiency (50.58%) was CC, with values for
BZMA and LMA at the outlet of 2596.68 and 2626.98 s, respectively. There was
another model with ventilation efficiency above 50%, B1, with 50.07%, due to a
BZMA value of 5807.18 s and a LMA at the outlet value of 5815.73 s. The cases
with the lowest efficiency (below 47%) were B4, B5, and D2, the latter had the
worst efficiency, with 46.59%.
Figure 13 shows the relationship between the ventilation efficiency in both BZ
and the whole room. In 24 of the 52 cases considered, the efficiency of the air
exchange in BZ was better than the one corresponding to the total volume of the
bedroom; in 5 of them (A1, B1, C1, E1, and F1), this improvement was above 2%
Hormigos-Jimenez et al. 17

Figure 10. Continued


18 Journal of Building Physics 00(0)

Figure 10. (a) Age-of-the-air distribution for the single bedroom configurations (A–D). (b)
Age-of-the-air distribution for the bedrooms configurations (E, F, and AA–CC).
Hormigos-Jimenez et al. 19

Figure 11. BZMA and LMA at the outlet of each of the 52 case studies.

Figure 12. Ventilation efficiency in the breathing zone of each of the 52 case studies.

Figure 13. Difference (%) in ventilation efficiency between BZ and the whole room.
20 Journal of Building Physics 00(0)

Figure 14. Variation (%) in the ventilation efficiency between arrangements per each
configuration, organized from the best to the worst case.

being, in case B1, higher than 3% (3.29%). Therefore, arrangements 1, except for
D1, had the highest efficiency improvement in BZ with respect to the entire space
of the bedroom. In case C5 was where the most pronounced worsening occurred
(22.83%); additionally, a deterioration in the efficiency in BZ occurred in the cases
of the double bedroom BB2 (21.92%) and AA2 (21.28%).

Discussion
Figure 14 shows the difference in ventilation efficiency in BZ between the different
arrangements within the same configuration. The data are shown following the
order from better to worse efficiency; so that the best cases by each configuration
serve as reference to establish the variation of the efficiency, in percentage, with
respect to the following cases, ordered from less to more variation. According to
this criterion, all configurations, except D, had as a common feature, the layout 1
as the best efficiency case within its category. This was so because, considering the
placement of the furniture in the space according to model 1, there was more unoc-
cupied air volume (free of obstacles) in the area close to the outlet; therefore, the
air exhaust was enhanced.
Figure 14 shows the difference between efficiencies, according to each configura-
tion, with respect to the best cases being A1, B1, C1, D6, E1, F1, AA1, and BB1;
the CC case is not shown as there is no possible comparison within its configura-
tion, since it is the only case. In all configurations except B, the difference in ventila-
tion efficiency between the best and the worst case was approximately 3%–4%. In
cases B4 and B5, the area occupied by the furniture (bed and bedside table) was the
corner shared by the wall where the inlet was located and the adjacent one, where
the outlet was placed (Figure 5). In this zone, and according to the geometric char-
acteristics of the objects including in the room, air stagnation zones were generated
on the bed (Figure 10(a)); this fact caused a greater difference, over 6.5%, in the
Hormigos-Jimenez et al. 21

Figure 15. Arithmetic mean of the ventilation efficiencies per configuration (A–F; AA–CC).

ventilation efficiency values with respect to the rest of the arrangements of the con-
figuration B.
With regard to the possible influence of the addition of furniture on the ventila-
tion efficiency in BZ, it was observed that the fact of adding furniture to an enclo-
sure did not imply an improvement of ventilation efficiency. Figure 15 shows the
arithmetic mean of the ventilation efficiencies in BZ for each category (A–F and
AA–CC). In the individual bedroom, the configuration A (with a dingle bed) was
the most efficient one and the configuration D (with the bed, the bedside table, the
closet and the shelving), the least efficient one.
Furthermore, for the double bedroom configurations (AA, BB, and CC), there
was an improvement in the efficiency in BZ as furniture was added. However, this
fact was not considered significant, since only five cases were assessed due to the
physical space considered, which did not allow the addition of more furniture ele-
ments according to acceptable architectural criteria. In addition, the average values
in the efficiencies of AA and BB were almost identical, of 47.99% and 48.07%,
respectively. The efficiencies were close but below to 50% (perfect mixing flow),
except in the cases of B1 and CC that exceeded this value. Therefore, they followed
the short circuit model (efficiency less than 50%), which is the most common flow
model especially in houses, which indicates the existence of an insufficient level of
ventilation and, therefore, the need to increase the fresh air supply (Meiss and
Feijó, 2011).
This research focuses on assessing the efficiency of ventilation in the BZ during
sleep hours, in a single space and with fixed inlet and outlet, but varying the num-
ber and the arrangement of furniture elements. In this way, the influence of the fur-
niture on the air exchange efficiency in a person’s breathing area during sleep is
analyzed, since this is where the exposure is concentrated and, therefore, has a
higher influence on health. In this particular realm, there are no examples in the
22 Journal of Building Physics 00(0)

literature, since most of the existing studies are mainly focused on the assessment
of different types of spaces (Awwad et al., 2017; Fan and Ito, 2012; Meiss, 2009;
Naboni et al., 2017; Sherman and Walker, 2009) and various ventilation systems
(Cao et al., 2014; CEN 15251, 2007; CEN 1752, 1998; ISO-7730, 2005; Li et al.,
2012; Sherman and Walker, 2009; Tomasi et al., 2013) including the location of
openings (ANSI/ASHRAE Standard-113, 2013; ASHRAE standard 70, 2011; Chu
et al., 2015; Liu et al., 2016; Meiss, 2009; Meiss et al., 2013).

Conclusion
This research consists of the use of the CFD technique, validated by the tracer gas
(SF6) concentration decay experimental method, to assess the influence of furniture
on the efficiency of the air exchange in BZ during sleep hours, since the absence of
movement of the sleeper makes that the air is exchanged mainly by the interior flow
pattern. For this purpose, 52 case studies are analyzed corresponding to the same
space but varying the number and the layout of the furniture in its interior. In total,
47 of the 52 cases belong to a single bedroom and the remaining 5 to a double bed-
room. From the assessment developed, the following conclusions are obtained:

 Is it advisable to place the furniture in the space to that there is more vol-
ume of unoccupied air, free of obstacles, in the area of the outlet; thus, the
exhaust of the air is facilitated. In addition, the airflow obstruction should
be avoided in the area near the inlet, such as placing a bed under the supply,
because areas of stagnation are generated. Zones of stale air also appear if
the area occupied by the furniture is the corner shared by the wall where the
supply is located and the contiguous wall, where the exhaust is placed.
Thus, according to this study, improvements of up to approximately 6.5%
in the ventilation efficiency can be achieved.
 Whenever the area close to the inlet is not obstructed, the wider the volume
of the fluid (area free of obstacles) close to the outlet is, the higher the
improvement in the ventilation efficiency in BZ will be compared to the ven-
tilation efficiency of the entire enclosure. Improvements of up to 3.3% can
be achieved, according to the results obtained in this research.
 If the space close to the inlet is obstructed, such as by placing a bed under
the window (air supply), there is generally a worsening of the ventilation
efficiency in BZ compared to the ventilation efficiency of the entire enclo-
sure. In this study, a reduction of efficiency of up to approximately 2.8% is
observed.
 If low-height furniture elements are placed in the center of the room, such as
a bed, and others of higher height are located in the wall opposite to the inlet
but contiguous to the outlet, such as a closet, there are areas of stagnation in
BZ. This occurs because the movement of the air is hampered by the addi-
tion of elements in the wall opposite to the supply; therefore, air in BZ takes
Hormigos-Jimenez et al. 23

longer to be exchanged, since the age of the air will be lower in higher areas
of the room. In these cases, there is a worsening of the ventilation efficiency
in BZ of up to 1.7% with respect to the efficiency of the ventilation consider-
ing the complete enclosure.
 Adding furniture to an enclosure does not result in improved ventilation effi-
ciency in BZ. In this research, the configuration A, which includes a single
bed, has the highest efficiency in BZ on average, while the worst one corre-
sponds to the configuration D, which contains a bed, bedside table, closet,
and shelving. However, in the double bedroom, there is an improvement in
the ventilation efficiency as furniture elements are introduced; though, this
fact is not considered significant, since only five cases have been evaluated
due to the geometrical characteristics of the space considered.

In this study, a model with isothermal conditions has been considered, in which
the temperature of the fluid (the air) and the walls are in thermal equilibrium. The
temperature influences the airflow due to the effects of convection; therefore, the
distribution of the age of air will be modified. For future research lines, it is pro-
posed to evaluate winter and summer conditions, together with the influence of dif-
ferent air conditioning systems on the ventilation efficiency in the BZ.
In addition, in order to continue with the research developed, it is intended to
consider the convective effects derived from the presence of people and their breath-
ing process when assessing the airflow patterns inside an enclosure. In this sense, it
is of special interest to evaluate the airflow in the BZ of a person, according to the
activity that is being developed, which in turn will be directly related to the type of
space that is occupied.

Acknowledgements
The authors thank Arie Group from E.P.S. Universidad CEU San Pablo; G.I.R.
Arquitectura & Energı́a from E.T.S. Arquitectura, Universidad de Valladolid; and the
International Center for Indoor Environment and Energy (ICIEE) from DTU for the gui-
dance provided.

Declaration of conflicting interests


The author(s) declared no potential conflicts of interest with respect to the research, author-
ship, and/or publication of this article.

Funding
The author(s) disclosed receipt of the following financial support for the research, author-
ship, and/or publication of this article: The authors thank San Pablo CEU University
Foundation for the pre-doctoral scholarship granted to co-author Susana Hormigos-Jimenez
in its FPI Program. Finally, authors thank both CEINDO CEU and Banco Santander for
24 Journal of Building Physics 00(0)

the international mobility scholarship granted to co-author Susana Hormigos-Jimenez to


develop an external stay at ICIEE.

References
ANSI/ASHRAE Standard 113 (2013) Method of Testing for Room Air Diffusion. Atlanta,
GA: American Society of Heating, Refrigerating and Air-Conditioning Engineers.
ANSYS (2013) ANSYS Fluent User’s Guide Release 15.0. Canonsburg, PA: ANSYS.
ASHRAE Standard 70-2006 (2011) Method of Testing the Performance of Air Outlets and
Air Inlets. Atlanta, GA: American Society of Heating, Refrigerating and Air-
Conditioning Engineers.
Awbi HB (2003) Ventilation of Buildings. Abingdon: Taylor & Francis.
Awwad A, Mohamed MH and Fatouh M (2017) Optimal design of a louver face ceiling
diffuser using CFD to improve occupant’s thermal comfort. Journal of Building
Engineering 11: 134–157.
Bartak M, Cermak M, Clarke J, et al. (2001) Experimental and numerical study of local
mean age of air. In: Proceedings of the 7th international building performance simulation
association (IBPSA) conference, Rio De Janeiro, Brazil, 13–15 August.
Borsboom W, De Gids W, Logue J, et al. (2016) Technical note AIVC 68 residential
ventilation and health. Brussels: INIVE.
Camino Olea MS, Feijo Muñoz J, Basterra Otero A, et al. (2005) Diseño y construccion de
un laboratorio para el analisis de la ventilación en edificios de viviendas, realizado en la
Universidad de Valladolid. In: Proceedings of the jornadas de investigacion en, Instituto
de Ciencias de la Construcción Eduardo Torroja CSIC, Madrid, Spain, 2–4 June, 2005.
Serrano: CSIC.
Cao G, Awbi H, Yao R, et al. (2014) A review of the performance of different ventilation
and airflow distribution systems in buildings. Building and Environment 73: 171–186.
CEN 15251 (2007) Indoor Environmental Input Parameters for Design and Assessment of
Energy Performance of Buildings Addressing Indoor Air Quality, Thermal Environment,
Lighting and Acoustics. Brussels: CEN.
CEN 1752 (1998) Ventilation for Buildings: Design Criteria for the Indoor Environment.
Brussels: CEN.
Chang H, Kato S and Chikamoto T (2003) Room air distribution and indoor air quality of
hybrid air conditioning system based on natural and mechanical ventilation in an office.
International Journal of Ventilation 2: 65–75.
Chanteloup V and Mirade P (2009) Computational fluid dynamics (CFD) modelling of local
mean age of air distribution in forced-ventilation food plants. Journal of Food Engineering
90: 90–103.
Chen Q (1996) Prediction of room air motion by Reynolds-stress models. Building and
Environment 31: 233–244.
Chen Q and Srebric J (2002) A procedure for verification, validation, and reporting of
indoor environment CFD analyses. HVAC&R Research 8: 201–216.
Chu C, Chiu Y, Tsai Y, et al. (2015) Wind-driven natural ventilation for buildings with two
openings on the same external wall. Energy and Buildings 108: 365–372.
Chung K and Hsu S (2001) Effect of ventilation pattern on room air and contaminant
distribution. Building and Environment 36: 989–998.
Hormigos-Jimenez et al. 25

Código Técnico de la Edificación (CTE) (2009) Documento Básico HS3: Calidad del Aire
Interior. Madrid: Ministerio de Fomento del Gobierno de España.
Davidson L and Olsson E (1987) Calculation of age and local purging flow-rate in rooms.
Building and Environment 22: 111–127.
Fan Y and Ito K (2012) Energy consumption analysis intended for real office space with
energy recovery ventilator by integrating BES and CFD approaches. Building and
Environment 52: 57–67.
Gan G and Awbi H (1994) Numerical prediction of the age of air in ventilated rooms. In:
Proceedings of the 4th international conference on air distribution in rooms. Krakow, 15–
17 June.
Gan GH (2000) Effective depth of fresh air distribution in rooms with single-sided natural
ventilation. Energy and Buildings 31: 65–73.
Hu SC and Chuah YK (2003) Deterministic simulation and assessment of air-recirculation
performance of unidirectional-flow cleanrooms that incorporate age of air concept.
Building and Environment 38: 563–570.
ISO 15202-1:2012 (2012). Workplace air–determination of metals and metalloids in airborne
particulate matter by inductively coupled plasma atomic emission spectrometry–part 1:
sampling.
ISO 16000-8:2007 (2007). Indoor air part 8: determination of local mean ages of air in
buildings for characterizing ventilation conditions.
ISO 7730:2005 (2005) Ergonomics of the thermal environment e analytical determination
and interpretation of thermal comfort using calculation of the PMV and PPD indices and
local thermal comfort criteria.
Klepeis NE, Nelson WC, Ott WR, et al. (2001) The national human activity pattern survey
(NHAPS): a resource for assessing exposure to environmental pollutants. Journal of
Exposure Science and Environmental Epidemiology 11: 231–252.
Li X, Cai H, Li R, et al. (2012) A theoretical model to calculate the distribution of air age in
general ventilation system. Building Services Engineering Research and Technology 33:
159–180.
Li Y, Leung GM, Tang JW, et al. (2007) Role of ventilation in airborne transmission of
infectious agents in the built environment—a multidisciplinary systematic review. Indoor
Air 17: 2–18.
Liddament M (2000) A review of ventilation and the quality of ventilation air. Indoor Air 10:
193–199.
Liu W, Jin M, Chen C, et al. (2016) Optimization of air supply location, size, and parameters
in enclosed environments using a computational fluid dynamics-based adjoint method.
Journal of Building Performance Simulation 9: 149–161.
Meiss A (2009) Estudio de la eficiencia de la ventilación en viviendas a partir de parámetros de
diseño arquitectónico. Valladolid (Spain): Universidad de Valladolid.
Meiss A and Feijó J (2011) Influencia de la ubicación de las aberturas en la eficiencia de la
ventilación en viviendas. Informes de la Construcción 63: 53–60.
Meiss A, Feijo-Munoz J and Garcia-Fuentes MA (2013) Age-of-the-air in rooms according
to the environmental condition of temperature: a case study. Energy and Buildings 67:
88–96.
Naboni E, Lee DS and Fabbri K (2017) Thermal Comfort-CFD maps for Architectural
Interior Design. Procedia Engineering 180: 110–117. 10.1016/j.proeng.2017.04.170
26 Journal of Building Physics 00(0)

Nielsen P (2000) Velocity distribution in a room ventilated by displacement ventilation and


wall-mounted air terminal devices. Energy and Buildings 31: 179–187.
Noh K, Han C and Oh M (2008) Effect of the airflow rate of a ceiling type air-conditioner
on ventilation effectiveness in a lecture room. International Journal of Refrigeration 31:
180–188.
Roos A (1998) The air exchange efficiency of the desk displacement ventilation concept
theory, measurements and simulations. In: Proceedings of the 6th international conference
on air distribution in rooms, Stockholm, 14–17 June.
Seppänen O (1999) Estimated cost of indoor climate in finish buildings. Proceedings of Indoor
Air 3: 13–18.
Seppänen O (2008) Ventilation strategies for good indoor air quality and energy efficiency.
International Journal of Ventilation 6(4): 297–306.
Sherman MH and Walker IS (2009) Measured air distribution effectiveness for residential
mechanical ventilation. HVAC&R Research 15: 211–229.
Tomasi R, Krajcik M, Simone A, et al. (2013) Experimental evaluation of air distribution in
mechanically ventilated residential rooms: thermal comfort and ventilation effectiveness.
Energy and Buildings 60: 28–37.
Universidad de Valladolid (n.d.) Web page of the ventilation laboratory HS3 and in the
school of architecture. Available at: http://www.ventilacion.uva.es/ (accessed 25
September 2017).
Valkeapaa A and Siren K (2010) The influence of air circulation, jet discharge momentum
flux and nozzle design parameters on the tightness of an upwards blowing air curtain.
International Journal of Ventilation 8: 337–346.
Wargocki P (2015) What are indoor air quality priorities for energy-efficient buildings?
Indoor and Built Environment 24: 579–582.
Wargocki P (2016) Ventilation, indoor air quality, health, and productivity. In: Wargocki P
(ed.) Ergonomic Workplace Design for Health, Wellness, and Productivity. Boca Raton,
FL: CRC Press, 39–72.
Wilcox DC (1998) Turbulence Modeling for CFD. La Canada Flintridge, CA: DCW
industries.

You might also like