You are on page 1of 11

Energy and Buildings 147 (2017) 155165

Contents lists available at ScienceDirect

Energy and Buildings


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

Effect of furniture and partitions on the surface temperatures in an


ofce under forced ow
F.A. Dominguez Espinosa , Leon R. Glicksman
Department of Mechanical Engineering, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139, USA

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

Article history: Computational Fluid Dynamics (CFD) simulations were performed to study the effects of surfaces typi-
Received 4 November 2016 cally found in an ofce, such as partitions and desks, on the temperatures of the surfaces surrounding
Received in revised form 10 March 2017 the occupants. An accurate estimation of these temperatures is needed to assess the thermal comfort
Accepted 21 April 2017
conditions in a space, thus ensuring that the occupants are satised with their thermal environment. A
Available online 3 May 2017
parameter quantifying the area available to convect heat to the air was proposed to account for the effect
of surfaces of different sizes and positions. A higher amount of available convective area was found to
Keywords:
decrease the temperatures of the surfaces in the room, and vice versa. Although the air temperature also
Room surface temperatures
CFD
increased with additional available convective area, this change was found to be small (typically below
Mean radiant temperature 15%). Moreover, the shape of the air temperature prole far from the surfaces remained unchanged when
Operative temperature changing the amount of available convective area. Therefore, the operative temperature that the occu-
Thermal comfort pants experience is lower in spaces with a large amount of available convective area. Expressions to
estimate the temperatures of the surfaces, generally within 10% of the air temperature rise in the space,
were proposed. These expressions can be useful during the early stages of the design process, when CFD
simulations are impractical and the exact location and type of furniture in a space are not well known.
Together with a model to predict the air temperature prole, these correlations can improve the accuracy
of thermal comfort assessment.
2017 Elsevier B.V. All rights reserved.

1. Introduction pute the operative temperature experienced by the occupants.


However, the geometries of these models rarely include surfaces
The thermal environment conditions that ensure that occupants found in a real ofce such as partitions and desks (e.g. [2]). These
are comfortable in a space refer not only to the air temperature surfaces can have an important effect on the temperature of all the
or velocity near them, but also to the temperatures of the sur- surfaces in the room, including the ceiling, oor and walls, because
faces that surround the occupants. These surfaces exchange heat they facilitate the transfer of heat from the sources (e.g. lighting,
with the occupants via radiation, so surfaces that are too hot or equipment and occupants) to the air.
cold can result in unacceptable thermal conditions in the space, Ultimately, the temperatures of the surfaces in an ofce depend
even if the air temperature is by itself acceptable [1]. Asymmetry on several factors, including convective heat transfer coefcients
in the temperatures of these surfaces can also result in thermal associated with the surfaces and the view factors between them,
dissatisfaction [1]. The Graphical Comfort Zone Method for Typi- among others. Taking into account all these parameters is time-
cal Indoor Environments and the Optional Method for Determining consuming and might not be appropriate during the early stages
Acceptable Thermal Conditions in Naturally Ventilated Spaces the of the design process, when the exact position of the furniture and
adaptive comfort method of the ASHRAE Standard 55-2010 take partitions in an ofce is not well dened. Hence, this work presents
into account the radiation exchange between people and surfaces a simple method to estimate the temperatures of the surfaces in a
by means of the operative temperature [1]. room that does not require exact knowledge of the position of the
Models commonly used to assess the thermal conditions in furniture or the partitions.
ofces, including CFD and multi-zone models, can be used to com-
1.1. Heat transfer paths in an ofce

Corresponding author. In a general sense the heat transfer processes in a space transport
E-mail address: alonso@mit.edu (F.A. Dominguez Espinosa). thermal energy from the source(s) to the heat sink(s). Heat is trans-

http://dx.doi.org/10.1016/j.enbuild.2017.04.067
0378-7788/ 2017 Elsevier B.V. All rights reserved.
156 F.A. Dominguez Espinosa, L.R. Glicksman / Energy and Buildings 147 (2017) 155165

Nomenclature
volumetric thermal expansion coefcient of air
Tr air temperature rise across room
 dimensionless temperature

Roman symbols
A tting constant
Aconv unheated area available for convection heat transfer
Aconv dimensionless unheated convective area
Arhw Archimedes number based on inlet height
Ad dimensionless desk area
Ap dimensionless partition area
Aw dimensionless inlet area Fig. 1. Thermal circuit representation of a space with heat gains generated at ceiling
B tting constant level (a) without partition and (b) with a partition.

C tting constant
cp air heat capacity at constant pressure air temperature prole, instead they rely on the assumption of per-
D tting constant fect mixing of the air, so the temperature in a zone is uniform and
E tting constant equal to the temperature of the outlet. This so-called well-mixed
F tting constant temperature, Tair , can be obtained using the steady-state energy
G tting constant conservation equation:
H room height
Q
hc convective heat transfer coefcient Tair = Toutlet = Tinlet + (1)
mcp
hr radiative heat transfer coefcient
h* dimensionless height coordinate where Toutlet is the outlet temperature, Tinlet is the inlet tempera-
hw inlet height ture, Q are the total heat gains in the room, m is the mass ow rate
hw dimensionless inlet height of air through the space and cp is the heat capacity of air at constant
I tting constant pressure. The well-mixed temperature serves as a convenient ref-
J tting constant erence for the present work because its value is independent of
K tting constant the amount of furniture or the distribution of heat gains in a room
L room length under forced ow (wind-driven natural ventilation or mechanical
l* dimensionless length coordinate ventilation). Moreover, this temperature can be calculated using
m airow rate multi-zone models according to interior and, in the case of natural
Q total heat gains ventilation, exterior conditions.
Rfc oor convective resistance A thermal resistor network is a useful tool to represent the heat
Rpc partition convective resistance transfer paths in a room. As a simple example, consider a very long
r
Rc,f radiative resistance from the ceiling to the oor and wide space, where the view factors from the ceiling to the walls
r
Rc,p radiative resistance from the ceiling to a partition are small enough to be neglected. Assume also that the tempera-
r
Rp,f radiative resistance from the partition to the oor ture of the surfaces are uniform and that the only heat source is
due to lighting (or solar heat gains), which is located at the ceiling
Ta mean air temperature
level, while the other surfaces are adiabatic. In a real room, except
Tair air well-mixed temperature
for other heat sources, the ceiling is generally the warmest surface
Tceiling ceiling temperature
in the space because hot air that rises due to buoyancy stagnates
Toor oor temperature
under it. As a consequence, the convective heat transfer coefcient
Tinlet inlet/supply temperature
associated with the ceiling is signicantly lower than that of the
To operative temperature
other surfaces in the room. The only effective mechanism available
Toutlet outlet temperature
to transfer heat from the ceiling is via radiation. The same holds
Tpartition partition temperature
for the simple example considered here, therefore it is possible to
Trad mean radiant
assume that the convective heat transfer from the ceiling to the
u inlet velocity
air is negligibly small. This is a common assumption in ventila-
x x coordinate
tion studies (e.g. [4]). Fig. 1a shows the thermal resistor network of
y y coordinate r is the radiation thermal resistance
this example. In this gure, Rc,f
z z coordinate
between the ceiling and the oor, and Rfc is the convective thermal
resistance from the oor to the air. Tceiling , Toor and Tair are the
temperatures of the ceiling, oor and the well-mixed temperature
ferred from active surfaces such as lights, equipment and people of the air, respectively. Fig. 1b shows the same space but with a
r and Rr are the radia-
thin partition in the room. In this gure, Rc,p
by convection to the air and by radiation to the walls, oor, parti- p,f
tions and furniture, which is then transferred to the surrounding tion thermal resistances between the ceiling and the partition, and
air. In this work, it is assumed that lights, equipment and occu- between the partition and the oor, respectively. Rfc is the convec-
pants are the only heat sources in the room, while the air in the tive thermal resistance from the partition to the air and Tpartition
space is the only heat sink in steady state. The temperature of the is the temperature of the partition. The well-mixed temperature of
air in the space is not uniform, instead it is higher near the ceiling the air does not change when adding the partition. Now some of the
in spaces with either displacement or mixing ventilation systems radiation from the ceiling is directed to the partition, reducing the
[3]. However, multi-zone models cannot be used to compute the direct radiation from the ceiling to the oor. The total surface area
F.A. Dominguez Espinosa, L.R. Glicksman / Energy and Buildings 147 (2017) 155165 157

Fig. 2. A naturally ventilated ofce with partitions in between rows of occupants. Fig. 3. A naturally ventilated ofce with desks. The area of the desks, including the
The area of the partition is lp hp , which in the example shown in this gure is equal top and bottom surfaces, is measured with respect to the oor area, W L. In this
to 60% of the ofces side walls area of L H. case, the tables cover 80% of this area.

in the room below the ceiling is increased. This area absorbs radia- the centers of two adjacent inlets is 2 m. The outlet is a 0.1 m height
tion from the ceiling and transfers the heat by convection to the air. opening near the ceiling that extends the entire width of the room.
Since the total area is increased, the average surface temperatures The center of the outlet is located at 2.8 m from the oor.
are reduced. Each occupant in the space is represented by a rounded-top
Similar consequences are expected when adding horizontal sur- cylinder, a simple shape that was found to replicate the behavior of
faces, for example desks. The desks reduce the view factor between the convective ow generated by a sitting human and the radiation
the ceiling and the oor, but allow an indirect connection through exchange between a sitting person and the surfaces of a room like
them. A fraction of the heat gains radiated onto the top surfaces of the one considered here [3,5]. Each occupant is 0.31 m radius and
the desks will be conducted to their bottom surfaces and then radi- 1.13 m high. The separation between occupants is 2 m (centerline to
ated to the oor. The net result is an increase in the area available centerline); the rst occupant is located at 2 m from the inlet wall.
to convect heat to the air. Increasing this area reduces the effective There are ve occupants per row, located at 1 m from the partitions
thermal resistance between the heat sources and the heat sink. or, equivalently, exactly in the middle between inlets.
It is possible to reduce the effective thermal resistance by means Each desk between occupants is 75 cm long with a thickness
other than adding partitions or furniture. For a given room height, of 7.5 cm, while the large desk at the end of the room is 3 m long
length and heat gain intensity, a thinner room has surfaces with and of the same width. The desks of contiguous rows touch each
temperatures closer to the well-mixed temperature than wider other, so their width is the same as the width of the ofce. The
rooms, because the walls have a greater participation in the radia- number of desks were varied so their surface area (including top
tion exchange with the heat sources. and bottom surfaces) covers either 40% or 80% of the area of the
Adding convective area inuences the local air temperature dis- oor (L W). As an example, Fig. 3 shows an ofce where the
tribution within the room. This, in turn, could alter the operative desks cover 80% of the oor area. Desks of three materials with
temperature experienced by the occupants, an effect that is not con- different thermal conductivities were used: wood (kt = 0.17 W/mK)
sidered in zonal models. The following sections of this work present and metal (kt = 17.3 W/mK), as well as a case with adiabatic desks
a method to estimate the effect of additional convective area on the (kt = 0 W/mK). Each simulation had desks of only one material.
temperature of the air and the surfaces of a space, thereby allowing The size of the partitions was also varied. The area of each one,
a more accurate assessment of thermal comfort when CFD is not lp hp , was changed so that it would be equal to 0% (no partitions),
practical. 20%, 40%, 60% and 100% of the ofces side wall area, L H. The
length of the partitions was varied while their height was kept con-
2. CFD model stant at 2.25 m (except for the 0% and 100% cases). The location of
the center of the partitions was kept at 7.5 m from the inlet.
Different CFD domains were used to assess the effect of available All these domains are representative of very wide spaces, so
convective area on the temperature of the surfaces surrounding the the L H walls have no effect on the heat exchange of the room.
occupants to test the hypothesis proposed in the previous section. To study the effect of these walls two additional domains were
Fig. 2 shows a naturally ventilated ofce with partitions between simulated. These domains were rooms with a width of 4 m (Fig. 4)
rows of occupants, while Fig. 3 shows the same ofce with desks and 6 m. Together with the cases with partitions covering 0% and
or tables instead of partitions. The geometry of the CFD simulations 100% of the side wall area, the simulations cover a wide range of
is based on an actual ofce building in Tokyo served by a hybrid space widths, from very thin rooms to very wide open-plan ofces.
naturalmechanical ventilation system, but it is intended to be rep- The boundary conditions of the CFD model were dened as fol-
resentative of open plan ofces with few obstacles to the airow lowing:
between the inlets and the outlet. The domain is 15 m long and 3
m height, L and H, respectively, in these gures. The width of the 1 Inlets A uniform air velocity was specied at each one of the
ofce, W, is large compared to its length and height, so Figs. 2 and 3 inlets, the value of which was modied for different simula-
show only a fraction of the width of the ofce. The space has inlets tions. The air temperature at the inlets was dened to be 295 K
along one of the walls, each measuring 0.15 m 0.5 m and with (21.85 C).
their centers located at 1 m from the oor. The distance between 2 Outlet The outlet was dened as a zero gauge pressure boundary.
158 F.A. Dominguez Espinosa, L.R. Glicksman / Energy and Buildings 147 (2017) 155165

transparent to radiation, which is a valid approximation for air with


low humidity content [10,11]. The emissivity of all the surfaces was
dened to be 1. This simplication is justied given that typical
materials in an ofce, except for polished metals, have emissivities
above 0.9 [12]. The radiative view factors between all the surfaces
in the room were calculated using ray tracing.
The cases were run as transient simulations and stopped once
steady state was reached. Steady state was dened to be reached
when both the airow rate at the outlet and the air temperature
rise across the space stopped changing signicantly in value, and
when the value of the temperature rise across the space was within
2% of the value calculated using the energy conservation equation
with the selected airow rate.

2.1. Validation of the CFD model

The CFD code used here was validated by the following studies:

1 Grid independence Results with four different grid sizes


Fig. 4. A naturally ventilated ofce with a width of 4 m. (107,364, 593,598, 726,882 and 2,332,858 elements) were com-
pared against each other in terms of the shape and magnitude
of the resulting vertical air thermal prole. The prole was not
found to change signicantly among the four grid sizes. In partic-
ular, the average difference in temperature between the largest
two meshes at every height was 3%, with a largest deviation of 6%.
Additionally, the mean temperature of the surfaces in the room,
the ceiling, oor and the walls, were compared. The difference in
the temperature of the surfaces between the four grid sizes was
less than 0.4 C, even though the grid size changes by more than
Fig. 5. Mesh used in the study seen on a plane that extends along the length of 2000%. Finally, following the procedure in Ref. [13], the extrap-
the room through the occupants. The inset shows the ne ination layer over the olated relative error and the ne-grid convergence index for the
surfaces in the room (in this case over an occupant and the oor). This is a 2D section average temperature of the ceiling and the oor were computed
of a 3D mesh. for the three largest meshes, nding them all below 0.1%. The
largest mesh size was used for the simulations presented in this
3 Ceiling and occupants These surfaces generated a uniform work and is shown in Fig. 5.
amount of heat ux to represent the heat gains due to lighting 2 Turbulent air jet development The same CFD code of this study
(in the ceiling) and due to the occupants and equipment. Approx- was used to replicate experimental results of axisymmetric and
imately 25% of the total heat gains were dened to be evenly 3D turbulent jets. Centerline velocity, volume ow rate and jet
generated at the ceiling; the rest, by the occupants. The appro- width were compared against axisymmetric jet theory and, when
priateness of this division has been demonstrated before [6]. The possible, experimental data [6]. For the specic case of a rectan-
magnitude of the heat gains were also modied in different sim- gular opening, the shape of the inlet used in this work, the error
ulations. in the dimensionless centerline velocity of the CFD simulation
4 Tables At the surfaces of the tables an energy balance was is less than 10% when compared to theoretical models. Similar
enforced. Heat was allowed to be radiated to and from the sur- magnitude of errors (less than 10%) were found for the dimen-
faces of the tables, convected to the air or conducted through the sionless ow rate and dimensionless jet radius at a distance of
tables. one diameter from the inlet [6].
5 Other surfaces The rest of the surfaces, including the partitions, 3 Turbulent thermal plumes The CFD code was compared against
were dened as adiabatic with no-slip. An adiabatic, no-slip sur- theory and experiments of thermal plumes in terms of their cen-
face interacts with the ow of air, so heat can be transferred to terline velocity, excess temperature, ow rate, and inertial and
and from the surface by convection and radiation. There is also thermal plume width [6]. The difference between CFD and the-
non-zero shear stress at these surfaces. ory and experiments was also found to be below 10% in all the
variables. Additionally, CFD simulations were compared against
Air was modeled using the Boussinesq approximation, with a experimental measurements of centerline velocity of thermal
reference density of 1.14 kg/m3 . The viscosity and volumetric ther- plumes generated by a standing person nding a largest differ-
mal coefcient of expansion of the air were set at 1.8 kg/ms and ence of 7% between the simulations and the experiments [5].
3.3 103 1/K, respectively. 4 Thermal stratication in water scale models and full-sized
All the cases were simulated using Fluent (part of the ANSYS 14.5 rooms The CFD code used here was compared against the exper-
suite), solving the RANS equations for turbulent ow and the energy iments of Nansteel and Greif [14], Olson et al. [15] and Ray [16].
conservation equation. The turbulence model used is the RNG k. The CFD simulations replicated, both in shape and temperature
This model has been shown to accurately simulate room airow magnitude, the experiments by Nansteel and Greif with an R2
dynamics [7]. The near-wall regions were meshed using very ne of 0.97 and by Olson et al. with an R2 of 0.84 [17]. Airow rates
elements (y+ 5), as required by the enhanced wall treatment under buoyancy-driven ventilation were compared to the scale
used in this study [8,9]. Fig. 5 shows an example of a mesh used in experiments by Ray [16], nding an average deviation of 13.5%.
these simulations. In addition, CFD simulations using a similar geometry than in
Thermal radiation exchange was included using the surface- this study were compared against experiments by Chen and
to-surface model. This radiation model treats the air as perfectly Glicksman [18] using a climate chamber. The maximum devi-
F.A. Dominguez Espinosa, L.R. Glicksman / Energy and Buildings 147 (2017) 155165 159

ation between the simulation and the experiments far from the room itself. Some simplications are needed to make the daunting
oor was approximately 13% [6]. task of analyzing the effect of all these parameters manageable.
The aspect ratios of the inlet and outlet were found to have a
A detailed description and complete set of results of the val- marginal effect on the air temperature prole, nevertheless their
idation studies related to jet and plume development, as well as vertical positions are important [6]. In this work, the vertical posi-
thermal stratication previously mentioned, can be found in Refs. tion of the outlet is dened to be located near the ceiling, a common
[5,6,17]. assumption in ventilation studies, as this is generally the case in
real commercial buildings. The vertical position of the inlet was
kept constant at 1 m from the oor. Although this height is not
3. Dimensional analysis
uncommon of windows in some naturally ventilated buildings, it
is not representative of the full range of inlet heights in natu-
The Buckingham  theorem was used to determine a set of
rally ventilated buildings, nor of other ventilation systems, such
dimensionless numbers that determine the temperature of the sur-
as mechanically driven mixing ventilation where the air supply is
faces in the room according to the ow and heat load conditions in
located closer to the ceiling. The effect of the inlet height has only
the space. The dimensionless numbers related to the ow dynam-
been studied qualitatively, so additional research to properly estab-
ics in the room obtained by this method have also been obtained by
lish the effect of this parameter on the temperature of the surfaces
nondimensionalizing the set of equations that govern room airow
in a room is needed. The effects of the length of the room, its height,
[19]. These dimensionless groups are the Archimedes, Reynolds and
and the number, position and geometry of the heat sources have
Prandtl numbers, and the nondimensional temperature.
been studied in exploratory work before [3,6]. Although the effects
The Archimedes number, Arhw , is dened in this work according
of these parameters were found to be generally negligible, more
to the following equation:
exhaustive work is necessary to conrm this nding [3].
gT r h3w The air temperature prole far from the heat sources has been
Arhw = (2) shown to be affected marginally by the exact division of the total
u2 Aw
heat gains between lighting (and/or solar gains) at the ceiling and
where g is the acceleration of gravity, is the volumetric ther- occupancy [6]. The only exception is when the entire amount of
mal expansion coefcient of air, Tr is the air temperature change heat gains in the room are due to lighting or solar gains at the ceil-
across the room (Tr = Toutlet Tinlet ), hw is the height of the inlets, ing [6]. In addition, further tests carried out for a different study
u is the inlet velocity, Aw is the inlet area. The Buckingham  theo- showed that the way in which the lighting heat gains are modeled
rem allows for the arbitrary selection of the normalizing factors (e.g. (as a uniform heat source in the ceiling, as a thin strip along the ceil-
characteristic lengths, areas, velocities, etc.) when constructing a ing or as a separate body that represents suspended ofce lighting
dimensionless number. In this work the form of the Archimedes xtures) does not have a signicant impact on the temperature pro-
number presented in Eq. (2) is used, instead of an alternative ver- le of the air far from these heat sources. In particular, for the same
sion presented before (e.g. [19]), because it has been shown that total amount of heat gains, the temperature prole of the air in the
the expressions that correlate the temperature of the surfaces in occupant zone changed by less than 5% (and commonly below 2%)
the room, as well as its air thermal stratication prole, to its of the room temperature change when changing either the way the
Archimedes number in this form are simple [3,6]. heat sources are modeled, or when changing the percentage of the
Conceptually, the Archimedes number compares the relative total heat gains that were generated by the occupants (with the rest
strengths of the buoyancy forces in the room and the momentum generated by the lighting sources).
of the supply air. For this reason, the Archimedes number is related Furthermore, it is not necessary to determine the explicit depen-
to the path that the supply air jet follows [3,20]. Moreover, it has dence of the stratication prole on the Reynolds and Prandtl
been shown that the air stratication prole of the room depends numbers. Because the Prandtl number is a very weak function of
on its Archimedes number. Rooms with high Archimedes number temperature under typical room conditions (for room air this num-
tend to develop a temperature prole of displacement ventilation. ber changes by less than 1% under a change in temperature of
Low Archimedes numbers are related to more turbulent mixing in 25 C [11]), the only practical way to change the Prandtl number
the lower part of the room, but less mixing in the upper part of to analyze its effects is by changing the uid that is being studied.
the room than in displacement ventilation [3,6]. There is also more Therefore, in the present study that deals only with air in room ven-
variation of the air temperature along the length of the room in a tilation, the Prandtl number was considered to be constant and the
space with a low Archimedes number [3]. dependence of the stratication on it was thus not studied.
The Reynolds number, ReAw , compares the importance of iner- Nielsen [19] has shown experimentally the negligibly small
tial and viscous effects, while the Prandtl number, Pr, compares effect of the Reynolds number on the airow dynamics of rooms
momentum and thermal diffusivities in a uid [11]. under displacement ventilation, due to similar ow structures at
The dimensionless temperature, , is given by: the very high Reynolds numbers expected in realistic spaces. Sev-
T Tinlet eral other experimental works have shown the small effect of the
= (3) Reynolds number on the ow dynamics in rooms, over a wide
Toutlet Tinlet
range of Reynolds numbers, particularly far from the walls [21] and
where T is the temperature to be normalized, Tinlet is the tem- when compared to the signicant effect of the Archimedes number
perature of the air at the inlet, and Toutlet is the temperature of [22,23]. This negligible effect is expected as long as the Reynolds
the air at the outlet. The well-mixed temperature in dimensionless number is large enough to be in the fully turbulent regime, regard-
form is  a,wm = 1. All the temperatures in this work are expressed less of the driving force of the ow. The Reynolds number based
as dimensionless temperatures, including the air temperature and on inlet conditions, ReAw , used in this study always exceeded 3900,
the temperatures of the surfaces in the room (ceiling, oor, walls, which is in the fully turbulent jet regime [24]. Additionally, the
partitions, tables, etc.). independence of the thermal prole with respect to the Reynolds
The boundary conditions of the governing equations result in number was tested and conrmed using the CFD model of this work
dimensionless groups related to the geometry and position of the [3].
inlets and outlets, the strength of the heat sources, the geometry Finally, several dimensionless groups are related to the posi-
and location of the tables and partitions and the geometry of the tion, orientation and size of the partitions and tables, which include
160 F.A. Dominguez Espinosa, L.R. Glicksman / Energy and Buildings 147 (2017) 155165

the radiative view factors and convective heat transfer coefcient


associated with these surfaces in their denition. However, the
objective of this work is to provide a simple parameter that can
be used to estimate the effect of these surfaces on the tempera-
ture of all the surfaces in the room. Moreover, the exact position
of the surfaces might not be known during the design process of
the space. For these reasons, a simple dimensionless number that
is intended to capture all these parameters is proposed in this work.
This dimensionless number measures the amount of area available
to convect heat to the air. For the simulations with tables, for exam-
ple, the convective area of the entire room includes the oor, the
walls and the surface area of the tables, including the top and bot-
tom ones. The dimensionless available convective area is dened
as:
Aconv
Aconv = (4)
WL
where Aconv is the unheated area that is in contact with air, and W
Fig. 6. A section of an ofce (includes only one row of occupants), showing the
and L are the width and the length of the room, respectively. In the locations of the measurement planes. In all the different domains used in this work,
case of a very wide space, it is still possible to use this parameter the measurement planes extend along the entire width of the space.
with a simple modication. In this case, W refers to the width of the
repeating geometry. For example, W refers to the distance between due to occupancy and equipment, and due to lighting, respectively.
partitions in Fig. 2, or between occupants in Fig. 3. The temperature rise across the room, Tr , was kept between 2.5 C
Then, under the previous simplications, the dimensionless and 13 C. The inlet velocity was kept between 0.1 m/s and 6 m/s,
temperature can be expressed as: with approximately 75% of the velocities used below the value of
  2 m/s. The upper limit for the inlet velocity was chosen to be 6 m/s,
 = f Arhw , hw , Aw , Aconv (5)
because at this velocity the local air velocity near the occupants was
where hw is the dimensionless inlet height (normalized by the at most 0.7 m/s, thus staying under the limit value of 0.8 m/s for
height of the room) and Aw is the dimensionless inlet area (nor- sedentary activities of the Graphical Elevated Air Speed Method to
malized by the area of the wall on which the inlets are located). determine comfort conditions of the ASHRAE 55-2010 standard [1].
This dimensionless temperature is given at a certain location in the Nevertheless, by working with dimensionless numbers, the exact
room, described by two dimensionless coordinates: numerical values used in the simulations are less important, as the
z results for a given set of dimensionless numbers is valid for all com-
h = (6) binations of parameters that result in the same dimensionless set,
H
as long as there is not a signicant change in the ow and ther-
y
l = (7) mal dynamics in the room (e.g. as long as the ow in the room is
L
turbulent, et cetera) [26].
where z and y are the z (along the height) and y (along the length) Air and surfaces temperatures are reported. Air temperatures
coordinates at which the temperature is being measured. The x were measured at the planes labeled Measurement planes in
coordinate (across the room width) was not considered, as the air Fig. 6, which are located at a distance from the inlet of 5 m, 9 m,
temperature was not found to vary signicantly in this direction and 12 m, or equivalently at l* = 1/3, 3/5 and 4/5. Although Fig. 6
far from surfaces (by less than approximately 4%) [3]. shows only one row of occupants with no desks or partitions, the
air temperature was measured at these planes in simulations with
4. Methodology partitions and tables, and in the wider ofces. These planes were
chosen to be located between occupants near the inlet, near the
Simulations of the domains presented in Section 2 were per- middle of the length of the room, and near the outlet. Additionally,
formed. To explore the effect of a given dimensionless group, the average temperatures were obtained for the ceiling, the oor, the
value of this group was varied, while the rest of the dimensionless walls and the partitions.
numbers was kept the same. The Archimedes number was varied For all the cases, the resulting operative temperature expe-
by changing the inlet velocity and/or the internal heat gains. The rienced by the occupants was also computed. The operative
dimensionless free convective area, Aconv , was varied by changing temperature, To , is dened as:
the number of tables or the size of the partitions as explained before.
hc Ta + hr Trad
The dimensionless inlet area was also changed. To = (8)
hc + hc
The parameters of the simulations were chosen to be repre-
sentative of naturally ventilated spaces. The Archimedes number where hc is the convective heat transfer coefcient associated with
considered ranged from a low value of 0.07, up to 7.91. Additional the occupants, hr is the radiative heat transfer coefcient, Ta is the
simulations were run with Archimedes numbers of 63 and 100. mean air temperature near the occupants and Trad is the mean radi-
The dimensionless inlet areas used are 6.25 103 (0.0375 m2 each ant temperature [1]. Mathematically, the operative temperature is
inlet), 1.25 102 , 5 102 and 6.25 102 , although for some the weighted average of the mean radiant temperature and the
cases additional area values of 2.5 102 and 3.75 102 were also air temperature around the occupants with the radiative and con-
simulated. These dimensionless areas are typical of openings in nat- vective heat transfer coefcients as the weights. The mean radiant
urally ventilated buildings, although in natural ventilation there temperature is a function of the view factors between the occupant
is more variability in size [25]. The heat gains were kept below and the surfaces surrounding said occupant [1]. Results are pre-
125 W/m2 , four times the heat gains in a typical ofce in the United sented as plots of dimensionless operative temperature (calculated
States (approximately 30 W/m2 ) [12]. The heat gains were divided using Eqs. (3) and (8)) against dimensionless available convective
between the occupants and the ceiling, to represent the heat load area.
F.A. Dominguez Espinosa, L.R. Glicksman / Energy and Buildings 147 (2017) 155165 161

Fig. 7. Dimensionless air temperature,  a , as a function of dimensionless height from the oor, h* , for spaces (a) with different widths, (b) with partitions, and (c) with desks,
at the measurement plane located at 9 m (l* = 3/5) from the inlet.

In this work, the air temperature surrounding the occupants was all the measurement planes. However, the deviation near the oor
estimated using the measurement planes at 5 m and 9 m from the is as high as 40%, while near the ceiling is between 26% and 30%.
inlet because these planes are located between the occupants. The The deviation in the air temperature between the cases with
air temperature surrounding the occupants was approximated as different partition areas is substantial (up to 40%) near the oor
the average air temperature at these measurement planes, from the and the ceiling. However, the deviation due to the partition size in
oor up to a height of 1.35 m (l* = 0.45), i.e. approximately 20 cm the temperature of the air between the dimensionless heights of
above the head level of the occupants. The mean radiant tem- 0.1 h* 0.9 is generally below 14%, corresponding to 0.7 C. The
perature was calculated using the average temperature of all the air temperature prole in the space with desks show an impor-
surfaces (including partitions and tables, if applicable) and the view tant difference near the tables (h* 0.2). The temperature of the
factors were computed using ray tracing software. The accuracy of air surrounding the tables is signicantly higher than at other loca-
the ray tracing software has been validated before [3]. The convec- tions near them, however the prole seems to be insensitive to
tive and radiative heat transfer coefcients for the occupants where the thermal conductivity of the desks. The high temperature of the
found to be similar with a value of approximately 5 W/m2 K. Thus, desks is due to the radiation heat transfer they receive from the
they were assumed to be equal. active sources, particularly from the ceiling. This heat is then trans-
ferred to the surrounding air by convection, thereby resulting in the
increased air temperatures seen. Nevertheless, the air temperature
5. Results and discussion prole is similar in shape and the deviation of the dimensionless
temperature of the air is generally below 15% far from the tables
5.1. Effect of convective area on the temperature of the air and far from the ceiling and the oor.
This means that it is possible to use thermal stratication models
Fig. 7 shows the vertical temperature stratication prole of for rooms with no additional convective area to estimate the air
the air at the measurement plane at 9 m from the inlet, for the temperature prole (e.g. [3,6]) in rooms of different widths, with
spaces with different widths (Fig. 7a), with partitions (Fig. 7b) and partitions and desks, as long as the results are used for the air region
with desks (Fig. 7c). The spaces with partitions and desks are very away from the ceiling and the oor (0.1 h* 0.9) as well as the
wide. Except for the change in the width, the addition of parti- desks. The error incurred by using these stratication models would
tions or desks, the spaces are identical, including equal inlet areas be approximately 15%.
and velocities, and heat gains. The vertical axis of the plots shows
the nondimensional temperature of the air,  a , as a function of the
dimensionless vertical height from the oor, h* . Each plot has sev- 5.2. Surface temperatures
eral curves identied by different colors. In Fig. 7a the different
curves show the air temperature prole in spaces with different Determining the shape of the temperature of the surfaces as a
dimensionless widths, W* W/L. The plot in Fig. 7b shows the function of the dimensionless length, l* , or dimensionless height,
results for different partition areas, Ap , in a wide ofce. The par- h* , is of dubious usefulness due to the dependence of these tem-
tition areas are normalized with respect to the rooms side area, i.e. peratures on the exact location of the occupants, the stagnation
Ap (lp hp )/(L H). Finally, the different curves in Fig. 7c repre- point of the cool air jet from the inlet, and/or the furniture and
sent different surface areas (including the top and bottom surfaces) partitions [3]. For this reason, the following discussion is limited
and thermal conductivities of the desks in wide spaces. The dimen- to the average temperatures of the surfaces in the room. Average
sionless desk area, Ad , is normalized with respect to the room oor temperatures can still provide valuable information for the opera-
area, W L. tive temperature to the building designer to assess thermal comfort
Although as expected the air temperature increases with conditions.
increasing convective area (higher air temperature in thinner Fig. 8 shows the average dimensionless temperature of the
spaces and with more partition and desk area), the temperature ceiling, the oor and the walls as a function of the dimension-
prole of all the spaces is similar in magnitude and shape. Similar less convective area in the room, Aconv . All these cases share the
results were found for the other measurement planes. The devia- same Reynolds number, ReAw = 35, 379, dimensionless inlet area,
tion in air temperature between the rooms with different widths is Aw = 1.25 102 , and inlet height, hw = 1/3. Different curves show
generally below 15%, corresponding to less than 1 C for the param- the results for different values of the Archimedes number, as indi-
eters used to create Fig. 7, for the region between 0.1 h* 0.9 in cated in the gure.
162 F.A. Dominguez Espinosa, L.R. Glicksman / Energy and Buildings 147 (2017) 155165

Fig. 8. Average dimensionless temperature of the (a) ceiling, (b) oor, and (c) walls as a function of the convective area, Aconv . Different Archimedes numbers, Arhw , results
in different curves. All simulations had an inlet area of Aw = 1.25 102 .

Fig. 8 shows that the temperatures of all the surfaces decrease With a higher Archimedes number, the cold air does not mix sig-
with increasing area available to convect heat to the air. For nicantly with the warm air in the upper part of the room, as
example, for an Archimedes number of Arhw = 0.6 the temper- buoyant forces keep it at oor level. The centerline temperature of
ature of the ceiling decreases from being twice the well-mixed the jet at the stagnation point, for Arhw = 100, was found to be only
temperature for the innitely wide room, to being 1.5 times the jet,h =0 0.6. In addition, air velocities were higher in the lower
well-mixed temperature for the 2 m wide space, a difference of part of the room with higher Archimedes number, because the cold
2.5 C and 5 C for typical room air temperature rises of 5 C and jet experiences a larger acceleration due to buoyant forces [3,27].
10 C, respectively. The temperatures of the walls and the oor The centerline velocity of the supply jet at 5 mm from the stag-
also decrease with increasing convective area. For Arhw = 0.6, the nation point over the oor was 2.3 times higher in the case with
reduction corresponds to 2 C for a room temperature rise of 5 C. Arhw = 100 than with Arhw = 0.6, when the velocities were nor-
When the dimensionless convective area in the room is Aconv = 4.3, malized with respect to their value at the inlet. The same higher
the temperatures of the walls and the ceiling are just 1 C above air velocities in the lower part of the room with high Archimedes
the well-mixed temperature each, for a room temperature rise number were found when comparing two simulations with the
of 5 C. This gure also shows that both horizontal and verti- same inlet velocity. Similar ndings have been reported in exper-
cal surfaces (partitions and tables) approximately lie on a single imental work that measured the air velocity distribution from air
temperatureconvective area curve. The only substantial exception diffusers and its dependence on Archimedes number [28]. Thus,
is seen in the ceiling temperature for the table that covers 40% of lower jet temperatures and higher air velocities cause a decrease in
the oor area. The temperature of the ceiling for this case is above the temperature of the oor when the Archimedes number is high.
the trend of the rest of the cases, signaling that this horizontal sur- The ceiling and walls are also cold because they are connected via
face is less efcient at convecting heat from the sources to the air. radiation to the cold oor.
Vertical surfaces are more efcient at convecting heat to the air Fig. 8 demonstrates that as the Archimedes number increases,
because they do not block the direct radiation from the ceiling to the effect of the convective area decreases and the dimensionless
the oor. Moreover, the convective heat transfer coefcient associ- temperatures of all the surfaces tend to get closer to a value of 1,
ated with the surface of the desk facing downwards is signicantly the well-mixed temperature.
lower than that of the vertical surfaces. This lower convective coef- The effect of the inlet area can be seen by comparing Figs. 8 and 9.
cient is due to the lower air velocities under a warm surface facing The former shows the results for a dimensionless inlet area of
downwards, as the buoyant forces keep a layer of near stagnant Aw = 1.25 102 , while the latter is for Aw = 5.00 102 . Sim-
warm air in contact with the surface. ulations for a dimensionless partition area of Ap = 0.4 were not
Fig. 8 also shows that the dimensionless temperatures of all the performed for the dimensionless area of Aw = 5.00 102 because,
surfaces decrease with increasing Archimedes number. It was seen as seen in these gures, it is not needed to establish a trend. At a
in the results of the CFD simulations that with a low Archimedes xed Archimedes number, the dimensionless temperatures of all
number, the supply air jet throw is longer, so the jet touches the the surfaces increase with increasing inlet area. The dimensionless
oor farther from the inlet and continues moving along the oor temperature of the ceiling increased the most when increasing the
towards the outlet wall [3,27]. The high supply momentum causes inlet area, with an increase between 13% and 70%. The oor had the
the air that reaches the outlet wall to rise before returning towards lowest temperature change, between 0.5% and 32%, with respect to
the inlet at a higher height [3,27]. This air is entrained by the sup- a change in the inlet area.
ply jet and, because its temperature is higher than the supply jet, The trend seen in Fig. 8a and 9a for the dimensionless tem-
it increases the temperature of the jet [3,27]. For example, the perature of the ceiling, for example, can be approximated as the
centerline dimensionless temperature of the jet when it reached
the oor is jet,h =0 1.3 for an Archimedes number of Arhw = 0.6.
F.A. Dominguez Espinosa, L.R. Glicksman / Energy and Buildings 147 (2017) 155165 163

Fig. 9. Average dimensionless temperature of the (a) ceiling, (b) oor, and (c) walls as a function of the convective area, Aconv . Different Archimedes numbers results in
different curves. All simulations had an inlet area of Aw = 5.00 102 .

Table 1 Table 3
Values of the tting constants in Eq. (9). Values of the tting constants in Eq. (11).

A B C D E F Parameter Expression r2

0.33 0.61 0.39 1.63 0.52 2.25 J 2.43Aw + 0.83 0.97


K 24.14Aw 2 1.01Aw + 0.07 0.92

Table 2
Values of the tting constants in Eq. (10). dashed lines in this gure encompass a 10% deviation from the t.
Parameter Expression r2 The largest deviation from this line is seen in the simulation with
the lowest Archimedes number. Table 3 shows the functional forms
G 9.73Aw + 1.04 0.99
I 13.51Aw + 0.61 0.94 to obtain J and K as a function of the inlet area.
A simplied correlation for the temperature of the oor and
the walls that also ignores the effect of the inlet area is shown
multiplication of two exponential decays, one with respect to the schematically in Fig. 10b. In this gure, the data points show the
Archimedes number, Arhw , and one for the convective area, Aconv : dimensionless temperature of the oor and walls, regardless of
       Archimedes number, convective area and inlet area, as a function
 c = A exp BArhw +C D exp EAconv +F (9) of the ceiling dimensionless temperature. The dashed lines in this
where A, B, C, D, E, F are tting parameters that, in general, depend gure encompass an area with a deviation of 10% from the tting
on the inlet area, Aw . This equation captures the fact that the dimen- line. Because the effect of the inlet area has been ignored, more data
sionless temperature decreases both with increasing Archimedes points lie outside of the 10% region, and the correlation factor of
number, Arhw , and increasing convective area, Aconv , and that these the t, at r2 = 0.95, is lower than when using Eq. (11). Neverthe-
changes decay approximately exponentially. Table 1 shows the less, most of the data points are still reasonably represented by the
values for the tting constants for the case with an inlet area of simplied t. This simplied correlation is given by:
Aw = 1.25 102 . This t was obtained using the Curve Fitting  f
=  w = 0.67 c + 0.21 (12)
Toolbox for MATLAB and has a correlation factor of r2 = 0.97.
To account for different inlet areas, a trend that relates the result
5.3. Operative temperature
in Eq. (9) with the dimensionless ceiling temperature of rooms with
different inlet areas is given next:
Fig. 11a shows a plot of dimensionless mean radiant tempera-
 c (Aw ) = G c (Aw = 1.25 102 ) + I (10) ture against dimensionless available convective area, while Fig. 11b
shows a plot of dimensionless air temperature. The mean air tem-
where G and I are tting parameters obtained using the expressions perature was estimated using the average air temperature from the
in Table 2, and  c (Aw = 1.25 102 ) is given by Eq. (9) and Table 1. oor to a height of 1.35 m, 25 cm above the head of the occupants,
A similar expression as that in Eq. (9) can be constructed to esti- at the 5 m and 9 m planes. These two quantities were then used to
mate the dimensionless temperature of the oor and the walls. compute the operative temperature that the occupants experience.
However, by taking advantage of the fact that the dimensionless The results are shown in dimensionless form in Fig. 11c. These g-
temperatures of the oor and the walls are approximately equal, ures are all for an inlet area of Aw = 1.25 102 , however similar
and that they were found to be linearly related to the ceiling tem- trends were seen for other values of the inlet area. Fig. 11a rein-
perature, regardless of surface type and/or Archimedes number, forces the results of the previous sections by demonstrating that the
(Fig. 10a), a simpler expression that depends on a lower number of mean radiant temperature of the occupants decreases with increas-
tting constants can be used instead: ing convective area. Fig. 11b shows that the air temperature does
not vary signicantly with respect to the convective area, how-
 f
=  w = J  c + K (11)
ever it is dependent on the Archimedes number. It is important to
where J and, K are tting parameters that, in general, depend on mention that the air temperature around the occupants is lower
the inlet area, Aw . This line is compared to the data in Fig. 10a. The than the well-mixed temperature of the space at  wm = 1. For the
164 F.A. Dominguez Espinosa, L.R. Glicksman / Energy and Buildings 147 (2017) 155165

Fig. 10. Average dimensionless temperature of the oor,  f , or walls,  w , as a function of the average dimensionless temperature of the ceiling,  c . The dashed lines encompass
a 10% range from the line of Eq. (11). (a) Plot for spaces with different surfaces (marker shape and border color) and different Archimedes numbers (marker ll color) and
with dimensionless inlet area of Aw = 1.25 103 . (b) Plot that includes the results of all the inlet areas considered in this work.

Fig. 11. Average dimensionless (a) mean radiant, (b) mean air, and (c) operative temperature against dimensionless convective area, Aconv . The inlet area of the simulations
in this plot was Aw = 1.25 102 .

range of Archimedes numbers used in this study, the mean tem- fer coefcients of the surfaces typically found in an ofce, such as
perature decreased with increasing Archimedes number, being as partitions and desks, was proposed. Correlations to determine the
low as 64% of the well-mixed temperature. Thus, approximating temperatures of the surfaces surrounding an occupant in a room
the air temperature as the well-mixed temperature might result (ceiling, oor, walls) as a function of the Archimedes number, the
in an inaccurate estimation of the operative temperature. Previ- inlet area and the free convective area were developed. It was found
ous research has called attention to this problem of the well-mixed that regardless of the orientation or position of these surfaces, the
assumption (e.g. [3,6,29]). Finally, Fig. 11c demonstrates that the average temperatures of the surfaces in the room depend on this
decrease in mean radiant temperature is important, as it causes a parameter.
reduction in the operative temperature, even if the air temperature The temperatures of the room surfaces were found to decrease
does not follow a simple trend, as hypothesized in this work. with increasing Archimedes number. A high Archimedes number
was related to higher air velocities and lower supply jet temper-
6. Conclusions ature at the stagnation point. The high velocity is a consequence
of the larger acceleration experienced by the supply air due to the
The addition of partitions and horizontal surfaces substantially higher buoyant forces. Cold oors and walls are related to cold ceil-
reduces the average radiant temperature while only moderately ings and vice versa, as these surfaces are all connected via radiation
changing the air temperature near the occupants. This results in exchange. Indeed, a simple linear relation was found between the
a decrease in the operative temperature between 10% and 21% oor and walls temperatures, and the ceiling temperature. In addi-
of the overall temperature difference between the inlet and the tion, it was found that for a given Archimedes number, a larger
outlet. This corresponds to a temperature difference of up to 2 C supply area meant that the temperature of all the surfaces were
between the spaces with few and several additional surfaces in higher than with smaller inlet areas. However, it is also possible to
typical naturally ventilated spaces. Due to the small effect of the change the inlet height while keeping the same Archimedes num-
convective surfaces on the shape of the air temperature prole, it is ber. This option was not explored in this work. Future work can
still possible to use thermal stratication models for rooms without help expand this study by analyzing the effect of the inlet height
additional surfaces to estimate this prole away from the surfaces on the temperatures of the surfaces in a space.
in the room with an error below approximately 15%. The correlations proposed in this work, together with air strati-
The free convective area, a parameter that effectively captures cation proles, can help designers to assess comfort conditions in
the effect of the radiation view factors and convective heat trans- buildings more accurately than is currently possible using models
F.A. Dominguez Espinosa, L.R. Glicksman / Energy and Buildings 147 (2017) 155165 165

that rely exclusively on the well-mixed temperature. The correla- [13] I.B. Celik, U. Ghia, P.J. Roache, C.J. Freitas, H. Coleman, P. Raad, Procedure for
tions presented here are simple enough to be used in multi-zone estimation and reporting of uncertainty due to discretization in CFD
applications, J. Fluids Eng. Trans. ASME 130 (2008).
models, given that its inputs are all available in multi-zone models. [14] M.W. Nansteel, R. Greif, Natural convection in undivided and partially divided
rectangular enclosures, J. Heat Transf. 103 (1981) 623629.
Acknowledgments [15] D.A. Olson, L.R. Glicksman, H.M. Ferm, Steady-state natural convection in
empty and partitioned enclosures at high Rayleigh numbers, J. Heat Transf.
112 (1990) 640647.
This research was supported by the Mexican Council for Science [16] S.D. Ray, Modeling Buoyancy-Driven Airow in Ventilation Shafts,
and Technology (CONACYT), the U.S. China Clean Energy Research Massachusetts Institute of Technology, 2012 (Ph.D. thesis).
[17] M.-A. Menchaca-Brandan, L.R. Glicksman, The importance of accounting for
Center (CERC) and the Martin Family Society of Fellows for Sustain- radiative heat transfer in room airow simulations, in: Proceedings of
ability Roomvent 2011 The 12th International Conference on Air Distribution in
Rooms, SCANVAC (Scandinavian Federation of Heating, Ventilating and
Sanitary Engineering Association), 2011.
References [18] Q. Chen, L. Glicksman, System Performance Evaluation and Design Guidelines
for Displacement Ventilation, 2003.
[1] Thermal Environmental Conditions for Human Occupancy, American Society [19] P.V. Nielsen, Displacement ventilation in a room with low-level diffusers
of Heating, Refrigerating and Air-Conditioning Engineers, 2010. Indoor Environmental Technology, vol. R8836, Department of Building
[2] T. Catalina, J. Virgone, F. Kuznik, Evaluation of thermal comfort using Technology and Structural Engineering, Aalborg University, 1988.
combined cfd and experimentation study in a test room equipped with a [20] T.-G. Malmstrm, Archimedes number and jet similarity, in: Proceedings of
cooling ceiling, Build. Environ. 44 (2009) 17401750. the 5th International Conference on Air Distribution in Rooms ROOMVENT,
[3] F.A. Dominguez Espinosa, Determining thermal stratication in rooms under 1996, pp. 415422.
mixing and displacement ventilation, Massachusetts Institute of Technology, [21] H.B. Awbi, M.M. Nemri, Scale effect in room airow studies, Energy Build. 14
2016 (Ph.D. thesis). (1990) 207210.
[4] E. Mundt, Displacement Ventilation Systems convection ows and [22] D. Etheridge, M. Sandberg, Building Ventilation: Theory and Measurement,
temperature gradients, Build. Environ. 30 (1995) 129133. John Wiley & Sons, 1996.
[5] M.-A. Menchaca Brandan, L.R. Glicksman, How to replicate computationally a [23] C. Walker, G. Tan, L. Glicksman, Reduced-scale building model and numerical
human plume with a simple geometry, and how not to, in: Proceedings of investigations to buoyancy-driven natural ventilation, Energy Build. 43
Roomvent 2011 The 12th International conference on air distribution in (2011) 24042413.
rooms, SCANVAC (Scandinavian Federation of Heating, Ventilating and [24] J.H. Lee, V.H. Chu, Turbulent jets and plumes, in: A Lagrangian Approach,
Sanitary Engineering Association), 2011. Springer Science & Business Media, 2012.
[6] M.A. Menchaca Brandan, Study of Airow and Thermal Stratication in [25] A. Roetzel, A. Tsangrassoulis, U. Dietrich, S. Busching, A review of occupant
Naturally Ventilated Rooms, Massachusetts Institute of Technology, 2012 control on natural ventilation, Renew. Sustain. Energy Rev. 14 (2010)
(Ph.D. thesis). 10011013.
[7] Q. Chen, Comparison of different k- models for indoor air ow computations, [26] A.A. Sonin, The Physical Basis of Dimensional Analysis, Technical Report,
Numer. Heat Transf. B Fundam. 28 (1995) 353369. Massachusetts Institute of Technology, 2001.
[8] ANSYS, FLUENT Theory Guide, ANSYS, 2011. [27] F. Dominguez Espinosa, L.R. Glicksman, Determining thermal stratication in
[9] H.B. Awbi, Calculation of convective heat transfer coefcients of room rooms with high supply momentum, Build. Environ. 112 (2017) 99114.
surfaces for natural convection, Energy Build. 28 (1998) 219227. [28] P.V. Nielsen, Velocity distribution in a room ventilated by displacement
[10] J.H. Lienhard V, J.H. Lienhard IV, A Heat Transfer Textbook, Dover Publications ventilation and wall-mounted air terminal devices, Energy Build. 31 (2000)
Inc., 2011. 179187.
[11] A.F. Mills, Basic Heat and Mass Transfer, 2 ed., Prentice Hall, 1999. [29] P.F. Linden, G.F. Lane-Serff, D.A. Smeed, Emptying lling boxes: the uid
[12] 2009 ASHRAE Handbook Fundamentals, American Society of Heating, mechanics of natural ventilation, J. Fluid Mech. 212 (1990) 309335.
Refrigerating and Air-Conditioning Engineers, 2009.

You might also like