Anales AFA - XVI Meeting on Recent Advances of Physics of Fluids and its Applications 46-51
IMPACT OF DESIGN PARAMETERS ON THE NATURAL VENTILATION OF A FAMILY
HOUSE ATTIC FROM SANTA FE (ARGENTINA)
M. Berli*1, A. Brondino1y J. Di Paolo 1
1Grupo de Investigación en Mecánica de Fluidos (GIMEF) Universidad Tecnológica Nacional Facultad Regional Santa Fe,
Lavaisse 610 (3000) Santa Fe Santa Fe Argentina.
Recibido: 27/10/2021 ; Aceptado: 15/11/2021
The roof of a family home introduces important amounts of heat into the house. In Santa Fe city (Argentina), the attic
of social family houses is not designed with thermal optimization purposes. In this work, a low Reynolds number k-e
turbulent model is computationally implemented to show that a proper selection of constructive parameters of a vented
attic leads to a great reduction of the heat transferred from the roof to the ceiling. The main physical consequence of a
vented attic is the existence of a convective flow barrier close to the roof that carries important amounts of heat from
the attic out to the environment. It is shown that the heat transferred through the ceiling to the interior of the house can
be reduced in more than 70% for summer conditions. Roof pitch and attic volume are also explored.
Keywords: vented attic, computer fluid dynamic, buoyancy flow.
https://doi.org/10.31527/analesafa.2022.fluidos.46 ISSN 1850-1168 (online)
I. INTRODUCTION
A family home roof is usually the component that trans-
fers more heat than any other wall of the house. Conse-
quently, most design building codes suggest passive ventila-
tion as an alternative to produce important energy savings,
i.e., the International Residential Code [1] demands attics
ventilation unless local codes state the contrary, generally
due to atmospheric or climatic conditions.
Most attics of social family houses in Argentina are built
without connections to the exterior. In some cases, only
ventilation grills are used to allow pressure homogenization
with external environment and to evacuate moisture, but the
vent size, position, and orientation are not specifically de-
signed to generate an important flow rate. Passive cooling
isn’t recommended by any national building code and is
rarely implemented in local government social houses pro-
grams. The advantages and disadvantages of the application
of attic ventilation compared to the traditional closed attic
need to be evaluated using local climate data bases before
recommending it as requirement in model house schemes
[2]. Therefore, the impact of natural convection airflow on
the heating load and its relationship with standard construc-
tion parameters (roof inclination angle, air gap cavity volu-
me, etc.) can offer helpful information for sustainable roof
design, especially in hot climate zones like Santa Fe city
(Argentina) with temperatures that can exceed 43ºC. In this
work, a trapezoidal-shaped cavity is used to describe the
geometry of a single pitch roof attic. This geometry allows
the possibility to study different roof pitches yet maintai-
ning the attic volume constant and vice versa [3].
Recent works [4] have proved Computer Fluid Dynamics
(CFD) as a valuable tool to estimate the impact of attic ven-
tilation in roof thermal performance. Wang et al. [5] studied
the natural convection effects on vented attics performan-
* marcelo.berli@uner.edu.ar
ce under winter conditions employing a two-dimensional
steady-state finite volume model on a triangular air cavity
by changing the vent size, ambient air temperature and cei-
ling insulation. On later papers, Wang and Shen [6,7] em-
ployed a 2D unsteady CFD model assuming a mid-plane
symmetry to study the impacts of ventilation ratio, vent ba-
lance and roof pitch on air flow and heating load of both
sealed and vented attic for summer conditions. Recently, If-
fa and Tariku [8] studied the effects on temperature profile
and airflow patterns inside the cavity by varying the gap size
between roof sheathing and ceiling insulation and the loca-
tion of the vent area under both summer and winter condi-
tion. The same authors used transient boundary conditions
[9] to study the dynamic response of the roof thermal per-
formance. All mentioned works showed the benefits of a
vented attic in reducing the cooling loads, but they were
mostly based on strict building codes guidelines and focu-
sed mostly in the vent size influence on the thermal per-
formance, minimizing the importance of other constructive
parameters that could improve the thermal performance of
the attic of social homes of our zone.
The aim of this work is to obtain basic guidelines for se-
lecting a few construction parameters that can optimize the
attic thermal performance, i.e. vent openings size, attic vo-
lume and roof pitch. The turbulent air flow and natural con-
vection heat transfer inside the attic are modeled in terms
of the low Reynolds number kεturbulent model which is
validated using both experimental and numerical works.
II. MATHEMATICAL MODEL AND NUMERICAL
TECHNIQUE
Fig. 1sketches the typical parts of a single pitch roof that
usually cover a low-cost family house, where thermal loads
and the proposed vent openings on the front and back walls
are shown. The air cavity is bounded by the inclined roof,
the ceiling and the walls. The high temperatures of air and
©2022 Anales AFA 46
FIG. 1: Low-cost family house vented shed roof (attic), with ther-
mals loads scheme.
solar radiation that impact on the roof, are transmitted to
the house indoor passing through this cavity. The received
thermal load will be considered as the combined effect of
both solar radiation and heat exchange by natural convec-
tion between the roof and the environment. The front and
back walls are assumed adiabatic (we neglect the heat flow
through these walls). The ceiling consists of a thin cove-
ring material with negligible thermal resistance; therefore,
we assume that it has the same temperature as that inside
the house. The attic is located on a square room of 4 m long
each side. The set of governing equations for natural con-
vection turbulent flow is given by the Reynolds-averaged
Navier-Stokes (RANS) equations of continuity (1), momen-
tum (2) and thermal energy (3):
·(ρu) = 0 (1)
(u·)u=·[PI + (µ+µT)(u+ (u)T)
2
3(µ+µT)(·u)I
2
3ρkI]+(ρρ0)g(2)
ρCpu·T=·(λT)(3)
where uis the velocity field vector, ρthe temperature
dependent air density, ρ0the air density evaluated at a re-
ference temperature (ambient), Pthe pressure, gthe gravity
acceleration field, µthe air dynamic viscosity, µTthe eddy
viscosity, kthe turbulent kinetic energy, Cpthe air heat va-
lue, λthe air thermal conductivity and Tthe temperature
field.
The low Reynolds number kεturbulence model em-
ployed in this work is a modification of the well-known
kεmodel, widely used in analyzing ventilation flow in-
side houses and buildings [10]. The model introduces two
additional transport equations and two dependent variables:
the turbulent kinetic energy, k, and the turbulent dissipa-
tion rate εto the above RANS equations [11]. In order to
describe the flow in the wall region, where viscous effects
dominate, the employed low Reynolds number kεmo-
del adapts the turbulence transport equations with the AKN
(initials of its developers: Abe, Kondoh and Nagano) model
[12] by introducing damping functions.
The equations system was solved using the Non-
Isothermal Turbulent Flow module of the commercial fini-
te element software COMSOL Multiphysics 4.4. To have a
good convergence scheme, the solutions were obtained by
following two steps: in the first step, the viscosity was set
to a value approximately 10 times higher than the expected
physical viscosity inside the attic, in the second step, the ac-
tual physical viscosity is calculated using the first step result
as initial guess.
The nonlinearity introduced by the Navier-Stokes
(RANS) and turbulence transport equations were solved
using a segregated approach [13]: Navier-Stokes equations
in one group and the turbulence transport equations in
another. For each iteration of the Navier-Stokes group, three
iterations for the turbulence transport equations were nee-
ded. A pseudo-time stepping approach was used to obtain
steady state solutions.
III. RESULTS AND DISCUSSION
Model validation
The Numerical model was validated with Ampofo and
Karaviannis [14] experimental measurements, who studied
the low turbulence air flux driven by natural convection insi-
de a closed cavity of square cross section (0.75 m each side),
under a temperature gradient of 40 K between the lateral
walls (323 K on the left wall and 283 K on the right one),
with insulated top and bottom walls. The cavity was deep
enough (1.5 m) to assure a 2D flux at any cross section far
from anterior and posterior walls. Velocity and temperature
profiles of moving air were measured at different positions
in a cross section placed at the middle distance between an-
terior and posterior walls. Our numerical results show good
agreement with experimental measurements (see Fig. 2).
Moreover, the agreement between the temperature profiles
is very precise in the boundary layer region where heat flu-
xes need to be computed in this work.
The model was also validated with the numerical work
by Wang et al. [5], who developed a V2f turbulent model
to analyze the buoyancy-driven air flux inside a vented attic
of triangular shape, under a temperature gradient of 20 K
between the base and the inclined top side, for winter con-
ditions. The geometry of the 2D triangular cross section of
the attic was 8 m long in the base and had a roof inclination
ratio of 5/11. The streamlines and isotherms predictions ob-
tained with our model reproduce the same behavior, espe-
cially for the case of a vented attic with 2 cm openings width
and a R-20 insulation type for a 267 K ambient temperatu-
re. But the best agreement is observed when comparing two
variables relevant for our study at summer conditions: heat
transferred through the ceiling (HTTC) and mass flow. The
HTTC predicted by Wang et al. [6] is 48.5 W/m while ours
is 46.8 W/m, computing a percentage difference of about
3.5%, while the mass flow difference between both models
is about 1% (0.0236 kg/s for Wang’s model and 0.0238 kg/s
for our model).
Summer boundary condition: Roof constant temperatu-
re vs solar radiation.
The influence of geometric dimensions on the attic ther-
mal performance is very dependent on the thermal boun-
dary conditions. A 295.15 K temperature is desired to have
a comfortable indoor environment, being this the prescribed
temperature on the ceiling. But it is not fully understood
what types of thermal conditions can mimic the real roof
conditions for any situation.
For winter conditions and snowfall zones, constant roof
Berli et al. / Anales AFA - XVI Meeting on Recent Advances of Physics of Fluids and its Applications 46-51 47
FIG. 2: Comparison of the present model results and experimental
measurements made by Ampofo and Karaviannis14. Red: Vertical
velocity. Orange: Temperature.
temperature is an appropriate condition since snow accumu-
lation imposes its temperature and insulates the roof from
outside atmosphere. For summer conditions, the roof is ex-
posed to sun radiation, different ambient air temperature
and convective dissipation, so it is not clear which roof tem-
perature should be used as a boundary condition. For this
reason, we compared two cases: a) a prescribed constant
roof temperature (CRT), used widely in the literature [5-8]
and b) a prescribed constant solar radiation (CSR) coupled
to convective heat exchange with external environment. For
both cases, a closed cavity without conection with exterior
is used.
The value of 345.15 K was used for the CRT, as a ty-
pical summer roof temperature [6]. For the CSR case both
solar radiation (heat flux) of 1000 W/m2and a convective
heat flux from the Non-Isothermal Turbulent Flow module
of COMSOL was used under different summer air tempera-
tures.
Table 1compares the HTTC between CRT and CSR for
a varying ambient temperature and summarizes the abso-
lute percentage differences (APD) between them. It can be
noted that the APD minimum difference is computed for
305.15 K, and can be as greater as 22.5%. This means that
TABLE 1: Constant roof temperature and constant solar radiation
results comparison for summer conditions.
Ambient Tempe-
rature (K)
Roof
Average
Tempera-
ture (K)
Heat
transfer
to the
ceiling
(MW/m)
Absolute
percen-
tage
difference
(APD)
(%)
CRT Case 345.15 65.6 -
CSR
300.15 340.75 55.3 15.8
305.15 345.47 64.6 1.5
310.15 350.99 72.0 9.7
315.15 355.51 80.3 22.5
prescribing a constant roof temperature with a single value
could not be an appropriate boundary condition for diffe-
rent ambient scenarios. The roof temperature will be hardly
constant during summer for a wide ambient temperature va-
riation and it is expected that the attic thermal performance
will depend on the heat dissipation due to atmosphere con-
ditions changes.
Radiation values are known from daily weather reports
and convective coefficient can be easily obtained from bi-
bliography, but roof temperature is not always known. Mo-
reover, for vented configurations, the air exchange between
the attic and the atmosphere could affect the average roof
temperature due to an increased heat exchange by convec-
tion. Thus, the following results were obtained using radia-
tion and convective exchange as the roof thermal boundary
conditions.
Open vs closed cavity
In this section, we compare the HTTC between open
(vented, see Fig. 1) and closed attic. For a 2D analysis, it
is assumed that the openings go through the entire thick-
ness of the walls. In this section, a constant height (vent
size, see Fig. 1) of 5 cm is selected for the openings. Com-
parison is made for a proposed ambient temperature ran-
ge corresponding to a hot summer day. As Fig. 3shows,
a constant HTTC difference between open and closed attic
can be observed for the full ambient temperature range (Ta-
ble 2). This difference is about 43 W/m for the selected ope-
ning height and is an important reduction of heat that does
not enter the house. It can be noted that the HTTC increa-
ses with the ambient temperature for any case. However, for
the highest ambient temperature (315.15 K), the computed
HTTC for the vented attic is 50% lower than that for the
closed attic. The lower the heat transferred through the cei-
ling, the lower the energy needed to maintain a comfortable
indoor temperature.
Air flow streamlines inside the cavity for the open cavity
with h=5 cm can be observed in the second attic from
above in Fig. 4, where an air flow near to the roof running
from inlet to outlet vents can be observed. This flow creates
a convective barrier that carries important amounts of heat
from the attic out to the external ambient, thus reducing the
heat transferred to the ceiling. This socalled flow barrier is
a buoyancy-driven convective flux created by the inlet air in
contact with the roof, being the first cooler than the second.
Berli et al. / Anales AFA - XVI Meeting on Recent Advances of Physics of Fluids and its Applications 46-51 48
FIG. 3: Heat transferred through the ceiling for both closed and
open cavity. Mass flow through the attic is plotted for the open
cavity.
TABLE 2: Impact of ambient temperature.
Ambient
Tempera-
ture (K)
Closed Attic Open Attic
Heat transfer to
the ceiling (MW/m)
Mass flow
(kg/ms)
300.15 55.3 13.4 86.4
305.15 64.6 21.4 82.5
310.15 72.0 30.0 78.0
315.15 80.3 37.3 72.9
Influence of vent (openings) size.
In the previous section, it was shown that the cavity with
openings greatly improves the thermal performance of the
attic. It is expected that the wider the openings, the higher
the air flow rate through the attic, thus optimizing the con-
vective heat exchange with the external environment. In this
section, the attic thermal performance is explored for the fo-
llowing openings heights: 2 cm, 5 cm, 8 cm and 11 cm (see
Fig. 4). The ambient air temperature was set to 305.15 K.
The wider the openings, the greater the convective barrier
near the ceiling and therefore the greater the heat dissipated
convectively to the outside of the cavity, as can be seen in
Fig. 4.
Fig. 5shows that even for a small opening height of 2 cm,
the HTTC is reduced up to 58% respect to the closed cavity,
while for the widest vents (h=11 cm) the reduction exceeds
73%, meaning that the wider the openings the better the
attic thermal performance. This great reduction in HTTC
is generated by the existence of a flow barrier close to the
roof that gets thicker as openings size is increased (see Fig.
4), which transports convective heat to the outside of the
attic, as was explained in the previous section. However, the
HTTC reduction seems to be kept under 80% for h>11 cm,
being this the maximum opening size we recommend.
Influence of roof pitch.
The roof pitch not only involves the possibility of thermal
optimization, but also implies different material costs and
water drainage performances. The minimum recommended
pitch is and this value should not exceed 20º. In this sec-
tion, the thermal performance will be explored for a roof
pitch ranging from to 17.5º. For this purpose, the attic
height at the center of the base (parameter Am in Fig. 1) will
be kept constant at 0.8 m, to keep the attic volume constant.
FIG. 4: Streamlines (black lines) and temperature field for an open
cavity. Legends on the right show temperatures in K.
FIG. 5: Dashed curve: HTTC for every explored opening height.
Solid curve: Percentage reduction of computed HTTC for the open
attics with respect to that of the closed one.
The opening height was set to 5 cm.
Table 3shows that the HTTC increases with roof inclina-
tion. One could expect that the higher the roof inclination
the enhanced buoyancy effect and the greater the mass flow
that carries convective heat out of the attic, but the reason
for HTTC to be increased is that the higher the roof incli-
nation the wider the roof area that is exposed to solar radia-
tion. Although the mass flow increases, the proportion of
heat transferred to the ceiling also increases. In this sense,
a pitch of seems to be the better choice from a thermal
viewpoint, and it is the cheaper choice too. It can also be
noted that for any roof inclination, the vented attic shows a
very high reduction of the HTTC compared to a closed one,
being this reduction about 63% for a pitch of and 67 %
for 17.5º.
Influence of Attic volume.
Attic volume was also explored keeping both the roof
pitch and the opening height constants, with values of θ=
Berli et al. / Anales AFA - XVI Meeting on Recent Advances of Physics of Fluids and its Applications 46-51 49
TABLE 3: Impact of roof pitch.
Roof
Pitch
(º)
Closed Attic Open Attic
Heat transfer to
the ceiling (MW/m)
Mass flow
(kg/ms)
7 40.6 16.2 64.8
10.5 52.6 17.8 69.6
14 64.6 21.4 82.5
17.5 77.0 25.0 92.1
14and h=5 cm, respectively and varying the middle
height of the attic (Am). The inclination was selected by
considering the abundant rains in summer days of our zone.
In this sense, roof drainage is critical.
Table 4shows that the HTTC is reduced as Am increa-
ses, an expected trend because the taller the attic the greater
the air volume that generates a natural insulation between
the roof and the ceiling. For a vented attic, the HTTC re-
duction between the extreme volume sizes is 32 %, while
this reduction is about 23% for a closed one. Even though
these reductions are important, the attic cannot be as tall as
desired due to construction costs. Even for the lowest ven-
ted attic (Am = 0.7 m) the HTTC is 56% lower than that for
the tallest closed one (Am = 1.0 m). When comparing these
results with the ones observed for different vent sizes, we
can conclude that a vented attic needs an optimization of
its openings instead of the roof height to greatly reduce the
HTTC without requiring higher construction costs. Due to
the lack of ventilation, the attic volume is more important
to be higher for a closed cavity.
TABLE 4: Impact of attic middle height.
Attic Middle
Height (m)
Heat transfer to the ceiling (MW/m)
Closed Attic Open Attic
0.7 73.6 23.9
0.8 64.6 21.5
0.9 59.3 19.2
1.0 55.2 17.0
IV. CONCLUSIONS
A low Reynolds number kεturbulent model was nu-
merically solved to analyze the natural ventilation influence
on an open cavity which resembles the attic of a low-cost
family house. The model was validated with experimental
and numerical models, showing good agreements in both
cases. The analysis was focused on studying how simple
construction parameters can reduce the heat transferred to
the interior of the house, i.e., openings sizes, roof pitch and
attic volume. By exploring these parameters, we found that:
The wider the openings (inlet and outlet), the better the
attic thermal performance.
The lower the roof pitch, the lower the HTTC due to a
lower surface exposed to solar radiation.
For a closed attic, the greater its inner volume the bet-
ter the thermal performance. For a vented attic, the vo-
lume is not as important as the openings size to have
an important reduction of the HTTC.
Opening’s existence is the most important feature that
can enhance the attic thermal performance. It was shown
that a vented attic can reduce the HTTC in more than 70%
with respect to a closed one for summer conditions of Santa
Fe (Argentina). As limitations, we can mention that winds,
winter conditions, ceiling insulations and possible 3D ef-
fects of different internal attic geometries were not conside-
red in this work.
FUNDING
This work was supported by Universidad Tecnológica
Nacional through Project AMUTIFE 3457.
REFERENCES
[1] I. C. Council. International Residential Code, R806 Roof
Ventilation 2015.
[2] A. F. Rudd y J. W. Lstiburek. Vented and sealed attics in hot
climates. Transactions-American Society of Heating Refri-
gerating and Air Conditioning Engineers 104, 1199-1210
(1998).
[3] M. E. Berli, J. D. Paolo y F. A. Saita. Heat transfer on a
naturally cross-driven ventilated triangular cavity with ope-
nings. Journal of Physics: Conference Series 166, 012019
(2009).
[4] S. C. Saha y M. Khan. A review of natural convection and
heat transfer in attic-shaped space. Energy and Buildings
43, 2564-2571 (2011).
[5] S. Wang, Z. Shen y L. Gu. Numerical simulation of
buoyancy-driven turbulent ventilation in attic space un-
der winter conditions. Energy and Buildings 47, 360-368
(2012).
[6] S. Wang y Z. Shen. Impacts of Ventilation Ratio and Vent
Balance on Cooling Load and Air Flow of Naturally Ven-
tilated Attics. Energies 5, 3218-3232 (ago. de 2012).https:
//doi.org/10.3390/en5093218.
[7] S. Wang y Z. Shen. Effects of Roof Pitch on Air Flow and
Heating Load of Sealed and Vented Attics for Gable-Roof
Residential Buildings. Sustainability 4, 1999-2021 (2012).
[8] E. Iffa y F. Tariku. Attic baffle size and vent configuration
impacts on attic ventilation. Building and Environment 89,
28-37 (2015).
[9] F. Tariku y E. D. Iffa. Temperature and Air Flow Patterns
in Attic Roofs. Journal of Architectural Engineering 23,
04017014 (sep. de 2017).https://doi.org/10.1061/(asce)ae.
1943-5568.0000261.
[10] C. Ghiaus y L. Roche. Whole life costing of ventilation op-
tions (Earthscan, 2005).
[11] L. Ignat, D. Pelletier y F. Ilinca. A universal formulation
of two-equation models for adaptive computation of turbu-
lent flows. Computer Methods in Applied Mechanics and
Engineering 189, 1119-1139 (2000).
[12] K. Abe, T. Kondoh e Y. Nagano. A new turbulence model
for predicting fluid flow and heat transfer in separating and
reattaching flows—I. Flow field calculations. International
Journal of Heat and Mass Transfer 37, 139-151 (1994).
[13] M. Vázquez, M. Ravachol, F. Chalot y M. Mallet. The
robustness issue on multigrid schemes applied to the Na-
vier–Stokes equations for laminar and turbulent, incom-
pressible and compressible flows. International Journal for
Numerical Methods in Fluids 45, 555-579 (2004).
Berli et al. / Anales AFA - XVI Meeting on Recent Advances of Physics of Fluids and its Applications 46-51 50
[14] F. Ampofo y T. Karayiannis. Experimental benchmark data
for turbulent natural convection in an air filled square ca-
vity. International Journal of Heat and Mass Transfer 46,
3551-3572 (2003).
Berli et al. / Anales AFA - XVI Meeting on Recent Advances of Physics of Fluids and its Applications 46-51 51