You are on page 1of 25

applied

sciences
Article
Analysis of Heat and Smoke Propagation and
Oscillatory Flow through Ceiling Vents in
a Large-Scale Compartment Fire
Claudio Zanzi 1, * , Pablo Gómez 1 , Joaquín López 2 and Julio Hernández 1, *
1 Department de Mecánica, ETSII, Universidad Nacional de Educación a Distancia (UNED),
E-28040 Madrid, Spain
2 Department de Ingeniería Mecánica, Materiales y Fabricación, ETSII, Universidad Politécnica de Cartagena
(UPCT), E-30202 Cartagena, Spain
* Correspondence: czanzi@ind.uned.es (C.Z.); jhernandez@ind.uned.es (J.H.)

Received: 10 July 2019; Accepted: 9 August 2019; Published: 12 August 2019 

Abstract: One question that often arises is whether a specialized code or a more general code
may be equally suitable for fire modeling. This paper investigates the performance and capabilities
of a specialized code (FDS) and a general-purpose code (FLUENT) to simulate a fire in the commercial
area of an underground intermodal transportation station. In order to facilitate a more precise
comparison between the two codes, especially with regard to ventilation issues, the number of factors
that may affect the fire evolution is reduced by simplifying the scenario and the fire model. The codes
are applied to the same fire scenario using a simplified fire model, which considers a source of
mass, heat and species to characterize the fire focus, and whose results are also compared with
those obtained using FDS and a combustion model. An oscillating behavior of the fire-induced
convective heat and mass fluxes through the natural vents is predicted, whose frequency compares
well with experimental results for the ranges of compartment heights and heat release rates considered.
The results obtained with the two codes for the smoke and heat propagation patterns and convective
fluxes through the forced and natural ventilation systems are discussed and compared to each
other. The agreement is very good for the temperature and species concentration distributions
and the overall flow pattern, whereas appreciable discrepancies are only found in the oscillatory
behavior of the fire-induced convective heat and mass fluxes through the natural vents. The relative
performance of the codes in terms of central processing unit (CPU) time consumption is also discussed.

Keywords: fire modeling; compartment fire; heat and smoke propagation; oscillatory flow; large
eddy simulation; FDS; FLUENT

1. Introduction
Performance-based design (PBD) principles are commonly used to draw up fire protection
plans [1]. In light of the multiplicity of ventilation and fire-fighting systems that may be operational,
both theoretical and empirical studies are frequently made (see, e.g., [2]). The results of the
corresponding tests will help in drawing up protection, extinction and evacuation plans, perhaps
even suggest improvements in the infrastructures themselves [3–5]. As may be expected, the main
inconvenience of this approach is the expense involved in full-scale experiments.
Computational tools have contributed much to PBD in the case of confined areas such as stations,
road tunnels, underground carparks shopping centres and stairwells of high-rise buildings [6–10].
Examples of reviews on field and zone approaches, turbulence, combustion, radiation, soot production
and fire modeling in such scenarios can be found in [11,12].

Appl. Sci. 2019, 9, 3305; doi:10.3390/app9163305 www.mdpi.com/journal/applsci


Appl. Sci. 2019, 9, 3305 2 of 25

However, despite the advances made in computational sciences, numerically modeling fires
in buildings is still far from being an easy task because of the multiple combinations of scenarios
and materials involved, the resulting chemical reactions, turbulent transport of smoke and heat, etc.
As a consequence, numerical models are invariably simplified, although even such simplification
still requires vast computational resources because the wide ranges of scales involved and the
multi-dimensional and unsteady nature of the problem cannot be avoided when analyzing complex
real scenarios.
To simulate compartment fires, the computational codes used include Fire Dynamics Simulator
(FDS), FireFOAM, OpenFOAM, PHOENICS, FLUENT and CFX, among others (see, e.g., [7,10,12–17]).
All these codes have been extensively validated for a great variety of fire scenarios in the case of the first
two and for a very large number of applications in the field of fluid mechanics in the case of the others.
Despite the exhaustive validation of the codes, the uncertainties involved in experimental studies of
complex large-scale fires (associated with data collection and the specification of the fire scenario and
detailed experimental conditions), the variety of simplifications introduced in the governing equations
for modeling turbulent transport and combustion processes, and the multiplicity of possible scenarios
where a fire can occur, among other factors, introduce uncertainty in the assessment of the relative
merits of different simulation tools, whether specialized or general-purpose.
In this work we compare the performance and capabilities of a specialized code (FDS) [18,19],
which is a code specifically designed for low-speed fire-driven flows, developed by the National
Institute of Standards and Technology (NIST), and a general-purpose code (FLUENT) [20] for the
numerical simulation of a fire in the commercial area of an underground intermodal transportation
station. Details on the verification and validation of FDS can be found, respectively, in Volumes 2 and
3 of the FDS Technical Reference Guide [21,22]. Both codes, which are among the most frequently used
in the literature to analyze, respectively, fire-related and general thermo-fluid dynamics problems,
have been thoroughly validated against experimental results (see, e.g., [23–25]). However, the above
mentioned uncertainties involved in large-scale tests, along with difficulties in obtaining well resolved
solutions because of practical computer limitations and inappropriate specification of the numerical
and boundary conditions settings, make it often difficult to assess the accuracy and efficiency of the
codes used to simulate a fire in a particular complex facility. Furthermore, fires can involve a wide
variety of scenarios, with different possible burning materials, which represents a huge challenge
for simulation programs. Precisely for this reason, and in order to make a more general comparison
between the codes considered, less dependent on specific sub-models, and more particularly focused
on the oscillatory behavior of the flow through openings, we have tried to simplify as much as possible
some aspects that determine the evolution of the fire. To this end, the scenario and the fire model
are simplified by reducing the number of factors that may affect the fire evolution. More specifically,
a simplified fire model, which considers a source of mass, heat and species to characterize the fire
focus, is used. In order to investigate whether the simplified model can reproduce the essential aspects
of the fire, it has been first validated with a combustion model using the FDS code. Then, the results
for the smoke and heat propagation patterns and convective heat and mass fluxes through the natural
vents, obtained with the two codes using the simplified fire model, are compared and discussed in
detail. The predicted oscillating behavior of the fire-induced convective fluxes through the open vents
is compared with experimental results obtained by other authors. The CPU times consumed by the
two fire models and codes are also compared.

2. Numerical Models
The mass, momentum and internal energy conservation equations, along with appropriate
turbulence and fire models, are solved using the FDS (v. 6.1.2) and FLUENT (v. 6.3.26) codes, which use
finite-differences and finite-volume discretization schemes, respectively, and are second-order accurate.
Details on the discretization schemes can be found in [19,20].
Appl. Sci. 2019, 9, 3305 3 of 25

2.1. Turbulence Model


In a previous work [26] we compared the results obtained using LES (Large Eddy Simulation) and
RANS (Reynolds Averaged Navier Stokes equations) models for a scenario similar to that considered
here, showing the advantage of using LES. Therefore, all the results of the present work were
obtained using LES to simulate turbulence. In FDS, the Deardorff eddy viscosity subgrid-scale
(SGS) model [27] (default choice in version 6.1.2 of the code) was used in most of the simulations.
p
In this model, the turbulent viscosity is expressed as µt = ρCv ∆ ksgs , where Cv = 0.1, the filter
width ∆ = (∆x ∆y ∆z)1/3 and ksgs is the subgrid-scale kinetic energy [18,19]. The turbulent Schmidt
and Prandtl numbers are set equal to the default value of 0.5, although the effects of using other
values will also be investigated. The influence of different SGS models on the numerical results
was assessed by using, as an alternative, the Smagorinsky model with different values of the model
constant Cs [18,19]. In FLUENT, we used a LES model with the default options (Smagorinsky-Lilly
SGS model, with Cs = 0.1).

2.2. Combustion Model


One aspect that can substantially and in many ways determine the evolution of a fire is the type
of burning material. Bearing in mind that it is not the purpose of this paper to focus attention on
complex combustion and soot formation processes, and despite the fact that fires in public spaces
will inevitably involve many different materials, we consider heptane as the burning fuel, as it has
been in other fire simulation studies (see, e.g., [7,25,28]). Although the combustion of this substance
involves a large number of chemical reactions and species, we only consider the production and
transport of those resulting from the reaction C7 H16 + 11 O2 → 7 CO2 + 8 H2 O. Other products such
as CO and soot have not been considered, although the transport of other species (except, for example,
soot, which would require a more complex treatment) could also be easily handled. Not considering
soot production is an additional simplification in the fire model aimed at minimizing the differences
between the FDS and FLUENT settings, so that the comparison between the results of the two codes
can focus primarily on aspects such as those related to the oscillatory flow through open vents.
One of the two fire models considered in this work is the EDC (Eddy Dissipation Concept)
non-premixed combustion model implemented by default in version 6.1.2 of FDS [18,29], based on the
infinitely fast, mixing-limited reaction approximation, in which fuel is locally consumed at a rate which
is proportional to both the mixing rate and the limiting concentration of reactant, and the production
rates of species mass times the respective heats of formation are summed to determine the heat release
per unit volume. The model for the mixing time scale is discussed in [29].

2.3. Simplified Fire Model


As an alternative to the EDC combustion model, we have modeled the fire by directly introducing
the gases generated from combustion, at temperature Tin , through an inlet section of area Ain , with a
mass flow rate ṁin .
The combustion products constitute a mixture that is assumed to behave as a conserved scalar,
whose properties (viscosity, thermal conductivity, mass diffusivity and molecular weight) are the
average of those of its constituent species. The distributions of mass fractions of species are determined
from the conserved scalar concentration and mass fractions Yin,CO2 = 0.1912, Yin,H2 O = 0.0894,
and Yin,N2 = 0.7194 at the fire focus. For a different composition of the combustion gases, the transport
of CO and other species could be carried out in a similar way. Thus, the problem would be reduced
to knowing the amount of CO and other species produced at the fire focus, which, in turn, would
depend on the type of material burned and the ventilation conditions. As indicated above, in order
to be realistic, the complex analysis of the influence of different types of burning materials on the
evolution of the fire has been left outside the scope of this work.
Appl. Sci. 2019, 9, 3305 4 of 25

In the simulations involving both FDS and FLUENT the temperature dependency of the
specific heats of carbon dioxide, water vapor and nitrogen, cip , was expressed by a 4th order
polynomial function (cip ( T ) = ∑4j=0 kij T j ) and, in the case of air, by a 7th order polynomial function
7
(cair air j
p ( T ) = ∑ j=0 k j T ). This is the default option in FDS. The data in the FLUENT and FDS databases
are very similar, although the default option in the former is to set the specific heat of any species
4
as constant. The molecular weight, Min = 28.57 g mol−1 , and specific heat, cin in j
p = ∑ j=0 k j T , of the
mixture are calculated as
N N
Min = ∑ Xin,i Mi ; cin
p (T ) = ∑ Yin,i cip (T ), (1)
i =1 i =1

where N = 3 is the number of species and Xin,i the molar fraction of species i.
The rate of heat convected through the inlet section is calculated as
Z Tin
HRR = ṁin cin
p ( T ) dT, (2)
Tref

where the ambient temperature, Ta = 300 K, is set equal to the reference temperature Tref ,
and ṁin = ρin vin Ain . Here, ρin = p a Min /( Ru Tin ) is the density and vin the velocity of the mixture
of gases at the inlet section (p a is the ambient pressure). It is assumed that the temperature, density,
velocity and composition of the gas entering the domain are time-independent and uniform across the
inlet section. Note that, once heat release rate (HRR) and Ain are fixed, the choice of Tin determines vin
and, thus, the mass flow rate, ṁin . One of the goals of this paper is to determine whether the effects of
turbulent mixing and flame radiation on fire evolution, in regions near or far from the fire focus, are
well reproduced for different combinations of Tin and ṁin values.

2.4. Radiation Model


FDS uses techniques for radiative transport similar to those used in finite volume methods for
convective transport in fluid flow [30]. When using the combustion model, the fraction of the energy
released in the fire that is emitted as thermal radiation may be specified as a source term in the
radiation transport equation through parameter RADIATIVE FRACTION, or calculated by the radiation
model implemented in FDS [19] by setting RADIATIVE FRACTION = 0. We have tested both options,
with a specified fraction of the HRR emitted as thermal radiation of 0.35 in the first case, and the results
obtained did not differ significantly. Therefore, only the results obtained with the second option will
be shown below. In FLUENT, the default P1 radiation model was used [20]. The actual amount of
radiation absorbed or emitted by the hot gases in the simulations we run, which depended on the local
mass fractions of the gas mixture considered and the properties of each gas, is expected to be lower
than that obtained if the production of soot were taken into account.

3. Fire Scenario

3.1. Description of the Facility


In a previous contribution we studied the bus maneuvering and parking zones of the underground
transportation hub “Avenida de América”, in Madrid, Spain. Here, we focus on a fire that starts
in a commercial premises (one of twenty) situated on level −3 (Figure 1a). The different levels are
connected by escalators. The computational domain considered in all the simulations, shown in
Figure 1b, only includes the commercial area at level −3. The overall dimensions of the zone studied
are 66.66 × 39.86 m, with a usable floor area of 2085 m2 . In the geometric models created in FDS and
FLUENT, the x and y axes are horizontal, and the z axis is vertical. The origin of the coordinate system
is located in the S-W corner of the commercial area. It is assumed that the floor and ceiling surfaces are
flat and horizontal and that the height is uniformly 3 m. As mentioned above, the origin of the fire
is a commercial premises in the N-W corner of the shopping area. The danger is greater at this site
Version July 31, 2019 submitted to Appl. Sci. 5 of 24
Appl. Sci. 2019, 9, 3305 5 of 25

because of the proximity to the stairs and consequent risk of spreading vertically and because of the
greater affluence of people, which would hinder evacuation. There are two escalators connecting level
−3 with level −2, and 11 1200 × 600 mm air extraction grilles installed in the ceiling, which form part
of the forced ventilation system.

Floor −2 (passenger waiting area)

Floor 0 (entrance building)

Floor −1 (commercial area)

(a)
Floor −2

Floor − 3 (commercial area)

Fire
focus

N Air extraction grilles


(b) W E
S West East
vent vent
y

z Air extraction grilles


x

FigureFigure 1. (a) Schematic


1. (a) Schematic view
view of the of the transportation
transportation station
station building building
(excluding (excluding
platforms). (b)platforms).
Top view (b) Top
of the view of the
shopping shopping
area area
where the firewhere the fire
originates originates
(level (leveland
−3): natural -3):forced
natural and forced
ventilation ventilation
systems, fire systems,
fire focus
focus and andsystem.
reference reference system.

In a previous work [26], upper levels of the station (Figure 1a) were included in the computational
160 danger is greater at this site because of the proximity to the stairs and consequent risk of spreading
domain in order to study the smoke and heat propagated not only in the commercial area at level −3,
161 vertically and because of the greater affluence of people, which would hinder evacuation. There are
but also in the stairwell that connects levels from −3 to 0 and through the natural ventilation outlets
162 two escalators connecting level -3 with level -2, and 11 1200 × 600 mm air extraction grilles installed
installed at the ceiling of level −1. The evolution of the fire throughout level −3 when the upper levels
163 in the ceiling, which form part of the forced ventilation system.
were included and when the stairwells of level −3 were assumed to be open to the atmosphere did not
164 In a previous work [26], upper levels of the station (Figure 1a) were included in the
show significant variations. Therefore, in order to better focus on the fire spread through level −3 only,
165
where computational domain
the most relevant in were
effects orderobserved,
to study we
the will
smoke and heat
assume propagated
that the commercialnotarea
onlyisin the commercial
directly
area at level -3, but also in the stairwell that connects levels from -3 to 0 and
open to the atmosphere through two outlet sections (see Figure 1b), located 70 cm above the ceiling
166 through the natural
of
167 ventilation
level −3, with outlets
the same installed
areas ofat81.37
the m
ceiling
2 (westofvent)
leveland
-1. 60.06
The evolution of the
m2 (east vent) as fire throughout level -3
the horizontal
pass-through sections of the stairwells located on the W and E sides, respectively. As the aim of theto be open
168 when the upper levels were included and when the stairwells of level -3 were assumed
169 to the atmosphere did not show significant variations. Therefore, in order to better focus on the fire
170 spread through level -3 only, where the most relevant effects were observed, we will assume that the
Appl. Sci. 2019, 9, 3305 6 of 25

present work is to analyze the evolution of the fire using different codes and fire models rather than
evaluating the effectiveness of particular measures to limit the spread of the fire to the top floors of
the building, no smoke barriers were included around the outlets. Note that, since there are not any
additional natural vents or other means for introducing fresh air from outside, the necessary makeup
air will inevitably be drawn through one or both existing vents, as will be shown below.

3.2. Boundary Conditions


Boundary conditions at the fire focus are applied on a square floor surface area Ain (z = 0) of
the premises where the fire focus is located. In most of the simulations presented in this work a total
heat release rate HRR = 2.5 MW is assumed, which lays in the range of typical HRR values (270 to
1200 kW m−2 ) for fires that may occur in shops [31]. When using both the simplified fire model and
the EDC combustion model, we assumed that the indicated HRR value (or the mean value around
which HRR oscillates in the combustion model) is reached almost instantaneously. Although we could
have used a more realistic curve to describe the evolution of HRR, this simplification will facilitate the
comparison between codes (while avoiding the introduction of additional parameters) and will not
hinder the study of the oscillatory flow through the ceiling vents, a process that is usually settled long
before the fire starts to decay, at least for fire scenarios similar to that explored in this paper.
When the combustion model is used, we assume that the burning fuel is heptane, with a heat
of combustion Ẇhept = 44.566 MJ kg−1 [32]. The default values for all other chemical and physical
properties set by the FDS code are assumed. The heptane evaporation speed and fuel burning rate
will not be a part of the solution; instead, the above mentioned HRR value is imposed. As mentioned
before, the parameter RADIATIVE FRACTION is set equal to 0.
In the case of the simplified fire model (Section 2.3), inflow boundary conditions are applied at
the inlet section of the fire focus. A mixture of gases with the composition indicated in Section 2.3
enters the domain with a mass flow rate ṁin , at temperature Tin and velocity vin , yielding a convective
HRR of 2.5 MW. The emissivity of the inlet surface is set to 0 to avoid any additional contribution to
the heat release rate due to thermal radiation.
To keep the thermal boundary conditions as similar as possible for both the combustion and
simplified fire models, zero emissivity and a very small liquid conductivity value at the heptane free
surface are imposed.
To select a suitable value for Ain , the area to which the fire is expected to spread should be
taken into account. A suitable value of the inlet velocity vin must also be selected to ensure a realistic
ratio between inertial and buoyancy effects [33]. However, such parameters are expected to be less
important the further one moves away from the focus of the fire, although their values must be chosen
so that the dimensionless HRR (source Froude number [33]) remains representative of buoyant fires:
q
HRR∗ = HRR/(ρ a c p Ta D2f gD f ), (3)

where subscript a denotes ambient value, c p is the specific heat of air, g is the gravity acceleration and
D f is an equivalent fire focus diameter. In all the simulations we will assume an area of the fire focus
inlet section Ain = 3 m2 , for which the above data provide a value of around 0.42 for HRR∗ , which is
within the range considered normal for buoyant fires [33].
Table 1 shows the details of the four different combinations (labeled as cases G-1 to G-4) of Tin
and ṁin , which yield a convective HRR of 2.5 MW. The temperature range considered is intended to
include the typical average temperatures of the combustion products. The mass flow rate of carbon
dioxide is ṁin,CO2 = Yin,CO2 ṁin .
Appl. Sci. 2019, 9, 3305 7 of 25

Table 1. Conditions simulated with the simplified fire model (cases G-1 to G-4), run with Fire Dynamics
Simulator (FDS) and FLUENT.

Tin ṁin vin ρin


Case [K] [kg s−1 ] [m s−1 ] [kg m−3 ]
G-1 2436 0.8826 2.064 0.1429
G-2 2124 1.052 2.145 0.1639
G-3 1812 1.297 2.256 0.1921
G-4 1500 1.677 2.416 0.2320

In order to reproduce approximately the real operating conditions of the facility, we assume
that the air flow rate extracted through the forced ventilation system on level-3 is constant from the
outset of the fire (40,000 m3 h−1 ) and uniformly distributed among all the extraction grilles, assuming
that the fire does not affect the operation of the extraction units. The indicated value approximately
corresponds to a number of 6 to 7 air changes per hour, which is in accordance with the Spanish
regulations [34].
At the west and east natural vents, which are both assumed to be open during the simulations,
the gauge pressure is set to 0, a condition which is imposed through OPEN and PRESSURE OUTLET
boundary conditions in FDS and FLUENT codes, respectively. Details about the implementation of the
two types of boundary conditions, which are essentially the same, can be found in [19,20]. The results
obtained with FDS were very sensitive to the inadequate implementation of the inflow and outflow
boundary conditions, especially when a grid partition exists in the area closed to the fire focus and the
open vents, and so special care was taken to avoid these situations.
At all the walls, zero heat flux and no-slip boundary conditions are applied, without modeling the
conjugate heat transfer problem. Although the assumption of adiabatic walls might not be considered
sufficiently realistic, such an assumption allows us to reduce the number of factors to be considered
that affect the fire evolution, and thus to carry out a more precise comparison between FDS and
FLUENT performance in the prediction of the basic thermo-fluid dynamic aspects that determine the
fire evolution.

3.3. Computational Grid


Four different hexahedral computational grids of about 1017 k, 3358 k, 7074 k and 19,864 k cells
were used to run the simulations. The grid adaptation capabilities of FLUENT, which facilitate the
meshing of complex geometries and local refinement, were not used so that the grid characteristics
would be similar when comparing FDS. A uniform cell size in all directions was used for all grids in the
region 0 ≤ z ≤ 2 m, whereas a grid refinement in z direction was applied in the vicinity of the ceiling
(2 < z ≤ 3 m) in order to better resolve the flow near the wall. To choose an appropriate order of
magnitude for the size of the grid cells, the range of scales that must be resolved to obtain an acceptable
accuracy in the results was determined following the criterion that the estimated characteristic length
of the fire plume [35],
 2/5
HRR
L= √ , (4)
ρ a c p Ta g
should be spanned by at least ten computational cells [28]. For the general scenario conditions
described above (HRR = 2.5 MW, ρ a = 1.17 kg m−3 , c p = 1011 J kg−1 K−1 and Ta = 300 K), we obtain
L = 1.380 m. In the region 2 < z ≤ 3 m, the cell size in z direction, ∆z, gradually shrinks to ∆zBL at
the ceiling (z = 3 m) in order to ensure suitable values of the parameter y+ . Table 2 shows, for each
computational grid, the total number of cells, the approximate cell size in the region 0 ≤ z ≤ 2 m,
δ = ∆x = ∆y = ∆z, the approximate number of grid cells, L/δ, spanning the characteristic length L,
and the grid size in z direction at the ceiling, ∆zBL .
Appl. Sci. 2019, 9, 3305 8 of 25

Table 2. Grid characteristics: total number of cells, approximate cell size (δ), approximate number
of cells (L/δ) spanning the characteristic length L, and grid size (∆zBL ) in z direction at the ceiling
(z = 3 m).

Grid Total # of Cells δ [mm] L/δ ∆zBL [mm]


I 1,017,000 220 6 37
II 3,358,000 138 10 33
III 7,074,000 100 14 30
IV 19,864,000 74 18.5 25

Most of the results presented in the following section were obtained using grid II (L/δ ≈ 10),
whereas grids I (coarsest), III and IV (finest) were used for testing grid independence.
FDS uses an explicit scheme in time, and automatically adjusts the time step ∆t required for
numerical stability during the simulation, so that the following condition is met:
 
|u| |v| |w|
CFL = ∆t max ; ; < 1. (5)
∆x ∆y ∆z

The default maximum and minimum CFL numbers used by FDS are, respectively, 1 and 0.8. Mean
values of ∆t = 2.1 and 1.8 ms were obtained from Equation (5) when using the simplified fire model
and the EDC combustion model, respectively.
On the other hand, a fully-implicit scheme was used for time advancement in FLUENT, and no
stability criterion was taken into account to calculate ∆t. Instead, we used a constant time step
∆t = 50 ms during the simulations. All residuals diminished by at least two orders of magnitude after
around 40 to 50 iterations per time step, and then remained almost constant. The scaled residuals
defined in FLUENT [20] for the mass, species, three components of momentum, and energy and
radiation equations decreased to around 10−2 , 10−3 , 10−4 and 10−6 , respectively.
It should be pointed out that, for the same scenario and equivalent simulation settings, FLUENT
is considerably more time consuming than FDS, and obtaining strictly time step-independent solutions
may require substantially higher CPU times. In fact, reducing the time step used with FLUENT
to obtain the results presented in the next section would still produce small changes in the results,
although not be very significant for the purposes of comparison with FDS.

4. Results and Discussion


This section looks at the results obtained from the simulation of the scenario described in
Section 3.2 using the FDS and FLUENT codes with the simplified fire model and the default
EDC combustion model implemented in FDS. The results obtained are the temperature and CO2
concentration contours in horizontal and vertical sections of the domain, heat flow rates and CO2
mass flow rates extracted from the ventilation systems, and mass flow rates leaving and entering the
compartment through the ceiling vents.
First, we discuss the sensitivity of the results obtained with FDS to the grid size and different
settings of the LES turbulence model. Then, we analyze the predictions of FDS with the simplified
fire model, considering four combinations of the inlet temperature, Tin , and mass flow rate, ṁin , of the
mixture of gases introduced in the computational domain (cases G-1 to G-4 of Table 1). The results
obtained for the oscillatory flow through vents are compared with experimental results available in the
literature. Then, we compare the results for the fire evolution obtained with the simplified fire model
(Case G-1 of Table 1) and the EDC combustion model (with RADIATIVE FRACTION = 0) using the FDS
code. Finally, the results obtained with the simplified fire model using the FDS and FLUENT codes are
compared to each other.
Except in Section 4.1, the respective default settings for the LES model were used in each code,
and a computational grid with L/δ ≈ 10 was used in all the simulations described below.
Appl. Sci. 2019, 9, 3305 9 of 25

4.1. Preliminary Tests


Several tests were carried out to investigate the grid independence of the FDS results using the
four computational grids of Table 2. Using the simplified fire model with Tin = 1500 K (Case G-4 in
Table 1), results for CO2 mass flow rate extracted by the ventilation systems, shown in Figure 2a, can
be considered
Version July 31, 2019fairly grid to
submitted independent
Appl. Sci. even for the relatively coarse grid II, as well as results for heat 9 of 24
flow rate and temperature and CO2 mass fraction contours on vertical and horizontal sections of the
domain (not shown here).

Grid I - L/δ ≈ 6
Grid II - L/δ ≈ 10
Grid III - L/δ ≈ 14
Grid IV - L/δ ≈ 18.5
CO2 mass flow rate [kg s−1 ]

0.20 West vent

(a) 0.15

0.10

0.05
Forced extraction grilles
0.00
0 20 40 60 80 100 120 140
20
West vent
Mass flow rate [kg s−1 ]

10

0
Net

(b)
-10
East vent

-20

0 20 40 60 80 100 120 140

Time [s ]
Figure 2. (a) Mass flow rates of CO2 through
Figure 2. (a) Mass flow rates of CO2 through
the ventilation systems and (b) mass flow rates through
the ventilation systems and (b) mass flow rates through the
thewest
westandand east
east vents
vents (positive
(positive outwards)
outwards) and
and net netflow
mass mass flow
rate rate the
leaving leaving the compartment,
compartment, obtained
obtained with
with FDS, the simplified fire model with
FDS, the simplified fire model with Tin = 1500T 1500 K (Case G-4 in Table 1) and grids with
in K (Case G-4 in Table 1) and grids with L/δ ≈ 6, 10, 14L/δ ≈
=
6, 10,
and1418.5.
and 18.5.

Figure 2b shows the mass flow rate (positive outwards) through the west and east natural vents
270 the FDS code. Finally, the results obtained with the simplified fire model using the FDS and FLUENT
and the net mass flow rate leaving the compartment.
271 codes are compared to each other.
The results depicted in Figure 2b show an acceptable degree of grid independence for grids II
272 Except
and in Subsection
finer, especially 4.1,ofthe
in terms therespective default
net mass flow settings for
rate. However, notethe LES
that the model were
oscillating used
mass flowin each
273 code, and
rates a computational
through grid
the open vents arewith
veryL/δ ≈ 10 to
sensitive was
theused in alland
grid size, thedifferences
simulations described
remain below.
between the
results obtained with grids III and IV. Indeed, the dependence of these unsteady results on grid size
274 4.1.could
Preliminary tests
not be completely avoided unless an extremely fine grid were used. However, the temperature
275
and CO2 mass
Several testsfraction contours,
were carried outthe
to heat and CO2the
investigate mass
gridflow rates throughof
independence the forced
the FDS and natural
results using the
276
ventilation systems (Figure 2a) and the net mass flow rate leaving the compartment (Figure
four computational grids of Table 2. Using the simplified fire model with Tin = 1500 K (Case 2b) do notG-4 in
appreciably change when using grids with L/δ ≈ 10, 14 or 18.5, and, therefore, the computational grid
277 Table 1), results for CO2 mass flow rate extracted by the ventilation systems, shown in Figure 2a, can
278 be considered fairly grid independent even for the relatively coarse grid II, as well as results for heat
279 flow rate and temperature and CO2 mass fraction contours on vertical and horizontal sections of the
280 domain (not shown here). Figure 2b shows the mass flow rate (positive outwards) through the west
281 and east natural vents and the net mass flow rate leaving the compartment.
Appl. Sci. 2019, 9, 3305 10 of 25

with L/δ ≈ 10 (grid II) can be considered sufficiently fine to reach grid independency for the most
relevant results. As for the FLUENT code, it was shown in [26] that, for the current scenario, a grid
with L/δ ≈ 10 is also fine enough to reach grid independency for the most relevant results. However,
the comparison between Figure 2a and the equivalent results shown in [26] shows that FLUENT, when
using LES, is more sensitive to the grid size than FDS in predicting, for example, the CO2 mass flow
rate through the forced extraction grilles.
The Deardorff and Smagorinsky SGS models are the default options for LES models in FDS and
FLUENT, respectively [19,20]. We ran simulations with FDS, using both SGS models, two values
for the coefficient Cs of the Smagorinsky model (0.2 and 0.1, the latter being the default value in
FLUENT) and two combinations for the turbulent Prandtl and Schmidt numbers (Prt |Sct = 0.5|0.5 and
0.85|0.7, which are the default values in FDS and FLUENT, respectively). As in the grid independence
test, the results obtained for the mass flow rates through the west and east vents (not shown here)
appear to be the most sensitive ones to the settings used for the SGS model (differences for other
quantities such as the mass flow rate through the forced extraction grilles and net mass flow rate
leaving the compartment are much smaller). Moreover, differences are significant only between the
results obtained with FDS and FLUENT, whereas using different settings for the SGS model in FDS
does not affect the results substantially.

4.2. Results of the Simplified Fire Model


Figure 3 shows the results for the heat flow rate through the ventilation systems, obtained with
Version Julyusing
FDS the
31, 2019 simplified
submitted fire model
to Appl. Sci. with the four combinations of Tin and ṁin of cases G-1 to G-4
10 of 24
(Table 1).

Tin = 2436 K Tin = 1812 K


Tin = 2124 K Tin = 1500 K

1.6
West vent
Heat flow rate [MW ]

1.2

0.8

0.4
Forced extraction grilles

0.0
0 20 40 60 80 100

Time [s ]
Figure 3. Heat
Figure flowflow
3. Heat raterate
through thethe
through forced
forcedextraction
extraction grilles andwest
grilles and westvent,
vent, predicted
predicted by FDS
by FDS andand
the the
simplified fire model with T = 2436, 2124, 1812 and 1500 K (cases G-1 to G-4 in Table
simplified fire model withinTin = 2436, 2124, 1812 and 1500 K (cases G-1 to G-4 in Table 1). 1).

Note that, despite the broad range of inlet temperatures considered (2436–1500 K), maintaining
289 not appreciably change when using grids with L/δ ≈ 10, 14 or 18.5, and, therefore, the computational
the same convective heat release rate (calculated from Equation (2)) in all the cases yields results that
290 grid with L/δ ≈ 10 (grid II) can be considered sufficiently fine to reach grid independency for the
compare very well with each other. The results for the normalized CO2 mass flow rate (ṁCO2 /ṁin,CO2 ,
291 mostwhere
relevant results.
ṁin,CO is the As
massforflow
therate
FLUENT code, it was
of CO2 introduced shown
in the domainin [26]
at thethat, for the
fire focus) arecurrent scenario,
qualitatively
2
292 a grid with
very L/δ to
similar 10 isfor
≈those also
thefine
heatenough
flow ratetoshown
reachingrid independency
Figure 3. Although not for shown
the most relevant
here, results.
the results
293 However,
for thethe
mass comparison between
flow rate extracted byFigure 2a and
the forced the equivalent
ventilation system and results
net mass shown in [26]
flow rate shows
leaving the that
294 FLUENT, when using LES, is more sensitive
compartment are practically independent of Tin . to the grid size than FDS in predicting, for example, the
295 CO2 massTable
flow3 rate
shows the fractions
through of the heat
the forced and COgrilles.
extraction 2 mass introduced in the domain from the beginning
296 The Deardorff and Smagorinsky SGS models aresystems
of the fire that had been extracted by the ventilation at t =options
the default 90 and 720 fors,LES models in FDS and
297 FLUENT, respectively [19,20]. We ran 1 simulations
Z t with FDS,1using Z tboth SGS models, two values for
Ih (t) =
the coefficient Cs of the Smagorinsky model q̇ dt; Ic (t) = latter being ṁ 2 dt.default value in FLUENT) (6)
298
(HRR )t 0 (0.2 and 0.1, the ṁin,CO2 t 0 COthe
299 and two combinations for the turbulent Prandtl and Schmidt numbers (Prt |Sct = 0.5|0.5 and 0.85|0.7,
300 which are the default values in FDS and FLUENT, respectively). As in the grid independence test, the
301 results obtained for the mass flow rates through the west and east vents (not shown here) appear to
302 be the most sensitive ones to the settings used for the SGS model (differences for other quantities
303 such as the mass flow rate through the forced extraction grilles and net mass flow rate leaving
Appl. Sci. 2019, 9, 3305 11 of 25

Note the low dependence of Ih and Ic on the inlet temperature Tin .


Figure 4 shows the temperature contours at t = 90 s on horizontal planes at heights z = 2.95
and 2.05 m, and at t = 720 s and z = 2.95 m, predicted by FDS using the simplified fire model with
Tin = 2436 and 1500 K. It can be observed that, sufficiently far from the fire focus, the temperature
distributions obtained using different Tin values compare very well, behavior that is consistent with
the very slight dependence of the results of Figure 3 and Table 3 on Tin . A similar behavior is observed
for the distributions of the normalized CO2 concentration and velocity magnitude, which are not
shown for brevity.

Table 3. Fractions (in %) of heat (Ih ) and mass of CO2 (Ic ) extracted by the natural and forced
ventilation systems (Equation (6)). In parentheses, variation (in %) with respect to Case G-1. Values
obtained with FDS and the simplified fire model after the first 90 s (cases G-1 to G-4) and 720 s (cases
G-1 and G-4) of the fire.

West Vent Forced Extraction Grilles


Case Ih (∆) Ic (∆) Ih (∆) Ic (∆)
t = 90 s
G-1 38.90 40.16 17.76 19.19
G-2 38.87 (−0.1) 40.01 (−0.4) 17.77 (+0.0) 19.22 (+0.2)
G-3 39.20 (+0.8) 40.07 (−0.2) 17.46 (−1.7) 18.70 (−2.6)
G-4 39.37 (+1.2) 40.12 (−0.1) 17.31 (−2.6) 18.54 (−3.4)

t = 720 s
G-1 61.20 58.36 28.70 30.18
G-4 60.67 (−0.9) 58.28 (−0.1) 28.06 (−2.2) 29.86 (−1.0)

4.2.1. Oscillatory Flow through Ceiling Vents


Figure 5a shows, for Case G-4 (Tin = 1500 K), the FDS results for the mass flow rates leaving
and entering the compartment through the west vent (ṁw,out and ṁw,in , respectively), the net mass
flow rate through this vent (ṁw,out − ṁw,in ) and the mass flow rate entering the compartment through
the east vent (ṁe,in ). During the whole simulation, the outward mass flow rate through the east vent
is negligible compared to the inward one. On the other hand, hot gases flow outward through the
north side of the west vent (closer to the fire focus; see Figure 1b) and cool air at ambient conditions
(Ta = 300 K) flows inward through the south side. The bidirectional flow predicted through the
west vent is widely described in the literature (see, for example, [36–38]). In the problem considered,
whether the flow across the vents is induced by buoyancy or by a fire-generated pressure difference is
determined by the Richardson number, which can be expressed as [38,39]

Ri = Gr/Re2 = gD∆ρ/∆p,

where Gr and Re are the Grashof and Reynolds numbers, D is the vent size, ∆ρ is the variation of
density across the vent and ∆p is the pressure increment due to thermal expansion, calculated as
 2
1 HRR
∆p = ,
2ρe c p Te ACd

where ρe and Te are the density and temperature in the vicinity of the vent, Cd is a discharge coefficient
and A is the area of the vent. In the scenario of Figure 5a, using the mean values of density and
temperature obtained from the simulation at t = 45 s over the west vent (of area A = 11.38 × 7.15 m2 )
yields ∆p p ≈ 6 mPa, ∆ρe ≈ 30 g m−3 and Ri ≈ 500, a value indicating that the flow is dominated
Appl. Sci. 2019, 9, 3305 12 of 25

by buoyancy. As the mean temperature around the vent increases, Ri reaches values above 103 for
t > 135 s.
It can be observed from Figure 5a that the inward and outward mass flow rates across the west
vent oscillate. Note that the outward mass flow rate oscillation wave has a slight and approximately
constant delay of about 10 s with respect to the inward flow wave. Also note that the resulting
oscillating net mass flow rate across the west vent determines the oscillating behavior of the inward
flow through the east vent, with the same frequency (of about 0.03 Hz, calculated over a simulation of
750 s) and approximately the same wave shape and time-dependent amplitude. This behavior was
Version July 31, 2019 submitted to Appl. Sci. 12 of 24
observed for all the simulations run.

250 100
228 93
205 85
183 78
161 71
(a) 138 (c) 64
116 56
94 49
72 42
49 34
27 27
250 100
228 93
205 85
183 78
161 71
(b) 138 (d) 64
116 56
94 49
72 42
49 34
27 27

t = 90 s, z = 2.95 m t = 90 s, z = 2.05 m

250
228
205
183
161
(e) 138
116
94
72
49
27
250
228
205
183
161
(f) 138
116
94
72
49
27

t = 720 s, z = 2.95 m

Figure 4. Temperature
Figure 4. Temperature contours (◦ C)
contours ( C) atinstant
◦ at 90 ss and
instant tt = 90 andzz==2.95
2.95and
and2.05
2.05
mm (top
(top pictures),
pictures), andand
at at
t =t720
= 720 s and z = 2.95 m (bottom pictures). Results obtained with FDS, a grid with L/δ ≈ 10 and the
s and z = 2.95 m (bottom pictures). Results obtained with FDS, a grid with L/δ ≈ 10 and
simplified fire model
the simplified with (a)
fire model = T2436
Tin(a)
with in = K and
2436 K,(b) = 1500
(b)TTinin = 1500 K
K,(cases
(c) Tin G-1 andK,
= 2436 G-4
(d)inTin
Table 1). K,
= 1500
(e) Tin = 2436 K and (f) Tin = 1500 K (cases G-1 and G-4 in Table 1).
332 It can be observed from Figure 5a that the inward and outward mass flow rates across the west
Figure 5a also shows the results for the evolution of the CO2 mass flow rate through the west vent
333 ventdivided
oscillate. Note that the outward mass flow rate oscillation wave has a slight and approximately
by the constant C = 7 × 10−3 (triangle symbols). The approximate coincidence between these
334 constant delay
results and the of outward
about 10mass
s with
flowrespect to thethe
rate through inward flowmeans
west vent wave.thatAlso note that theofresulting
the concentration CO2
335 oscillating
in the hotnet mass
gases flow rate
leaving across the west
the compartment vent constant,
is fairly determinesandthe oscillating
varies behavior of 7the
from approximately inward
× 10 −3
336 flow through
during the the
firsteast
two vent, with
minutes thethe
after same frequency
instant at which(ofthe
about
gases0.03 Hz,tocalculated
begin overt =
flow (around a simulation
10 s) to
337 of 750 s) and
− 3 approximately the same wave shape
8.5 × 10 once the simulation reaches a steady state. and time-dependent amplitude. This behavior was
338 observed for all the simulations run.
339 Figure 5a also shows the results for the evolution of the CO2 mass flow rate through the west vent
340 divided by the constant C = 7 × 10−3 (triangle symbols). The approximate coincidence between these
341 results and the outward mass flow rate through the west vent means that the concentration of CO2
342 in the hot gases leaving the compartment is fairly constant, and varies from approximately 7 × 10−3
Appl. Sci. 2019, 9, 3305 13 of 25
Version July 31, 2019 submitted to Appl. Sci. 13 of 24

West vent East vent


Outside
Compartment
ṁw,out ṁw,in ṁe,in

50
West vent, ṁ w,out West vent, ṁCO2 /C
40 West vent, ṁ w,in
Mass flow rate [kg s−1 ] West vent, net
30 East vent, net (≈ ṁ e,in )

20

10
(a)
0

-10

-20

-30
0 20 40 60 80 100 120 140 160 180 200

1500 K 2436 K
50 West vent, ṁ w,out
West vent, ṁ w,in
Mass flow rate [kg s−1 ]

40
West vent, net
30
East vent, net (≈ ṁ e,in )

20

10
(b)
0

-10

-20

-30
0 25 50 75 100 125 150 175 200 225 250 275 300

Time [s ]
Figure
Figure 5. (a)Mass
5. (a) Massflowflowrates
ratesthrough
throughthe thewest
westand andeast
eastvents obtained
vents obtainedwith FDS
with forfor
FDS Case G-4G-4
Case (Tin =
1500 −3 .
(Tin =K).
1500Triangle symbols:
K). Triangle COCO
symbols: 2 mass
2 massflow
flowrate
ratethrough
through the westvent
the west ventdivided
divided
bybyC= C 7= 7 ×310
× 10 − .
(b) Comparison
(b) betweenFDS
Comparison between FDSresults
resultsfor
for the
the mass
mass flow
flow rates
rates through
through thethe west
west andand
easteast vents
vents for cases
for cases
G-1
G-1 and G-4
G-4 (T
(Tinin = 2436and
=2436 and1500
1500K).K).

Figure 5b compares the results of Figure 5a, for Tin = 1500 K, with those obtained with Tin =
351 with Tin = 2436 K. Therefore, the influence of Tin on the overall oscillatory behavior of the flow at the
2436 K. Note that the frequency of the mass flow rate oscillations is only slightly smaller and their
352 open vents is apparently small.
amplitude remains more constant with time for the higher Tin value. The absolute values of the net
353
mass Comparison
4.2.2. flow rates across
withthe west and east vents obtained with both Tin values are in phase to each
experiments
other and oscillate with very similar amplitudes. Also note that the mentioned delay of about 10 s
The the
between dependence of the
outward and frequency
inward of therate
mass flow oscillating flow
oscillation through
waves the open
also exists vents
in the wasobtained
results found to be
well
withdescribed
Tin = 2436by
K.expressions of influence
Therefore, the the form of Tin on the overall oscillatory behavior of the flow at the
open vents is apparently small.
f = k(HRR)n (7)
4.2.2. Comparison with Experiments
(see [37] and references therein for a review of results available in the literature). From experimental
The dependence of the frequency of the oscillating flow through the open vents was found to be
and numerical results it was shown in [40,41] that the flow oscillation frequency in a compartment
well described by expressions of the form
fire with two open vents is proportional to (HRR)1/3 . A similar problem but for a compartment with
a single horizontal ceiling vent was studied f =numerically
k (HRR)n in [42], where the flow oscillation frequency (7)
was found to be proportional to (HRR) 0.29 for heat release rates below a critical value. Figure 6a
(see [37]
shows and
the references
numerical therein
results forforthe
a review of results
oscillation available
frequency as ain the literature).
function Fromrelease
of the heat experimental
rate (in the
and numerical results it was shown in [40,41] that the flow oscillation frequency in a compartment fire
Appl. Sci. 2019, 9, 3305 14 of 25

with two open vents is proportional to (HRR)1/3 . A similar problem but for a compartment with a
single horizontal ceiling vent was studied numerically in [42], where the flow oscillation frequency was
found to be proportional to (HRR)0.29 for heat release rates below a critical value. Figure 6a shows the
numerical results for the oscillation frequency as a function of the heat release rate (in the interval 500
to 2500 kW) obtained in the present work using FDS and the simplified fire model, with Tin = 1500 K
and L/δ ≈ 10. The figure also shows the correlation

f = 1.562 × 10−4 (HRR)0.345 , (8)

which fits the FDS results with R2 = 0.9922, along with correlations of the form of Equation (7),
with the values of the exponent n proposed in [40–42]. Note that the results for the oscillation
frequency obtained in the
Version July present
31, 2019 worktoexhibit
submitted Appl. Sci.a dependence on the HRR that is very close to that of 14 of 24
the experimental results of [40,41].

4
Present FDS results
Correlation f = k(HRR) n
This work, Equation (7), n = 0.345
3 Refs. [40,41] (exp. & num.), n = 1/3
Frequency [Hz ]

Ref. [42] (numerical), n = 0.29

-2
(a) 2x10

0.01 3 4 5 6 7 8 9 2 3 4
6
10

HRR [W ]

1.0

Present FDS results


0.8 Experiments [40,41]

0.6
f H3

(b) 0.4

0.2

0.0
0.0 0.2 0.4 0.6 0.8 1.0

k(HRR)1/3

Figure 6. (a) Oscillation frequency


Figure 6. (a) of the
Oscillation fire-induced
frequency of theflow through the
fire-induced flowopen ventsthe
through as open
a function
ventsofas a function of
heat release rate (HRR).
HRR. Comparison
Comparison of present
of the the present numerical
numerical results
results withwith correlations
correlations of the
of the formform of
of Equation (7) with
Equation (7) with values
values of the
of the exponent
exponent n deduced
n deduced from
from experimental
experimental and and numerical
numerical results
results [40–42].(b) Collapse of
[40–42].
3 3= k (HRR)1/3 −3−
(b) Collapse of FDS and experimental
experimentalresults
resultsononthe
thecorrelation
correlationf H fH = k(HRR)1/3with withk= k=4.97
4.97
××1010 . 3.

Figure 6b shows a comparison between the present numerical results and the experimental results
interval 500 to 2500 kW) obtained in the present work using FDS and the simplified fire model, with
of [40,41], obtained in a cubic enclosure of side 0.6 m for thermal powers of 1 and 8 kW, for which the
Tin = 1500 K and L/δ ≈ 10. The figure also shows the correlation

f = 1.562 × 10−4 (HRR)0.345, (8)

354 which fits the FDS results with R2 = 0.9922, along with correlations of the form of Equation (7), with
355 the values of the exponent n proposed in [40,41] and [42]. Note that the results for the oscillation
Appl. Sci. 2019, 9, 3305 15 of 25

measured frequencies were 0.3 and 0.6 Hz, respectively. Although the experimental results available
for comparison are scarce, it is interesting to point out that numerical and experimental results collapse
very well after scaling on the curve f H 3 = k(HRR)1/3 , despite the wide ranges of the compartment
height, H, and HRR values considered.

4.3. Combustion Model vs. Simplified Fire Models


Figure 7 shows the HRR as a function of time in FDS simulations that use the simplified fire
model (cases G-1 to G-4) and the EDC combustion model. In the first case, the energy enters the
domain at a constant rate, whereas in the second one the HRR exhibits high frequency fluctuations,
with an amplitude that approximately varies in the range 1.0 to 1.5 MW. As will be shown below,
theseJuly
Version fluctuations produce
31, 2019 submitted high-frequency
to Appl. Sci. fluctuations of small amplitude in the heat and mass 15
flow
of 24
rates through the ventilation systems, especially through the west vent.
4.0
Simplified fire model
3.5 Combustion model
HRR [MW]

3.0

2.5

2.0

1.5
0 10 20 30 40 50 60 70 80 90
Version July 31, 2019 submitted to Appl. Sci. Time [s ] 15 of 24

Figure7.7.Heat
Figure Heatrelease
releaserate
rate(HRR)
(HRR) in
in the
the simulations cases G-1
simulations of cases G-1 to
toG-4
G-4(Table 1),compared
(Table1), comparedwith
withthat
that
ofofthe
theEDC
EDCcombustion
combustionmodel.
model.
4.0
Simplified fire model
100 3.5 Combustion model
80
HRR [MW]

Figure 8 shows the temperature and


93 CO2 mass fraction72
64
contours
at instant t = 90 s and vertical
85 3.0
planes at x = 11, 17, 23, 29 and 36 m, obtained with the56 combustion model and the simplified fire
78
71 2.5 48
model (Case(a)G-1). Figure 9 shows the temperature
64 (a)and40 velocity magnitude contours at t = 90 s,
56 32
on the horizontal plane at z =2.02.95 m, obtained with FDS24 using the EDC combustion model and the
49
42 16
simplified fire model (Case G-1).
34
1.5
0
Figure
10
10 shows
20 30 40
the50CO8 2 mass fraction contours at t = 90 s, on the
60 70 80 90
0
horizontal planes at z = 2.95 m (left column) and z = 2.05
27
m (right column), obtained as in Figure 9.
100 Time [s ] 80
It can be observed93from the figures the very good overall 72agreement between the three types of results
85 64
7. Heat
Figurewith
obtained the78release rate (HRR)
combustion and in the simulations
simplified models.of cases
56 G-1 to G-4 (Table 1), compared with that

of the EDC combustion


71 model. 48
(b) 64 (b) 40
56
100 8032
49
93 7224
42
85 6416
34
78 568
27
71 480
(a) 64 40
56 Temperature (◦ C) 32 YCO2 × 104
49 24

Figure 8. Temperature contours (◦ C) (left column) and CO mass fraction contours (YCO2 × 104 ) (right
42 16
8 2
34
column) at instant t = 90 s and x = 11, 17, 23, 29 and 36 m, obtained with FDS using a grid with
27 0

100 80
L/δ ≈ 10. (a) Combustion
93 model with RADIATIVE FRACTION 72 = 0 and (b) simplified fire model with
85 64
Tin = 2436 K (Case
78
G-1). 56
71 48
(b) 64 40
32
364 4.3. Combustion model vs. simplified fire models
56
49 24
42 16

365 Figure 7 shows the HRR as a function of time in FDS simulations that use the simplified fire
34
27
8
0

366 model (cases G-1 to G-4) and the EDC(◦ C)


Temperature combustion model. In the Yfirst case,
CO2 × 10
4 the energy enters the
367 domain at a constant rate, whereas◦ in the second one the HRR exhibits high frequency4 fluctuations,
Figure8.8. Temperature
Figure Temperaturecontours
contours ((◦C)
C)(left
(leftcolumn)
column) and
andCO
CO22 mass
massfraction
fractioncontours
contours (Y
(YCO × 10
CO22 × 104))(right
(right
368 with an amplitude
column)
thatt approximately varies in29the range 1.0 to 1.5 with
MW.FDS Asusing
will be shown below,
column) at instant t = 90 s and x = 11, 17, 23, 29 and 36 m, obtained with FDS using a grid with
at instant = 90 s and x = 11, 17, 23, and 36 m, obtained a grid with
369 these L/δ
fluctuations produce high-frequency fluctuations of small amplitude in the heat and mass flow
L/δ ≈ ≈ 10. (a) Combustion
10. (a) Combustion model
model with
with RADIATIVE
RADIATIVE FRACTION
FRACTION = = 00 and (b) simplified
and (b) simplified fire
fire model
model with
with
370 rates through
TTin = 2436the ventilation
K (Case G-1). systems, especially through the west vent.
in = 2436 K (Case G-1).
371 Figure 8 shows the temperature and CO2 mass fraction contours at instant t = 90 s and vertical
372
364 planes at x = 11,
4.3. Combustion 17, 23,
model vs. 29 and 36fire
simplified m,models
obtained with the combustion model and the simplified fire
373 model (Case G-1). Figure 9 shows the temperature and velocity magnitude contours at t = 90 s, on
365 Figure 7 shows the HRR as a function of time in FDS simulations that use the simplified fire
374 the horizontal plane at z = 2.95 m, obtained with FDS using the EDC combustion model and the
366 model (cases G-1 to G-4) and the EDC combustion model. In the first case, the energy enters the
375 simplified fire model (Case G-1). Figure 10 shows the CO2 mass fraction contours at t = 90 s, on the
367 domain at a constant rate, whereas in the second one the HRR exhibits high frequency fluctuations,
Appl. Sci. 2019, 9, 3305 16 of 25
Version July 31, 2019 submitted to Appl. Sci. 16 of 24

Version July 31, 2019 submitted to Appl. Sci. 16 of 24


250 3.50
228 3.15
205 2.80
250
183 3.50
2.45
228
161 3.15
2.10
(a) 205
138 2.80
1.75
183
116 2.45
1.40
161
94 2.10
1.05
(a) 138
72 (a) 1.75
0.70
116
49 1.40
0.35
94
27 1.05
0.00
72 0.70
250 3.50
49 0.35
228 3.15
27 0.00
205 2.80
250
183 3.50
2.45
228
161 3.15
2.10
(b) 205
138 2.80
1.75
183
116 2.45
1.40
161
94 2.10
1.05
(b) 138
72 (b) 1.75
0.70
116
49 1.40
0.35
94
27 1.05
0.00
72 0.70
49 Temperature (◦ C) 0.35 Velocity magnitude (m s−1 )
27 0.00

Figure
Figure 9. Temperature contours
9. Temperature contours (left
(left(◦column)
Temperature column)
C) and
and velocity
velocity magnitude
magnitude contours
contours(m(right
Velocity magnitude s−1 ) column)
(right column) at
at
tt = 90 s, on the horizontal plane at z = 2.95 m, obtained as in Figure 8.
= 90 s, on the horizontal plane at z = 2.95 m, obtained as in Figure 8. (a) Combustion model with
Figure 9. Temperature contours (left column) and velocity magnitude contours (right column) at
RADIATIVE FRACTION
150
= 0 and (b) simplified fire model with
t = 90 s, on the135horizontal plane at z = 2.95 m, obtained as
30
Tin = 2436 K (Case G-1).
27 in Figure 8.
120 24
150
105 30
21
135
90 27
18
(a) 120
75 (a) 24
15
105
60 21
12
90
45 18
9
(a) 75
30 15
6
60
15 12
3
45
0 90
30 6
150 30
15 3
135 27
0 0
120 24
150
105 30
21
135
90 27
18
(b) 120
75 (b) 24
15
105
60 21
12
90
45 18
9
(b) 75
30 15
6
60
15 12
3
45
0 90
30 6
15 z = 2.95 m 3 z = 2.05 m
0 0
Figure 10. Contours of CO2 mass fraction (YCO2 × 104 ) at t = 90 s and heights z = 2.95 m (left column)
z = 2.95 m z = 2.05 m
and z = 2.05 m (right column), obtained as in Figure4 8.
Figure 10.Contours
Figure10. Contours of of
COCO 2 mass fraction (YCO2 × 10×) 10
2 mass fraction (Y CO2
at4t) =at90
t s=and
90 heights z = 2.95z m=(left
s and heights 2.95column)
m (left
and z =
column) 2.05
and mz (right
= column),
2.05 m obtained
(right as
column), in Figure
obtained8. as in Figure 8. (a) Combustion model with
385 through the west vent obtained with FDS and the two fire models. This can be seen with more detail
386
RADIATIVE FRACTION = 0 and (b) simplified fire model with T = 2436 K (Case G-1).
in Figure 12, which compares results similar to those presented in Figure 5b, obtained with FDS using
in
385 through the west vent obtained with FDS and the two fire models. This can be seen with more detail
387 the combustion model and the simplified fire model. Note that the outward mass flow rate through
386 Figure
in Figure 12, 11 shows
which the results
compares forsimilar
results the mass of CO
to those 2 and heat
presented flow rates
in Figure throughwith
5b, obtained the ventilation
FDS using
388 the west vent predicted with the combustion model oscillates with a slightly smaller amplitude than
387 systems
the obtained
combustion withand
model FDStheusing the simplified
simplified fire model.fire Note
model (Case
that G-1) andmass
the outward the combustion model.
flow rate through
389 when using the simplified fire model, an effect that is also observed in the net mass flow rate through
388 Note
the thatvent
west the combustion
predicted withmodel
the predicts
combustionmassmodel
of COoscillates
2 flow rates through
with the forced
a slightly smallerventilation
amplitudegrilles
than
390 the west vent and, as a consequence, in the inward mass flow rate through the east vent.
389 slightly
when larger
using thethan those obtained
simplified fire model, withan the simplified
effect fireobserved
that is also model, a inbehavior that could
the net mass be explained
flow rate through
391 Table 4 shows the relative CPU time consumed by FDS using the EDC combustion model and the
390 by the
the westhigher CO2 as
vent and, concentrations
a consequence, at the ceiling
in the in the
inward vicinity
mass flowof thethrough
rate forced extraction grilles, as shown
the east vent.
392 simplified fire model. Note that the combustion model consumes a CPU time around 40% larger than
391 in Figure
Table 10. Alsothe
4 shows note the very
relative CPUgood
time degree
consumed of agreement
by FDS using between
the EDC thecombustion
results for model
the flow andrates
the
393 that consumed when the simplified fire model is used, with no appreciable change in the results, as
392 through the
simplified firewest ventNote
model. obtained with
that the FDS and the
combustion two consumes
model fire models. This time
a CPU can be seen with
around 40% more
largerdetail
than
394 shown above. Using the default value of 0.35 instead of zero for the parameter RADIATIVE FRACTION
393 in Figure
that 12, which
consumed when compares resultsfire
the simplified similar
model to those presented
is used, with noin Figure 5b, obtained
appreciable change inwith FDS using
the results, as
395 reduces the CPU time consumed by the combustion model by about 7%, with very little difference in
394 the combustion
shown above. Usingmodeltheand the simplified
default value of 0.35fireinstead
model.of Note
zerothat
for the outward
parametermass flow rate
RADIATIVE through
FRACTION
396 the overall results.
395 the westthe
reduces ventCPUpredicted with the by
time consumed combustion modelmodel
the combustion oscillates with a7%,
by about slightly
with smaller amplitude
very little difference than
in
396
397
when
the
4.4. FDSusing
overall the simplified fire model, an effect that is also observed in the net mass flow rate through
vs.results.
FLUENT
the west vent and, as a consequence, in the inward mass flow rate through the east vent.
398
397 Figures
4.4. FDS 13 to 18 show comparisons between temperature, velocity magnitude and CO2 mass
vs. FLUENT
399 fraction contours obtained with FDS and FLUENT at several vertical and horizontal planes and two
Figures 13 to 18 show comparisons between temperature, velocity magnitude and CO mass
different instants. It can be observed from the figures that the overall degree of agreement 2is very
398
400
399 fraction contours obtained with FDS and FLUENT at several vertical and horizontal planes and two
401 good. The most noticeable differences between FDS and FLUENT results are systematically observed
400 different instants. It can be observed from the figures that the overall degree of agreement is very
401 good. The most noticeable differences between FDS and FLUENT results are systematically observed
Appl. Sci.July
Version 3305 submitted to Appl. Sci.
9, 2019
2019,31, 17 of 25
17 of 24
Version July 31, 2019 submitted to Appl. Sci. 17 of 24

Simplified fire model (Case G-1)


Simplified fire modelmodel
EDC combustion (Case G-1)
EDC combustion model

CO2 mass flow rate [kg s−1 ]


0.14 West vent

CO2 mass flow rate [kg s−1 ]


0.14
0.12 West vent
0.12
0.10
0.10
(a) 0.08
(a) 0.08
0.06
0.06
0.04
Forced extraction grilles
0.04
0.02 Forced extraction grilles
0.02
0.00
0.00 0 100 200 300 400
0
2.00 100 200 300 400
West vent
Heat flow rate [MW]

2.00
West vent
Heat flow rate [MW]

1.60
1.60
1.20
1.20
(b) 0.80
(b) 0.80
0.40 Forced extraction grilles
0.40 Forced extraction grilles
0.00
0.00 0 100 200 300 400
0 100 200 300 400
Time [s ]
Time [s ]
Figure11.
11.(a) Massflow
(a)Mass flow rates
rates ofof
CO CO 2 , and
(b)(b) heat flow rates through the ventilation systems, obtained
Figure 2 , and heat flow rates through the ventilation systems, obtained
Figure 11.
withFDS (a) Mass flow
FDSasasininFigure
Figure rates of CO 2 , and (b) heat flow rates through the ventilation systems, obtained
with 8. 8.
with FDS as in Figure 8.
West vent East vent
West vent OutsideEast vent
Outside
Compartment
ṁw,out Compartment
ṁw,in ṁe,in
ṁw,out ṁw,in ṁe,in

Simplified model Combust. model


50 West vent, ṁ w,out Simplified model Combust. model
50 West
West vent,
vent, ṁ w,in
ṁ w,out
Mass flow rate [kg s−1 ]

40
West vent,
West ṁ w,in
vent, net
Mass flow rate [kg s−1 ]

40
30
East
West vent,
vent, net net (≈ ṁ e,in )
30
East vent, net (≈ ṁ e,in )
20
20
10
10
0
0
-10
-10
-20
-20
-30
0 20 40 60 80 100 120 140 160 180 200 220 240
-30
0 20 40 60 80 100 120 140 160 180 200 220 240
Time [s ]
Time [s ]
Figure 12. Results similar to those of Figure 5b, but obtained as in Figure 8.
Figure 12. Results
Figure12. Results similar to those
similar to thoseof
ofFigure
Figure5b,
5b,but
butobtained
obtainedasas
in in Figure
Figure 8. 8.
402 in the distribution of the CO2 concentration in the vicinity of the ceiling, as can be observed from
402
403
the Table
inFigures 4 shows the
distribution
17 and 18,ofand therelative CPU time consumed
CO2 concentration
pictures
by FDS using
in the vicinity
at the right column of Figuresof thethe EDC combustion
ceiling,
13 and
model and
14. as can be observed from
the simplified
Figures 17inand fireand
18, model. Note at
pictures that the combustion model consumes a14.
CPU time around 40% larger
403
404 As previous sections, thethe right
most column
sensitive of
resultsFigures
to the13useand
of FDS or FLUENT codes are related
404
thanAsthat
in consumed
previous when
sections,the simplified
the most fire model
sensitive is
resultsused,
to with
the use no
ofappreciable
FDS or changecodes
FLUENT in theare
results,
related
405 to the flow rates through the ventilation systems, and, therefore, we will discuss in the following
405 toas shown
the flow above.
rates Using
throughthe default
the value
ventilationof 0.35 instead
systems, of
and, zero for the
therefore, parameter
we will RADIATIVE
discuss in FRACTION
the following
406 the main
reduces
differences consumed
the CPU time
found in the by
results for thesemodel
the combustion
quantities when
by about 7%,
using
with
the
very
two codes.
little
Figures
difference
19a
in19a
406
407
the main differences
andoverall
19b compare found in the results for these quantities when using the two
the results obtained with FDS and FLUENT for the evolution of the CO2 mass codes. Figures
the
and results.the
407
408 and heat flow rates,results
19b compare obtained
respectively, with FDS
through and FLUENT
the open vents andforforced
the evolution
extractionofgrilles,
the COusing
2 mass the
408
409
and heat flow
simplified firerates,
model respectively,
with Tin = through the open
1500 K (Case G-4).vents and forced
The degree extraction
of agreement grilles, the
between using
FDSthe
and
409
410
simplified
FLUENTfire model
results with good,
is very 1500 K (Case
Tin = especially G-4).
for the heatThe degree
flow of Note
rates. agreement betweenpredicts
that FLUENT the FDSvalues
and
410 FLUENT results is very good, especially for the heat flow rates. Note that FLUENT predicts values
Appl. Sci. 2019, 9, 3305 18 of 25

Table 4. CPU time (relative to Case G-1) in FDS simulations.

Case Fire Model Relative CPU Time


G-1 Gas release (Tin = 2436 K) 1
G-2 Gas release (Tin = 2124 K) 1.019
G-3 Gas release (Tin = 1812 K) 1.023
G-4 Gas release (Tin = 1500 K) 1.021
Combustion model (rad. frac.: 0) 1.462

4.4. FDS vs. FLUENT


Figures 13–18 show comparisons between temperature, velocity magnitude and CO2 mass fraction
contours obtained with FDS and FLUENT at several vertical and horizontal planes and two different
instants. It can be observed from the figures that the overall degree of agreement is very good. The most
noticeable differences between FDS and FLUENT results are systematically observed in the distribution
of the CO2 concentration in the vicinity of the ceiling, as can be observed from Figures 17 and 18,
Version July 31, 2019
and pictures submitted
at the to Appl. Sci.
right column of Figures 13 and 14. 18 of 24
Version July 31, 2019 submitted to Appl. Sci. 18 of 24

100 80
100
93 80
72
93
85 72
64
85
78 64
56
78
71 56
48
(a) 71
64 48
40
(a) 64
56 (a) 40
32
56
49 32
24
49
42 24
16
42
34 16
8
34
27 80
27 370.0 0 0.0
100 80
370.0 0.0
100
93 80
72
93
85 72
64
85
78 64
56
78
71 56
48
(b) 71
64 48
40
(b) 64
56 (b) 40
32
56
49 32
24
49
42 24
16
42
34 16
8
34
27 80
27 0
Temperature (◦ C) YCO × 104
Temperature (◦ C) YCO22 × 104
Temperature contours ◦ C) (left column) and CO mass fraction (Y 4 ) (right column)
Figure 13.
Figure 13.
Figure 13. Temperature contours (((◦◦C)
Temperature contours C) (left
(left column)
column) and
and CO CO × 10
2 mass fraction contours
CO22 mass (Y4CO
fraction (YCO22 × 10
4 ) (right
× 10column)
) (right
2
at instant
column) t = 90 s and x = 11, 17, 23, 29 and 36 m obtained using a grid with L/δ ≈ 10, the simplified
at instant at
t =instant t =x 90
90 s and s and
= 11, x =
17, 23, 29 11,
and17,3623,
m 29 and 36using
obtained m obtained using
a grid with a grid
L/δ ≈ 10,with L/δ ≈ 10,
the simplified
fire
the model withfire
simplified = 2436with
Tinmodel K (Case
T = G-1)
2436 andK (a) FDS
(Case and
G-1) and FLUENT
(b)(a) FDS andcodes.
(b) FLUENT codes.
in
fire model with Tin = 2436 K (Case G-1) and (a) FDS and (b) FLUENT codes.
100 80
100
93 80
72
93
85 72
64
85
78 64
56
78
71 56
48
(a) 71
64 (a) 48
40
(a) 64
56 40
32
56
49 32
24
49
42 24
16
42
34 16
8
34
27 80
27 370.0 0 0.0
100 80
370.0 0.0
100
93 80
72
93
85 72
64
85
78 64
56
78
71 56
48
(b) 71
64 (b) 48
40
(b) 64
56 40
32
56
49 32
24
49
42 24
16
42
34 16
8
34
27 80
27 0
Temperature (◦ C) YCO × 104
Temperature (◦ C) YCO22 × 104
Figure 14. Temperature contours (◦◦ C) (left column) and CO2 mass fraction (YCO2 × 1044 ) (right column)
Figure
Figure
at 14.tTemperature
14.
instant Temperature
=
contours ( C)
240 s and x contours
= 11, 17, 23,
(left
(◦ C) column)
29 (left
and 36column)andand
CO2CO
m, obtained
mass fraction
as 2inmass
(Y 2contours
fraction
Figure 13. CO
× 10 ) (right
(YCOcolumn)
(a) FDS; (b) FLUENT.
2
× 104 )
at instant
(right t = 240
column) at sinstant
and x =t =11,
24017,s23,
and29xand 36 17,
= 11, m, 23,
obtained
29 andas36inm,
Figure
obtained asFDS;
13. (a) (b) FLUENT.
in Figure 13. (a) FDS;
(b) FLUENT.
for the CO mass flow rates extracted through the forced extraction grilles and through the west vent
for the CO22 mass flow rates extracted through the forced extraction grilles and through the west vent
411
411
412 that are higher and lower, respectively, than those predicted by FDS, which can be attributed to the
412 that are higher and lower, respectively, than those predicted by FDS, which can be attributed to the
higher CO concentrations at the ceiling in the vicinity of the forced extraction grilles, as shown in
higher CO22 concentrations at the ceiling in the vicinity of the forced extraction grilles, as shown in
413
413
414 Figures 17 and 18. However, despite the differences in the values predicted by FLUENT and FDS for
414 Figures 17 and 18. However, despite the differences in the values predicted by FLUENT and FDS for
415 the mass of CO2 extracted over 750 s through the forced (83.27 vs. 73.20 kg) and natural (134.5 vs.
415 the mass of CO2 extracted over 750 s through the forced (83.27 vs. 73.20 kg) and natural (134.5 vs.
Appl. Sci. 2019, 9, 3305 19 of 25
Version July 31, 2019 submitted to Appl. Sci. 19 of 24
Version July 31, 2019 submitted to Appl. Sci. 19 of 24
Version July 31, 2019 submitted to Appl. Sci. 19 of 24
250 3.50
228
250 3.15
3.50
250
205
228 3.50
2.80
3.15
228
183
205 3.15
2.45
2.80
205
161
183 2.80
2.10
2.45
(a) 183
138
161 2.45
1.75
2.10
(a) 161
116
138 (a) 2.10
1.40
1.75
(a) 138
94
116 (a) 1.75
1.05
1.40
116
72
94 1.40
0.70
1.05
94
49
72 1.05
0.35
0.70
72
27
49 0.70
0.00
0.35
49
27 0.35
0.00
250 3.50
0.00
27
228
250 3.15
3.50
250
205
228 3.50
2.80
3.15
228
183
205 3.15
2.45
2.80
205
161
183 2.80
2.10
2.45
(b) 183
138
161 2.45
1.75
2.10
(b) 161
116
138 (b) 2.10
1.40
1.75
(b) 138
94
116 (b) 1.75
1.05
1.40
116
72
94 1.40
0.70
1.05
94
49
72 1.05
0.35
0.70
72
27
49 0.70
0.00
0.35
49
27 0.35
0.00
27 Temperature (◦ C) 0.00 Velocity magnitude (m s−1 )
Temperature (◦ C) Velocity magnitude (m s−1 )
Temperature (◦ C) Velocity magnitude (m s−1 )
Figure
Figure 15. Temperature contours
15. Temperature contours (left
(left column)
column) and
and velocity
velocity magnitude
magnitude contours
contours (right
(right column)
column) at
at
Figure 15. Temperature contours (left column) and velocity magnitude contours (right column) at
tt = 90
Figure s, on
15. the horizontal
Temperature plane
contoursat z
(left
= 2.95 m,
column) obtained
and as
velocityin Figure 13.
magnitude (a) FDS; (b)
contours FLUENT.
(right
= 90 s, on the horizontal plane at z = 2.95 m, obtained as in Figure 13. (a) FDS; (b) FLUENT. column) at
t = 90 s, on the horizontal plane at z = 2.95 m, obtained as in Figure 13. (a) FDS; (b) FLUENT.
t = 90 s, on the250horizontal plane at z = 2.95 m, obtained as 3.50in Figure 13. (a) FDS; (b) FLUENT.
228
250 3.15
3.50
250
205
228 3.50
2.80
3.15
228
183
205 3.15
2.45
2.80
205
161
183 2.80
2.10
2.45
(a) 183
138
161 (a) 2.45
1.75
2.10
2.10
(a) 161
116
138 1.40
1.75
(a) 138
94
116 (a) 1.75
1.05
1.40
116
72
94 1.40
0.70
1.05
94
49
72 1.05
0.35
0.70
72
27
49 0.70
0.00
0.35
49
27 0.35
0.00
250 3.50
0.00
27
228
250 3.15
3.50
250
205
228 3.50
2.80
3.15
228
183
205 3.15
2.45
2.80
205
161
183 2.80
2.10
2.45
(b) 183
138
161 (b) 2.45
1.75
2.10
2.10
(b) 161
116
138 1.40
1.75
(b) 138
94
116 (b) 1.75
1.05
1.40
116
72
94 1.40
0.70
1.05
94
49
72 1.05
0.35
0.70
72
27
49 0.70
0.00
0.35
49
27 0.35
0.00
27 Temperature (◦ C) 0.00 Velocity magnitude (m s−1 )
Temperature (◦ C) Velocity magnitude (m s−1 )
Temperature (◦ C) Velocity magnitude (m s−1 )
Figure 16. Temperature contours (left column) and velocity magnitude contours (right column) at
Figure 16. Temperature contours
contours (left column)
column) and velocity
and velocity magnitude
velocity magnitude contours
magnitude contours (right
contours (right column)
(right column) at
column) at
at
tFigure
Figure 16.
= 24016.
s, obtained as in Figure
Temperature 13. (left
contours (a) FDS; (b) FLUENT.
column) and
tt =
= 240
240 s,
s, obtained
obtained as
as in
in Figure
Figure 13.
13. (a)
(a) FDS;
FDS; (b)
(b) FLUENT.
FLUENT.
t = 240 s, obtained
150 as in Figure 13. (a) FDS; (b) FLUENT. 30
135
150 27
30
150
120
135 30
24
27
135
105
120 27
21
24
120
90
105 2418
21
(a) 105
75
90 (a) 2115
18
(a) 90
60
75 (a) 1812
15
(a) 75
45
60 15
912
60
30
45 12
69
45
15
30 936
30
015 603
15
0 30
0150 030
135
150 27
30
150
120
135 30
24
27
135
105
120 27
21
24
120
90
105 2418
21
(b) 105
75
90 (b) 2115
18
(b) 90
60
75 (b) 1812
15
(b) 75
45
60 15
912
60
30
45 12
69
45
15
30 936
30
015 603
15
0 30
0 z = 2.95 m 0 z = 2.05 m
z = 2.95 m z = 2.05 m
z 2.95 m 4 z = 2.05
Figure 17. Contours of CO2 mass fraction (YCO2 × 10 ) at t = 90 s and heights
= z=m 2.95 m (left column)
Figure 17. Contours of CO2 mass fraction (YCO2 × 1044 ) at t = 90 s and heights z = 2.95 m (left column)
and 2.05
z =17.
Figure m (rightof
Contours
Contours column),
CO
CO22massobtained
fractionas
fraction (Yin
COFigure
× 10 )13. =FDS;
at t(a) (b) FLUENT.
90 s and heights z = 2.95 m (left column)
2 × 104 ) at t = 90 s and heights z = 2.95 m (left column)
Figure
and z =17.
2.05 m (rightofcolumn),
massobtained as(Yin
COFigure
2 13. (a) FDS; (b) FLUENT.
and z = 2.05 m (right column), obtained as in Figure 13. (a) FDS; (b)
and z = 2.05 m (right column), obtained as in Figure 13. (a) FDS; (b) FLUENT. FLUENT.
419 a higher CO2 concentration near the ceiling, resulting in a higher CO2 mass flow rate through the
419 a higher CO2 concentration near the ceiling, resulting in a higher CO2 mass flow rate through the
419
420 aforced
higher CO2 concentration
ventilation grilles andnear theextraction
a lower ceiling, resulting in a higher
rate through CO
the west 2 mass
vent. flow
Also noterate
thatthrough the
the slightly
420 forced ventilation grilles and a lower extraction rate through the west vent. Also note that the slightly
420 forced ventilation grilles and a lower extraction rate through the west vent. Also note that the slightly
Appl. Sci. 2019, 9, 3305 20 of 25
Version July 31, 2019 submitted to Appl. Sci. 20 of 24

150 30
135 27
120 24
105 21
90 18
(a) 75 15
60 12
45 9
30 6
15 3
0 0

150 30
135 27
120 24
105 21
90 18
(b) 75 15
60 12
45 9
30 6
15 3
0 0

z = 2.95 m z = 2.05 m
Figure
Figure 18. ContoursofofCO
18. Contours COmass
mass fraction
fraction (YCO
(YCO ×210 ) at t at
= t240×4104 )
= s240
andsheights
and heights z =
z = 2.95 2.95 column)
m (left m (left
22 2
column) and z = 2.05 m (right column), obtained as in Figure 13. (a) FDS;
and z = 2.05 m (right column), obtained as in Figure 13. (a) FDS; (b) FLUENT. (b) FLUENT.

421 lowerAs in previous


outward masssections,
flow ratethe most sensitive
through the westresults to the useby
vent predicted ofFLUENT,
FDS or FLUENT as will be codes
shownare related
below,
422 to thealso
may flow rates through
contribute the ventilation
to explain systems, and, therefore, we will discuss in the following the
the differences.
423 mainFigure
differences found inthe
19c compares theFDS
results
and for these quantities
FLUENT results for when
the netusingmassthe flow tworatescodes.
throughFigurethe 19a,b
west
424 compare
and the results
east vents, and obtained
the net mass with flow
FDS andrate FLUENT
leaving the for compartment,
the evolution ofobtainedthe CO2with massthe andsimplified
heat flow
425 rates,
fire respectively,
model with Tinthrough
= 1500the K open
(Casevents
G-4).andTheforced extraction
agreement betweengrilles,
bothusing
types theof simplified
results for firethe
model
net
426 with T
mass flow in = 1500 K (Case G-4). The degree of agreement between the
rate leaving the compartment is very good. Note that approximately 12 minutes after FDS and FLUENT results is very
427 good,
the especiallyoffor
beginning thethefire
heat theflow
netrates.
massNoteflowthatrateFLUENT
leaving predicts values for vanishes,
the compartment the CO2 mass andflow rates
the flow
428 extracted
pattern in through the forcedbegins
the compartment extraction grilles around
to oscillate and through a steadythe state.
west vent
On the thatother
are higher
hand, the andresults
lower,
429 respectively,
obtained withthan
FDSthose
showpredicted by FDS,
larger outward andwhich
inward canmass
be attributed
flow ratestothrough
the higher CO2 concentrations
the west and east vents,
430 at the ceilingwith
respectively, in the vicinity
greater of the forced
oscillation extraction
amplitudes. Both grilles,
codesas shownthat
predict in Figures
the mass17flow andrates
18. However,
through
431 despite the differences in the values predicted by FLUENT and FDS
both natural vents oscillate with a frequency of about 0.03 Hz for HRR = 2.5 MW. The 2results for for the mass of CO extracted
432 overmass
the 750 sflow
through the forced
rate through the(83.27
forced vs.extraction
73.20 kg) grilles
and natural
obtained(134.5 vs.both
with 143.6codeskg) ventilation
(not shown systems,
in the
433 the difference
paper) compareinvery the total
well. CO2 mass extracted over that time period is less than 1%. The reasons for
434 theseFigure
relatively low discrepancies
20 shows the comparison are not completely
presented clear. 19c,
in Figure Notebut that the more
with higherdetailSchmidt and number used
for a shorter
435 by default
period in FLUENT
of time. Besideswould the netproduce
mass flow a lower
rates,diffusion
Figure 20and also therefore
compares a higher CO2 concentration
the outward and inward
436 near the
mass flow ceiling, resultingthe
rates through in awest
higher
vent.CONote
2 mass flow
the rate
fairly through
good the
agreement forced
betweenventilation
the grilles
results forand
thea
437 lower extraction
outward mass flow raterate
through
throughthe west vent.vent
the west Alsoobtained
note thatwiththe slightly
FDS and lower
FLUENT. outward Thismass flow that
implies rate
438 through
the largerthenetwest
mass vent
flowpredicted by FLUENT,
rate through the west as vent
will be shown with
obtained below, FDSmay is also
mainly contribute
due to to explain
a smaller
439 the differences.
inward mass flow rate. This, in turn, determines the larger inward mass flow rate through the east
440 Figure 19cby
vent predicted compares
FDS. the FDS and FLUENT results for the net mass flow rates through the west
441 and Despite
east vents, theand the net mass
discrepancies flow rateabove,
mentioned leaving thethe compartment,
overall agreementobtained
betweenwith FDS the andsimplified
FLUENT
442 fire model with
is reasonably good. T in The higher discrepancies between the results of the two codes reportedthe
= 1500 K (Case G-4). The agreement between both types of results for net
in the
443 mass flowfor
literature rate leaving
some the compartment
fire simulations, whichisdo very
notgood.
seem to Note that approximately
be justified when similar 12mathematical
minutes after fire the
444 beginning
models areof the firemight
solved, the net in mass
part beflow rate leaving
attributed the compartment
to underresolved vanishes,
numerical and theinflow
solutions space pattern
and/or in
445 the compartment
time obtained with begins
FLUENT,to oscillate
whicharound
requiresa steady state. Onmore
substantially the other hand, the results
computational resourcesobtained
than FDSwith
446 FDSa show
for given larger outward and
grid resolution. Theinward mass flow rates
main disadvantage of through
FLUENTthe overwestFDS,andateastleastvents,
for therespectively,
scenarios
447 with greater
simulated andoscillation
code versionsamplitudes.
used inBoththiscodes
work,predict
is the that the mass higher
substantially flow rates CPU through both natural
times required (as
448 vents as
much oscillate
twice or with
even a frequency
more) when ofstandard
about 0.03 Hz forare
settings HRRused,= a2.5 MW. The
drawback thatresults
is notfor the mass flow
accompanied by
449 rate throughdifferences
appreciable the forced in extraction grillespredictions.
the numerical obtained with both codes
However, (not shown
differences in CPU in the
time paper) compare
consumption
450 very well.
should be taken with care since changes in certain settings may produce substantial variations. For
451 example, using the PISO algorithm instead of SIMPLEC for solving the coupling between mass and
452 momentum conservation equations through pressure increases CPU time consumption in FLUENT
453 by a factor of the order of 20%. Choosing different convergence criteria and numerical settings may
454 also result in appreciable variations in CPU time consumption. On the other hand, FLUENT offers
Appl.
Version Sci.31,
July 2019, 9, 3305
2019 submitted to Appl. Sci. 21 of 25 21 of 24

FDS FLUENT

CO2 mass flow rate [kg s−1 ]


0.25 West vent

0.20

0.15
(a)
0.10
Forced extraction grilles
0.05

0.00
0 200 400 600 800
2.00
Heat flow rate [MW]

1.60

1.20 West vent

0.80
(b)
0.40 Forced extraction grilles

0.00
0 200 400 600 800
20 West vent
Mass flow rate [kg s−1 ]

10

0
Net

-10
(c)
East vent
-20

-30
0 200 400 600 800

Time [s ]
Figure 19. (a) Mass flow rates of CO2 , heat
Figure 19. (a) Mass flow rates of CO2 , (b)
(b) heat flow rates, and (c) mass flow rates (including the
flow rates, and (c) mass flow rates (including the net
net mass flow rate leaving the compartment) through the ventilation
mass flow rate leaving the compartment) through the ventilation systemssystems
obtainedobtained
with FDSwith
and FDS and
FLUENT
FLUENTusing
usingthe
thesimplified firemodel
simplified fire modelwith
with
TinT= = 1500
in 1500 K (Case
K (Case G-4). G-4).

Figure 20 shows the comparison presented in Figure 19c, but with more detail and for a shorter
455 much more flexibility in grid generation for arbitrary geometries and postprocessing of results, and
period of time. Besides the net mass flow rates, Figure 20 also compares the outward and inward mass
456 makes
flowsimulation
rates through settheup easier
west vent.than
Note FDS, which
the fairly goodrequires
agreementa between
more careful specification
the results of boundary
for the outward
457 conditions
mass flowand rategrid partitions.
through the west It should
vent alsowith
obtained be FDS
pointed out thatThis
and FLUENT. more recent
implies thatreleases of ANSYS
the larger
458 FLUENT [43] are likely to improve the performance of the code, specially in terms of the amount of
net mass flow rate through the west vent obtained with FDS is mainly due to a smaller inward mass
459 CPUflow
timerate. This, in turn, determines the larger inward mass flow rate through the east vent predicted
consumed.
by FDS.
460 5. Conclusions
461 The merits of a specialized and a general-purpose code to simulate a fire in the commercial
462 area of an underground intermodal transportation station are investigated using a simplified fire
463 model, whose results have been previously compared with those of an EDC combustion model. The
464 following conclusions can be drawn from the results presented above.

465 1. The simplified fire model yielded results that, far enough away from the fire focus, are in
466 excellent agreement with those obtained using an EDC combustion model, while requiring
Appl. Sci. 2019, 9, 3305 22 of 25
Version July 31, 2019 submitted to Appl. Sci. 22 of 24

West vent East vent


Outside
Compartment
ṁw,out ṁw,in ṁe,in

FDS FLUENT
50 West vent, ṁ w,out
West vent, ṁ w,in
40
West vent, net
Mass flow rate [kg s−1 ]

30
East vent, net (≈ ṁ e,in )

20

10

-10

-20

-30
0 25 50 75 100 125 150 175 200 225 250 275 300

Time [s ]
Figure 20. Same results as in Figure 5b, but obtained with FDS and FLUENT for Case G-4 (Tin =
Figure 20. Same results as in Figure 5b, but obtained with FDS and FLUENT for Case G-4
1500
(Tin K).
= 1500 K).

468 Despite the


simplified discrepancies
fire model are fairlymentioned above, the
independent ofoverall agreement
the inlet velocitybetween FDS and FLUENT
and temperature of the hot
469 is reasonably good. The
gases introduced higher
at the discrepancies between the results of the two codes reported in the
fire focus.
470 2.literature for some
The results fire simulations,
obtained with FLUENT which do not
using theseem to be justified
simplified fire modelwhen similar
show mathematical
a reasonable degree
471 fireof
models are solved,
agreement might obtained
with those in part bewith
attributed to underresolved
FDS, except regarding the numerical solutions
mass flow rates in
ofspace
cold air
472
and/or time through
entering obtainedthe with FLUENT,
natural ventswhich
and CO requires substantially more computational resources
2 mass flow rates through the west vent and forced
473
than FDS for a given grid resolution. The main disadvantage
ventilation grilles, for which some quantitative differences were of FLUENT over FDS, at least for
observed.
474 3.the The
scenarios
features simulated
that areandmost code versions
sensitive used in
to using thethis work, is
different the substantially
codes, fire models, higher CPU times
computational grids
475
required (as much as twice or even more) when standard settings are used,
and turbulence model settings are those describing the fire-induced oscillatory flow through a drawback that is
476
notthe
accompanied by appreciable differences in the numerical predictions. However,
open vents, particularly the mass flow rate of cool air entering the domain. To the best differences in
CPU time consumption should be taken with care since changes in certain settings may produce
477 of our knowledge, neither FDS nor FLUENT have previously been used to predict the type of
substantial variations. For example, using the PISO algorithm instead of SIMPLEC for solving the
478 fire-induced oscillatory flows through openings described in this paper, about which there is
coupling between mass and momentum conservation equations through pressure increases CPU time
479 relatively little information in the literature.
consumption in FLUENT by a factor of the order of 20%. Choosing different convergence criteria
480 4. The dependence of the oscillation frequency on the HRR was in reasonable agreement with the
and numerical settings may also result in appreciable variations in CPU time consumption. On the
481 experimental and numerical results reported in the literature.
other hand, FLUENT offers much more flexibility in grid generation for arbitrary geometries and
482 5. For the fire scenarios considered and versions of the codes used in this work, the main advantage
postprocessing of results, and makes simulation set up easier than FDS, which requires a more careful
483 of FDS over FLUENT was found to be its substantially lower CPU time consumption. On the
specification of boundary conditions and grid partitions. It should also be pointed out that more recent
484 other hand, FLUENT allows much more flexibility in grid generation for arbitrary geometries,
releases of ANSYS FLUENT [43] are likely to improve the performance of the code, specially in terms
485 andamount
of the the simulation
of CPU time set up can be done more easily in FLUENT than in FDS.
consumed.
486 The results presented in this paper provide insight into the relative merits of specialized and
5. Conclusions
487 general-purpose codes and different fire models, which are frequently used to assess fire safety levels
The merits of a specialized and a general-purpose code to simulate a fire in the commercial area
488 in buildings in order to reduce risk to human life in the case of fire. The future work based on the
of an underground intermodal transportation station are investigated using a simplified fire model,
489 presented results may be of help in devising fire safety solutions, aimed at limiting the zones of
whose results have been previously compared with those of an EDC combustion model. The following
490 possible escape routes where people survival could be compromised in the event of a fire.
conclusions can be drawn from the results presented above.
491 Author Contributions: Conceptualization, C.Z., P.G., J.L. and J.H.; Investigation, C.Z., J.L.; Methodology, C.Z.,
492 P.G.
1. and J.H.;
The Visualization,
simplified J.L.; Supervision,
fire model J.H.;that,
yielded results Writing–original draft,from
far enough away C.Z. the
andfire
J.H.focus, are in excellent
493 agreement withThe
Acknowledgments: those obtained
support of using an EDC combustion
the Consorcio Regional demodel, while requiring
Transportes de Madrid appreciably
is gratefully
494 acknowledged.
lower CPU times. It was also found that, for a given HRR, the results of the simplified fire model
495 Conflicts of Interest: The authors declare no conflict of interest.

496 References
Appl. Sci. 2019, 9, 3305 23 of 25

are fairly independent of the inlet velocity and temperature of the hot gases introduced at the
fire focus.
2. The results obtained with FLUENT using the simplified fire model show a reasonable degree of
agreement with those obtained with FDS, except regarding the mass flow rates of cold air entering
through the natural vents and CO2 mass flow rates through the west vent and forced ventilation
grilles, for which some quantitative differences were observed.
3. The features that are most sensitive to using the different codes, fire models, computational grids
and turbulence model settings are those describing the fire-induced oscillatory flow through
the open vents, particularly the mass flow rate of cool air entering the domain. To the best
of our knowledge, neither FDS nor FLUENT have previously been used to predict the type of
fire-induced oscillatory flows through openings described in this paper, about which there is
relatively little information in the literature.
4. The dependence of the oscillation frequency on the HRR was in reasonable agreement with the
experimental and numerical results reported in the literature.
5. For the fire scenarios considered and versions of the codes used in this work, the main advantage
of FDS over FLUENT was found to be its substantially lower CPU time consumption. On the
other hand, FLUENT allows much more flexibility in grid generation for arbitrary geometries,
and the simulation set up can be done more easily in FLUENT than in FDS.

The results presented in this paper provide insight into the relative merits of specialized and
general-purpose codes and different fire models, which are frequently used to assess fire safety levels
in buildings in order to reduce risk to human life in the case of fire. The future work based on the
presented results may be of help in devising fire safety solutions, aimed at limiting the zones of possible
escape routes where people’s survival could be compromised in the event of a fire.

Author Contributions: Conceptualization, C.Z., P.G., J.L. and J.H.; Investigation, C.Z., J.L.; Methodology, C.Z.,
P.G. and J.H.; Visualization, J.L.; Supervision, J.H.; Writing–original draft, C.Z. and J.H.
Funding: This research was funded by Consorcio Regional de Transportes de Madrid, grant number OTRI-FUNED
5510003256.
Acknowledgments: The support of the Consorcio Regional de Transportes de Madrid is gratefully acknowledged.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Custer, R.L.P.; Meacham, B.J. SFPE Engineering Guide to Performance-Based Fire Protection Analysis and Design
of Buildings, 2nd ed.; National Fire Protection Association: Quincy, MA, USA, 2007.
2. Wang, Y.; Jiang, J.; Zhu, D. Full-scale experiment research and theoretical study for fires in tunnels with roof
openings. Fire Saf. J. 2009, 44, 339–348. [CrossRef]
3. National Fire Protection Association. The SFPE Handbook of Fire Protection Engineering, 4th ed.; NFPA: Quincy,
MA, USA, 2008.
4. Beard, A.; Carvel, R. The Handbook of Tunnel Fire Safety; Thomas Telford Publishing: London, UK, 2005.
5. Morgan, H.P. Design Methodologies for Smoke and Heat Exhaust Ventilation; Technical Report BRE368;
Building Research Establishment: Watford, UK, 1999.
6. Gao, R.; Li, A.; Hao, X.; Lei, W.; Zhao, Y.; Deng, B. Fire-induced smoke control via hybrid ventilation in a
huge transit terminal subway station. Energy Build. 2012, 45, 280–289. [CrossRef]
7. Migoya, E.; Crespo, A.; García, J.; Hernández, J. A quasi-one-dimensional model of fires in road tunnels.
Comparison with three-dimensional models and full-scale measurements. Tunn. Undergr. Space Technol.
2009, 24, 37–52. [CrossRef]
8. Deckers, X.; Haga, S.; Tilley, N.; Merci, B. Smoke control in case of fire in a large car park: CFD simulations
of full-scale configurations. Fire Saf. J. 2013, 57, 22–34. [CrossRef]
9. Tilley, N.; Rauwoens, P.; Merci, B. Verification of the accuracy of CFD simulations in small-scale tunnel and
atrium fire configurations. Fire Saf. J. 2011, 46, 186–193. [CrossRef]
Appl. Sci. 2019, 9, 3305 24 of 25

10. Zhang, J.; Weng, J.; Zhou, T.; Ouyang, D.; Chen, Q.; Wei, R.; Wang, J. Investigation on Smoke Flow in
Stairwells induced by an Adjacent Compartment Fire in High Rise Buildings. Appl. Sci. 2019, 9, 1431.
[CrossRef]
11. Novozhilov, V. Computational fluid dynamics modeling of compartment fires. Prog. Energy Combust. Sci.
2001, 27, 611–666. [CrossRef]
12. Olenick, S.M.; Carpenter, D.J. An Updated International Survey of Computer Models for Fire and Smoke.
J. Fire Prot. Eng. 2003, 13, 87–110. [CrossRef]
13. Boroń, S.; Wȩgrzyński, W.; Kubica, P.; Czarnecki, L. Numerical Modelling of the Fire Extinguishing Gas
Retention in Small Compartments. Appl. Sci. 2019, 9, 663. [CrossRef]
14. Wang, Y.; Chatterjee, P.; de Ris, J.L. Large eddy simulation of fire plumes. Proc. Combust. Inst. 2011,
33, 2473–2480. [CrossRef]
15. Van Maele, K.; Merci, B. Application of RANS and LES field simulations to predict the critical ventilation
velocity in longitudinally ventilated horizontal tunnels. Fire Saf. J. 2008, 43, 598–609. [CrossRef]
16. Vilfayeau, S.; Ren, N.; Wang, Y.; Trouvé, A. Numerical simulation of under-ventilated liquid-fueled
compartment fires with flame extinction and thermally-driven fuel evaporation. Proc. Combust. Inst. 2015,
35, 2563–2571. [CrossRef]
17. Jain, S.; Kumar, S.; Kumar, S.; Sharma, T. Numerical simulation of fire in a tunnel: Comparative study of
CFAST and CFX predictions. Tunn. Undergr. Space Technol. 2008, 23, 160–170. [CrossRef]
18. McGrattan, K.; Hostikka, S.; McDermott, R.; Floyd, J.; Weinschenk, C.; Overholt, K. Fire Dynamics Simulator
User’s Guide (FDS Version 6.1.2); National Institute of Standards and Technology (NIST): Gaithersburg, MD,
USA, 2014.
19. McGrattan, K.; Hostikka, S.; McDermott, R.; Floyd, J.; Weinschenk, C.; Overholt, K. Fire Dynamics Simulator
Technical Reference Guide. Volume 1: Mathematical Model (FDS Version 6.1.2); National Institute of Standards
and Technology (NIST): Gaithersburg, MD, USA, 2014.
20. FLUENT. User’s Guide 6.3; Fluent Inc.: New York, NY, USA, 2003.
21. McGrattan, K.; Hostikka, S.; McDermott, R.; Floyd, J.; Weinschenk, C.; Overholt, K. Fire Dynamics Simulator
Technical Reference Guide. Volume 2: Verification (FDS Version 6.1.2); National Institute of Standards and
Technology (NIST): Gaithersburg, MD, USA, 2014.
22. McGrattan, K.; Hostikka, S.; McDermott, R.; Floyd, J.; Weinschenk, C.; Overholt, K. Fire Dynamics Simulator
Technical Reference Guide. Volume 3: Validation (FDS Version 6.1.2); National Institute of Standards and
Technology (NIST): Gaithersburg, MD, USA, 2014.
23. Hadjisophocleous, G.; Jia, Q. Comparison of FDS Prediction of Smoke Movement in a 10-Storey Building
with Experimental Data. Fire Technol. 2009, 45, 163–177. [CrossRef]
24. Xiao, B. Comparison of Numerical and Experimental Results of Fire Induced Doorway Flows. Fire Technol.
2012, 48, 595–614. [CrossRef]
25. Blanchard, E.; Boulet, P.; Desanghere, S.; Cesmat, E.; Meyrand, R.; Garo, J.; Vantelon, J. Experimental and
numerical study of fire in a midscale test tunnel. Fire Saf. J. 2012, 47, 18 – 31. [CrossRef]
26. Zanzi, C.; Mozas, A.; Hernández, J.; García-Hortelano, A.; Aldecoa, J. Smoke and Heat Propagation Modeling
in Transportation Interchange Stations Using LES and RANS Methods. In Proceedings of the ASME 2013
International Mechanical Engineering Congress & Exposition, San Diego, CA, USA, 15–21 November 2013.
27. Deardorff, J.W. Numerical Investigation of Neutral and Unstable Planetary Boundary Layers. J. Atmos. Sci.
1972, 29, 91–115. [CrossRef]
28. McGrattan, K.B.; Baum, H.R.; Rehm, R.G. Large eddy simulations of smoke movement. Fire Saf. J. 1998,
30, 161–178. [CrossRef]
29. Magnussen, B.; Hjertager, B. On mathematical modeling of turbulent combustion with special emphasis on
soot formation and combustion. Symp. (Int.) Combust. 1977, 16, 719–729. [CrossRef]
30. Raithby, G.D.; Chui, E.H. A Finite-Volume Method for Predicting Radiant Heat Transfer in Enclosures with
Participating Media. J. Heat Transf. 1990, 112, 415–423. [CrossRef]
31. Hopkin, C.; Spearpoint, M.; Hopkin, D. A Review of Design Values Adopted for Heat Release Rate Per Unit
Area. Fire Technol. 2019. doi:10.1007/s10694-019-00834-8. [CrossRef]
32. Haynes, W.E. CRC Handbook of Chemistry and Physics, 94th ed.; CRC Press LLC: Boca Raton, FL, USA, 2014.
33. Cox, G.; Kumar, S. Modelling Enclosure Fires Using CFD. In SFPE Handbook of Fire Protection Engineering,
3rd ed.; NFPA: Quincy, MA, USA, 2002; Chapter 3–8, pp. 194–218.
Appl. Sci. 2019, 9, 3305 25 of 25

34. Smoke and Heat Control Systems—Part 3: Specification for Powered Smoke and Heat Control Ventilators (Fans);
Norma Europea UNE-EN 12101-3:2016; Asociación Española de Normalización: Madrid, Spain, 2016.
35. Baum, H.R.; McCaffrey, B.J.F. Fire Induced Flow Field—Theory and Experiment. In Fire Safety Science,
Proceedings 2nd International Symposium; Hemisphere: New York, NY, USA, 1989; pp. 129–148.
36. Cooper, L.Y. Combined Buoyancy and Pressure-Driven Flow Through a Shallow, Horizontal, Circular Vent.
J. Heat Transf. 1995, 117, 659–667. [CrossRef]
37. Chow, W.K.; Gao, Y. Oscillating behaviour of fire-induced air flow through a ceiling vent. Appl. Therm. Eng.
2009, 29, 3289–3298. [CrossRef]
38. Chow, W.K.; Gao, Y. Buoyancy and inertial force on oscillations of thermal-induced convective flow across a
vent. Build. Environ. 2011, 46, 315–323. [CrossRef]
39. Tan, Q.; Jaluria, Y. Mass flow through a horizontal vent in an enclosure due to pressure and density
differences. Int. J. Heat Mass Transf. 2001, 44, 1543–1553. [CrossRef]
40. Satoh, K.; Matsubara, Y.; Kumano, Y. Flow and Temperature Oscillations of Fire in a Cubic Enclosure with a
Ceiling and Floor Vent. Part I. In Technical Report 55, Report of Fire Research Institute of Japan; Fire Research
Institute: Tokyo, Japan, 1983.
41. Satoh, K.; Lloyd, J.R.; Yang, K.T. Flow and Temperature Oscillations of Fire in a Cubic Enclosure with a
Ceiling and Floor Vent. Part III. In Technical Report 57, Report of Fire Research Institute of Japan; Fire Research
Institute: Tokyo, Japan, 1984.
42. Kerrison, L.; Galea, E.R.; Patel, M.K. A two-dimensional numerical investigation of the oscillatory flow
behaviour in rectangular fire compartments with a single horizontal ceiling vent. Fire Saf. J. 1998, 30, 357–382.
[CrossRef]
43. ANSYS Fluent. User’s Guide, Release 2019 R2; ANSYS, Inc.: Canonsburg, PA, USA, 2019.

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like