You are on page 1of 11

Case Studies in Thermal Engineering 33 (2022) 101969

Contents lists available at ScienceDirect

Case Studies in Thermal Engineering


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

Enhancement of smoke extraction in tunnels in the case of fire by


using L-shape curtain near smoke extraction vents
Taher Halawa a, b, *
a
Mechanical Power Department, Faculty of Engineering, Cairo University, Giza, 12613, Egypt
b
Mechanical Engineering Department, Faculty of Engineering, The British University in Egypt (BUE), El-Sherouk City, Cairo, 11837, Egypt

A R T I C L E I N F O A B S T R A C T

Keywords: One of the challenges in the case of fire in tunnels is how to limit the propagation of smoke
Tunnels efficiently. Solid curtains which are released from the tunnel ceiling at the time of fire can prevent
Fire the smoke from further spread in the tunnel. The main issue with the traditional straight shape
Smoke curtains is that there is no guarantee for providing a complete safe environment as enough vis­
L-shape curtain ibility, temperature, and smoke concentration for the people at the region where the smoke is
Straight curtain
trapped behind the curtain. This paper introduces an idea to make a modification of the straight
curtain shape to be L-shape and to use it at a location close to the smoke extraction vent. Results
showed that in the case of straight curtain, part of the smoke moves downward after hitting the
curtain and then diffuses at low elevations. On the other hand, the L-shape curtain redirects the
smoke stream moving downward about 90◦ before being captured by the smoke extraction vent.
The L-shape curtain increased the visibility and provided lower temperatures and smoke con­
centration compared with the straight curtain shape. Moreover, L-shape curtain created better
interaction with the vent resulting in higher smoke flow extraction rate.

Nomenclature

Dc L-shape curtain depth (m)


Hc Curtain height (m)
Hsd Smoke extraction duct height (m)
Ht Tunnel height (m)
t Time (sec)
Wt Tunnel width (m)
z Distance measured from the solid curtain towards fire source (m)
δ Solid curtain thickness (m)

* Mechanical Power Engineering Department, Cairo University, 12613, Giza, Egypt.


E-mail address: taher.halawa@bue.edu.eg.

https://doi.org/10.1016/j.csite.2022.101969
Received 1 January 2022; Received in revised form 11 March 2022; Accepted 22 March 2022
Available online 24 March 2022
2214-157X/© 2022 The Author. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
T. Halawa Case Studies in Thermal Engineering 33 (2022) 101969

Fig. 1. Description of the tunnel selected for simulation and solid curtain models.

1. Introduction
Smoke control in tunnels in the case of fire is important for providing escape routes for people to evacuate safely and to avoid
exposure to harmful smoke effects.
One of the efficient methods in limiting the smoke propagation in tunnels is using curtains. These curtains can be formed using a
fluid as air and water or they can be in the form of solid obstacles made of fire resisting materials.
Air curtains are formed by injecting air with a specific speed and angle at certain locations in the tunnel with respect to the fire.
The air curtain has two effects, the first is to act as a resistance to limit smoke propagation, and the second is a cooling effect in
which the temperature is reduced [1]. Many parameters are linked to the performance of air curtains as air jet velocity, air jet direction,
thickness of the nozzle, smoke flow rate, and fire heat release rate [2–6]. Viegas and Cruz [7] referred to that the smoke flow rate is the
key parameter for the achievement of smoke capture by the air curtain.
The direction of the curtain air jet has a significant effect on the air curtain performance. The best air injection angle that will
optimize air curtain performance is about 30◦ [8]. The curtain air jet should be directed outward in order to achieve the best effect in
controlling the smoke [9].
In addition to having the injection angle optimized, the air released from the jet should have enough momentum to block the
smoke. Gao et al. [10] found that air curtain jet velocity should be at least 12 m/s for 1 MW fire, while it should be increased to 16 m/s
for 2 MW fire in order to stop the smoke propagation.
It is important to mention that the air jet velocity can be reduced by increasing the air curtain width. However, changing the width
of the air curtain does not bring a significant effect on the air curtain efficiency compared to the air injection angle variation [8].
Water curtain proved its successfulness in preventing smoke propagation and in reducing temperatures under specific conditions
[11–15].
Water droplets size as well as water pressure are two important factors related to the water curtain effectiveness. Yang et al. [16]
noticed that as the droplets size of the water mist is increased, the impact of water droplets on the smoke becomes higher. The effi­
ciency of the water mist system in blocking smoke can be increased by increasing water pressure and by using more nozzle rows [17].
If the working pressure is low, water curtain can reduce temperature effectively in spite of that the ability to block smoke is reduced
[18].
There are some issues related to the performance of water curtains and the ventilation type used in tunnels. Sun et al. [19] indicated
that water curtain is able to block smoke for tunnels with natural ventilation only and is not efficient for tunnels with longitudinal
ventilation. Li et al. [20] confirmed the poor performance of the water mist system in blocking smoke for longitudinal ventilation
tunnels.
The idea of adding solid curtain composed of strips made of a fire resisting material in order to block the smoke while having the
allowance of passing persons and vehicles in tunnels was introduced by Öttl et al. [21]. Using six curtains, the flow velocity was

2
T. Halawa Case Studies in Thermal Engineering 33 (2022) 101969

Fig. 2. Smoke-free layer height at three different locations from the fire for various mesh configurations.

decreased from 5 m/s to 1.1 m/s and in general, solid curtains were efficient in reducing the longitudinal air velocity.
Seike et al. [22] made a numerical investigation about the use of solid barriers at tunnel ceiling to improve the conditions of
evacuation during the fire event. Results showed that the solid curtain height should be at least 2 m and the tip velocity of the smoke
can be decreased by increasing the curtain height. Chaabat et al. [23] concluded that the pressure losses are much higher in the case of
large barriers compared to small barriers. Results revealed also that the curtain height is the key factor related to the effectiveness of
the curtain. AlAkam et al. [24] confirmed that the solid curtains are effective in limiting the smoke spread for a tunnel with transverse
ventilation system for three different fire sizes. Halawa, T. and Safwat, H [25]. studied the effect of adding number of solid barriers
with different geometrical properties at the tunnel ceiling to stop the motion of smoke in the case of fire. Results showed that smoke
barriers have a great influence on preventing smoke from spreading and this can be achieved either by increasing solid barriers height
while increasing the distance between them or by increasing the number of solid barriers with lower heights.
However, there are some problems associated with the using of solid curtains. The accumulated smoke behind solid barriers in
tunnels may lead to the reduction of visibility in addition to the increase of smoke concentration and temperature especially when the
curtain height is increased for the purpose of creating compartmentalization [23,26]. Halawa, T [27]. found that increasing the solid
curtain height more than 40% of the tunnel height results in spreading of smoke at low elevations near the people in the tunnel leading
to visibility deterioration and temperature increase.
In the light of the presented previous researches regarding using curtains to control smoke in road tunnels, it may be concluded that
air and water curtains are more complex and expensive compared with solid curtains which are efficient in blocking smoke but there is
a big concern regarding the safety level provided by solid curtains for the people at locations behind the curtain. This research gap
about how to provide a simple method to control smoke motion like the solid curtains but with high degree of safety triggered the idea
of the currently presented research.
The aim of this research is to investigate the enhancement of solid curtain performance by changing the traditional straight shape of
the curtain to a new suggested L-shape in addition to studying the effect of positioning the curtain close to the smoke extraction vent.
Comparisons between the performance of straight curtain and L-shape curtain in terms of smoke concentration, height of smoke
free layer, visibility conditions, and temperature levels are provided in this paper with detailed analysis.

2. Description of the selected case study and solid curtain models


The case study selected for simulation is the tunnel presented in the study of Liang et al. [11] as shown in Fig. 1(a). This tunnel has a
length of 600 m and a width (Wt) of 10 m. The tunnel height (Ht) is 5 m, while the height of the smoke extraction duct (Hsd) is 2 m. The
straight curtain and L-shape curtain models are presented in Fig. 1(b) and (c) respectively. Each curtain was positioned at a distance of
4 m from the smoke extraction vent edge in order to achieve the best interaction between the curtain and the vent; this was assumed
based on the findings from a previous research [27]. The solid curtain thickness (δ) was assumed to have a fixed value of 0.1 m. Both of
straight and L-shape curtains were assumed to have the same width of the tunnel. The depth of the L-shape curtain (Dc) was assumed to
be 1 m based on several simulation results which revealed that this is the minimum value under which, the L-shape curtain model will
have results close to that for the straight curtain.
The range of the curtain height (Hc) values selected in the current study is from 1.5 m to 3 m. The reason for choosing this range is
that if Hc is smaller than 1.5 m, curtain fails to stop smoke propagation, while Hc was avoided to be larger than 3 m to allow the passage
of people from undernaeth the curtain.

3. Numerical modeling
Fire simulations were conducted using Fire Dynamics Simulator [FDS] software [28], while the evacuation simulations were done
using Pathfinder software [29]. The fire dynamics simulation results were coupled with the evacuation model and the motion of people

3
T. Halawa Case Studies in Thermal Engineering 33 (2022) 101969

Fig. 3. CO Yield concentration contours at a vertical plane located at the mid-tunnel width at t = 600 s for both of straight shape and L-shape curtains with various
curtain heights.

Fig. 4. CO Yield concentration contours at different times for the straight shape curtain and the L-shape curtain with Hc = 3 m

during evacuation was taken into consideration to be sensitive to the smoke distribution and visibility according to the correlations
presented in the study [30].
A bus fire was assumed to take place at the mid-length of the tunnel as shown in Fig. 1(a). The validation analysis of the tunnel
model used in the current research was discussed in previous published papers [25,27].
Large-eddy simulation (LES) model was implemented in the numerical model. The time step used in the simulations is 0.01 s. Fans

4
T. Halawa Case Studies in Thermal Engineering 33 (2022) 101969

Fig. 5. Height of the smoke-free layer for the straight curtain case in comparison with L-shape curtain case at two vertical lines located in the region between curtain
and smoke extraction vent.

are attached to the ends of the smoke exhaust duct with a flow rate of 120 m3/s. Open boundary condition was assigned to the tunnel
inlet and exit. The ambient temperature was taken as 20 ◦ C. Tunnel walls material was assumed to be gypsum board which has a
specific heat of 900 J/kgK, and thermal conductivity of 0.16 W/mK.
The variation of heat release with time related to a bus fire reported by Refs. [31,32] was used as an input to the numerical model.
Also, the fire characteristics included in the model are the fire peak heat release rate of 30 MW, total fire calorific content of 41 GJ, and
ratio of the convective heat to the total heat release of 67% [33,34].
The activated reaction in the FDS model is polyurethane which represents most of the bus seating material and thus, the soot yield
was set to be 0.19 kg of soot/kg of fuel [35,36].
The algorithm used in the movement of people in the evacuation model is based on the Helbing’s group method [37,38].
The governing equations used to determine the motion paths of people in the evacuation model include three groups of forces. The
first group accounts for the social and physical interaction between people. The second group involves the people interaction with
surrounding walls while the third group includes forces between people and environment.
Grid independence test was done to obtain the suitable mesh configuration which provides accurate results. Fig. 2 indicates the
variation of the smoke-free layer height with the number of mesh elements at three different locations from the fire. It may be noted
from Fig. 2 that the variations in the smoke-free height values become small when the number of mesh elements is larger than 1.75
million cells. Other parameters as the temperature and smoke concentration were monitored and confirmed the small variation in
results for number of mesh elements higher than 2 million cells. Based on these results, the final mesh configuration selected for
simulations contain 2886350 elements.

4. Results
4.1. Smoke characteristics
4.1.1. Carbon oxide [CO] yield concentration
Fig. 3 shows the CO Yield concentration contours for the straight and L-shape curtain designs with different values of Hc at a time
corresponding to the fire peak heat release rate (at t = 600 s). For all shown cases in Fig. 3, the curtain was able to prevent smoke from
exceeding the curtain location but the interaction between smoke and curtain differs from case to another depending on the curtain
shape and curtain dimensions. It may be noted from Fig. 3 for the straight curtain results that as Hc increases, high concentration layers

5
T. Halawa Case Studies in Thermal Engineering 33 (2022) 101969

Fig. 6. Variation of smoke mass flow rate passing through the vent with time for various values of Hc.

of smoke (0.004–0.007) approach the tunnel ground. On the other side, for the L-shape curtain results, high concentration smoke
regions drop slightly (by 0.1 m) below the curtain lower tip level and never reach the region where the people are existing (up to 1.9 m
from the tunnel ground), and this provides more safe conditions for the people compared with the straight curtain results. The L-shape
curtain provided trapping of smoke in such a way that the smoke spread in both of the forward and downward directions are limited.
It is important to study the development of smoke concentration with time in order to know how the final situation is reached and at
which time there will be a significant difference between the smoke distributions of the two curtain models. The case with Hc = 3 m
was selected for this purpose, and the smoke concentration contours at five different times are shown in Fig. 4. It is important to
mention that similar analysis for the other three cases (Hc = 1.5, 2, and 2.5 m) will bring the same concept introduced by analyzing the
case of Hc = 3 m. However, the case of Hc = 3 m shows the concept more clearly compared to the other three cases because as the
curtain height increases, the performance gap between the two curtain designs becomes larger.
The mass flow rate of smoke produced by the fire increases with time; this leads to the accumulation of smoke layers and increasing
of smoke concentration with time at the region behind the curtain. What makes a difference between the two curtain models is that
how the accumulated smoke is distributed over the tunnel volume behind the curtain and whether if the smoke approaches tunnel
ground or not.
It may be noted from the smoke concentration contours comparison shown in Fig. 4 that in the case of straight curtain, the smoke
moves towards the ground after hitting the curtain. In the case of L-shape curtain, the smoke returns to the vent after hitting the curtain
and forms like a closed loop connected to the vent. This occurs slightly at t = 360 s and becomes more clear starting from t = 420 s.
At t = 480 s (Fig. 4), the high concentration area appears near the ground for the case of straight curtain. On the other hand, for the
L-shape curtain, the area between the ground and the curtain lower end is nearly free of smoke.
Part of the area behind the curtain is divided into 3 regions (R1, R2, and R3) for the aid of providing clarifications related to smoke
diffusion at t = 540 s as shown in Fig. 4. The three regions differ from each other in terms of smoke concentration levels. For the region
R1, the smoke concentration is very high for both curtain models. However, smoke concentrations at regions R2 and R3 for the straight
curtain are not the same as for L-shape curtain. The smoke concentrations at regions R2 and R3 can be related to the direction of smoke
diffusion (mostly from R1 to R3 for the straight curtain case and from R1 to R2 for the L-shape curtain case). This confirms the suc­
cessfulness of the L-shape curtain in confining the smoke in the area between the curtain and the smoke extraction vent.
The boundary separating the area with high smoke concentration and the area clear of smoke underneath the curtain is shifted
upward for the case of L-shape curtain compared with the straight curtain as can be seen from Fig. 4 at t = 600 s.

4.1.2. Height of smoke-free layer


The height of smoke-free layer is an important measure of the curtain performance in providing areas free of smoke above the
tunnel ground. The variation of the height of this layer with time was monitored at various locations during simulations. Samples of

6
T. Halawa Case Studies in Thermal Engineering 33 (2022) 101969

Table 1
Average values of vent mass flow rate for the straight and L-shape curtains corresponding to different values of Hc.

Hc Average vent mass flow rate (kg/s) [for Average vent mass flow rate (kg/s) [for L- Relative % increase in the average vent mass flow rate when
(m) straight curtain] shape curtain] using L-shape curtain

1.5 65.03 67.62 3.98


2 61.45 66.44 8.12
2.5 53.65 65.05 21.25
3 47.78 61.85 29.45

Fig. 7. Visibility contours at different times for the straight and L-shape curtains for Hc = 2.5 m

these results are displayed in Fig. 5 where the smoke free layer height values are reported at two vertical lines (A & B). Line-A passes
through the smoke extraction vent, while line-B is located at the mid distance between the curtain and the smoke extraction vent as
presented in the graphical representation in Fig. 5(a). Time averaged values of the smoke free layer height corresponding to different
Hc values are shown in Fig. 5(b) for vertical line-A and in Fig. 5(c) for vertical line-B.
It can be concluded that as curtain height increases, there is a drop in the smoke free layer height for both curtain designs but L-
shape curtain provides higher values of time averaged smoke free layer height compared to straight curtain for all cases as may be
noted from Fig. 5(b) and (c).
The maximum percentage increase in the time averaged smoke free layer height provided by the L-shape curtain compared to the
straight curtain is 45% at line-A corresponding to Hc = 2 m (marked by a dotted rectangle in Fig. 5(b)), while the maximum percentage
increase at line-B is 106% corresponding to Hc = 2.5 m (marked by a dotted rectangle in Fig. 5(c)).
The variation of smoke free layer height with time for the cases with maximum improvement (marked in Fig. 5(b) and (c)) are
shown in Fig. 5(d) and (e) respectively.
The fluctuations in curves shown in Fig. 5(d) are higher than those in Fig. 5(e) because at line-A, there is more turbulent flow than
at line-B due to the interaction between the incoming smoke stream and the smoke reflected by the curtain and also the flow moving
through the smoke extraction vent at line-A.
Fig. 5(e) indicates that the smoke free layer height for the straight curtain starting from t = 260 s drops mostly below 1 m which is
lower than the average people height and this is a signal of visibility reduction for the people in this case. The L-shape curtain provides
more stable and higher values (around 2 m) of smoke free layer height in comparison with the traditional curtain design as can be seen
from Fig. 5(e).

7
T. Halawa Case Studies in Thermal Engineering 33 (2022) 101969

Fig. 8. Smoke distribution at t = 600 s for Hc = 2.5 m

4.2. Smoke extraction efficiency


In order to have an accurate vision about the smoke extraction efficiency for the straight and L-shape curtains, the smoke flow rate
of the smoke extraction vent was monitored in all cases. Fig. 6 shows a comparison between the straight and L-shape curtains in terms
of the variation of smoke flow rate through the vent with time for various Hc values.
It may be noted from Fig. 6 that the smoke extraction rate for L-shape curtain is higher than that for the straight curtain for all Hc
cases.
The time-averaged mass flow rate values as well as the improvement rates provided by the L-shape curtain are reported in Table 1.
When Hc is increased from 1.5 m to the double of its value, the average mass flow rate is decreased sharply from 65.03 kg/s to 47.78
kg/s in the case of straight curtain, while a slight decrease from 67.62 kg/s to 61.85 kg/s occur in the case of L-shape curtain. The
maximum increase in the smoke flow rate extracted through the vent when using L-shape curtain was 29.45% at Hc = 3 m.
The main reason for the sharp decline in the smoke flow rate as Hc is increased for the straight curtain shape is that as the curtain is
elongated, the smoke becomes more aligned downward after its impact with the curtain. Consequently, the smoke is distracted away
from the smoke extraction vent and diffuses in the air at low elevations before being partially extracted through the vent.

4.3. Visibility distribution


The case of Hc = 2.5 m was selected for presenting the evacuation analysis among all the cases which prove the same concept.
However, the case of Hc = 2.5 m may be better for practical implementation compared to other cases because it allows more space for
the people to pass from underneath the curtain comfortably compared to higher Hc values and in the meantime provides more efficient
smoke capturing compared to cases with lower Hc values. Visibility contours at different moments of time for the case of Hc = 2.5 m are
shown in Fig. 7. At t = 300 s, the smoke is relatively away from the curtain and there is no interaction between them at this moment,
and so the visibility contours are nearly identical for the two curtain models. At t = 360 s, the interaction between smoke and curtain is
in its earliest stages, and there is no big difference between the visibility contours for straight and L-shape curtains. The only difference
is that for the case of straight curtain, a layer with intermediate visibility values (6–12 m) appears slightly below the curtain lower tip
and extends from the curtain to the smoke extraction vent.
Starting from t = 420 s, there is a remarkable difference between the visibility distributions of the two curtain models as shown in
Fig. 7. For the straight curtain at t = 420 s (Fig. 7), low visibility region (2–5 m) is formed vertically between the ground level and close
to the curtain lower edge, and horizontally from the curtain to about 10 m in the reverse direction of the main smoke motion. This low
visibility region expands horizontally as the time passes; this can be noted by comparing the visibility contours at t = 420 s with that at
t = 480 s for the straight curtain case.
For the straight curtain at t = 540 & 600 s, visibility conditions are deteriorated compared with conditions at earlier times because
the smoke generation rate increases with time resulting in more accumulation of smoke behind the straight curtain which directs more
smoke downward in this case. Very low visibility regions (less than 2 m) are nearly covering the entire areas at low elevations including
vehicles and people.
The L-shape curtain provides much better visibility conditions for people at all times compared to the straight curtain. This can be
seen by comparing the two magnified views at t = 600 s in Fig. 7 as the straight curtain effect resembles as if the curtain is extended to
touch the ground causing the trapping of people in a region with very low visibility values. In spite of the curtains are not physically
preventing the flow of people, but the straight curtain caused decreasing of people speed compared to the L-shape curtain.
The smoke distribution shown in Fig. 8 corresponding to t = 600 s for the case of Hc = 2.5 m adds more clarification about the
difference between straight and L-shape curtains. This scene depicts the degree of severity related to smoke distribution where the
smoke is nearly everywhere behind the straight curtain except for a thin layer close to the ground while for the L-shape curtain, the

8
T. Halawa Case Studies in Thermal Engineering 33 (2022) 101969

Fig. 9. Visibility distribution at a horizontal line located at 1.5 m from the ground t = 600 s.

Fig. 10. Temperature values at different distances from the curtain and at a height of 1.5 m from the ground for various Hc cases corresponding to t = 600 s.

9
T. Halawa Case Studies in Thermal Engineering 33 (2022) 101969

smoke layer thickness is slightly exceeding the curtain height.


For the purpose of providing an accurate estimate of visibility values at level close to the people sight region, the visibility at a
horizontal line located at 1.5 m from the ground for the cases of Hc = 2, 2.5, 3 m at t = 600 s is presented in Fig. 9. It may be noted that
the visibility values ranges from 3.5 to 15 m for the L-shape curtain, while it ranges from 0.5 to 5 m for the straight curtain. Moreover,
for any selected curtain height value, the L-shape curtain provides higher visibility values than the straight curtain at any location
measured from the curtain position (at any value of z). It is important to highlight that as Hc is increased the visibility gap is increased
between the straight and L-shape curtains. The visibility conditions provided by the L-shape curtain are much closer to the tenability
limit of 10 m [39] compared to the straight curtain.

4.4. Temperature distribution


One of the critical issues related to the accumulated smoke in a specific area is the excessive temperature levels which may hurt
people waiting for the fire to be extinguished or trying to change their locations and possibly passing through dangerous high tem­
perature zones. Sample of the temperature results extracted from simulations are shown in Fig. 10 in which the temperature values at
four points having different z values with the same elevation from the ground (1.5 m) at t = 600 s are presented.
As Hc is increased, temperature is increased for both curtain designs but the rate of temperature increase provided by the L-shape
curtain is lower than that for the straight curtain as may be seen from Fig. 10.
The minimum and maximum temperature values for the L-shape curtain for all Hc values are 28 ◦ C and 45 ◦ C respectively, while in
the case of straight curtain these values are 31 ◦ C and 68 ◦ C. Keeping the smoke away from diffusion at lower elevations in the case of L-
shape curtain is the key factor behind having much lower temperatures than for the straight curtain. It can be concluded that tem­
peratures in the case of straight curtain at some locations exceed the tenability limit of 60 ◦ C [39] when Hc is higher than or equal to
2.5 m, while the L-shape curtain provides temperatures much lower the tenability limit.

5. Conclusions
The idea of using a curtain with L-shape at a location close to the smoke extraction vent is proposed in this paper. The performance
of straight and L-shape curtains with various heights is investigated numerically and the following are the most important findings:
• Using L-shape curtain extended the main purpose of the traditional straight curtain in limiting the propagation of smoke in the
tunnel to include two additional functions; improving the efficiency of smoke extraction and providing safer environment for the
people in the tunnel.
• The mass flow rate of smoke extracted through the smoke extraction vent is increased when using L-shape curtain compared to the
straight curtain for all tested curtain heights.
• Significant improvement in the visibility conditions is achieved when replacing the straight curtain with the L-shape curtain. At 1.5
m from the ground, visibility values ranged from 0.5 m to 5 m for the straight curtain design, while in the case of L-shape curtain,
the range is from 3.5 m to 15 m.
• Extremely high temperatures (ranged from 60 ◦ C to 70 ◦ C) appear at people zone for the straight curtain. This is completely avoided
when using L-shape curtain which provided much lower temperature ranges.

6. Recommendations for future studies


It is recommended for future researches to make investigations on the performance of more shapes of the solid curtain like the L-
shape with curved lower boundary. Also, it is recommended to study the idea of placing the smoke extraction vent between two solid
curtains with different shapes for the purpose of capturing the smoke between the two curtains which resembles a storage room for the
smoke.

Author agreement statement


We the undersigned declare that this manuscript is original, has not been published before and is not currently being considered for
publication elsewhere.
We confirm that the manuscript has been read and approved by all named authors and that there are no other persons who satisfied
the criteria for authorship but are not listed. We further confirm that the order of authors listed in the manuscript has been approved by
all of us.
We understand that the Corresponding Author is the sole contact for the Editorial process. He is responsible for communicating
with the other authors about progress, submissions of revisions and final approval of proofs.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

References
[1] K. Lu, H. Xu, C. Shi, Z. Wang, J. Wang, Y. Ding, Numerical investigation of air curtain jet effect on the upper layer temperature evolution of a compartment fire
and its transition, Appl. Therm. Eng. 197 (2021) 117409.

10
T. Halawa Case Studies in Thermal Engineering 33 (2022) 101969

[2] J. Elicer-Cortes, N. Molina, G. Severino, A. Fuentes, P. Rojas, Turbulent transport mechanisms on the heat confinement in tunnels by using low-velocity air
curtains, Appl. Therm. Eng. 181 (2020) 115852.
[3] U.-H. Jung, S. Kim, S.-H. Yang, J.-H. Kim, Y.-S. Choi, Numerical study of air curtain systems for blocking smoke in tunnel fires, J. Mech. Sci. Technol. 30 (11)
(2016) 4961–4969.
[4] M. Juraeva, K.J. Ryu, S.-H. Jeong, D.J. Song, Numerical optimization study to install air curtain in a subway tunnel by using design of experiment, J. Mech. Sci.
Technol. 28 (1) (2014) 183–190.
[5] Li, T., Yang, Z., and Chen, L., "Numerical simulation study on smoke control in urban bifurcated tunnel under the effect of air curtain," Proc. 2019 9th
International Conference on Fire Science and Fire Protection Engineering (ICFSFPE), IEEE, pp. 1-4.
[6] N. Luo, A. Li, R. Gao, W. Zhang, Z. Tian, An experiment and simulation of smoke confinement utilizing an air curtain, Saf. Sci. 59 (2013) 10–18.
[7] J.C. Viegas, H. Cruz, Air curtains combined with smoke exhaust for smoke control in case of fire: full-size experiments, Fire Technol. 55 (1) (2019) 211–232.
[8] L.-X. Yu, F. Liu, T. Beji, M.-C. Weng, B. Merci, Experimental study of the effectiveness of air curtains of variable width and injection angle to block fire-induced
smoke in a tunnel configuration, Int. J. Therm. Sci. 134 (2018) 13–26.
[9] Z. Zhou, Y. Lu, Y. Cui, Study on the effect of jet direction of compound air curtain on smoke control, Energies 14 (21) (2021) 6983.
[10] D. Gao, T. Li, X. Mei, Z. Chen, S. You, Z. Wang, K. Wang, P. Lin, Effectiveness of smoke confinement of air curtain in tunnel fire, Fire Technol. 56 (5) (2020)
2283–2314.
[11] Q. Liang, Y. Li, J. Li, H. Xu, K. Li, Numerical studies on the smoke control by water mist screens with transverse ventilation in tunnel fires, Tunn. Undergr. Space
Technol. 64 (2017) 177–183.
[12] Z. Lin, R. Bu, J. Zhao, Y. Zhou, Numerical investigation on fire-extinguishing performance using pulsed water mist in open and confined spaces, Case Stud.
Therm. Eng. 13 (2019) 100402.
[13] P. Luo, W.F. Liu, W.P. Han, J.J. Xia, K. Yao, Influence of nozzle height on water mist fire extinguishing system in railway tunnel rescue station, in: Proc. 2017
3rd International Forum on Energy, Environment Science and Materials, IFEESM, 2017, pp. 1677–1682. Atlantis Press.
[14] Xia, F., and Yang, Y., "Research progress on smoke control technology for highway tunnel fire," Proc. IOP Conference Series: Materials Science and Engineering,
IOP Publishing, p. 012112.
[15] P. Zhang, X. Tang, X. Tian, C. Liu, M. Zhong, Experimental study on the interaction between fire and water mist in long and narrow spaces, Appl. Therm. Eng. 94
(2016) 706–714.
[16] Y.-b. Yang, B. Yang, Z. Bing, The study on influence of water mist particle size on fire smoke migration with longitudinal ventilation in road tunnel, Procedia
Eng. 211 (2018) 917–924.
[17] Q. Li, Z. Tang, Z. Fang, J. Yuan, J. Wang, Experimental study of the effectiveness of a water mist segment system in blocking fire-induced smoke and heat in mid-
scale tunnel tests, Tunn. Undergr. Space Technol. 88 (2019) 237–249.
[18] Z. Wang, X. Wang, Y. Huang, C. Tao, H. Zhang, Experimental study on fire smoke control using water mist curtain in channel, J. Hazard Mater. 342 (2018)
231–241.
[19] J. Sun, Z. Fang, Z. Tang, T. Beji, B. Merci, Experimental study of the effectiveness of a water system in blocking fire-induced smoke and heat in reduced-scale
tunnel tests, Tunn. Undergr. Space Technol. 56 (2016) 34–44.
[20] H. Li, X. Chen, X. Shi, Y. Wang, R. Zhu, W. Jiang, An experimental study of fire protection and control of low pressure water mist in highway tunnel, Tunn.
Construct. 38 (2018) 73–78.
[21] D. Öttl, P. Sturm, R. Almbauer, W. Öttl, A. Thurner, G. Seitlinger, A new system to reduce the velocity of the air flow in the case of fire, in: Proceedings Tunnel
Safety and Ventilation, ISBN, 2002, pp. 3–901354.
[22] Seike, M., Kawabata, N., and Hasegawa, M., "The effect of fixed smoke barriers on evacuation environment in road tunnel fires with natural ventilation," Proc.
7th International Conference ‘Tunnel Safety and Ventilation, pp. 126-132.
[23] F. Chaabat, M. Creyssels, A. Mos, J. Wingrave, H. Correia, M. Marro, P. Salizzoni, The effects of solid barriers and blocks on the propagation of smoke within
longitudinally ventilated tunnels, Build. Environ. 160 (2019) 106207.
[24] H. AlAkam, A.A. Fahim, E.E. Khalil, Numerical investigations of smoke management in road tunnel with and without curtains, in: Proc. AIAA Scitech 2020
Forum, 1936.
[25] T. Halawa, H. Safwat, Fire-smoke control strategies in road tunnels: the effectiveness of solid barriers, Case Stud. Therm. Eng. 27 (2021) 101260.
[26] M. Bettelini, S. Rigert, N. Seifert, Flexible devices for smoke control in road tunnels, in: 6th International Conference ‘Tunnel Safety and Ventilation’Graz, 2012.
[27] T. Halawa, On the use of solid curtains near smoke extraction vents to control smoke spread resulting from fire in road tunnels, J. Therm. Sci. Eng. Appl. 13 (3)
(2021), 031025.
[28] R. McDermott, K. McGrattan, S. Hostikka, Fire Dynamics Simulator (Version 5) Technical Reference Guide, vol. 1018, NIST Special Publication, 2008, p. 5.
[29] C. Thornton, R. O’Konski, B. Hardeman, D. Swenson, Pathfinder: an agent-based egress simulator, in: Pedestrian and Evacuation Dynamics, Springer, 2011,
pp. 889–892.
[30] Fridolf, K., Nilsson, D., Frantzich, H., Ronchi, E., and Arias, S., "Walking speed in smoke: representation in life safety verifications," Proc. 12th International
Performance-Based Codes and Fire Safety Design Methods Conference, Oahu, Hawaii, pp. 1-6.
[31] H. Ingason, S. Gustavsson, M. Dahlberg, Heat release rate measurements in tunnel fires, Brandforsk Proj. (1994) 723–924.
[32] H. Ingason, Y.Z. Li, A. Lönnermark, Tunnel fire ventilation, in: Tunnel Fire Dynamics, Springer, 2015, pp. 333–369.
[33] H. Ingason, A. Lönnermark, Heat release rates in tunnel fires: a summary, in: Handbook of Tunnel Fire Safety, ICE publishing, 2012, pp. 309–327.
[34] Y.Z. Li, H. Ingason, Overview of research on fire safety in underground road and railway tunnels, Tunn. Undergr. Space Technol. 81 (2018) 568–589.
[35] Y.-J. Chen, C.-M. Shu, S.-P. Ho, H.-C. Kung, S.-W. Chien, H.-H. Ho, W.-S. Hsu, Analysis of smoke movement in the Hsuehshan tunnel fire, Tunn. Undergr. Space
Technol. 84 (2019) 142–150.
[36] G.W. Mulholland, Smoke production and properties, in: SFPE Handbook of Fire Protection Engineering, vol. 2, 1995, p. 2.
[37] D. Helbing, I.J. Farkas, P. Molnar, T. Vicsek, Simulation of pedestrian crowds in normal and evacuation situations, Pedestrian Evacuation Dynam. 21 (2) (2002)
21–58.
[38] Werner, T., and Helbing, D., "The social force pedestrian model applied to real life scenarios," Proc. Pedestrian and Evacuation Dynamics 2003. Proceedings of
the 2nd International Conference on Pedestrian and Evacuation Dynamics, CMS Press, University of Greenwich, pp. 17-26.
[39] H. Ingason, UPTUN-Workpackage 2 Fire development and mitigation measures-Target criteria, Upgrading Existing Tunn. Fire Saf. (2005).

11

You might also like