You are on page 1of 26

JP Journal of Heat and Mass Transfer

© 2019 Pushpa Publishing House, Prayagraj, India


http://www.pphmj.com
http://dx.doi.org/10.17654/HM017020451
Volume 17, Number 2, 2019, Pages 451-476 ISSN: 0973-5763

EXPERIMENTAL STUDY OF THE INFLUENCE OF


BLOCKAGE ON CRITICAL VELOCITY AND
BACKLAYERING LENGTH IN A LONGITUDINALLY
VENTILATED TUNNEL

1 2 3
Razieh Khaksari Haddad , Zambri Harun , Cristian Maluk and
4
M. Rasidi Rasani
1,2,4
Faculty of Engineering and Built Environment
The National University of Malaysia
43600, Bangi
Malaysia
3
School of Civil Engineering
The University of Queensland
St Lucia, QLD 4072
Australia

Abstract

This study has been conducted to analyze the effect of vehicular


blockage on critical velocity and backlayering length experimentally
through a 3 m model tunnel (1:50 of a standard tunnel length). Three
vehicle sizes in two or three arrays, occupying 3-15% of tunnel cross-
section in six scenarios, were used. Based on this study, critical
velocity decreases in the case that ventilation velocity reaches the fire
source directly. However, critical velocity increases when the path of
ventilation flows are blocked by blockages due to an increase in heat
release rate through vehicular obstructions. In addition, the influence
Received: February 25, 2019; Revised: March 13, 2019; Accepted: April 25, 2019
Keywords and phrases: backlayering length, critical velocity, heat release rate.
452 R. K. Haddad, Z. Harun, C. Maluk and M. R. Rasani
of blockages on backlayering length and its dimensionless analysis by
the ratio of longitudinal ventilation velocity to critical velocity is also
presented. It is found that the backlayering length follows an
exponential relation with and without blockages. Moreover, the
analysis of backlayering length with respect to the blockage section
length shows that as backflow occurs along the ceiling and flows over
blockage section, its heat transfer to surrounding increases leading to
shorter backlayering length due to the enhanced inertial force and heat
loss.

Nomenclature
2
AT Cross-sectional area of tunnel, m
Ablkg 2
Cross-sectional area of the vehicular blockage, m
Cp –1 –1
Specific heat capacity, kjkg K
–2
g Gravitational force, ms
H Tunnel height, m
H Tunnel mean hydraulic diameter, m
Lb Backlayering length, m

L∗b Dimensionless Backlayering length


Q Heat release rate of the fire source, kW
Q∗ Dimensionless heat release rate
Q′ Fire convective heat release rate per tunnel unit width, kWm
–1

T Smoke temperature, K
Ta Ambient temperature, K
–1
uc Critical ventilation velocity, ms

uc∗ Dimensionless critical velocity


–1
V Longitudinal ventilation velocity, ms

V ∗∗ The ratio of longitudinal ventilation velocity to critical velocity

uc∗, b Dimensionless critical velocity with blocks


Experimental Study of the Influence of Blockage … 453

uc, b –1
Critical velocity with blocks, ms
–3
ρa Ambient air density, kgm
ε experimental constants in Eq. 10

1. Introduction

Although the construction of tunnels in urban transport system can help


reducing traffic congestion and traveling time within cities, the potential
hazards towards the lives of users of these urban structures should be
investigated in all perspectives. One of these dangers is a fire that could
spread from one point to the next vehicles easily. In addition to fire itself, fire
by-products such as smokes and poisonous fumes could also threaten the
lives of drivers and firefighters. Different ventilation methods including
longitudinal and transversal ventilation are used to direct any potential smoke
away from crowded areas. When the ventilation system’s efficiency to
control the flow of smoke is not sufficient, a stream of smoke flows in the
section between the fan and the source of fire (upstream section), in the
opposite direction along the tunnel ceiling and causes that passengers to get
stuck in a dangerous area. This corresponding extent of smoke is known as
backlayering. When backlayering flow disappears in the upstream section,
the longitudinal air flow which effectively eliminates backlayering distance
is called the critical velocity. This parameter is a significant criterion when
designing tunnel ventilation systems. As a result, the upstream section is
converted into an area free from smokes and toxic gases so that firefighters
can rescue occupants who get stuck in the fire. Furthermore, the areas free
from smokes have less intensity of the fire, is out of the danger of
asphyxiation. Although studies have focused extensively on critical velocity
and backlayering length, studies on critical velocity are more ancient and
widespread [1-5].
Critical velocity model
Previous models on critical ventilation velocity have been derived based
on the Froude modeling and experimental correlations from small- or large-
454 R. K. Haddad, Z. Harun, C. Maluk and M. R. Rasani
scale model tunnels. The first analytical study about the correlation between
critical velocity and the Froude number was carried out by Thomas [6]. This
relationship was based on a comparison between the buoyant force of the fire
plume and inertial force of ventilation velocity (the Froude number). Based
on Thomas’ study, smoke buoyancy force and the inertial force of ventilated
air should be equal under the critical conditions. This scheme was followed
by many investigators. The following model was proposed by Thomas based
on the Froude number [6]:
1
⎛ gQ ′ ⎞3
U c = k ⎜⎜ ⎟ ,
⎟ (1)
⎝ ρ0C pT ⎠

where k is a coefficient and it was found to be equal to unity only for some
corresponding experimental set-ups.
Oka and Atkinson [7] further carried out a dimensionless analysis for
critical velocity, taking different geometries of the fire source into account as
follows:
Q
Q∗ = 1 5
, (2)
ρ0C pTa g 2H 2

Uc
U c∗ = . (3)
gH

They suggested the following replacement for Eq. 1

⎧⎪ 1 1
− ∗
U c∗ =⎨ 0. 35( 0 . 124 ) 3 (Q ) 3 for Q∗ ≤ 0.124, (4)
⎪⎩0.35 for Q∗ > 0.124.

Wu and Bakar [3] carried out another series of small-scale experiments


to find out how U c is affected by the tunnel cross-sectional shape both
experimentally and numerically using a CFD computer simulation. Tunnel
hydraulic diameter was determined as characteristic length instead of the
tunnel height and new correlations were identified to determine U c [3]:
Experimental Study of the Influence of Blockage … 455
Q
Q∗ = 1 5
, (5)
ρ0C pTa g 2H 2

Uc
U c∗ = . (6)
gH

Wu and Bakar [3] proposed a simple one-dimensional correlation to


predict critical ventilation velocity for tunnels according to their
experimental results. The suggested equations are:

⎧⎪ 1 1
− ∗
U c∗ = ⎨0.4(0.2) 3 (Q ) 3 for Q∗ ≤ 0.2, (7)
⎪⎩0.4 for Q∗ > 0.2.

1.1. Backlayering Length Model


Based on the theoretical model, the backlayering length has been
investigated quite extensively [2, 5, 8]. In addition to theoretical analyses, the
thermal buoyant smoke backlayering flow length beneath the tunnel ceiling
has also been conducted experimentally [9-11]. A theoretical analysis of
backlayering length based on the Froude number in a longitudinally
ventilated tunnel was also proposed by Thomas [8]. He suggested that the
critical Froude number is equal to unity, causing backlayering distance to
disappear. The proposed relation can be expressed as follows [8]:

L gHQ
L∗b = b ∝ . (8)
H ρ0C pTV 3 A

Full-scale experiments were carried out by Hu et al. [2] in addition to


FDS simulations and they proposed the following model for backlayering
length:

⎡ Ck H ⎤
⎢⎣ K 2 2 ⎥⎦
Lb = ln V , (9)
0.019

where Ck is the empirical constant, and K 2 is expressed as follows:


456 R. K. Haddad, Z. Harun, C. Maluk and M. R. Rasani
ε
⎛ Q∗ 2 3 ⎞
K 2 = gγ ⎜⎜ 1 3 ⎟⎟ , (10)
⎝ Fr ⎠
where γ is heat loss coefficient.

With the help of dimensionless analysis and small-scale experiments, Li


et al. [5] further proposed a well-known prediction model for the
dimensionless backlayering length as a function of the dimensionless heat
release rate and the dimensionless ventilation velocity:
⎧ ⎛ 180Q∗1 3 ⎞ ∗
⎜ ⎟



18. 5 ln ⎜ V ∗ ⎟ for Q ≤ 0.15
Lb = ⎨ ⎝ ⎠ (11)
⎪18.5 ln ⎛ 0 . 43 ⎞
⎜ ∗ ⎟ for Q∗ > 0.15.
⎪⎩ ⎝V ⎠

1.2. Blockage effect


The blockage effect section discusses critical velocity and smoke
backlayering distance in a tunnel when blockages and vehicles are
introduced. In a real tunnel fire scenario, vehicles often get stuck in the
tunnel at the upstream of the fire location, making the ventilation situation
more complex. Oka and Atkinson [7] investigated the effect of a solid
obstruction on critical velocity in a small-scale model tunnel. They found
that the reduction ratio of U c was 15% when a vehicle occupied roughly
12% of the tunnel cross-sectional area and was 40-45% when a vehicle
occupied roughly 32% of the tunnel cross-sectional area. A model vehicle
occupied 20% of the tunnel cross-sectional area, which is placed 40 mm
above the ground, and its influence on critical velocity and backlayering
length was investigated by Li et al. [5]. They demonstrated that U c
decreases about 23%, which was slightly greater than the ratio of the cross-
sectional area of a vehicle to the tunnel cross-sectional area. The following
equation was proposed for critical velocity with vehicular blockage [5]:
⎧ 1
⎪0.63Q∗ 3
∗ for Q∗ ≤ 0.15,
V =⎨ (12)
⎪⎩0.33 for Q∗ > 0.15.
Experimental Study of the Influence of Blockage … 457

Li et al. [5] also defined V ∗∗ , which is a ratio of the longitudinal


velocity, V, to its corresponding critical velocity, U c . The dimensionless

backlayering distance was found to be a function of V ∗∗. The following


empirical correlation for backlayering length with vehicular blockage was
proposed [5]:

V ∗∗ = e − 0.074 Lb . (13)
Small-scale experiments and numerical simulations were carried out to
study the influence of vehicular blockage on the tunnel fire behavior and
critical velocity in the tunnel with a longitudinal ventilation system [12]. Lee
and Tsai [12] applied different arrangement of vehicles which occupied 3-
31% of the cross-section area. They found that when ventilation flow could
reach the fire source, critical velocity decreased with the existence of
blockages. Whereas, in the case that vehicles prevent direct ventilation flow
to the heat source, critical velocity increased. It was also concluded that the
backlayering length became shorter with an increase in blockage ratio
compared to the situations of no blockages in the tunnel. However,
backlayering length is larger with the fire source located directly at the
downstream of the blockage. Alva et al. [13] produced buoyant plume by a
helium-air mixture to investigate the effect of vehicles on longitudinal
ventilation control in tunnel fires. Blockage ratios between 0.1 and 0.41 were
utilized and it was concluded that the backlayering length could be
influenced by the vehicular blockage and the relative distance between the
fire source and vehicular blockage. A quantitative data in the case of the
blockage influence was provided; however, the mechanism of vehicular
blockage effect and different kinds of vehicles in previous studies [5, 7, 12]
has not been discussed adequately and needs more investigation about the
impact of various vehicle arrangements on critical velocity or the length of
obstacle on backlayering length. This study investigates the effect of
percentage occupancy of tunnel cross section and different vehicle layout on
critical velocity and the influence of blockage and its length on backlayering
length as well as the relationship between the dimensionless backlayering
length and dimensionless critical velocity experimentally.
458 R. K. Haddad, Z. Harun, C. Maluk and M. R. Rasani
2. Reduced-Scale Experiment

2.1. Model tunnel system


A series of experiments are conducted in a 1:50 reduced scale tunnel
with 3 m × 0.6 m × 0.95 m (length × width × height ). This model follows the
1:50 scale model of the Resalat transportation tunnel in Tehran, Iran. The
Froude number modeling is used to relate the model to the full-scale tunnel
in reduced- scale experiments. Froude modeling has been widely utilized in
reduced-scale model tunnel fires [7]. Whereas, with a smaller scale like the
one used in this experiment, Froude and Richardson numbers and high
Reynolds number will be approximately preserved. The Froude number
modeling is utilized because inertial and buoyancy forces are two influential
forces in tunnel fire (the Froude number is the ratio of inertial forces to
gravity forces). The scaling of the heat release rate and smoke velocity
between reduced-scale and full-scale based on the Froude modeling could be
presented as follows [7]:

VM = VF LM LF , (14)

52
⎛L ⎞
QM = QF ⎜ M ⎟ , (15)
⎝ LF ⎠

where the subscripts M and F indicate model scale and full scale
respectively. The tunnel ceiling, floor and sidewalls are made by fireproof
boards and half of the side wall is made of tempered glass in order to the fire
and smokes movements were visible during experiments. Pool fires are used
in this study to simulate real tunnel fires. The square burner, with three
different dimensions of 8, 10, and 13 cm are used, and the fuel was gasoline.
These sizes of pools simulate fires with heat release rates of 39 MW, 74.5
MW, and 158.3 MW in a typical tunnel. The fuel pans, which are placed on
the floor of the tunnel, are 2 cm deep and the fuel in each pan is 1 cm deep.
The longitudinal ventilation is produced by an axial fan installed at the left
end. Two wire mesh screens with sizes of 1 mm and 0.5 mm for coarse and
fine screens and one honeycomb mesh with an aspect ratio of 6.5 are placed
Experimental Study of the Influence of Blockage … 459
between the fan and the test section to straighten flow and maintaining a
nearly homogeneous airflow throughout the test section, which is critical to
ensure the accuracy of the tests.

2.2. Measurement system


Upstream backlayering flow of hot gases is detected by using one array
of eight K-type stainless-steel sheathed thermocouples of 0.3 mm diameter.
Chain of temperature measurement included K-type thermocouples, Arduino
MEGA 2560, and MAX6675K amplifier. The thermocouple error is 2% for
sheathed thermocouple with twisted wire and it is approximately 0.29% and
0.25% for MAX6675K amplifier and Arduino MEGA 2560, respectively.
Smoke temperature distribution beneath the tunnel decreases sharply at the
end of smoke reverse flow which can be used to measure backlayering
length. Thermocouples are placed 10 mm below the tunnel ceiling with 17
cm intervals in the center line of the ceiling. One of the eight thermocouples,
T1, is installed above the fire source and the others, T2-T8, are installed after
T1 in the upstream direction (Figure 1). The temperature read by T1 is the
temperature above the fire plume and the distance of backlayering thermal
fumes is defined as the maximum distance between T1 and the thermocouple
which recorded quasi-steady ambient temperature. Figure 1 shows the layout
of the tunnel, the fire source, and the eight thermocouples.

Figure 1. Cross section of model tunnel.

Table 1 summarizes the uncertainties of this experiment.

Table 1. The uncertainties of this experiment


Instrument Uncertainties
K-type Thermocouple 2%
Arduino MEGA 2560 0.25%
Amplifier MAX6675K 0.29%
460 R. K. Haddad, Z. Harun, C. Maluk and M. R. Rasani
2.3. Experimental set-ups
To simulate the influence of vehicles, experimental tests with various
sizes of blockages are carried out. The vehicle models consisted of three
sizes, representing the sedan, bus, and truck. The dimensions of sedan are
0.38 m (L ) × 0.15 m ( W ) × 0.12 m (H ). For the bus: 0.54 m × 0.17 m × 0.21 m
and for the truck: 0.61 m × 0.19 m × 0.20 m. This represents a 1:12 model of
typical vehicles. The purpose of this dimension (1:12 scale) is to examine the
effect of different cross-sectional occupancy percentages of the tunnel cross-
sectional area rather than to simulate conditions close to actual conditions.
Although, further investigations are needed to explore the behavior of the
smoke flow in conditions close to real scenarios for safe and efficient design
of tunnel fire. The blockages are positioned in one, two, and three arrays on
the floor of the tunnel to simulate a tunnel with various lanes of vehicles. The
impact of vehicles in the tunnel on critical velocity and backlayering length
is investigated in six scenarios. The model vehicles occupy approximately
from 3 to 15% of the tunnel cross-sectional area. The position of blockages
related to the fire source is defined through two expressions; “obstructed”
when ventilation velocity cannot reach the flame directly and “non-
obstructed” means the direct effect of ventilation on the fire source. Figure 2
shows schematic diagrams of vehicle blockage in the tunnel cross-section.
Moreover, Table 2 summarizes the blockage ratio of all scenarios.

Figure 2. Blockage locations.


Experimental Study of the Influence of Blockage … 461
Table 2. Blockage ratio of various scenarios in this study
Scenario Vehicles Lanes Occupied Blockage Ratio Obstruction
1 Sedan 1 3.4% Non-obstructed
2 Sedan 1-3 6.8% Non-obstructed
3 Sedan 1-2 6.8% Obstructed
4 Sedan-Bus 1-3 9.6% Non-obstructed
5 Bus-Truck 1-3 12% Non-obstructed
6 Bus-Truck-Sedan 1-2-3 15% Obstructed

3. Results and Discussion

Before examining the influence of blockages on critical velocity and


backlayering length, it is necessary to define the parameters which are used
in the various stages of this study.

⎛ AT − Ablkg ⎞
Blockage percentage (α% ) = ⎜⎜ ⎟⎟ . (16)
⎝ AT ⎠

Figure 3 shows the cross-sectional area of the tunnel and blockage where
AT is the cross-sectional area of the tunnel and Ablkg is the total cross-
sectional area of blockages.

Figure 3. The cross-sectional area of the tunnel and blockage.


Tunnel blockage ratio is defined as the ratio of the cross-sectional area of
the fire source to that of the tunnel [14].

Ablkg
Tunnel blockage ratio (ϕ ) = . (17)
AT
462 R. K. Haddad, Z. Harun, C. Maluk and M. R. Rasani
Local ventilation velocity in the vicinity of the fire source would be
changed through the cross-sectional area of blockages as the following
equation [14]:

⎛ Ablkg ⎞
ulocal = (1 − ϕ) u = ⎜⎜1 − ⎟⎟ u. (18)
⎝ AT ⎠

3.1. Critical velocity

3.1.1. Influence of tunnel blockage percentage on critical velocity


Figure 4-a shows that critical velocity decreases with increasing tunnel
blockage ratio when the fire source is in the middle lane and obstacles are
located on the lanes on the sides. This change is due to the decrease in the
cross-sectional area for air entertainment, which reduces the fire size. Li et al.
[5] and Oka and Atkinson [7] also obtained that critical velocity experiences
a reduction when a vehicle occupied the tunnel cross-sectional area (Figure
4-b).

Figure 4. Effect of tunnel blockage ratio on the change ratio of U c . (Dashed


lines are polynomial second order fitting lines).
The change in critical velocity as a result of increasing the cross-
sectional area of blocks can also be investigated by the dimensionless critical
velocity ratio, which is a ratio between the dimensionless critical velocity in
cases with vehicular blockage and the corresponding dimensionless critical
velocity in the same tunnel without vehicular blockage (Eqs. 19-20), as a
function of vehicle blockage ratio (1 − ϕ ) (Figure 4).
Experimental Study of the Influence of Blockage … 463
uc , b
uc∗, b = , (19)
gH

uc
uc∗ = . (20)
gH

Figure 5. Dimensionless critical velocity as a function of 1 − ϕ (Dashed


lines show second order polynomial fitting lines for non-obstructed flows
and solid lines show the fitting lines for obstructed flows).

In Figure 5, the dimensionless critical velocity ratio (uc∗, b uc∗ ) decreases


with increasing vehicular blockage ratio when the longitudinal air stream is
not affected directly by vehicular blockage and the ventilation velocity
reaches the fire plume. The reduction ratio of the dimensionless critical
velocity is approximately the same as vehicle blockage ratio for various fire
sizes in this situation (non-obstructed). On the other hand, when the obstacles
prevent direct access of ventilation flow to the fire source (2.21 kW
obstructed and 4.22 kW obstructed in Figure 5, the dimensionless critical
velocity ratio increases with increasing vehicle blockage ratio. It reported
that the critical velocities increase in cases with blockage ratio up to 20-30%,
as discussed by Lee and Tsai [12].

3.1.2. Reduction ratio


To further analyze the impact of vehicles on critical velocity, the
464 R. K. Haddad, Z. Harun, C. Maluk and M. R. Rasani
reduction ratio of critical velocity due to the obstruction is defined as
follows:

(uc − uc, b )
ε= . (21)
uc

This parameter is plotted versus the dimensionless ratio of the


obstruction area to the tunnel cross-sectional area. As shown in Figure 6, it is
clear that for 1 − α < 10%, the declination rate of critical velocity is
approximately equal to the tunnel blockage ratio and when 1 − α > 10%, the
change ratio of U c due to vehicle obstruction is slightly greater than the ratio
of cross-sectional area of obstructions to the tunnel cross-sectional area. The
critical 1 − α = 10% is shown by the dashed line in Figure 6.

Figure 6. The reduction ratio of critical velocity due to obstruction versus


the dimension ratio of obstruction.

3.1.2. Influence of blockage position on the critical velocity


Figure 7 shows the influence of the location and the cross-sectional
percentage occupied by obstacles on critical velocity for the occupancy rate
from 3 to 15%. In this analysis, the relationship between the location of
blockages and the location of the fire source is defined by two terms;
“obstructed” is the situation that the ventilation velocity cannot reach directly
the fire source because the obstacles which barrier ventilation flow (Scenario
3 and 6 in Table 2) and “non-obstructed” means the direct effect of
Experimental Study of the Influence of Blockage … 465
ventilation on the flame, in fact, the vehicles are located on the sidelines. In
Figure 7, in the obstructed mode, critical velocity increases with increasing
occupancy of the tunnel cross-sectional area by vehicles. In this case, the fire
flame is not exposed by the obstacles of the ventilation flow, and as a result,
smokes rapidly diffuse in this area and more ventilation flow is required to
prevent backlayering effect since part of the air flow is blocked as was found
out by Alva et al. [13]. However, critical velocity decreases with increasing
blockage ratio when the air flows directly to the flame. This decrease is due
to the reduction in the amount of air entering into the vicinity of fire source,
which reduces the burning rate, fire size and therefore the buoyancy force.
The data for 2.21 kW, 4.22 kW, and 8.98 kW fires show similar. Therefore,
the impact of the blockages on critical velocity depends on the location of the
obstacles and spatial relationship between the blocks and the source of the
fire.

Figure 7. Blockage position effect on critical velocity; (a: case 1, 2, 4, 5 and


b: case 3 and 6).

3.2. Backlayering length


The influence of obstacles and the fire source location relative to
vehicular blockages were also considered when analyzing backlayering
length was studied.

3.2.1. Effect of ventilation velocity and heat release rate on backlayering


length
Backlayering length in a tunnel with vehicular blockages is influenced
by the longitudinal ventilation velocity as well as the heat release rate as
466 R. K. Haddad, Z. Harun, C. Maluk and M. R. Rasani
shown in Figure 8. When ventilation velocity increases, the backlayering
length decreases. For instance, when the fire size is 2.21 kW and velocity is
0.2 m/s, the backlayering length is 84 cm while it is 16 cm at 0.8 m/s. On the
other hand, backlayering length is longer with an increase in the heat release
rate of less than 4.22 kW. For example, backlayering length is almost 51 cm
at 4.22 kW heat release rate and at 0.6 m/s whereas it is 17 cm at 2.21 kW
and 0.6 m/s. Because the larger fire produces more buoyancy force, which
causes more smoke spreads along the ceiling of the tunnel. While, when
HRR is more than 4.22 kW, backflow length is almost independent of HRR
and ventilation velocity changes only influence backlayering length.

Figure 8. The effect of ventilation velocity and the heat release rate on the
backlayering length.

3.2.2 Effect of blockages on backlayering length


When an obstacle is placed inside the tunnel, less amount of ventilating
air reaches the vicinity of fire source. As a result, more heat is transmitted
through convection and radiation, this reduces the length of backflowing
smoke. This decrease in backlayering length in Figure 9 is clearly visible. In
other words, by placing vehicles in the tunnel, the convection heat transfer
increases and the reducing of the cross-sectional area of the tunnel by
blockages causes that local velocity near the fire source increases. Moreover,
the body of obstacles also helps heat transfer. All these factors lead to less
ventilation to prevent smoke emissions.
Experimental Study of the Influence of Blockage … 467

Figure 9. The effect of blockages on the backlayering length. (a: 2.21 kW, b:
4.22 kW and c: 8.98 kW).

3.2.3. Dimensionless analysis of backlayering length


For further investigation of the impact of obstacles on backlayering
distance, a dimensionless analysis is also carried out. Parameters which effect
on backlayering length are the heat release rate, ventilation velocity, tunnel
geometry, ambient properties and gravitational acceleration. Therefore, based
on the simple dimensional analysis presented in Li et al. [5], the
dimensionless backlayering length can be calculated through the
dimensionless critical velocity. In this section, the change in the
dimensionless backlayering length is plotted against the dimensionless
critical velocity. The dimensionless critical velocity concept is chosen to
468 R. K. Haddad, Z. Harun, C. Maluk and M. R. Rasani
investigate the dependency of the longitudinal velocity to the backlayering
distance. The backlayering length is studied regarding the dimensionless
critical velocity in two cases; without blockages and with blockages. The
dimensionless backlayering length is defined as follows:
L
L∗b = b . (22)
H

The dimensionless critical velocity (V ∗∗ ) is:

u∗ u
V ∗∗ = ∗ = . (23)
uc u c

Figure 10 shows the relationship between the dimensionless backlayering


length and the dimensionless critical velocity in both with and without
blockages situations. It can be seen that the obtained data follow the same
exponential trend as Li et al. [5] have observed.

Figure 10. Dimensionless critical velocity vs. dimensionless backlayering


length; Dashed lines show the exponential fitting lines.
Experimental Study of the Influence of Blockage … 469
Experimental data of correlation between the dimensionless backlayering
length and the dimensionless critical velocity for tests without a vehicle can
be expressed into a universal form. The suggested equation is as follows:

V ∗∗ = 1.2363e −1.574 Lb . (24)

The correlation coefficient is 1.2363 for Eq. 24. Figure 10 also shows the
dimensionless backlayering length versus the dimensionless critical velocity
with blockages. Based on Figure 10, when the ventilation velocity increases,
the backlayering length decreases until it disappears as the ventilation
velocity approaches the critical velocity. Comparison of the dimensionless
backlayering length at certain dimensionless critical velocities indicates that
backlayering length is reduced through obstacles, which confirms the results
of the previous section. In other words, blockages show an influence on the
backlayering distance due to larger convection and radiation heat losses
when obstacles are placed in the tunnel. An equation that is able to predict
the dimensionless backlayering length with the presence of blockages is as
follows:

V ∗∗ = 1.15e −1.385 Lb . (25)

In Figure 11(a), the experimental results for the situation without


blockages are compared with values of Li et al. [5] model for the lower Q ∗
regime. It is clear that the obtained data follows roughly the same trend when
there are no vehicles in the tunnel. However, a certain degree of dispersion is
caused due to different experimental scales, 1:50 (in this experiment)
compared with 1:20 and 1:23 (in Li et al. experiment [5]). As the scale
becomes smaller, a higher degree of dispersion can be estimated and also the
results of this experiment underestimate the backlayering length.
The comparison between the obtained dimensionless backlayering
distance with blockage as a function of the dimensionless critical velocity of
this study and predictions of Li’s model [5] is represented in Figure 11. It can
be seen that there are differences between current experimental data and the
reference correlation due to different vehicle blockage ratios utilized in this
470 R. K. Haddad, Z. Harun, C. Maluk and M. R. Rasani
study and the relative position of blockages attributed to the fire source,
which are “obstructed” and “non-obstructed” as mentioned. This means that
the relationship presented in Li’s study [5] cannot be deduced for different
blockage ratios.

Figure 11. Dimensionless backlayering length as a function of dimensionless


critical velocity compared with Li’s study [14]; (a): Without blockages. (b):
With blockages.

3.2.4. Backlayering length analysis regarding its relationship with


blockage section
In the case of a tunnel fire, when smoke impinges on the ceiling, hot
gases accumulate and spread along the ceiling and smoke backlayering forms
Experimental Study of the Influence of Blockage … 471
if the longitudinal ventilation velocity is lower than the critical value. The
driving force of smoke is caused by the temperature difference between
smoke and ambient air, i.e., buoyancy. At the impinging point, smokes have
its highest temperature difference compared to ambient temperature. As
smokes travel backward, the temperature difference declines because of heat
transfer to ventilated air and tunnel ceiling, resulting in a decrease in
buoyancy. In other words, the temperature difference at the point of
impinging is ΔTmax with the highest driving force f max , and its heat release
rate can move smoke L meters long. At x meters distance from ΔTmax point,
the temperature difference between backlayering and ambient air is ΔTx and
the driving force is f x . This driving force causes that smoke spreads L-x
meters. When blockages are in the tunnel, the conditions are different. The
ventilation velocity increases due to the smaller open cross-sectional area of
the tunnel. Consequently, heat convection is enhanced through higher
ventilation velocity. When backflow smoke reaches blockages, in turn, leads
to a faster decay of temperature, buoyancy force and subsequently the
released heat rate. This region of fast decay is shown in Figure 12. On the
other hand, however, the presence of obstacles also influences the heat
transfer in the area before blockages, the region before obstacles have slower
temperature decay. Therefore, in order to study backlayering length
considering the impact of blockages more accurately, each scenario is
divided into two sections. The first part includes tests which smoke does not
reach the blockage region however the presence of obstacles causes that the
1
velocity of ventilated air will become times of that without blockage
1− ϕ
as explained at the beginning of Section 3. The smoke backlayering length
will be affected at different ventilation velocity. However, the enhanced heat
loss from smoke to ventilated air is lower than the case which smoke covers
the blockage section. Therefore, the second part involves tests that smoke
flows in the area of obstacles are also studied (Figure 12).
Figure 13 compares the relationship between the dimensionless
backlayering length and the dimensionless critical velocity in both fast decay
472 R. K. Haddad, Z. Harun, C. Maluk and M. R. Rasani
and slow decay areas. Since the fast decay area has a higher heat release rate,
backlayering length is shorter than slow decay area, which is clear from the
exponential fitting equation of both graphs. As shown in Figure 13, decay
factor kb is smaller than the decay factor kT , which indicates a decrease in
the backlayering length in a tunnel with vehicles. In the second section, when
backflow smoke reaches the blockage region and covers both fast decay and
slow decay areas (both blockage and non-blockage sections fill with smoke),
the heat transfer is relatively stronger than the previous one due to higher
ventilation velocity, and as a result, the temperature reduction factor is
greater than the situation when backlayering does not reach the blockage
section.

Figure 12. Smoke backlayering length in the tunnel with long blockage.

Figure 13. The comparison of smoke backlayering in the fast decay section
with slow decay section.
Experimental Study of the Influence of Blockage … 473
The relationship of the backlayering length in slow decay area (smoke
backlayering shorter than blockage section) is also compared with the
relationship of the backlayering length in the tunnel without obstacles in
Figure 14. Figure 14 indicates that obstacles influence the backlayering
length in both fast and slow decay regions, although it does not reach the
blockage region in slow decay section.

Figure 14. The comparison of smoke backlayering in the slow decay area
with no-blockage situation.

4. Conclusion

Experimental model-scale tests were carried out to examine the impact of


vehicle blockages on critical velocity and backlayering length. Three
vehicles sizes in two or three arrays, occupying 3-15% of the tunnel cross-
sectional area were tested.
The fire sources were positioned at the tunnel center line and vehicles are
positioned in the middle lane or lanes on the sides within six different
scenarios. The fuel was gasoline put in square pans with dimensions of 8 × 8
cm 2 , 10 × 10 cm 2 , and 13 × 13 cm 2 . The following summary based on
experimental data:
474 R. K. Haddad, Z. Harun, C. Maluk and M. R. Rasani
• The effect of blockages on critical velocity is influenced by the
positions of the fire source and vehicles. The value of critical
velocity decreases when the fire source is located in the middle lane
and the ventilated air can reach the fire source directly. A decrease in
the cross-sectional area and air entertainment, which reduces the fire
size, causes a decrease in critical velocity.

• When the ventilation flow cannot reach the fires, critical velocity
increases as the heat release rate enhances and smokes rapidly
diffuse in this area. Therefore, more ventilation flow is required to
prevent backlayering length.

• Experimental results demonstrate that the length of backlayering


decreases in the presence of vehicles for certain ventilation
velocities.

• Experimental data also point out that there is an exponential


relationship between the ratio of longitudinal ventilation velocity to
critical velocity and the dimensionless backlayering length.

• The analyzing of the relationship between dimensionless critical


velocity and dimensionless backlayering considering the relationship
between backlayering length and the distance where occupied by
obstacles indicates that L∗b follows an exponential equation with
different decay factor in fast decay area and slow decay area which
presents the influence of both obstacles and heat release to
surroundings on backlayering length.

Acknowledgements

We are extremely grateful for the financial assistance provided by the


FRGS/1/2016/TK03/UKM/03/1 grant from the Ministry of Higher Education
Malaysia (MOHE) and GUP-2018-102 grant from Universiti Kebangsaan
Malaysia.
Experimental Study of the Influence of Blockage … 475
References

[1] L. H. Hu, Wei Peng and Ran Huo, Critical wind velocity for arresting upwind gas
and smoke dispersion induced by near-wall fire in a road tunnel, Journal of
Hazardous Materials 150(1) (2008), 68-75.
[2] L. H. Hu, Ran Huo and W. K. Chow, Studies on buoyancy-driven back-layering
flow in tunnel fires, Experimental Thermal and Fluid Science 32(8) (2008),
1468-1483.
[3] Y. B. M. Z. Wu and M. Z. Abu Bakar, Control of smoke flow in tunnel fires using
longitudinal ventilation systems-a study of the critical velocity, Fire Safety Journal
35(4) (2000), 363-390.
[4] W. Tang, L. H. Hu and L. F. Chen, Effect of blockage-fire distance on buoyancy
driven back-layering length and critical velocity in a tunnel: an experimental
investigation and global correlations, Applied Thermal Engineering 60(1-2)
(2013), 7-14.
[5] Ying Zhen Li, Bo Lei and Haukur Ingason, Study of critical velocity and
backlayering length in longitudinally ventilated tunnel fires, Fire Safety Journal
45(6-8) (2010), 361-370.
[6] Philip H. Thomas, The movement of smoke in horizontal passages against an air
flow, Fire Safety Science 723 (1968), 1-1.
[7] Yasushi Oka and Graham T. Atkinson, Control of smoke flow in tunnel fires, Fire
Safety Journal 25(4) (1995), 305-322.
[8] P. H. Thomas, The movement of buoyant fluid against a stream and the venting of
underground fires, Fire Safety Science 351 (1958), 1-1.
[9] Haukur Ingason and Ying Zhen Li, Model scale tunnel fire tests with longitudinal
ventilation, Fire Safety Journal 45(6-8) (2010), 371-384.
[10] Miao-Cheng Weng, Xin-Ling Lu, Fang Liu, Xiang-Peng Shi and Long-Xing Yu,
Prediction of backlayering length and critical velocity in metro tunnel fires,
Tunnelling and Underground Space Technology 47 (2015), 64-72.
[11] F. Tang, L. J. Li, F. Z. Mei and M. S. Dong, Thermal smoke back-layering flow
length with ceiling extraction at upstream side of fire source in a longitudinal
ventilated tunnel, Applied Thermal Engineering 106 (2016), 125-130.
[12] Yee-Ping Lee and Kuang-Chung Tsai, Effect of vehicular blockage on critical
ventilation velocity and tunnel fire behavior in longitudinally ventilated tunnels,
Fire Safety Journal 53 (2012), 35-42.
476 R. K. Haddad, Z. Harun, C. Maluk and M. R. Rasani
[13] W. Ulises Rojas Alva, Grunde Jomaas and Anne S. Dederichs, The influence of
vehicular obstacles on longitudinal ventilation control in tunnel fires, Fire Safety
Journal 87 (2017), 25-36.
[14] Liming Li, Xudong Cheng, Yu Cui, Sen Li and Heping Zhang, Effect of blockage
ratio on critical velocity in tunnel fires, Journal of Fire Sciences 30(5) (2012),
413-427.

You might also like