Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

International Journal of Pressure Vessels and Piping 187 (2020) 104159

Contents lists available at ScienceDirect

International Journal of Pressure Vessels and Piping


journal homepage: http://www.elsevier.com/locate/ijpvp

Numerical study on the impact distance of a jet fire following the rupture of
a natural gas pipeline
Ke Shan a, b, *, Jian Shuai a, Guang Yang b, Wei Meng b, Chen Wang b, Jixiang Zhou b, Xu Wu a,
Lei Shi c
a
College of Safety and Ocean Engineering, China University of Petroleum, Beijing, 102249, China
b
Shenzhen Gas Corporation Ltd., 518049, Guangdong, China
c
Dalian Research Institute of Petroleum and Petrochemicals, SINOPEC, 116045, Dalian, China

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

Keywords: As a result of the catastrophic damage to society, the economy and the environment induced by pipeline fire and
Jet fire explosion accidents, determining the impact distance of a natural gas pipeline has attracted increasing attention.
Gas pipeline In this paper, first the impact distances of some typical natural gas pipeline fire and explosion accidents are
Impact distance
statistically reviewed. Then, a 3D natural gas pipeline jet fire accident scenario model is established with FLACS
Accident scenario
software to simulate the entire dynamic process from gas leakage, diffusion, ignition to jet fire formation. The
Risk assessment
numerical results, such as the flame appearance, dimensions and thermal radiation, are compared with the
experimental results in the literature, and very good agreement is obtained. Thereafter, a parametric study is
performed by varying the pipe diameter, internal pressure, and wind speed. According to the thermal flux level, a
method is proposed for determining the thermal radiation impact distance of a jet fire as a function of the in­
ternal pressure, pipe diameter and wind speed. This method provides technical support for risk assessment and
accident consequence predictions for natural gas pipelines.

1. Introduction produce jet fires, vapor cloud fires, or fireballs [4]. This paper focuses on
discussing jet fires following the rupture of natural gas pipelines. A series
As an urban public infrastructure, gas pipelines are a double-edged of controlled experiments have been conducted to analyze the jet flame
sword for which the advantages outweigh the disadvantages. Pipelines length and thermal radiation [5–11]. However, most of the flame length
are regarded as one of the most practical and economical ways to and thermal radiation results from these studies are in the form of the
transport dangerous and combustible substances, such as oil and natural mass release rate rather than the basic parameters of the pipeline, such
gas, which are often impractical to transport via road or rail [1–3]. The as the internal pressure and pipe diameter. The relationship between the
increasing use of natural gas makes gas pipelines an indispensable life­ impact distance of a jet fire and the basic parameters of a pipeline has
line in the development of urban modernization. However, not been discussed in depth, which is a focus of pipeline operators.
high-pressure natural gas pipeline accidents, such as leakages, fires and Lowesmith has carried out many high-pressure natural gas pipeline
explosions, are major threats to residents around pipelines [2]. In recent jet fire experiments [12–14]. The key objective of their work was to
years, the endless stream of accidents associated with gas pipelines has provide large-scale experimental data on the characteristics of jet fires,
highlighted the importance of researching safe distances from pipelines. which were used to validate a mathematical model for predicting fire
Therefore, it is significant to study the impact distance of jet fires risks following high-pressure pipeline failures. Significant results were
following the rupture of high-pressure natural gas pipelines. achieved in these experiments [12]. These experimental data provided a
A major challenge in reducing the consequences of pipeline accidents validation basis for determining the impact distance of a high-pressure
is controlling the impacts from large, open natural gas fires. Depending natural gas pipeline. In addition, based on experimental research, an
upon the circumstances and conditions leading to such a pipeline acci­ empirical formula of the jet flame length for a natural gas pipeline
dent, the type of open fire may vary. For example, ignited releases can depending on the internal pressure and pipe diameter was proposed by

* Corresponding author. College of Safety and Ocean Engineering, China University of Petroleum, Beijing, 102249, China.
E-mail address: 965610446@qq.com (K. Shan).

https://doi.org/10.1016/j.ijpvp.2020.104159
Received 17 May 2020; Received in revised form 13 July 2020; Accepted 17 July 2020
Available online 2 August 2020
0308-0161/© 2020 Elsevier Ltd. All rights reserved.
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

Table 1
Statistics on the impact distance of typical gas pipeline accidents.
Country or Accident date Accident location Accident type Pipe Internal Impact distance and characteristic description
region diameter/ pressure/
mm MPa

America 08/19/2000 Carlsbad, New Natural gas pipeline 762 4.65 The victims were camped approximately 206 m from the
Mexico rupture and fire crater. Two of the pieces of pipe were thrown 71 and 87 m.
April 05, Palm City, Florida Gas transmission pipeline 457 5.9 The potential impact radius was 112 m.
2009 rupture and natural gas
release
July 06, 2010 Cleburne, Texas Natural gas transmission 914 6.55 The eruption of high-pressure gas threw a 60-ton auger truck
pipeline rupture and fire more than 30 m and violently ignited.
September San Bruno, Natural gas transmission 762 2.7 The fire damage extended to a radius of approximately 183 m
09, 2010 California pipeline rupture and fire from the pipeline blast center.
November Sissonville, West Gas transmission pipeline 508 6.4 An area of fire damage approximately 250 m wide extended
12, 2012 Virginia rupture and fire nearly 335 m along the pipeline right-of-way.
Canada December 09, Swastika, Ontario Natural gas pipeline 914 6.87 Minor exterior damage occurred to a house located
2009 rupture, explosion and approximately 320 m north of the pipeline rupture.
fire
09/26/2009 Marten River, Natural gas pipeline 762 5.23 Pieces of pipe were found up to approximately 100 m away
Ontario rupture from the site of the rupture.
02/19/2011 Beardmore, Natural gas pipeline 914 6.62 One wrinkled pipe segment was ejected during the explosion
Ontario rupture, fire and and was found approximately 100 m north-east. The natural
explosion gas fire caused a secondary brush fire covering an area of
approximately 0.56 ha.
Europe 07/30/2004 Ghislenghien, Natural gas pipeline 1000 8 Vegetation was burned in a radius of 200–300 m around the
Belgium rupture, explosion and event. Flames were reported to reach a height of 200 m.
fire
China April 05, Qinhuangdao, Natural gas pipeline 1016 5 A pigsty approximately 100 m away was destroyed.
2010 Hebei rupture, explosion and
fire
February 07, Qinglong, Natural gas pipeline 1016 7 The glass of a house 290 m away was broken, and the leaves of
2017 Guizhou rupture, explosion and trees 350 m away were withered.
fire
October 06, Qinglong, Natural gas pipeline 1016 8 Several slab houses built 350 m away were damaged.
2018 Guizhou rupture and fire

Fig. 1. Gas leakage from a pipe.

Lowesmith [14]. This empirical formula can be applied to calculate the


jet flame length if the leakage hole is small because Lowesmith’s
calculation of the mass flow rate was based on a small-hole-leakage
Fig. 2. Selection of ray directions for DTM.
model. However, for large hole leakage and ruptured pipeline leakage,
this empirical formula over predicts the jet flame length. Large hole
leakage and ruptured pipeline leakage models have been widely
considered to be a regular circle because the wind around the pipeline
considered in the literature [15–20]. Montiel et al. [19] developed a
was not considered.
model for calculating gas discharge rates through a small hole or a full
Many process industry hazards occur in accident scenarios involving
borehole. In this paper, the mass flow rate of gas was determined using
fluid flow in complex, large-scale, three-dimensional (3D) geometries.
Montiel’s research results.
FLACS is a specialized CFD toolbox developed to address process safety
An empirical formula of the potential impact radius for natural gas
applications, such as gas explosions and jet fires [22]. Fire and explosion
pipeline accidents was proposed in ASME B31⋅8S-2016 [21]. This for­
accidents have been numerically analyzed in several studies [23–29].
mula was derived from the point source model of fire thermal radiation.
In this paper, a numerical method is used to study the impact dis­
However, in this model, the thermal radiation impact range was
tance of jet fires following the rupture of natural gas pipelines. A 3D

2
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

Fig. 3. Schematic of the jet fire model for a completely ruptured natural gas pipeline.

Fig. 4. Distribution of radiometers.

natural gas pipeline jet fire accident scenario model is established with fire and explosion accidents has been described in some accident
FLACS software, and the experimental data provided by Lowesmith are investigation reports. These statistical data can provide technical sup­
used to validate this jet fire model. Moreover, formulas for determining port for determining the safe distance from gas pipelines. Statistical
the thermal radiation impact distances with respect to different thermal results of the impact distance of typical gas pipeline accidents are shown
flux levels are proposed, depending on the variations in internal pres­ in Table 1 [30,31].
sure, pipe diameter and wind speed. An empirical formula for the potential impact radius r (m) of natural
gas pipeline accidents was proposed in ASME B31⋅8S-2016 [21]:
2. Problem description √̅̅̅
r = 99D P, (1)
In recent years, fire and explosion accidents from gas pipelines have
where P is the maximum allowable operating pressure of the pipeline
been common throughout the world. The impact distance of gas pipeline

3
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

Table 2 segment (MPa) and D is the pipe diameter (m).


Cases considered for the ruptured natural gas pipeline jet fire model. The formula proposed in ASME B31⋅8S-2016 is derived considering a
Case Internal Pipe Wind Mass flow Leak point source model for fire thermal radiation and does not account for
pressure/MPa diameter/m speed/ rate/(kg⋅s− 1) area/ wind around a pipeline. However, the thermal radiation impact distance
(m⋅s− 1) m2 increases significantly with the wind speed. In this paper, a formula is
1 1 0.8 5 176.14 0.51 proposed for determining the thermal radiation impact distance S (m)
2 2 0.8 5 352.28 1.33 for different thermal flux levels that depends on the internal pressure,
3 3 0.8 5 528.42 2.21 pipe diameter and wind speed u (m/s):
4 4 0.8 5 704.56 3.09
5 5 0.8 5 880.70 3.96 S = f (D, P, u). (2)
6 6 0.8 5 1056.85 4.80
7 6 0.1 5 5.07 0.20
8 6 0.2 5 30.43 0.30 3. Methodology
9 6 0.3 5 86.33 0.37
10 6 0.4 5 180.35 0.79 3.1. Hole-pipe model
11 6 0.5 5 319.39 1.42
12 6 0.6 5 508.33 2.29
13 6 0.7 5 752.63 3.41
The “hole model” is used to calculate the mass flow rate for a pipe
14 6 0.8 5 1056.85 4.80 with a small leakage hole [15–17]. The “pipe model” is used to calculate
15 6 0.8 1 1056.85 4.80 the mass flow rate for a completely ruptured pipe [15,16]. These two
16 6 0.8 2 1056.85 4.80 models are combined in the “hole-pipe model” [19], which reproduces
17 6 0.8 3 1056.85 4.80
the “hole model” prediction for small holes and the “pipe model” pre­
18 6 0.8 4 1056.85 4.80
19 6 0.8 5 1056.85 4.80 diction when the hole diameter is that of the pipe. This classical model is
20 6 0.8 6 1056.85 4.80 applicable to large hole leakage. Therefore, the “hole-pipe model” is
21 6 0.8 7 1056.85 4.80 used to calculate the mass flow rate in this study. Fig. 1 is a schematic of
22 6 0.8 8 1056.85 4.80
gas leakage from a pipe. P1, T1, u1 and ρ1 are the pressure (MPa), tem­
23 6 0.8 9 1056.85 4.80
24 6 0.8 10 1056.85 4.80
perature (K), velocity (m/s) and density (kg/m3) at the pipe entrance,
respectively; P2, T2, u2 and ρ2 are the corresponding parameters inside
the pipe, respectively; P3, T3, u3 and ρ3 are the corresponding parameters
at the hole, respectively; Q is the mass flowrate (kg/s); and Le is the
equivalent pipe length (m).
Table 3 Assuming that all of the gas flowing through the pipe also flows
Other parameters for the ruptured natural gas pipeline jet fire model. through the hole, the mass flow rate is given by Eq. (3) [19]:
Parameters Values
Q = Aρ1 u1 = Aρ2 u2 = Aor ρ3 u3 . (3)
Leak parameters Leak position (200, 200, − 0.35)
Leak direction +Z (Vertical upward) Applying the continuity law to the flow of gas through the pipe and
Start time 0.1 s the hole yields Eq. (4) [19]:
Duration 200 s √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
Outlet ( )k+1 √̅̅̅̅̅̅̅̅ √̅̅̅̅̅̅̅̅
Aor kM 2 k− 1
kM kM
a. Area In Table 2 (Leak area) G = ρ1 u1 = ρ2 u2 = P2 ⋅ ⋅ = Ma1 P1 ⋅ = Ma2 P2 ⋅
b. Mass flow rate In Table 2 A RT2 k + 1 RT1 RT2
c. Relative turbulence intensity 0.1 (4)
d. Turbulence length scale 0.08 m
e. Temperature 20 ◦ C √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )k+1
Initial conditions Characteristic velocity In Table 2 (Wind speed) Aor 2 k− 1
Ma2 = , (5)
Relative turbulence intensity 0.1 A k+1
Turbulence length scale 0.01 m
Temperature 20 ◦ C
where A is the cross-sectional area of the pipe (m2), Aor is the hole area
Ambient pressure 100 kPa
Ground roughness 0.0002 m (m2), G is the mass flux (kg/m2⋅s), k is the heat capacity ratio (− ), M is
Reference height 10.9 m the molecular weight (kg/mol), R is the ideal gas constant (J/mol⋅K),
Pasquill stability class F and Ma is the Mach number (− ).
Boundary conditions XLO Wind Applying the momentum and energy equations to adiabatic flow
XHI Nozzle
YLO Wind
through a pipe yields Eq. (6). This equation can be solved using an
YHI Wind iterative method:
ZLO Nozzle ( 2 ) ( )
ZHI Wind K+1 Ma2 Y1 1 1 4kfLe
ln − − + = 0. (6)
Ignition Ignition time 0.16 s
2 2 2
2 Ma1 Y2 Ma1 Ma2 D
Gas monitor region Size 400 m × 400 m × 300 m
where
k− 1
Yi = 1 + Ma2i (7)
2

Table 4
Parameters of the natural gas pipeline jet fire experiment.
Case Release diameter/mm Internal pressure/MPa Internal pressure state Mass flowrate/(kg⋅s− 1) Leakage duration/s Wind speed/(m⋅s− 1) Leakage direction

25 20 5.94 Constant 2.9 60 6.3 Horizontal


26 35 6.15 Constant 9.4 60 6.2 Horizontal
27 50 5.88 Constant 18.4 60 3.6 Horizontal
28 150 7 Nonconstant 215 (Maximum) 200 5.7 Vertically upward

4
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

Fig. 5. Comparison of the flame appearance between the jet fire experiments [13] and 3D simulations.

Table 5
Comparison of the flame dimensions between the jet fire experiments and 3D simulations.
Case Hole diameter/mm Internal pressure/MPa Mass flowrate/(kg⋅s− 1) Flame extent/m

Experimental results Simulation results in this paper Relative error

25 20 5.94 2.9 19.8 ± 1.6 24 21.21%


26 35 6.15 9.4 37.8 ± 2.9 35 − 7.41%
27 50 5.88 18.4 49.9 ± 2.9 45 − 9.82%

1
(
D
) the advantages of the Monte Carlo method, flux methods, and zonal
√̅̅̅ = 4 log 3.7 , (8) methods [32]. An arbitrary intensity distribution can be calculated for
f ε
complex 3D geometries. A detailed formulation of the DTM is given by
Eq. (7) defines the parameter Y (− ), f is the friction factor (− ), D is Deiveegan [33]. The radiative transfer equation (RTE) is solved for
the pipe diameter (m), and ε is the roughness (m). representative rays fired from the boundaries and solid surfaces in the
domain. Rays are fired from surface elements into a finite number of
solid angles that cover the radiating hemisphere around each element.
3.2. Radiation model
The main assumption of the DTM is that the intensity through a solid
angle can be approximated by that of a single ray. The number of rays
The discrete transfer method (DTM) is widely used to solve radiative
and the distribution of directions are chosen in advance [22]. Fig. 2
transfer problems with a participating medium. This method combines

5
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

Fig. 6. Comparison of the flame appearance between the jet fire experiments [12] and the 3D simulations with the complete rupture conditions (on the left are the
Lowesmith’s experimental images and on the right are the simulation results in this paper).

6
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

approximately 6 MPa. Another experiment involved complete rupture of


a 150-mm-diameter pipeline pressurized to approximately 7 MPa. Sig­
nificant results were achieved in these experiments. These experimental
data provided a validation basis for determining the safe distance from a
natural gas pipeline. The different conditions considered in the natural
gas pipeline jet fire simulation are summarized in Table 4. Some
experimental parameters, such as the internal pressure state and release
diameter, differ slightly from those presented in Sections 4.1 and 5.1.
The experiments nevertheless provide some verification of the accuracy
of the jet fire model presented in Section 4.1 because the worst case
scenario is considered in this study; that is, the maximum thermal ra­
diation impact distance is determined.

4.2.1. Verification of flame appearance and dimensions


The flame obtained from the 3D simulation is plotted for tempera­
tures of 600 ◦ C–2200 ◦ C. The simulation results of flame appearance are
compared with Lowesmith’s experimental results to verify the accuracy
of the jet fire model. For cases 25–27, the flame in Lowesmith’s exper­
iment looks very similar to the simulated flame in this paper, a high
Fig. 7. Comparison of the flame height between the jet fire experiment and 3D
simulation under complete rupture conditions. velocity, turbulent, luminous, yellow-colored jet fire with some eleva­
tion of the flame tip (Fig. 5). Close to the release point, a lift-off region
was apparent where the natural gas expanded, which was followed by a
nonluminous region that preceded the main body of the flame [13]. The
extents of the flame from the 3D simulation in this paper and Lowe­
smith’s experiment are provided in Table 5. For each release orifice size,
shows the ray selection and angular discretization. the numerical results, such as the flame appearance and dimensions, are
in good agreement with Lowesmith’s experimental results.
4. Numerical method and validation For case 28, the flame for the first 9 s in Lowesmith’s complete
rupture experiment looks very similar to the simulated flame in this
4.1. Jet fire model paper (Fig. 6). The ignited natural gas expanded quickly, forming a
fireball. The fireball then burned out quickly, but the flame length
In this paper, FLACS software is used to establish a 3D numerical continued to increase, reaching a maximum of approximately 110 m.
model for natural gas pipeline jet fire accident scenarios (Fig. 3). The Thereafter, a jet fire was formed that fluctuated sharply in length but
“hole-pipe model” is used to calculate the mass flow rate. Several radi­ overall gradually decreased in length over time. This phenomenon was
ometers are placed around the leakage point to measure the thermal due to the gradual decrease in the internal pressure of the pipe with
radiation around the jet fire (Fig. 4). The wind direction is from west to respect to time. The flame was tilted over from the vertical direction and
east. The DTM model is used to calculate the radiation levels in the 3D essentially coincident with the wind direction [12].
domain as well as for selected monitoring points. It should be noted that For case 28, the flame length with respect to time is shown in Fig. 7.
the thermal flux is calculated only at the wall surfaces and monitoring The flame length fluctuates sharply in the experiment, whereas the
points. simulation results are relatively flat. The overall trend of the flame
The main parameters in the ruptured natural gas pipeline jet fire length in both the experiment and simulation is gradually decreasing
simulation are the internal pressure, pipeline diameter, and wind speed. with time, which is due to the decrease in the internal pressure in the
Table 2 is a summary of the parameters used to simulate different cases. pipeline over time. In the experiment, the maximum flame length is 105
To consider the worst case scenario, the internal pressure of the pipe is m, whereas in the simulation results, the maximum flame length is 110
held constant, and the leakage diameter is set equal to the pipe diameter. m. Both of these maximum values occur at the time of the maximum
The other parameters are shown in Table 3. The simulation volume is leakage rate, which is 215 kg s− 1. For complete rupture conditions, the
considered to be 400 m × 400 m × 300 m. Around the leak location, the numerical results, such as the flame appearance and dimensions, are in
grid resolution is adjusted to Aleak < Acv< 2Aleak, while at the locations good agreement with Lowesmith’s experimental results.
far from this area, grids were stretched. Aleak is the area of the leak, and
Acv is the area of the grid cell. For these 24 cases, the total number of 4.2.2. Verification of incident thermal radiation
control volumes during the simulation are 50,000–415,800. For cases 25–27, the primary mechanism for injury from a jet fire is
thermal radiation. Fig. 8 shows the incident radiation with respect to the
north and south distance from the flame during the period of steady
4.2. Validation of the jet fire model conditions. The slight differences between the 3D simulation results and
Lowesmith’s experimental results are attributed to environmental
Experimental data provided by Lowesmith are used to validate the instability. These errors lie within an acceptable range.
jet fire model. Lowesmith carried out many large-scale high-pressure jet
fire experiments [12,13]. A series of three tests involved horizontal re­
leases from 20-, 35- and 50-mm-diameter holes at a gauge pressure of

7
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

Fig. 8. Comparison of incident radiation between the jet fire experiment and 3D simulation.

5. Results and discussion

Table 6 5.1. Analysis of the impact distance of jet fire thermal radiation under
Approximate levels of damage for different thermal fluxes. different working conditions
Thermal flux/ Effect
(kW⋅m− 2) In Section 4.2, the simulation results of the flame appearance, di­
35–37.5 Damage to process equipment; collapse of structures. mensions and incident thermal radiation are basically consistent with
25 Thin, insulated steel may lose its mechanical integrity. the experimental results, which verifies the accuracy of the 3D model in
12.5–15 Critical radiation intensity for wood (flame ignition without
this paper. In this section, the thermal radiation impact distance of a gas
contacting the surface); plastic insulation of electrical wires
melts; plastic tubing melts; 100% lethality.
pipeline jet fire under different internal pressures, pipe diameters, and
4.0 Enough to cause pain after an exposure of 20 s; blistering of the wind speeds is analyzed based on the model established in Section 4.1.
skin is likely; 0% lethality. The main effects of thermal flux on individuals are burns on the skin,
the severity of which depends on the intensity of the radiation (kW/m2)
and the dose received. Table 6 gives, in an approximate way, the levels

8
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

Fig. 9. Variations in the flame appearance and thermal radiation influence range with respect to the internal pressure.

Table 7
Variations in the flame height and thermal radiation impact distance with
respect to the internal pressure.
Internal Mass Flame Impact distance of different thermal
pressure/ flowrate/ height/ radiation fluxes/m
MPa (kg⋅s− 1) m
4 12.5–15 25 35–37.5
kW/ kW/m2 kW/ kW/m2
m2 m2

1 176.14 95 87 48 27 19
2 352.28 145 130 82 39 25
3 528.42 175 187 125 49 30
4 704.56 200 220 157 55 34
5 880.70 218 239 180 60 42
6 1056.85 248 251 185 73 46

of damage for different thermal fluxes [34]. According to the thermal


flux level, a method for determining the thermal radiation impact dis­ Fig. 10. Variation in the flame height with respect to the internal pressure.
tance of a jet fire is proposed as a function of the internal pressure, pipe
diameter and wind speed.
wind speed = 5 m/s, leakage direction = vertical upward, ignition time
5.1.1. Impact of the internal pressure = 0.16 s after rupture, simulation duration = 200 s, and ambient tem­
Six cases were simulated with internal pressures of 1 MPa, 2 MPa, 3 perature = 20 ◦ C.
MPa, 4 MPa, 5 MPa and 6 MPa. The other parameters are set as follows: The variations in the flame appearance and thermal radiation impact
pipe diameter = 800 mm, leakage aperture = full-diameter fracture,

9
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

Table 8
Variations in the flame height and thermal radiation impact distance with
respect to the pipe diameter.
Pipe Mass Flame Impact distance of different thermal
diameter/ flowrate/ height/ radiation fluxes/m
mm (kg⋅s− 1) m
4 12.5–15 25 35–37.5
kW/ kW/m2 kW/ kW/m2
m2 m2

100 5.07 26 27 12 8 6
200 30.43 49 61 28 13 8
300 86.33 76 96 49 22 15
400 180.35 110 110 59 33 24
500 319.39 152 130 82 36 29
600 508.33 183 189 111 47 33
700 752.63 210 222 140 59 43
800 1056.85 248 251 185 73 46

range with respect to different internal pressure conditions are shown in


Fig. 9. The vertical upward but slightly right inclined pattern is the jet
Fig. 11. Variation in the thermal radiation impact distance with respect to the fire, whereas the red part on the ground is the impact range of the
internal pressure. thermal flux (12.5–15 kW/m2). The detailed data of the simulation re­
sults are shown in Table 7. The variation trend of the flame height with
respect to the internal pressure is shown in Fig. 10. The variation trends
of the impact distance of different thermal radiation fluxes with respect

Fig. 12. Variations in the flame appearance and thermal radiation influence range with respect to the pipe diameter.

10
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

range with respect to different pipe diameter conditions are shown in


Fig. 12. The jet fire is represented by the region moving vertically up­
wards and slightly to the right, whereas the red region on the ground
represents the impact range of the thermal flux (12.5–15 kW/m2). The
detailed simulation data are shown in Table 8. The variation trend of the
flame height versus the pipe diameter is shown in Fig. 13. The variation
trends of the impact distance of different thermal radiation fluxes with
respect to the pipe diameter are shown in Fig. 14.
Figs. 13 and 14 show that the flame height and the thermal radiation
impact distance increase gradually with increasing pipe diameter, and
the trend of the two increases is basically the same, showing a linear
increasing trend.

5.1.3. Impact of the wind speed


Ten cases were simulated with wind speeds of 1 m/s, 2 m/s, 3 m/s, 4
m/s, 5 m/s, 6 m/s, 7 m/s, 8 m/s, 9 m/s and 10 m/s. The other param­
eters are set as follows: internal pressure = 6 MPa, pipe diameter = 800
mm, leakage aperture = full-diameter fracture, leakage direction =
Fig. 13. Variation in the flame height with respect to the pipe diameter. vertical upward, ignition time = 0.16 s after rupture, simulation dura­
tion = 200 s, and ambient temperature = 20 ◦ C.
The variations in the flame appearance and thermal radiation impact
range with respect to different wind speed conditions are shown in
Fig. 15. The jet fire is represented by the region moving vertically up­
wards and slightly to the right, whereas the red region on the ground
represents the impact range of the thermal flux (12.5–15 kW/m2). The
detailed simulation data are shown in Table 9. The variation trend of the
flame height versus the wind speed is shown in Fig. 16. The variation
trend of the impact distance versus the wind speed for different thermal
radiation fluxes is shown in Fig. 17.
Fig. 15 shows that the flame height and the thermal radiation impact
range are closely related to the wind speed. The flame is tilted over by
the wind from the vertical direction, and the tilt direction is essentially
coincident with the wind direction. Due to the action of wind, the
thermal radiation impact range on the ground is no longer a regular
circle but extends downwind. The higher the wind speed is, the greater
the amplitude of tilt and extension. Figs. 16 and 17 show that as the wind
speed increases, the flame height decreases, but the thermal radiation
impact distance increases significantly.

5.2. Formulas of the thermal radiation impact distance

Fig. 14. Variation in the thermal radiation impact distance with respect to the According to the research in Section 5.1, the thermal radiation
pipe diameter. impact distance of the gas pipeline jet fire has a linear relationship with
the pipe diameter and a power function with the internal pressure and
wind speed. Based on the research in Section 5.1, the formulas of the
to the internal pressure are shown in Fig. 11.
thermal radiation impact distance of a gas pipeline jet fire depending on
Figs. 10 and 11 show that the flame height and the thermal radiation
the internal pressure, pipe diameter and wind speed are proposed using
impact distance increase with increasing internal pressure; the
a fitting method. To ensure that the equation is homogeneous in
increasing trends of both are basically the same, and the increasing
dimension, two parameters, the standard atmospheric pressure Pa and
amplitudes gradually decrease.
speed of sound in the gas ua, are added to the formula. Formulas for
determining the thermal radiation impact distances with respect to
5.1.2. Impact of the pipe diameter
different thermal flux levels are shown in Table 10.
Eight cases were simulated with pipe diameters of 100 mm, 200 mm,
In Table 10, S is the thermal radiation impact distance (m), D is the
300 mm, 400 mm, 500 mm, 600 mm, 700 mm, and 800 mm. The other
pipe diameter (m), P is the internal pressure (MPa), Pa is the standard
parameters are set as follows: internal pressure = 6 MPa, leakage
atmospheric pressure (MPa), u is the wind speed (m/s), and ua is the
aperture = full-diameter fracture, wind speed = 5 m/s, leakage direc­
speed of sound in the gas (m/s); ua = 442 m/s for natural gas.
tion = vertical upward, ignition time = 0.16 s after rupture, simulation
duration = 200 s, and ambient temperature = 20 ◦ C.
The variations in the flame appearance and thermal radiation impact

11
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

Fig. 15. Variations in the flame appearance and thermal radiation influence range with respect to the wind speed.

Table 9
Variations in the flame height and thermal radiation impact distance with respect to the wind speed.
Wind speed/(m⋅s− 1) Mass flowrate/(kg⋅s− 1) Flame height/m Impact distance of different thermal radiation fluxes/m

4 kW/m2 12.5–15 kW/m2 25 kW/m2 35–37.5 kW/m2

1 1056.85 359 240 93 50 28


2 1056.85 349 243 118 55 32
3 1056.85 333 247 135 56 34
4 1056.85 280 250 160 70 43
5 1056.85 248 251 185 73 46
6 1056.85 214 260 190 80 49
7 1056.85 200 275 200 100 56
8 1056.85 189 295 202 110 60
9 1056.85 175 310 215 114 70
10 1056.85 169 351 250 140 84

5.3. Error analysis Therefore, the formula provides a relatively accurate prediction of the
jet fire thermal radiation impact distance.
To verify the accuracy of the formulas proposed in this paper, the
calculation results of the formulas are compared with the simulation
5.4. Comparative analysis
results, and the errors are analyzed. The relative error analysis results of
the thermal radiation impact distance (S4) are shown in Table 11. As
The formula of the potential impact radius for natural gas pipeline
shown in Table 11, the maximum relative error is 26.57%, whereas most
accidents proposed in ASME B31⋅8S-2016 is derived considering a point
of the relative errors are within 10%; the average relative error is 5.52%.
source model for fire thermal radiation and does not account for wind

12
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

around a pipeline. However, the thermal radiation impact distance in­


creases significantly with the wind speed. The formulas proposed in this
paper can calculate the thermal radiation impact distance under
different wind speed conditions and can also calculate the thermal ra­
diation impact distance under different damage levels. Table 12 is a
comparison of the thermal radiation impact distance calculated using
the formula in ASME B31⋅8S-2016 and formula (12) proposed in this
paper.

6. Conclusion

In this paper, a numerical method is proposed for determining the


impact distance of a jet fire following the rupture of a natural gas
pipeline. The major findings of this study can be summarized as follows:

(1) The impact distances of some typical natural gas pipeline fire and
explosion accidents are statistically reviewed. These statistical
data can provide technical support for determining the safe dis­
tance from gas pipelines.
Fig. 16. Variation in the flame height with respect to the wind speed. (2) A 3D natural gas pipeline jet fire accident scenario model is
established with FLACS software. The entire dynamic process
from the gas leakage, diffusion, ignition to the formation of the jet
fire is simulated. The numerical results, such as the flame
appearance, dimensions and thermal radiation, are compared
with the experimental results in the literature, and very good
agreement is obtained.
(3) The flame height and the thermal radiation impact distance in­
crease with increasing internal pressure and pipe diameter. The
flame height and the thermal radiation impact range are closely
related to the wind speed. The flame is tilted over by the wind
from the vertical direction, and the tilt direction is essentially
coincident with the wind direction. Due to the action of the wind,
the thermal radiation impact range on the ground is no longer a
regular circle but extends downwind. The higher the wind speed
is, the greater the amplitude of tilt and extension. As the wind
speed increases, the flame height decreases, but the thermal ra­
diation impact distance increases significantly.
(4) According to the thermal flux level, a method is proposed for
determining the thermal radiation impact distance as a function
of the internal pressure, pipe diameter and wind speed. Most
Fig. 17. Variation in the thermal radiation impact distance with respect to the relative errors between the simulation results and formula
wind speed. calculation results are within 10%. The formulas presented in this

Table 10
Formulas for calculating thermal radiation impact distances under different damage levels.
Thermal flux/(kW⋅m− 2) Effect Formula for the thermal radiation impact distance

35–37.5 Damage to process equipment; collapse of structures ( )0.53 [ ( )1.646 ]


P u
S1 = 17.624D 0.23 + 214.889 (9)
Pa ua

25 Thin, insulated steel may lose its mechanical integrity ( )0.554 [ ( )2.082 ]
P u
S2 = 68.78D 0.094 + 444.82 (10)
Pa ua

12.5–15 Critical radiation intensity for wood (flame ignition without contacting ( )0.573 [ ( )0.568 ]
P u
the surface); plastic insulation of electrical wires melts; plastic tubing melts; S3 = 2.91D 1.63 + 69.7 (11)
Pa ua
100% lethality

4.0 Enough to cause pain after an exposure of 20 s; blistering of the skin is likely; ( )0.54 [ ( )3.51 ]
P u
0% lethality S4 = 81.96D 0.4 + 1.05 × 105 (12)
Pa ua

13
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

Table 11
Error analysis.
Case Internal pressure/MPa Pipe diameter/m Wind speed/(m⋅s− 1) Mass flowrate/(kg⋅s− 1) Impact distance of thermal radiation/m

Simulation results Formula calculation Relative error

87 95.07 9.28%
2 2 0.8 5 352.28 130 138.23 6.33%
3 3 0.8 5 528.42 187 172.07 7.98%
4 4 0.8 5 704.56 220 200.99 8.64%
5 5 0.8 5 880.70 239 226.73 5.13%
6 6 0.8 5 1056.85 251 250.19 0.32%
7 6 0.1 5 5.07 27 31.27 15.83%
8 6 0.2 5 30.43 61 62.55 2.54%
9 6 0.3 5 86.33 96 93.82 2.27%
10 6 0.4 5 180.35 110 125.09 13.72%
11 6 0.5 5 319.39 130 156.37 20.28%
12 6 0.6 5 508.33 189 187.64 0.72%
13 6 0.7 5 752.63 222 218.91 1.39%
14 6 0.8 5 1056.85 251 250.19 0.32%
15 6 0.8 1 1056.85 240 239.34 0.27%
16 6 0.8 2 1056.85 243 239.74 1.34%
17 6 0.8 3 1056.85 247 241.12 2.38%
18 6 0.8 4 1056.85 250 244.28 2.29%
19 6 0.8 5 1056.85 251 250.19 0.32%
20 6 0.8 6 1056.85 260 259.94 0.02%
21 6 0.8 7 1056.85 275 274.75 0.09%
22 6 0.8 8 1056.85 295 295.94 0.32%
23 6 0.8 9 1056.85 310 324.94 4.82%
24 6 0.8 10 1056.85 351 363.25 3.49%

Table 12
Calculation results of different formulas.
Formula Parameters Calculation results
− 1
Internal pressure/MPa Pipe diameter/m Wind speed/(m⋅s ) Impact distance of thermal radiation/m

ASME B31⋅8S-2016: 6 0.8 – 194


√̅̅̅
r = 99D P
Formula (12) proposed in this paper 6 0.8 1 92.90
6 0.8 2 118.61
6 0.8 3 139.06
6 0.8 4 156.71
6 0.8 5 172.53
6 0.8 6 187.03
6 0.8 7 200.52
6 0.8 8 213.19
6 0.8 9 225.20
6 0.8 10 236.65

paper can calculate the thermal radiation impact distance under [2] K. Shan, J. Shuai, Statistical analyses of incidents on oil and gas pipelines based on
comparing different pipeline incident databases, Proceedings of the ASME 2017
different wind speed conditions and can also calculate the ther­
Pressure Vessels and Piping Conference, 16-20 July 2017, Hawaii, USA. DOI:
mal radiation impact distance under different damage levels. https://doi.org/10.1115/PVP 2017-65289.2017.
[3] A.J. Brito, A.T. Almeida, Multi-attribute risk assessment for risk ranking of natural
gas pipelines, Reliab. Eng. Syst. Saf. 94 (2009) 187–198.
Declaration of competing interest [4] The Society of Fire Protection Engineers, SFPE Handbook of Fire Protection
Engineering, Springer: National Fire Protection Association, 2002.
[5] T.A. Brzustowski, A new criterion for the length of a gaseous turbulent diffusion
The authors declare that they have no known competing financial flame, Combust. Sci. Technol. 6 (1973) 313–319.
interests or personal relationships that could have appeared to influence [6] T.A. Brzustowski, E.C. Sommer, Predicting radiant heating from flares, Proceedings
of the API Division of Refining 53 (1973) 865.
the work reported in this paper.
[7] W.R. Hawthorne, D.S. Weddell, H.C. Hottell, Mixing and combustion on turbulent
gas jet, in: Third Symposium (International) on Combustion, Combustion Institute,
Acknowledgments Pittsburgh, 1949, pp. 266–288.
[8] G.T. Kalghatki, The visible shape and size of a turbulent jet diffusion flame in a
crosswind, Combust. Flame 52 (1983) 91–106.
This research was financially supported by the National Natural [9] G.A. Chamberlain, Developments in design methods for predicting thermal
Science Foundation of China (grant No. 51874324) and the Key radiation from flares, Chem. Eng. Res. Des. 65 (1987) 299.
[10] A.D. Johnson, H.M. Brightwell, A.J. Carsley, A model for predicting the thermal
Technology Projects of PetroChina Co Ltd. (grant No. ZLZX2020-05). radiation hazards from large-scale horizontally released nature gas jet fires,
Hazards (1994) 123.
References [11] D.K. Cook, M. Fairweather, J. Hammonds, D.J. Hughes, Size and radiative
characteristics of natural gas flares, Chem. Eng. Res. Des. 65 (1987) 310–318.
[12] B.J. Lowesmith, G. Hankinson, Large scale experiments to study fires following the
[1] K. Shan, J. Shuai, K. Xu, W. Zheng, Failure probability assessment of gas rupture of high pressure pipelines conveying natural gas and natural gas/hydrogen
transmission pipelines based on historical failure-related data and modification mixtures, Process Saf. Environ. Protect. 91 (2013) 101–111.
factors, J. Nat. Gas Sci. Eng. 52 (2018) 356–366.

14
K. Shan et al. International Journal of Pressure Vessels and Piping 187 (2020) 104159

[13] B.J. Lowesmith, G. Hankinson, Large scale high pressure jet fires involving natural [25] T. Baalisampang, R. Abbassi, V. Garaniya, F. Khan, M. Dadashzadeh, Accidental
gas and natural gas/hydrogen mixtures, Process Saf. Environ. Protect. 90 (2012) release of Liquefied Natural Gas in a processing facility: effect of equipment
108–120. congestion level on dispersion behaviour of the flammable vapour, J. Loss Prev.
[14] B.J. Lowesmith, G. Hankinson, M.R. Acton, G. Chamberlain, An overview of the Process. Ind. 61 (2019) 237–248.
nature of hydrocarbon jet fire hazards in the oil and gas industry and a simplified [26] T. Baalisampang, R. Abbassi, V. Garaniya, F. Khan, M. Dadashzadeh, Modelling an
approach to assessing the hazards, Process Saf. Environ. Protect. 85 (2007) integrated impact of fire, explosion and combustion products during transitional
207–220. events caused by an accidental release of LNG, Process Saf. Environ. Protect. 128
[15] O. Levenspiel, Engineering Flow and Heat Exchange, Plenum Press, New York, (2019) 259–272.
1986, pp. 41–51. [27] T. Baalisampang, R. Abbassi, V. Garaniya, F. Khan, M. Dadashzadeh, Fire impact
[16] D.A. Crowl, J.F. Louvar, Chemical Process Safety: Fundamental with Applications, assessment in FLNG processing facilities using Computational Fluid Dynamics
Prentice Hall, NJ, 1990, pp. 110–130. (CFD), Fire Saf. J. 92 (2017) 42–52.
[17] J.L. Woodward, K.S. Mudan, Liquid and gas discharge rates through holes in [28] R. Yang, F. Khan, M. Taleb-Berrouane, D. Kong, A time-dependent probabilistic
process vessels, J. Loss Prev. Process. Ind. 4 (1991) 161–165. model for fire accident analysis, Fire Saf. J. 111 (2020) 102891.
[18] T.W. Cochran, Calculate pipeline flow of compressible fluid, Chem. Eng. 103 [29] R. Yang, F. Khan, M. Yang, D. Kong, C. Xu, A numerical fire simulation approach
(1996) 115–122. for effectiveness analysis of fire safety measures in floating liquefied natural gas
[19] H. Montiel, J.A. Vilchez, J. Casal, J. Arnaldos, Mathematical modeling of facilities, Ocean. Eng. 157 (2018) 219–233.
accidental gas releases, J. Hazard Mater. 59 (1998) 211–233. [30] Pipeline and Hazardous Materials Safety Administration, Pipeline Accident
[20] Y. Deng, H. Hu, B. Yu, D. Sun, L. Hou, Y. Liang, A method for simulating the release Reports, 2018-04-17. https://www.ntsb.gov/investigations/AccidentReports/
of natural gas from the rupture of high-pressure pipelines in any terrain, J. Hazard Pages/pipeline.aspx.2018.
Mater. 342 (2018) 418–428. [31] Transportation Safety Board of Canada, Pipeline Investigation Reports, 2018-04-
[21] The American Society of Mechanical Engineers, ASME Code for Pressure Piping, 17. http://www.tsb.gc.ca/eng/rapports-reports/pipeline/index.asp. 2018.
B31 Supplement to ASME B31.8, Managing System Integrity of Gas Pipelines, The [32] F.C. Lockwood, N.G. Shah, A new radiation method for incorporation in general
American Society of Mechanical Engineers, New York, 2016. combustion prediction procedures, Proceedings, 18th Symposium (International)
[22] A.S. GexCon, FLACS v10.4 User’s Manual, Norway, 2015. on Combustion (1981) 1405–1414.
[23] M. Dadashzadeh, F. Khan, K. Hawboldt, P. Amyotte, An integrated approach for [33] M. Deiveegan, Implementation of Radiative Transfer Calculations in FLACS,
fire and explosion consequence modeling, Fire Saf. J. 61 (2013) 324–337. Internal Report, Christian Michelsen Research (Confifidential), 2011.
[24] M.T. Amin, F. Khan, P. Amyotte, A bibliometric review of process safety and risk [34] J. Casal, Evaluation of the Effects and Consequences of Major Accidents in
analysis, Process Saf. Environ. Protect. 126 (2019) 366–381. Industrial Plants, 2008. Spain.

15

You might also like