Vous êtes sur la page 1sur 12

Computational uid dynamics analysis of liqueed natural gas

dispersion for risk assessment strategies


Biao Sun
a, b
, Ranjeet P. Utikar
a
, Vishnu K. Pareek
a,
*
, Kaihua Guo
b
a
Department of Chemical Engineering, Curtin University, GPO Box U1987, Perth, WA 6845, Australia
b
Engineering School, Sun Yat-Sen University, Guangzhou 510006, China
a r t i c l e i n f o
Article history:
Received 15 March 2012
Received in revised form
7 October 2012
Accepted 7 October 2012
Keywords:
LNG
CFD
Dense gas dispersion
Risk assessment
NFPA 59A
Impoundment
a b s t r a c t
Computational uid dynamics (CFD) simulations have been conducted for dense gas dispersion of liq-
ueed natural gas (LNG). The simulations have taken into account the effects of gravity, time-dependent
downwind and crosswind dispersion, and terrain. Experimental data from the Burro series eld tests,
and results from integral model (DEGADIS) have been used to assess the validity of simulation results,
which were found to compare better with experimental data than the commonly used integral model
DEGADIS. The average relative error in maximum downwind gas concentration between CFD predictions
and experimental data was 19.62%.
The validated CFD model was then used to perform risk assessment for most-likely-spill scenario at
LNG stations as described in the standard of NFPA 59A (2009) Standard for the Production, Storage and
Handling of Liqueed Natural Gas. Simulations were conducted to calculate the gas dispersion behav-
iour in the presence of obstacles (dikes walls). Interestingly for spill at a higher elevation, e.g., tank top,
the effect of impounding dikes on the affected area was minimal. However, the impoundment zone did
affect the wind velocity eld in general, and generated a swirl inside it, which then played an important
function in conning the dispersion cloud inside the dike. For most cases, almost 75% of the dispersed
vapour was retained inside the impoundment zone. The nding and analysis presented here will provide
an important tool for designing LNG plant layout and site selection.
2012 Elsevier Ltd. All rights reserved.
1. Introduction
Worldwide demand for liqueed natural gas (LNG) has
increased rapidly in recent years (ExxonMobil, 2008; Kumar et al.,
2011) because of expanding economies, growing awareness of
environmental protection, and enhanced transportation methods
for cryogenic tankers (by road, ships, and rail). A great number of
LNG receiving terminals and satellite stations have been planned
and constructed in recent years across many countries (Dorigoni &
Portatadino, 2008; Lin, Zhang, & Gu, 2010). Consequently, strong
emphasis has been placed on the safety of LNG stations in regards
to site selection, transportation and regasication (Cleaver,
Johnson, & Ho, 2007). For example, the exclusive distance from
LNG station to densely populated areas, in the event of LNG spill, is
one of the most important factors to be determined in risk
assessment. This paper shows application of computational uid
dynamics (CFD) for addressing such issues.
LNG has unique properties, cryogenic temperature, amma-
bility, and vapour dispersion characteristics, compared with other
fuels. Due to its cryogenic boiling point (111.7 K), once it is leaked
onto subsoil or water surface, it vaporizes very quickly, with
vaporization rate varying between 0.029 and 0.195 kg/(m
2
s)
(Luketa-Hanlin, 2006; TNO Report, 2005).
At atmospheric conditions, LNG vapour is 1.5 times heavier than
the ambient air (Koopman & Ermak, 2007), and therefore is called
as a dense or a heavy gas. This manifests special behaviour that
affects its dispersion. The schematic of dispersion process is shown
in Fig. 1 (Spicer & Havens, 1989; TNO Report, 2005). LNG dispersion
experiments show three distinct dispersion stages (Spicer &
Havens, 1989), namely negative-buoyancy-dominated, stably-
stratied and passive dispersion. Once the vapour is released, it
descends to the surface and spreads radially under the inuence of
gravity. The dispersed cloud then behaves as a stably-stratied
cloud embedded in the mean wind ow. Finally, the LNG vapour
is diluted as a neutrally buoyant cloud by turbulent air ow
(passive dispersion). Since the upper and lower ammable limits
(UFL and LFL) of natural gas are respectively 15% and 5% in volume,
for the safety consideration, according to NFPA 59A (NFPA, 2009),
* Corresponding author. Tel.: 61 8 9266 4687; fax: 61 8 9266 2681.
E-mail address: v.pareek@curtin.edu.au (V.K. Pareek).
Contents lists available at SciVerse ScienceDirect
Journal of Loss Prevention in the Process Industries
j ournal homepage: www. el sevi er. com/ l ocat e/ j l p
0950-4230/$ e see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jlp.2012.10.002
Journal of Loss Prevention in the Process Industries 26 (2013) 117e128
the exclusion distance in the downwind direction must be such
that the average volume concentration of methane is lower than
2.5%.
Numerical models for dense gas dispersion are mainly classied
into three types (Duijm, Carissimo, Mercer, Bartholome, &
Giesbrecht, 1997), phenomenological models, integral models and
computational uid dynamics (CFD) models. Phenomenological
models often simplify the physics of the ow and are written by
nomograms or simple correlations from the LNG eld tests to
describe the dispersion behaviour, for example, the B&M model
(Britter, 1988; Duijm et al., 1997). Integral models assume
a concentration prole for the downwind gas dispersion, such as
Gaussian proles, and then perform the integration along the
downwind distance. Two of the most commonly used integral
models DEGADIS (Havens, 1998; Spicer & Havens, 1989) and SLAB
(Ermak, 1990), are schematically shown in Fig. 2. One of the major
disadvantages of these two types of models is that they work well
only if the terrain is relatively at (Gavelli, Bullister, & Kytomaa,
2008). The CFD models solve mass, momentum and energy
conservation equations numerically. Contrary to the phenomeno-
logical and integral models, CFD models use physically complete
terrain description. While CFD models are computationally inten-
sive, they include the detailed ow physics and are capable of
modelling various phenomena adequately.
CFD models have received increasing interest for the simulation
of dense gas dispersion behaviour. Sutton, Brandt, and White
(1986) were among the rst to use CFD code to simulate the
dense gas dispersion in the presence of obstacles in a boundary-
layer wind tunnel (21.3 m long, 1.18 m wide, 1.68 m high). A two-
equation turbulence viscous model, k was employed in their
study. Several commercial CFD codes are now available (Coldrick,
Lea, & Ivings, 2010) for LNG vapour dispersion simulation. These
include, FEM 3 CFD (Luketa-Hanlin, Koopman, & Ermak, 2007),
FLACS (Dharmavaram, Hanna, & Hansen, 2005), Star-CD (Deaves,
Gilham, Mitchell, Woodburn, & Shepherd, 2001), FLUENT (Gavelli
et al., 2008; Tauseef, Rashtchian, & Abbasi, 2011; Tauseef,
Rashtchian, Abbasi, & Abbasi, 2011), and CFX (Cormier, Qi, Yun,
Zhang, & Mannan, 2009; Qi, Ng, Cormier, & Mannan, 2010;
Sklavounos & Rigas, 2006). Several LNG eld tests of 1980s
(Cornwell & Pfenning, 1987; Luketa-Hanlin, 2006; Mohan, Panwar,
& Singh, 1995) have formed the basis of validating CFD simulation.
Application of CFD for simulating LNG vapour cloud dispersion was
strongly recommended by the Sandia National Laboratories 2004
report (Hightower et al., 2004). Sklavounos and Rigas (2006) used
CFX to simulate the Coyote series trials and made a comparison
with the box-models. However, they used tetrahedral meshing,
which for the simple 3D box geometry (600 m long, 250 m wind,
50 m high) is less computational efcient than hexahedral (brick)
meshes (Gavelli et al., 2008). Luketa-Hanlin et al. (2007) compared
the FEM 3 codes with Burro series test. MonineObuhkov similarity
theory was applied in their study to describe the vertical behaviour
of air ow and turbulence properties within atmospheric surface
layer. One of Falcon series tests (Brown et al., 1990), Falcon-1, was
simulated by Gavelli et al. (2008) in order to nd the effects of
a billboard in the wind eld on the vapour cloud dispersion
behaviour. Standard k model together with Reynolds stress
model (RSM) were applied in their study. The RSM model was
computationally more costly and hard to convergence compared to
other turbulent models. Qi et al. (2010) used CFX code in
conjunction with the MonineObukhov similarity theory to simu-
late the LNG vapour dispersion of Brayton re Training Field tests.
Tauseef, Rashtchian, and Abbasi (2011), Tauseef, Rashtchian, Abbasi,
and Abbasi (2011) also considered the obstacles present in the
downwind direction, comparing with Trial 26 of the Thorney Island
series tests (Coldrick et al., 2010). Realizable k model was found
to give more accurate results than the standard k mode in
modelling LNG vapour dispersion.
The above simulation studies have been primarily conducted to
validate CFD models for LNG vapour dispersion. Most of these
simulations did not consider the transient behaviour of LNG vapour
or carry out any practical risk assessment of LNG stations. This
paper presents a detailed CFD model for LNG dispersion and risk
assessment. The commercial code FLUENT has been used to simu-
late two typical experiments of the Burro series eld test. The
dynamic simulation results were validated with the available
experimental data and DEGADIS calculations results. NFPA 59A has
been used to design the most likely spill scenarios in LNGstations.
The validated CFD simulation method was then used to carry out
the risk assessment under the designed spill scenario, in order to
nd the mitigation effect of the impoundments, and reduction of
exclusive distance.
Gravity Spreading
Air Entrainment
Vortex
Fig. 1. Gravity spreading of a dense gas cloud.
Fig. 2. Schematic diagram of integral dense gas dispersion model.
B. Sun et al. / Journal of Loss Prevention in the Process Industries 26 (2013) 117e128 118
2. Theoretical background
Present simulations were carried out using commercial software
FLUENT 12 (ANSYS, 2010), which utilizes the Finite Volume Method
(FVM) (Versteeg & Malalasekera, 1995) to discretize the computa-
tional domain and equations, combined with Reynolds averaged
NaviereStokes (RANS) equations and Reynolds stress models (RSM)
for calculating the processes of momentum. In this study, the
equations involve continuity, energy, momentum, turbulence
model and species model (ANSYS, 2010). These equations are given
below.
Continuity equation
vr
vt
V$r v
!
0 (1)
where r is the density of dense gas cloud, v
!
is velocity vector of
three dimension.
Momentum equation
vr v
!

vt
V$r v
!
v
!
Vp V$s r g
!
F
!
(2)
s m
_
_
V v
!
V v
!
T
_

2
3
V$ v
!
I
_
(3)
where p is pressure, s is stress tensor, F
!
is the sum of body forces,
r g
!
means gravity works on dense gas cloud in vertical direction, m
is dynamic viscosity.
Energy equation
vrE
vt
V$ v
!
rE p V$
_
_
k
eff
VT

j
h
j
J
j
!
s
eff
$ v
!

_
_
S
h
(4)
E h
p
r

v
2
2
(5)
where E is total energy, k
eff
is effective conductivity, S
h
is energy
source (if chemical reaction exist).
The choice of turbulence model is a key in dispersion simulation
using CFD codes. In this study, realizable k model (Shih, Liou,
Shabbir, Yang, & Zhu, 1995) is applied, because this model
satises certain mathematical constraints on Reynolds stress, and is
consistent with the physics of turbulent ows. Besides, this model
exhibits an excellent performance in capturing the phenomenon of
gravity slumping associated with dense gas dispersion, and is
superior to the standard k viscosity model (Shih et al., 1995). It
has been reported that realizable k predicts the spatial and
temporal concentration prole of the vapour cloud in the presence
of obstacles more accurately than the hitherto oft-used standard
k model (Launder & Spalding, 1972; Tauseef, Rashtchian, &
Abbasi, 2011; Tauseef, Rashtchian, Abbasi, & Abbasi, 2011). Realiz-
able k model differs from the standard k model in two ways.
1) The realizable k model contains a new formulation for the
turbulent viscosity and a neweddy-viscosity formula involving
a variable C
m
, shown in Equation (6), whereas C
m
is a constant
and equals 0.09 in standard k and RNG k model.
2) A new transport equation for the dissipation rate has been
derived from an exact equation for the transport of the mean-
square vorticity uctuation, shown in Equation (8).
m
t
rC
m
k
2

(6)
vrk
vt

v
_
rku
j
_
vx
j

v
vx
j
_
_
m
m
t
s
k
_
vk
vx
j
_
G
k
G
b
rY
M
S
k
(7)
vr
vt

v
_
ru
j
_
vx
j

v
vx
j
_
_
m
m
t
s

_
v
vx
j
_
rC
1
S rC
2

2
k

v
p
C
1

k
C
3
G
b
S

(8)
where m
t
turbulence viscosity, k is kinetic energy, is turbulence
eddy dissipation, G
k
and G
b
are the turbulence production due to
viscous and buoyancy forces, S
k
and S

are the source termfor k and


.
An additional transport equation is needed to calculate the
volume concentration of natural gas, because of the involvement of
the multi-component ow. Species transport equations are
considered in the following form.
v
vx
rY
i
V$r v
!
Y
i
V$ J
j
!
R
i
S
i
(9)
J
j
!

_
rD
i;m

m
t
Sc
t
_
VY
i
D
T;i
VT
T
(10)
where Y
i
is Mass fraction of species i, J
j
!
is Mass diffusion, R
i
is Net
rate of production of species, S
i
is the source term, D
i,m
is mass
diffusion coefcient.
3. The Burro series test
The Burro tests were performed by the Lawrence Livermore
National Laboratory (LLNL) at the Naval Weapons Center, China
Lake, California in 1980, and sponsored by the US DOE and the Gas
Research Institute (Ermak, Chan, Morgan, & Morris, 1982; Koopman
et al., 1982a, 1982b; Koopman, Cederwall, et al., 1982). The main
purpose was to obtain adequate samples of data under different
meteorological conditions. These experiments involved eight spills
of LNG and one of liquid nitrogen. A total of eight LNG spills onto
water surface were performed with spill volumes ranging from
24 m
3
to 39 m
3
, spill rates from 11.3 m
3
/min to 18.4 m
3
/min, wind
speeds from1.8 m/s to 9.1 m/s, and atmospheric stability conditions
being fromunstable (Class B) to slightly stable (Class E). Water pond
and instrument array (Koopman et al., 1982a, 1982b; Koopman,
Cederwall, et al., 1982) are shown in Fig. 3. There were 25 gas
sensor stations were placed in arcs in the downwind distance 57 m,
140 m, 400 m and 800 m. Meanwhile, 6 turbulence stations and 20
wind eld anemometer stations were arranged in both upwind and
downwind directions. The centreline of instrument array was
oriented at 225

from the southwest, in order to coincide with the


prevailing wind direction. Atmospheric conditions of Burro series
tests are listed in Table 1 (Koopman et al., 1982a, 1982b; Koopman,
Cederwall, et al., 1982). In this study, Burro 5 and Burro 8 (B5 and
B8) tests have been selected to validate the simulations, because of
the longest spill duration (190 s) in B5 test and the most stable
atmospheric conditions (atmospheric stability E) in B8 test.
4. Simulation approach
The computational domain and mesh are shown in Fig. 4.
X-direction has been assumed to be horizontal and parallel to the
wind, from 100 m upwind to 900 m downwind. Y-direction was
B. Sun et al. / Journal of Loss Prevention in the Process Industries 26 (2013) 117e128 119
horizontal and perpendicular to the wind, y 500 m. Moreover,
z-direction was vertical, z 50 m. The original point was placed at
the ground level in the centre of the water pond. The computational
domain was discretised in hexahedral cells, which are much more
computationally efcient than tetrahedral meshes. In the mass
inlet area, O-grid was applied in order to improve the mesh quality.
Overall, the mesh contained 420,080 cells inside the computational
domain.
The density of vapour cloud was calculated using the ideal gas
law. In order to judge whether the vapour was a dense gas or not,
the modied Richardson number (Britter, 1989), Ri
m
, was applied as
a criterion, expressed in Equation (11). g is gravity (9.81 m/s
2
). r and
r
a
are density of vapour cloud and air (kg/m
3
), respectively. Q
0
is
LNG spill rate (m
3
/s). D is characteristic length of LNG vapour cloud,
and U
0
is wind velocity (m/s). If Ri
m
0.15, the vapour cloud
behaves as dense gas. In this study, Ri
m
0.18 for B5 test, Ri
m
0.78
for B8 test, hence the gravity-induced effect must be taken into
consideration.
Ri
m

_
g
r r
a
r
a
Q
0
D
_1
3
U
0
0:15 (11)
4.1. Boundary conditions
At the air velocity inlet, a vertical velocity prole was applied as
described by Equation (12), where U(z) means the wind velocity at
the height of Z, U
0
is wind velocity, z
0
is the maximum height. The
power-lawexponent l depends upon the ground surface roughness
and the atmospheric stability. The exponent value can be analysed
through MonineObukhov similarity theory (Arys, 1999). The
equations for non-dimensional wind shear are shown in Equations
(13)e(16). A value l 0.007 was applied in this study, which
corresponds to slightly unstable in class C in PasquilleGifford
atmospheric stability (Essa, Embaby, & Etman, 2003; Mohan &
Siddiqui, 1998; Pontiggia, Derudi, Busini, & Rota, 2009).
Uz U
0
_
z
z
0
_
l
(12)
Table 1
Initial condition of Burro LNG dispersion series tests.
Trial No. B2 B3 B4 B5 B6 B7 B8 B9
Liquid pool diameter (m) 58 58 58 58 58 58 58 58
Spill volume (m
3
) 34.3 34.0 35.3 35.8 27.5 39.4 28.4 24.2
Spill duration (s) 173 167 175 190 128.9 174 107 79
Spill rate (m
3
/min) 11.9 12.2 12.1 11.3 12.0 13.6 16.0 18.4
Wind speed (m/s) 5.4 5.4 9.0 7.4 9.1 8.4 1.8 5.7
Relative humidity (%) 7.1 5.2 2.7 5.6 5.1 5.6 4.5 13.1
Atmospheric
temperature (

C)
37.6 33.8 35.4 40.5 39.2 33.7 33.1 35.4
Atmospheric stability B B C C C D E D
Fig. 4. Computational domain and boundaries (a) and hexahedral mesh rened near
the ground surface (b).
Fig. 3. Water pond (a) and instrumentation array (b) in Burro series dispersion tests
(overhead view).
B. Sun et al. / Journal of Loss Prevention in the Process Industries 26 (2013) 117e128 120
Non-dimensional wind shear
F
m

_
KZ
U
*
__
vU
vz
_
(13)
where F
m
is calculated differently upon different atmospheric
conditions, described as ows: For unstable conditions
F
m

_
1
15Z
L
_
0:25
(14)
For stable conditions
F
m
1
4:7Z
L
(15)
For neutral conditions
F
m
1 (16)
LNG evaporates very quickly fromthe water pond area. The heat
transfer between LNG and water is due to convection ow in the
water, with very minor quantities of ice formed (Opschoor, 1980).
The published LNG vaporization rate for leaking on water surface
varies between 0.029 and 0.195 kg/(m
2
s) (Luketa-Hanlin, 2006).
Since LNG spill rates are 0.032 and 0.045 kg/(m
2
s) in B5 test and B8
test, therefore, this paper uses the spill rate as mass ow rate in
boundary conditions. The outlet boundary condition was set as
pressure outlet, since ow and energy eld were not known at this
boundary. A fully-developed ow was assumed at this boundary.
A wall boundary condition was applied to ground surface, since
no ow or energy exchanging occurs in this boundary. For side and
top surfaces, zero gradients of ow, energy and species variables
were assumed, because these surfaces are far enough from the
mass ow area.
4.2. Solution method
Prior to injection of LNG, a steady-state oweld was calculated.
Once a converged solution was obtained for steady state ow eld,
the time-dependent simulations were performed using the steady-
state as the initial condition. At time t 0 s, the injection of
natural gas was switched on and the volume concentration of
natural gas was obtained as a function of time. For the transient
vapour dispersion simulation, the residuals convergence criterion
was set as 1.0 10
4
, and the time step to 0.5 s. Approximately 20
iterations per time step were required to reach the limited residuals.
The total execution time of each transient simulation was about 5 h
on a 2.93 GHz Intel

Core
TM
i7 processor with 8.00 GB RAM.
5. Results and discussion
5.1. Crosswind dispersion
Fig. 5 shows vertical concentration proles at a downwind
distance of 57 m at different times after the LNG spill, 20 s, 70 s and
130 s, respectively. It compares the lateral dispersion distance
between experiment and simulation results. At the initial stage,
negative-buoyancy was dominated, because of heavier density of
LNG vapour than ambient air. The width of vapour cloud was wider
than the source diameter under the effect of gravity leading to lateral
spread. From 20 s to 130 s, lateral spreading distance increased
farther. It is obvious that the vapour cloud was driven by the gravity
and it moved laterally in both experiment and simulations as shown
in Fig. 5. This demonstrates that the CFD simulations were in a good
agreement with the theory of dense gas dispersion and eld
Fig. 5. FLUENT simulation of vertical volume fraction contours at 1 m height for Burro 5 test.
B. Sun et al. / Journal of Loss Prevention in the Process Industries 26 (2013) 117e128 121
experiment. In B5 test, iso-concentration contour 1% (volume
concentration) could cover the lateral distance as far as 76 min 130 s,
with the source radius of 29 m. However, the simulation predicted
a higher spreading distance of about 110 m. This was due to atmo-
spheric turbulence, and the boundary conditions applied (Koopman
et al., 1982a, 1982b; Koopman, Cederwall, et al., 1982). While, in the
eld tests, the wind speed and the wind direction were non-
uniform, in the simulation, the wind speed and its direction were
assumed to be uniform based upon the average wind speed and the
prevailing wind direction of experiments.
5.2. Horizontal dispersion
The most valuable data for model validation are considered to
be the gas concentration plume parameters (Luketa-Hanlin et al.,
2007). The four recommended values by Ermak, Chapman,
Goldwire, Gouveia, and Rodean (1988) are maximum gas concen-
tration, average ground-level plume centreline concentration,
plume half-width and plume height, all as a function of downwind
distance. Fig. 5 shows the plume height and spread in a time
dependent manner.
Fig. 6. CFD prediction of horizontal gas concentration prole at 1 m height for B5 test experiment. (a) Experiment (b) simulation (c) comparison.
B. Sun et al. / Journal of Loss Prevention in the Process Industries 26 (2013) 117e128 122
Fig. 7. CFD prediction of horizontal gas concentration prole at 1 m height for B8 test. (a) Experiment (b) simulation.
B. Sun et al. / Journal of Loss Prevention in the Process Industries 26 (2013) 117e128 123
In the B5 test, spill duration was 190 s, with a constant spill rate
11.3 m
3
/min. According to the experimental results, steady state
reached at about 90 s. The simulations achieved the steady state at
around 80 s and the dimension of the vapour cloud did not change
signicantly afterwards. Fig. 6 illustrates the horizontal iso-
concentration contour at 1 m height level in the downwind direc-
tion and compares with the experimental data with simulation
results at 20 s, 50 s, 90 s and 190 s. The experimental data are shown
in Fig. 6(a); the simulation results are in Fig. 6(b) and comparison is
shown in Fig. 6(c). As the time elapsed, the gas cloud grewbigger in
both downwind and crosswind direction. Lower volume concen-
tration in the downwind direction and larger covered area by the
dispersion cloud was also observed. The CFD simulations were able
to realistically capture these features. A very good quantitative
agreement was observed between the CFD simulations and the
experimental data for vapour cloud dimensions, especially as the
steady state was approached with an average relative error of
60.30% compared with experiment as shown in Fig. 6(c). The
gravity-induced effect can also be identied in Fig. 6 with the cloud
lateral distance wider than the source dimension. Because of the
Fig. 9. Computational domain of case study, and wind prole before LNG releasing. (a)
Geometry (b) mesh (c) wind prole.
a
b
Fig. 8. Comparison of steady-state results for experiment results, FLUENT simulation
and DEGADIS. (a) B5 test (b) B8 test.
B. Sun et al. / Journal of Loss Prevention in the Process Industries 26 (2013) 117e128 124
changes in the wind direction during the execution of experiments,
the plume shape deviates from the centreline by a small angle
(within 5

), especially around the cloud front.


Fig. 7 shows the simulation results of B8 test under steady-state
dispersion. Because of the high wind speed in B5 test, with an
average of 7.4 m/s, downwind dispersion distance was as much as
350 m. It is clear that the CFDsimulations qualitatively captured the
vapour cloud. Under the more stable atmospheric conditions in B8
test, the downwind dispersion distance was much shorter and the
dense gas cloud also moved upwind a little, but the crosswind
dispersion was much wider as shown in Fig. 7. Furthermore, in the
relatively more stable atmospheric conditions, it was expected to
take longer time to reach the steady-state dispersion. In B8 test, the
steady-state dispersion reached at 140 s approximately, and the
spill duration was 107 s.
Fig. 8 compares the predicted maximum downwind gas
concentration proles of DEGADIS and CFDwith experimental data.
The relative errors of DEGADIS and FLUENT in comparison with
experimental data were 42.58% and 19.62% respectively, showing
an excellent consistency of the CFD simulations even when the
atmospheric conditions are changed. DEGADIS predicts the gas
dispersion consistently with experimental data, but when dealing
with stable conditions, it overestimates the maximum downwind
gas concentration. More importantly, DEGADIS cannot deal with
the case having complicated terrain. This is due to the concentra-
tion assumption of Gaussian prole in the vapour cloud, which
cannot take into consideration of the non-at terrain effects.
6. Risk assessment
From the comparison with Burro tests, it is clear that CFD
simulations give more accurate prediction on temporal and spatial
concentration proles of dense gas dispersion compared to other
models, such as DEGADIS. Therefore, CFD simulation can be effec-
tively used for risk assessment in LNG station siting, pre-accident
predicting and post-accident analysis.
NFPA 59A (2009 Edition) regulates that impoundments are
required for LNG tanks or containers for leak control. The minimum
volumetric liquid holding capacity must be 110% of the maximum
tank capacity (Raj & Lemoff, 2009). Once LNG spills, the dike would
prevent the liquid from spreading. For LNG gas dispersion, the
vapour cloud will be conned inside the impoundments because of
the dense gas effect. Exclusion zones must be built at the point of
volume concentration of 2.5% (half of the LFL).
In this study, the container capacity was assumed to be 2000 m
3
(7 m radius and 13 m high), and the impoundment dimension was
28 m long, 28 m wide and 3 m high. Geometry and mesh for this
case are shown in Fig. 9. The CFD simulation method used in Burro
LNG dispersion was applied for this case, and the boundary
conditions were set according to NFPA 59A (atmospheric stability F,
wind speed 2 m/s). Wind direction was parallel to the x-axis in
order to predict the maximum dispersion distance. As calculated in
Table 2, the spill rate was 18.5 kg/s (0.12 kg/(m
2
s)) from the tank
top surface according to the Design Spill in NFPA 59A and the spill
duration was 200 s.
Fig. 10. Time-depended vapour cloud footprint (volume concentration contour 2.5%) in the impoundment area at different time.
Table 2
Design spill in NFPA 59A (2009).
Design spill source Design spill criteria Design spill rate and volume
Containers with
penetrations below
the liquid level
without internal
shutoff valves
A spill through an assumed opening at,
and equal in area to,
that penetration below the liquid level
resulting in the largest ow an initially
full container
If more than one container in the
impounding area. Use the container with
the largest ow.
Use the following formula: q
4
3
d
2

h
p
Until the differential head acting on the opening is 0.
For SI units, use the following formula:
q
1:06
10; 000
d
2

h
p
Until the differential head acting on the opening is 0.
Containers with
penetrations below
the liquid level with
internal shutoff
valves.
The ow through an assumed opening at,
and equal in area to, that penetration below
the liquid level that could result in the
largest ow from an initially full container.
Use the following formula: q
4
3
d
2

h
p
For SI units, use the following formula:
q
1:06
10; 000
d
2

h
p
for 10 min.
Note: q ow rate [ft
3
/min, (m
3
/min)] of liquid; d diameter [in, (m)] of tank penetration below the liquid level, h height [ft, (m)] of liquid above penetration in the
container when the container is full.
B. Sun et al. / Journal of Loss Prevention in the Process Industries 26 (2013) 117e128 125
6.1. LNG spill with impoundment
The tank impoundment area plays an important role during LNG
spills. Liquid LNG and vaporized gas will be conned inside this
area. In the presence of impoundment walls, swirl and recirculation
will be formed inside the impoundment area as shown in Fig. 9
(right). Because of this kind of effect, when LNG spills, vapour
cloud will be dragged upwind. Fig. 10 shows the footprint of iso-
concentration contour of 2.5% at time stages in the dispersion,
from the beginning of spill to the end of spill. In the initial stage,
because of the dense gas effect, the vapour cloud falls down along
the tank outer wall into the impoundment area. Then, the cloud
conned in the area moves downwind over the front wall. Mean-
while, the cloud is dragged backward because of the wind swirl
inside the impoundment area. In the end, the impoundment is
almost full of vapour cloud. At T 100 s, the vapour dispersion
starts at steady-state, and the dispersion behaviour does not
change until the end of spill. Downwind distance and lateral
dispersion distance for iso-concentration 2.5% are 35 m and 75 m,
respectively. Almost 75% of the total dispersed vapour was conned
inside the impoundment area.
6.2. LNG spill without impoundment
Fig. 11 shows the spill case without the impoundment.
Compared to Fig. 10, there was a subtle change in the gas dispersion
behaviour. In this case, the steady-state dispersion reached much
earlier at 60 s. At the steady-state, for iso-concentration of 2.5%, the
downwind distance and lateral distance were 39.5 m and 63 m,
respectively. Thus the gas travelled 11% more in the downwind
direction whereas the spread was 16% less.
The comparison of the two cases shows impoundment is quite
important in controlling the effects of LNG spills. Within the
impoundment area, the wind swirl helps to drag the vapour cloud
upwind, and the dispersion in downwind direction is weakened.
Meanwhile, the impoundment walls strengthen the air ow
turbulence. It therefore, takes a longer time to arrive at the steady
state dispersion, 100 s compared with 60 s of the case without the
impoundment. Thus, LNG spill can be controlled more easily with
an impoundment.
However, further research is needed to evaluate the effect of
impoundment on the LNG evaporation rate. When liquid LNG spills
at the tank bottom and is conned inside the dike, the main heat
transfer to the cryogenic liquid is due to heat conduction from the
dike wall and the impoundment subsoil. At initial stage, the LNG
evaporation rate can be very high at around 0.19 kg/(m
2
s). As
time elapses, it would decrease to a very low value (less than
0.05 kg/(m
2
s)) (ioMosaic, 2007). If LNG spills without any dike,
then the liquid would ow freely without resistance. Evaporation
would be more severe, and an even larger area would be considered
hazardous.
7. Conclusion
The dispersion of LNG vapour has been simulated using
computational uid dynamics (CFD). Simulation results were
compared with the experimental results fromBurro series tests and
with integral model (DEGADIS). Compared to the integral models,
CFD simulations were in a much better agreement with the
experimental data, especially for the downwind dispersion. The
CFD simulations are also advantageous when dealing with
temporal and spatial dense gas dispersion, including three
dimensional analyses, turbulence modelling, gravity slumping
effect, complicated terrain and time depended effect. In the
prediction of downwind maximum concentration, relative error
between CFD simulations and experimental data was 19.62%. The
CFD model also took longer time to arrive at the steady state
compared to experimental results. These can be attributed to
averaging of the boundary conditions over the period of test. It was
also observed that stable atmospheric conditions take longer time
to arrive to the steady-state compared to the less stable conditions.
The inuencing area was also larger in the later case.
Simulations were then used for industrial risk assessment of
LNG spill as dened by NFPA 59A in order to determine the
exclusion distance from the most-likely-spill scenario. Dense gas
dispersion scenario with or without impoundment was examined.
In case of an impoundment, it was found that the dense vapour was
collected inside the impoundment area increasing the time to
achieve the steady state, which was quite helpful in conning the
liquid or vapour from spreading. Impoundment would also reduce
the evaporation rate to a very low amount, compared with the
scenario of without impoundment. These results and analysis are
important in evaluating and designing LNG regasication terminals
and LNG station sitting.
Acknowledgement
This work was supported from the Key Laboratory of LNG
Cryogenic Technology of Guangdong High Education Institute (No.
39000-3211101), the SYSU-BP LNG Centre (No. 99103-9390001)
and Key Laboratory of Fire Science and Technology of Guangdong
Fig. 11. Time-depended vapour cloud footprint (volume concentration contour 2.5%) at different time (spill duration 200 s).
B. Sun et al. / Journal of Loss Prevention in the Process Industries 26 (2013) 117e128 126
Province (No.2010A060801010). Author Sun also acknowledges the
support received from the Australia-China council scholarship.
Nomenclature
b(x) Homogeneous characteristic half width of gas plume in
DEGADIS model (m)
c(x, y, z) Local concentration in DEGADIS model (kg/m
3
)
c
c
(x) Centreline concentration in DEGADIS model (kg/m
3
)
C
1
Constant
C
2
Constant
C
1
Constant
C
3
Constant
C
m
Turbulence model constant
D Characteristic dimension of the source (m)
D
i,m
Mass diffusion coefcient (m
2
/s)
D
T,i
Turbulent diffusivity
E Total energy (J)
F
!
Force vector from source terms (N)
g
!
Gravitational acceleration (m/s
2
)
G
k
Generation of turbulence kinetic energy due to mean
velocity gradients (kg/(m s
3
))
G
b
Generation of turbulence kinetic energy due to buoyancy
(kg/(m s
3
))
h Cloud height in SLAB model (m)
h
j
Enthalpy of species j (J/kg)
I Unit tensor (1/s)
J
j
!
Mass diffusion (kg/(m
2
s))
k Turbulence kinetic energy (m
2
/s
2
)
k
eff
Effective conductivity (W/(m K)
L Cloud half width (m)
p Pressure (pa)
Q
0
Total vapour volumetric ow rate (m
3
/s)
R
i
Net rate of production of species i (kg/(m
3
s))
Ri
m
Modied Richardson number
Sc
t
Turbulent Schmidt number
S
h
Energy source (including chemical reaction)
S
i
User dened source (kg/(m
3
s))
S
k
User-dened source terms of turbulence kinetic energy
S
m
Mass added or any user-dened sources (kg/(m
3
s))
S
y
(x) Horizontal concentration scaling parameter in DEGADIS
model (m)
S
z
(x) Vertical concentration scaling parameter in DEGADIS
model (m)
S

User-dened source terms of turbulence dissipation rate


t Time (s)
T Temperature (K)
U
0
, U
a
Ambient air velocity (m/s)
U
t
, U
e
Horizontal and vertical entrainment rates (m/s)
U
x
Wind velocity along z-direction in DEGADIS model (m/s)
U(z) Wind speed at a given elevation (m/s)
v Kinematic viscosity (m
2
/s)
v
!
Overall velocity vector (m/s)
x, y, z Elevation along downwind, crosswind and vertical
direction (m)
Y
M
Fluctuating dilatation in compressible turbulence to the
overall dissipation rate (kg/(m s
3
))
Y
i
Mass fraction of species i
z
0
Height for detecting weed speed (8 m in this paper) (m/s)
Greek letters
a Constant in power law wind prole in DEGADIS model
Turbulent dissipation rate (m
2
/s
3
)
l Dimensionless parameter depended upon atmospheric
stability and surface roughness
m Dynamic viscosity (Pa s)
m
t
Turbulence viscosity (Pa s)
r Density of vapour (kg/m
3
)
r
a
Density of ambient air (kg/m
3
)
s
k
, s

Turbulent Prandtl numbers for k and


s Stress tensor (kg/(m s
2
)
References
ANSYS.. (2010). ANSYS FLUENT theory guide. USA: ANSYS Inc.
Arys, S. P. (1999). Air pollution meteorology and dispersion. New York-Oxford: Oxford
University Press.
Britter, R. E. (1988). Workbook on the dispersion of dense gases. HSE Contract
Research Report No. 17/1988.
Britter, R. E. (1989). Atmospheric dispersion of dense gases. Annual Review of Fluid
Mechanics, 21, 317e344.
Brown, T. C., Cederwall, R. T., Chan, S. T., Ermak, D. L., Koopman, R. P., Lamson, K. C.,
et al. (1990). Falcon series data report: 1987 LNG vapor barrier verication eld
trials. America: (LLNL) Lawrence Livermore Laboratory.
Cleaver, P., Johnson, M., & Ho, B. (2007). A summary of some experimental data on
LNG safety. Journal of Hazardous Materials, 140(3), 429e438.
Coldrick, S., Lea, C. J., & Ivings, M. J. (2010). Validation database for evaluating vapor
dispersion models for safety analysis of LNG facilities. Health & Safety Laboratory.
Cormier, B. R., Qi, R., Yun, G. W., Zhang, Y., & Mannan, S. M. (2009). Application
of computational uid dynamics for LNG vapor dispersion modeling: a study
of key parameters. Journal of Loss Prevention in the Process Industries, 22(3),
332e352.
Cornwell, J. B., & Pfenning, D. B. (1987). Comparison of thorney island data with
heavy gas dispersion models. Journal of Hazardous Materials, 16, 315e337.
Deaves, D. M., Gilham, S., Mitchell, B. H., Woodburn, P., & Shepherd, A. M. (2001).
Modelling of catastrophic ashing releases. Journal of Hazardous Materials,
88(1), 1e32.
Dharmavaram, S., Hanna, S. R., & Hansen, O. R. (2005). Consequence analysisdusing
a CFD model for industrial sites. Process Safety Progress, 24(4), 316e327.
Dorigoni, S., & Portatadino, S. (2008). LNG development across Europe: infra-
structural and regulatory analysis. Energy Policy, 36(9), 3366e3373.
Duijm, N. J., Carissimo, B., Mercer, A., Bartholome, C., & Giesbrecht, H. (1997).
Development and test of an evaluation protocol for heavy gas dispersion
models. Journal of Hazardous Materials, 56(3), 273e285.
Ermak, D. L. (1990). Users manual for SLAB: An atmospheric dispersion model for
denser-than-air releases. California 94550: Lawrence Livermore National
Laboratory.
Ermak, D. L., Chan, S. T., Morgan, D. L., & Morris, L. K. (1982). A comparison of dense
gas dispersion model simulations with burro series LNG spill test results.
Journal of Hazardous Materials, 6(1e2), 129e160.
Ermak, D. L., Chapman, R., Goldwire, H. C., Gouveia, F. J., & Rodean, H. C. (1988).
Heavy gas dispersion test summary report. Lawrence Livermore Laboratory,
UCRL-21210.
Essa, K. S. M., Embaby, M., & Etman, S. M. (2003). A notional variation of the wind
prole power-law exponent as a function of surface roughness and stability. In
4th Conference on nuclear and particle physics. Fayoum, Egypt.
ExxonMobil. (2008). Outlook for energy: A view to 2030.
Gavelli, F., Bullister, E., & Kytomaa, H. (2008). Application of CFD (uent) to LNG
spills into geometrically complex environments. Journal of Hazardous Materials,
159(1), 158e168.
Havens, J. (1998). A disperison model for elevated dense gas jet chemical releases.
USA: EPA.
Hightower, M., Gritzo, L., Luketa-Hanlin, A., Covan, J., Tieszen, S., Wellman, G., et al.
(2004). Guidance on risk analysis and safety implications of a large LNG spill over
water. Sandia Report SAND2004-6258. Sandia National Laboratories.
ioMosaic. (2007). Modeling LNG pool spreading and vaporization. An ioMosaic
Corporation Whitepaper.
Koopman, R. P., Baker, J., Cederwall, R. T., Goldwire, H. C., Hogan, W. J.,
Kamppinen, L. M., et al. (1982a). Burro series data report-LLNL/NWC 1980 LNG
spill tests, Vol. 1. Lawrence Livermore Laboratory.
Koopman, R. P., Baker, J., Cederwall, R. T., Goldwire, H. C., Hogan, W. J.,
Kamppinen, L. M., et al. (1982b). Burro series data report-LLNL/NWC 1980 LNG
spill tests, Vol. 2. Lawrence Livermore Laboratory.
Koopman, R. P., Cederwall, R. T., Ermak, D. L., Goldwire, H. C., Jr., Hogan, W. J.,
McClure, J. W., et al. (1982). Analysis of Burro series 40-m3 LNG spill experi-
ments. Journal of Hazardous Materials, 6(1e2), 43e83.
Koopman, R. P., & Ermak, D. L. (2007). Lessons learned from LNG safety research.
Journal of Hazardous Materials, 140(3), 412e428.
Kumar, S., Kwon, H. T., Choi, K. H., Cho, J. H., Lim, W., & Moon, I. (2011). Current
status and future projections of LNG demand and supplies: a global prospective.
Energy Policy, 39(7), 4097e4104.
Launder, B. E., & Spalding, D. B. (1972). Lectures in mathematical models of turbulence.
London: Academic Press.
B. Sun et al. / Journal of Loss Prevention in the Process Industries 26 (2013) 117e128 127
Lin, W. S., Zhang, N., & Gu, A. Z. (2010). LNG (liqueed natural gas): a necessary part
in Chinas future energy infrastructure. Energy, 35(11), 4383e4391.
Luketa-Hanlin, A. (2006). A review of large-scale LNG spills: experiments and
modeling. Journal of Hazardous Materials, 132(2e3), 119e140.
Luketa-Hanlin, A., Koopman, R. P., & Ermak, D. L. (2007). On the application of
computational uid dynamics codes for liqueed natural gas dispersion. Journal
of Hazardous Materials, 140(3), 504e517.
Mohan, M., Panwar, T. S., & Singh, M. P. (1995). Development of dense gas
dispersion model for emergency preparedness. Atmospheric Environment,
29(16), 2075e2087.
Mohan, M., & Siddiqui, T. A. (1998). Analysis of various schemes for the estima-
tion of atmospheric stability classication. Atmospheric Environment, 32(21),
3775e3781.
NFPA. (2009). Standard for the production, storage and handling of liqueed
natural gas.
Opschoor, G. (1980). The spreading and evaporation of LNG- and burning LNG-spills
on water. Journal of Hazardous Materials, 3(3), 249e266.
Pontiggia, M., Derudi, M., Busini, V., & Rota, R. (2009). Hazardous gas dispersion:
a CFD model accounting for atmospheric stability classes. Journal of Hazardous
Materials, 171(1e3), 739e747.
Qi, R., Ng, D., Cormier, B. R., & Mannan, M. S. (2010). Numerical simulations of LNG
vapor dispersion in Brayton Fire Training Field tests with ANSYS CFX. Journal of
Hazardous Materials, 183(1e3), 51e61.
Raj, P. K., & Lemoff, T. (2009). Risk analysis based LNG facility siting standard in
NFPA 59A. Journal of Loss Prevention in the Process Industries, 22(6), 820e829.
Shih, T. H., Liou, W. W., Shabbir, A., Yang, Z., & Zhu, J. (1995). A new k-[epsilon] eddy
viscosity model for high Reynolds number turbulent ows. Computers & Fluids,
24(3), 227e238.
Sklavounos, S., & Rigas, F. (2006). Simulation of Coyote series trials e part I: CFD
estimation of non-isothermal LNG releases and comparison with box-model
predictions. Chemical Engineering Science, 61(5), 1434e1443.
Spicer, T., & Havens, J. (1989). Users guide for the DEGADIS 2.1 dense gas dispersion
model. Report of U.S. Environmental Protection Agency.
Sutton, S. B., Brandt, H., & White, B. R. (1986). Atmospheric dispersion of a heavier-
than-air gas near a two-dimensional obstacle. Boundary-Layer Meteorology,
35(1), 125e153.
Tauseef, S. M., Rashtchian, D., & Abbasi, S. A. (2011). CFD-based simulation of dense
gas dispersion in presence of obstacles. Journal of Loss Prevention in the Process
Industries, 24(4), 371e376.
Tauseef, S. M., Rashtchian, D., Abbasi, T., & Abbasi, S. A. (2011). A method for
simulation of vapour cloud explosions based on computational uid dynamics
(CFD). Journal of Loss Prevention in the Process Industries, 24(5), 638e647.
TNO Report. (2005). Methods for the calculation of physical effects: Due to releases of
hazardous materials (liquids and gases).
Versteeg, H. K., & Malalasekera, W. (1995). An introduction to computational uid
dynamics e The nite volume method. England: Longman Group Ltd.
B. Sun et al. / Journal of Loss Prevention in the Process Industries 26 (2013) 117e128 128

Vous aimerez peut-être aussi