You are on page 1of 14

AN OVERVIEW OF THE NATURE OF

HYDROCARBON JET FIRE HAZARDS IN THE


OIL AND GAS INDUSTRY AND A SIMPLIFIED
APPROACH TO ASSESSING THE HAZARDS

B. J. Lowesmith1, , G. Hankinson1, M. R. Acton2 and G. Chamberlain3


1
Loughborough University, Loughborough, UK.
2
Advantica Limited (formerly British Gas Research and Development), Loughborough, UK.
3
Shell Global Solutions, Chester, UK.

Abstract: High pressure jet fires pose a serious hazard to offshore installations operated by the
oil and gas industry as demonstrated by the Piper Alpha incident. Following the Piper Alpha inci-
dent a major initiative by the offshore oil and gas industry operators led to the production of
Interim Guidance Notes which provided guidance to operators on how to assess jet fire hazards.
However, many areas of uncertainty were identified where no data was available. Areas of par-
ticular concern identified in the Interim Guidance Notes were two-phase jet fires, the effect of
confinement on jet fires and their behaviour with water deluge. Since that time a considerable
body of experimental research has been undertaken. Based on this recent data, this paper reas-
sesses jet fire hazards in an offshore environment and provides updated guidance on the
hazards they pose, including tabulated data and simple calculation techniques for predicting
jet fire hazards.

Keywords: jet fires; two-phase jet fires; offshore fire hazards; water deluge; thermal load.

BACKGROUND Prior to the Piper Alpha incident, there was


little public domain information concerning the
The oil and gas industry produces, processes quantification of fire hazards which could be

Correspondence to:
and conveys a wide range of hydrocarbon accessed by operators and the incident led
Dr B.J. Lowesmith, fuels from natural gas to heavy crude oil in to the initiation of a major Joint Industry Pro-
Department of Chemical large quantities and often at high pressures. ject (JIP) coordinated by the Steel Construc-
Engineering, Loughborough An ignited accidental release of pressurized tion Institute (SCI) entitled ‘Blast and Fire
University, Loughborough,
fuel can present a serious jet fire hazard to Engineering Project for Topside Structures’
Leicestershine, LE11 3TU,
UK. industry personnel and to the installation (BFETS) in 1990. Phase 1 reviewed the cur-
E-mail: b.j.lowesmith@ itself potentially leading to escalation of the rent understanding of fire and explosion
lboro.ac.uk incident due to further inventory loss or struc- hazards offshore, identified areas requiring
tural collapse. For example, the riser failure further research and provided Interim Gui-
DOI: 10.1205/psep06038
which occurred during the Piper Alpha Incident dance Notes (IGNs) for use by the offshore
0957–5820/07/ in 1988 (Cullen, 1990), contributed to the col- industry (Cowley and Johnson, 1992; SCI,
$30.00 þ 0.00 lapse of the platform and a major loss of life. 1992). In particular, the IGNs included gui-
Offshore installations present particular dance on the kinds of fires likely to occur off-
Process Safety and problems in relation to the protection of shore and, where possible, guidance on their
Environmental Protection
personnel from hazardous events. Space quantification, for example, in terms of fire
Trans IChemE, restrictions lead to closely packed process size and thermal loading. However, the gui-
Part B, May 2007 pipework and hence the likelihood of escala- dance recognized many areas of uncertainty
tion following a relatively small fire is where no reliable data could be provided.
# 2007 Institution
of Chemical Engineers
increased compared to an onshore facility. This led directly to the initiation of Phase 2
The close proximity of personnel and the diffi- of the project where major research projects
culties of escape are an important consider- were undertaken to obtain full scale data on
ation. Hence it is important for operators to the hazards presented by offshore fires and
have a good understanding of the fire hazards explosions to address the gaps in under-
posed by their installation and to be able to standing identified during Phase 1. As part
demonstrate that the risk is managed effec- of Phase 2, jet fires involving oil and gas mix-
tively as part of their Safety Case. tures were undertaken by Advantica Limited

207 Vol 85 (B3) 207–220


208 LOWESMITH et al.

(formerly British Gas Research and Development) and con- water deluge systems on fire behaviour and consequences
fined condensate jet fires were studied by SINTEF in are also summarized.
Norway (SCI, 1998; Advantica, 1997a). Guidance is then presented for a simplified approach to
Advantica were well qualified to undertake this work since quantifying jet fire hazards (in terms of values for flame
from the early 1980s Advantica had been studying jet fire size, temperature, heat loads, and so on) for the range of
hazards associated with natural gas, principally related to fire sizes likely to be considered in a quantified risk assess-
gas transmission pipelines. Recognizing the importance of ment (QRA). These typical values for jet fire hazards are
studying these hazards at a representative scale, large and based on the experimental data reviewed here sup-
full scale experiments were undertaken at their Spadeadam plemented with predictions by validated models (developed
Test Site in Cumbria to determine the fire characteristics by Shell and Advantica) and can be used to assess the
(Wickens and Lowesmith, 1993; Cook et al., 1987; potential hazard to personnel and the likely affect to fire
Hankinson et al., 2000), although much remains unpublished. impacted obstacles. The approach taken in defining jet fire
Other studies of jet fires (conducted jointly with Shell Global hazards is similar to, but more detailed than, that presented
Solutions) included other fuels such as propane and butane within the Energy Institute guidance on severe fires (Energy
of particular relevance to onshore storage facilities (Bennett Institute, 2003).
et al., 1991; Shell, 1992; Davenport, 1994a; Sekulin and As the focus is offshore operations, the fuels for which gui-
Acton, 1995). dance is required are principally pressurized gas, oil fractions
After Phase 2, Advantica continued work on gas jet fire and their mixtures, sometimes including water. The guidance
hazards (Gosse and Pritchard, 1995; Advantica, 1997b) does not extend to activities involving liquefied natural gas
and Shell undertook large scale experiments studying gas/ (LNG). Experimental studies generally used processed oil
kerosene jet fires (Davenport, 1994b). Attention was also fractions such as diesel or kerosene and these are con-
focused on fire mitigation using active water deluge. Advan- sidered to be representative of the behaviour of light crude
tica initiated two JIPs to study the effectiveness of water oils. Some experimental studies related to LPG (propane
deluge to mitigate jet and pool fire hazards (Advantica, and butane) and, although not generally experienced off-
1997c, 2000a), whilst Advantica and Shell jointly conducted shore except as part of the composition, where appropriate
a JIP to study jet fires involving oil/gas/water mixtures the results of these studies have been included.
(Advantica, 2000b). Only limited information from these
JIPs has been published thus far (Gosse and Hankinson,
NATURE AND CHARACTERISTICS OF JET FIRES
2001; Hankinson and Lowesmith, 2004; Hankinson et al.,
forthcoming). Experiments studying dedicated vessel Jet fires can be produced following the pressurized
deluge with jet fires have also been conducted at Spadea- release of a variety of fuel types. The simplest being a
dam and elsewhere (Shirvill and White, 1994; Shirvill, 2004; pressurized gas giving rise to a gas jet fire. A pressurized
Roberts, 2004). liquid/gas mixture (such as ‘live crude’ or gas dissolved in
The effect of confinement (with and without deluge) on fire a liquid) will give rise to a two-phase jet fire. The gas
behaviour was studied during the BFETS JIP experiments stream atomizes the liquid into droplets which are then
conducted at SINTEF and subsequently in further studies evaporated by radiation from the flame. However, a pressur-
by Shell (Chamberlain, 1994; Chamberlain and Brightwell, ized release of a liquid can also give rise to a jet fire in
1994). which two-phase behaviour is observed if the liquid is able
It was always the intention of the BFETS JIP that the to vaporize quickly. This is most likely to occur when a
IGNs would be updated after Phase 2 and this was liquid is released from containment at a temperature
completed initially in relation to explosion hazards. After above its boiling point at ambient conditions whereupon
some delay, the updating of the IGNs for fire hazards is flash evaporation occurs (e.g., propane, butane) and a
now in progress with the support of the UK Offshore flashing liquid jet fire results. Non-volatile liquids (e.g., kero-
Operators Association (UKOOA, 2006). sene, diesel, or stabilized crude) are unlikely to be able to
sustain a two-phase jet fire, unless permanently piloted by
an adjacent fire; even so, some liquid drop-out is likely
and hence the formation of a pool.
SCOPE OF PAPER
As part of the updating of the IGNs for fire hazards, infor-
Flame Stability
mation on the nature of fire hazards likely to be experienced
offshore and guidance on their quantification is being Whether or not a stable jet fire will arise following the
reviewed taking into account the wealth of large scale exper- releases of a pressurized hydrocarbon will depend principally
imental data that has been generated during and since upon the nature of the fuel, the size of the hole from which the
Phase 2 of the BFETS JIP, as described above. This paper release occurs and the geometry of the surroundings. In the
presents the results of this review in relation to jet fire case of natural gas, it has been found that for free jets (not
hazards which forms the basis of the guidance relating to impacting) some combinations of hole size and pressure
jet fires which is to be included within the updated industry cannot produce stable flames (Birch et al., 1988). Figure 1
guidance notes (UKOOA, 2006). shows that for hole sizes under 30 mm diameter, there is a
Based on data from the large scale experiments outlined lower bound pressure which vertical high pressure releases
above including much data which has not been previously must exceed to produce stable jet fires. In practice this
published, the nature and characteristics of hydrocarbon jet means that most small leaks will be inherently unstable and
fires, especially in relation to offshore oil and gas installa- will not support a flame without some form of flame stabiliz-
tions, are described. The effects of confinement and active ation, such as the presence of another fire in the vicinity to

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B3): 207– 220
HYDROCARBON JET FIRE HAZARDS IN THE OIL AND GAS INDUSTRY 209

the prevailing wind conditions except towards the tail of the


fire. By contrast, the generally lower exit velocities from flash-
ing liquid releases lead to flashing liquid jet fires with shorter
flame lift-offs and proportionately shorter and more buoyant
flames overall. These lower velocities also result in fires
that are more wind affected whilst the higher hydrocarbon
content of these fuels increases the flame luminosity. How-
ever, releases involving gas dissolved in, or mixed with, a
liquid can result in a two-phase jet fire which combines the
worst aspects of both the gas jet fire and the flashing liquid
jet fire, that is, high velocities and high flame luminosity.
Apart from providing flame stabilization, impact onto an
obstacle may modify the shape of a jet fire. Objects which
are smaller than the flame half-width at the point of impact
are unlikely to modify the shape or length of the flame to
any great extent. However, impact onto a large vessel may
significantly shorten the jet fire, and impact onto a wall or
roof could transform the jet into a radial wall jet where the
location and direction of the fire is determined by the
surface onto which it impacts and its distance from the
release point.
Figure 1. Stability of free (non-impacting) natural gas jet fires.

Combustion Characteristics and Smoke


provide a permanent pilot or stabilization as a result of impact In the case of high pressure releases of natural gas, the
onto an object such as pipework, vessels or the surrounding mixing and combustion is relatively efficient resulting in little
structure. soot (carbon) formation except for extremely large release
Figure 1 also includes data from horizontal free jet fires rates. Hence little or no smoke is produced by natural gas
with and without general area deluge at two different flow- jet fires (typically ,0.01 g m23). CO concentrations in the
rates (Advantica, 2000a) from which it can be seen that the region of 5– 7% v/v have been measured within a jet fire
unstable region is larger for horizontal jets and that deluge itself but this is expected to drop to less than 0.1% v/v by
further increases flame instability. However, in the highly con- the end of the flame. Considerably more soot is produced
gested environment offshore, impact within a short distance in jet fires involving higher hydrocarbons, although there is
is very likely, and hence small leaks are most likely to stabil- no available experimental data quantifying the difference.
ize on the nearest point of impact. Limited measurements in the smoke one flame length down-
stream of the end of a ‘live crude’ jet fire determined a per-
centage obscuration of typically 10% over a 200 mm path
length (Advantica, 2000a). This corresponds to a visibility
Flame Shape and Appearance
distance of about 5 m (Drysdale, 1985).
A jet fire is a turbulent diffusion flame produced by the com-
bustion of a continuous release of fuel. Except in the case of
Jet Fire Length
extreme confinement, which might give rise to extinguish-
ment of gas jet fires or liquid drop-out with two-phase jet Jet fire size is primarily related to the mass release rate.
fires, the combustion rate will be directly related to the For gaseous releases this, in turn, is related to the size of
mass release rate of the fuel. the leak (hole diameter) and the pressure (which may vary
In the offshore context, operating pressures are high so the with time as a result of emergency blowdown). For non-
flow of an accidental pressurized gas release into the atmos- gaseous fuels, the liquid content results in relatively higher
phere will be choked having a velocity on release equal to the release rates for a given aperture and pressure compared
local speed of sound in the fluid (called sonic releases). to gaseous releases and, when the release is two-phase
A factor which is often overlooked is the noise produced (such as may arise from a relatively long pipe connected to
by sonic gaseous releases. This is usually high pitched and a storage vessel containing a liquid above its boiling point),
so loud that it may prevent effective radio communication estimating the release rate is non-trivial. This is because it
between personnel. As a result emergency actions could is necessary to know what proportion of the fluid is in each
be hampered. phase and this proportion may be changing along the
For pressurized gaseous releases, following an expansion length of the pipe due to heat transfer through the pipe wall
region downstream of the release point, the flame itself com- and pressure changes.
mences in a region of sub-sonic velocities as a blue relatively Figure 2 shows jet fire lengths for a range of fuels plotted
non-luminous flame. Further air entrainment and expansion against the power of the flame in megawatts, Q ( ¼ mass
of the jet then occurs producing the main body of the gas release rate x net calorific value). The figure includes a corre-
jet fire as turbulent and yellow. In the absence of impact lation for jet fire length, L, based on the majority of the natural
onto an object, these fires are characteristically long and gas data and all the higher hydrocarbon data, which is
thin and highly directional. The high velocities within the
released gas mean that they are relatively unaffected by L ¼ 2:8893Q0:3728 (1)

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B3): 207 –220
210 LOWESMITH et al.

Figure 2. Jet flame length.

For high pressure gaseous releases, the mass release rate radiated, F, for such fires as can be seen in Figure 3. (F is
through an orifice, Ṁ, can be related to the area of the hole, defined as ‘(energy released as radiation from the flame
Ad, and pressure, P, using surface/net energy of combustion)’). As can be seen,
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi F increases with carbon number, reflecting the increased
u   ggþ1 ! radiative emissions from soot within the higher hydrocarbon
u g m 2 1
_ ¼ Cd Ad P10
M 6 t (2) jet fires.
RTg Z g þ 1
For gas– liquid fuel mixtures, interpolation based on the
percentage liquid within the mixture provides a reasonable
For natural gas, this can be approximated for circular holes estimate of F as can be seen in Figure 4 for butane/natural
of diameter, d (mm), by gas mixtures.
A special case of interest at some installations is ‘live’
_ ¼ d 2 P  103
M (3) crude that includes a significant quantity of water. Experi-
ments (Hankinson et al., 2007) have shown that mixtures
Substituting into equation (1), the jet fire length for natural with a ‘water cut’ [defined as (mass of water/mass of
gas can be approximated by fuel)  100%) of up to 125% remain flammable, although
not necessarily capable of supporting a stable flame in the
L ¼ 0:933d 0:746 P0:373 (4)
absence of some other supporting mechanism. The inclusion
of water also slightly increases flame length and flame
buoyancy, and significantly reduces the amount of smoke
Flame Radiative Emissions
produced. For water cuts less than 50% there is little
As noted above, the combustion process within a natural impact on the fraction of heat radiated but for higher water
gas jet fire is relatively efficient and produces little soot cuts the fraction of heat radiated is reduced (Figure 5).
(carbon). Consequently, these flames are not as luminous
as higher hydrocarbon flames. Radiation emissions from
Heat Loads and Flame Temperatures
natural gas jet fires arise mostly from water vapour and
carbon dioxide, except for very large releases where soot The thermal load to an engulfed object in a jet fire will be a
production starts to enhance the process. The long thin combination of radiative load and convective load from the
shape may also result in a flame path which may not be opti- hot combustion products passing over the surface. Clearly
cally thick. The net result is that the radiative heat transfer to the total heat flux which is imparted to an engulfed object
the surroundings is lower than for comparable higher hydro- will vary over the surface of the object. In addition, the relative
carbon jet fires. This is reflected in the fraction of heat proportions of convective and radiative heat flux will vary over

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B3): 207– 220
HYDROCARBON JET FIRE HAZARDS IN THE OIL AND GAS INDUSTRY 211

Figure 3. Fraction of heat radiated.

the surface, with the highest convective component likely to cylinder (pipe) impacted by a horizontal high pressure gas
be experienced close to the point of impact of a flame jet fire (Pritchard and Cowley, 1991). The cylindrical surface
where the highest velocities occur, whereas the highest is presented flat by cutting along the rear. When the pipe
radiative heat load will be experienced where the more was located at 21 m (towards the end of the flame) the
radiative part of the flame (usually towards the end of the maximum heat fluxes were experienced at the point of
flame) is viewed by the object. As the more radiative part of impact on the front of the pipe [Figure 6(a)]. At 15 m
the flame is closer to the tail, this can result in the highest [Figure 6(b)] the heat loads were relatively uniform around
overall heat fluxes being experienced on the rear surface of the pipe, but at 9 m [Figure 6(c)] the heat loads were greatest
an engulfed object which may seem counter-intuitive. to the rear of the pipe due to radiation from the tail of
Figure 6 shows total heat fluxes experienced by a horizontal the flame.

Figure 4. Fraction of heat radiated for fuel mixtures. Figure 5. Effect of water cut on fraction of heat radiated.

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B3): 207 –220
212 LOWESMITH et al.

Figure 6. Variation of total heat flux over the surface: (a) pipe at 21 m; (b) pipe at 15 m; (c) pipe at 9 m.

Due to the radiant soot emissions, the radiative heat trans- phenomenon can be explained by considering the maximum
fer from higher hydrocarbon jet fires is generally greater than time averaged flame temperatures measured in over 180
that from natural gas flames and the generally lower vel- large scale gas jet and flashing liquid jet fires (Bennett
ocities arising from flashing liquid releases (such as propane et al., 1991; Sekulin and Acton, 1995; Advantica, 1997b) pre-
or butane) result in a lower convective flux to engulfed sented in Figure 9. As can be seen, the maximum flame
objects. temperatures are higher for the gas jet fire (on average
For the rear surface of an engulfed object, Figure 7 shows 12808C and up to 15008C), compared to a flashing liquid
that the fraction of the heat flux which is radiative increases fire which has a lower flame temperature (typically 10508C).
from about 0.5 for natural gas to about 0.8 for fuels containing In the case of two-phase jet fires, the flame temperature is
a large proportion of higher hydrocarbons (data from Bennett dominated by the gas content but the flame emissivity will
et al., 1991; Davenport, 1994a, b; Advantica, 1997a). also be enhanced by the higher hydrocarbon content leading
In the case of a pressurized gas–liquid mixture (such as to overall higher radiative fluxes for such mixtures.
‘live’ crude), the high gas velocities may still occur and Figure 8 includes data with mass flowrate between 2.5 and
result in a high convective contribution, whilst the higher 5 kg s21 and shows that for a flashing liquid fuel the maxi-
hydrocarbon content maintains a high radiative contribution; mum heat flux is generally around 200 kW m22 whereas for
making these type of jet fires a ‘worst case’ in terms of total gas only (0% liquid), the maximum heat flux is typically
heat flux to engulfed obstacles. Experimental work 250 –275 kW m22. In experiments involving natural gas at
(Davenport, 1994a, b; Advantica, 1997a) suggests that the higher flowrates (up to 10 kg s21) maximum heat fluxes up
maximum combined fluxes occur for gas –liquid mixtures
which are about 60 –80% by mass of liquid (Figure 8). This

Figure 7. Radiative heat flux as a fraction of total heat flux to rear of Figure 8. Maximum heat fluxes to an engulfed object for gas– liquid
engulfed object. mixtures.

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B3): 207– 220
HYDROCARBON JET FIRE HAZARDS IN THE OIL AND GAS INDUSTRY 213

particularly if the fire is not impacting onto an obstacle


(Figure 1). However, in most practical cases at the industry
standard deluge rate of 12 l m22 min21, this undesirable
effect is unlikely to occur due to impact onto obstacles provid-
ing adequate flame stabilization.
Although there is some evidence that the deluge increases
combustion efficiency resulting in lower CO and increased
CO2 levels within the flame (Advantica, 2000a), in most
other respects deluge has little effect on the size, shape
and thermal characteristics of a high pressure gas jet fire.
Therefore, the heat loading to engulfed obstacles is not
diminished. The same is true for dedicated vessel deluge
systems; the water being unable to form a film over the
vessel in the presence of the high velocity jet, and so dry
patches form where the temperature rise is undiminished
Figure 9. Maximum flame temperatures.
by the action of deluge (Shirvill and White, 1994).
Compared to the situation with a gas jet fire, the use of
dedicated vessel deluge to protect a vessel against a flashing
to 320 kW m22 were measured. This was due to a higher liquid jet fire (e.g., propane, butane) can be more effective.
radiative component arising from increased flame thickness The water interacts with the flame to some extent; reducing
and soot production within the flame. Broadly speaking, the flame luminosity and the amount of smoke produced.
neglecting spatial variations, for a given location of an Nevertheless, at typical application rates (10 –15 l m22
object within a flame (as a proportion of flame length), the min21) it cannot be relied upon to maintain a water film
convective component for a gas jet fire is more or less con- over the vessel and hence to prevent vessel temperature
stant with increasing size of release, whereas the radiative rise in areas where dry patches form, although the rate of
component increases with release rate as the flame becomes rise may be expected to reduce to 20 –70% of the rate with-
optically thick and more smoky. Hence the relative proportion out deluge for a propane jet fire (Shirvill, 2004). However, in
of convective to radiative flux varies with fire size. tests with an increased water application of 40 l m22 min21,
For fuels containing water, water cuts under 50% result in a 2 tonne LPG tank was effectively protected when subjected
no significant reduction in heat fluxes to engulfed objects to a 2 kg s21 flashing propane jet fire (Roberts, 2004).
(,10%). However, over 50% the flames are significantly For two-phase jet fires of ‘live’ crude, dedicated deluge (at
less radiative, and the overall heat flux to an obstacle can 10 l m22 min21) offered limited protection and no reduction in
be reduced by 40% or more (Advantica, 2000b; Hankinson the rate of temperature rise in the area where the fire
et al., 2007) (Figure 10). impacted the obstacle (Advantica, 2000a; Hankinson and
Lowesmith, 2004). This work also showed that using area
Water Deluge of Jet Fires deluge alone at the ‘standard’ rate of 12 l m22 min21 is unli-
kely to modify the flame behaviour although there is some
The most usual forms of active water deluge are water evidence that a higher deluge rate (24 l m22 min21) can
curtains (used to protect an escape corridor), general area result in water interaction with the flame, resulting in a shorter
deluge and dedicated vessel deluge. The success or other- flame and some reduction in heat fluxes to certain areas of an
wise of active water deluge in mitigating the effect of jet engulfed object, notably the front (where flame impact
fires depends upon the nature of the jet fire (gas, liquid or occurs) and top areas. Since dedicated vessel deluge is
two-phase) and objective to be achieved (reduction of inci- more effective at reducing the radiative heat fluxes in the
dent thermal radiation or protection of engulfed objects). region to the rear of the vessel, the combination of area
As noted above, the activation of general area deluge can deluge (at the higher rate) and dedicated vessel deluge can
adversely affect the stability of high pressure gas jet fires, be effective in reducing overall heat fluxes to a vessel such
that the temperature rise is halted or at least the rate of temp-
erature rise is reduced. This may prevent vessel failure,
especially if combined with a blowdown strategy. The relative
performance of these deluge strategies for a two-phase ‘live’
crude jet fire impacting onto a vessel (Hankinson and
Lowesmith, 2004) are shown in Figure 11. Measurements in
the combustion products downstream of these ‘live crude’ jet
fires determined an obscuration factor due to smoke of typi-
cally 10% over a 200 mm path length without deluge (visibility
distance of about 5 m) and no significant change was noted in
the presence of area deluge at 12 l m22 min21. At 24 l m22
min21 the smoke changed from black to grey due to the
increased water vapour and this had the effect of increasing
the obscuration factor to about 20%, corresponding to a visi-
bility distance of between 1 m and 2 m.
Figure 10. Effect of water cut in the fuel on maximum heat fluxes to The major benefit of area deluge with all kinds of jet fires
an engulfed object. arises from the suppression of incident thermal radiation to

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B3): 207 –220
214 LOWESMITH et al.

Figure 11. Performance of different deluge strategies with a two-phase jet fire. (a) Top of vessel; (b) front; (c) rear; (d) bottom.

the surroundings, which protects adjacent plant and in par- Increased deluge rates can further reduce incident radi-
ticular, aids escape by personnel. Similarly water curtains ation levels: 60 –70% at 18 l m22 min21; 80– 90% at 24 l
can be used to protect escape corridors. Figure 12 presents m22 min21 for general area deluge. Nozzles producing smal-
large scale data (Advantica, 1997c; Advantica, 2000a) on the ler droplet sizes can have an enhanced mitigation effect, but
reduction of incident radiation due to water deluge by MV57 there is an increased risk that the droplets will be blown away
type nozzles spaced 2.84 m apart. For this arrangement, by the wind (Advantica, 2000a).
operating at 12 l m22 min21, incident radiation levels can The mitigation of incident radiation is due to attenuation of
be reduced by about 20% for a single row of nozzles, radiation by the water droplets in the atmosphere, that is,
30–40% for two rows and 40–60% for more than two rows effectively reducing the atmospheric transmissivity, t. The
(general area deluge). reduction in incident radiation, Is, with distance, s, due to
this attenuation is characteristically expressed by an equation
of the form

Is ¼ I0 eats (5)

When using deluge, additional water is present in the


atmosphere. For a deluge system with N nozzles over an
area Aw, each supplied with V̇w litres per minute and produ-
cing droplets with a velocity Uw, the water volume fraction,
Wf, is given by

N:V_ W 1
Wf ¼  (6)
AW :UW 60 000

From which it can be seen that as smaller droplets will


have a lower terminal velocity, the water volume fraction
will be comparatively higher. Multiplying Wf by the pathlength
through the deluged area (that is, the ‘thickness’ of the water
curtain) gives a measure of the average water pathlength.
Figure 13 presents data on the reduction in incident radi-
Figure 12. Reduction in incident radiation by water curtains of MV57 ation as a function of average water pathlength for a range
nozzles. of nozzle types being: TF12-170 producing low velocity

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B3): 207– 220
HYDROCARBON JET FIRE HAZARDS IN THE OIL AND GAS INDUSTRY 215

trapped and recirculate. This leads to increased heat fluxes


to objects and surfaces compared to an unconfined fire
(SCI, 1998; Chamberlain, 1994).
The location where combustion occurs and the hottest
parts of the flame may also shift due to the confinement. In
tests (Chamberlain, 1994) involving horizontal jet fires in a
compartment incorporating a single wall vent, where the jet
was directed away from the vent, increased temperatures
were seen at the interface between the smoke layer leaving
the compartment and the air layer entering the compartment,
most particularly in the area furthest from the vent.
Unlike unconfined fires, the behaviour of under-ventilated
confined fires changes with time as the air initially available
within the compartment is consumed, and this may lead to
‘external flaming’ after a period of time when the body of
flame moves through the vent in order to find the oxygen
Figure 13. Effectiveness of water deluge in reducing incident
radiation.
required for combustion. For gas jet fires, CO levels of up
to 5% v/v at the vent may occur but after the onset of external
combustion the CO levels drop to typically less than 0.5% v/v
by the end of the flame (Chamberlain, 1994). Soot production
droplets with a Sauter mean diameter of 300 micron; HV60
is related to the equivalence ratio f and hence the degree of
producing droplets in the range 670 –750 micron depending
ventilation and may range from about 0.1 g m23 at f ¼ 1.3 to
on the operating pressure; MV57 producing droplets
up to 2.5 g m23 at f ¼ 2 (Chamberlain and Brightwell, 1994).
640–870 micron; and a prototype Large Droplet Nozzle
The worst case condition in terms of fire severity is likely to
(LDN) producing droplets about 1890 micron. The best-fit
occur if the jet fire is slightly under-ventilated as this leads to
equation of the appropriate form fitted to the data is
high heat release rates and enhanced soot production.
y ¼ 100  (1 2 e 21005x). Therefore, the reduction in incident
Deluge of a confined gas jet fire may lead to flame extin-
radiation, RI (%), for a curtain of thickness Cw is
guishment and hence a serious explosion hazard from the
continuing release. In the case of a two-phase jet fire, extin-
RI ¼ 100  (1  e1005Wf Cw ) (7)
guishment may result in a mist-air explosion hazard and/or
the formation of a liquid pool. The likelihood of flame extin-
guishment is significantly increased if the surroundings are
Confined Jet Fires already hot at the time the deluge is activated as the main
The behaviour of a jet fire within a confined or partially con- mechanism which results in extinguishment of the jet fire is
fined area will depend upon the degree of confinement and ‘inerting’, that is evaporation of the water droplets leading to
the direction of the jet relative to the ventilation opening. It a mixture of gas/air/steam within the compartment which is
should be noted that certain ventilation patterns could lead outside the flammable limits. The water vapour may also con-
to flame instability and extinguishment. tribute to flame instability by reducing the burning velocity.
The amount of air required for complete stoichiometric However, if the deluge is activated at an early stage, prior
combustion of methane is about 17.2 times the fuel flowrate to the compartment walls becoming hot, then the fire may
on a mass basis. For higher hydrocarbons this figure is about not be extinguished and some benefit in terms of reduced
15. The equivalence ratio, f, for a confined fire in terms of the flame temperatures and wall temperatures may accrue
mass burning rate of the fuel and the mass flowrate of air (SCI, 1998; Chamberlain, 1994).
entrained (Ṁ and Ḃ) is defined as In tests (Advantica, 2000a) involving only partial confine-
ment around the upper area of a module, during which
M_ ‘dead spaces’ occurred close to the ceiling, area deluge
f¼r (8) was found to be beneficial in reducing heat fluxes to the ceil-
B_
ing surface, reducing the flame extent and the amount of
where r is the mass ratio of air to fuel required for stoichio- smoke produced.
metric burning (that is r  15). If f , 1 then the fire is well
ventilated (also termed ‘fuel controlled’) and if f . 1 the fire
A SIMPLIFIED APPROACH TO ASSESSING JET FIRE
is under-ventilated (or ‘ventilation controlled’).
HAZARDS
If ventilation is plentiful (f , 1) or the jet is directed out of a
confined region through a vent then there may be little differ- It is normal practice for a QRA of an oil and gas installation
ence in jet fire characteristics compared to an unconfined fire. to consider a range of sizes of event. The critical parameter in
However, if the release rate is large relative to the size of the determining fire size is the mass release rate, which in turn is
confinement or the ventilation openings are small then the fire related to the leak size and pressure. A leak classification
may not be able to entrain enough air for complete combus- system which is often used, considers holes of 3, 10, 30
tion inside the compartment (f . 1). This is likely to result in and 100 mm diameter which, for an operating pressure of
increased levels of incomplete combustion products such as about 100 bar, correspond to gaseous release rates of
CO, increased levels of smoke (soot) and increased flame about 0.1, 1, 10 and 100 kg s21.
temperatures, particularly in regions close to the ceiling of Ideally, during a QRA, the fire size and thermal loading
a compartment where hot combustion products may be from fires should be assessed using mathematical models

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B3): 207 –220
216 LOWESMITH et al.

which have been extensively validated against large scale heat released by combustion as
data. A range of such models are available on a licence or
consultancy basis. However, a simplified approach was pro- EAf

posed in the Interim Guidance Notes (SCI, 1992), whereby 1000 Q
correlations for flame dimensions were suggested for jet
and pool fires together with guidance on typical heat loadings The incident radiation received in the far field at a distance,
to engulfed objects. Similarly, guidance on typical heat load- s (m), from the fire, Is, is then expressed as
ings from fires was provided in the Energy Institute document _
for severe jet and pool fires (Energy Institute, 2003). 1000 tF MH
Is ¼ 2
(10)
In this paper, a similar approach is taken in order to provide 4 ps
tabulated guidance on assessing jet fire hazards, based on The atmospheric transmissivity will depend upon the pre-
recent knowledge and large scale experimental work. The vailing atmospheric conditions (absolute humidity) and the
guidance values (which are conservative) can then be used path length, but might typically be 0.8 on a dry day for dis-
to undertake an initial simplified QRA. If necessary, problem tances likely to considered. However, fog would significantly
areas can then be identified for more rigorous assessment. reduce this value.
Guidance values for jet fire hazards are provided for the fol- The activation of water deluge on an offshore installation,
lowing range of sizes of leak: producing water droplets in the air, effectively reduces the
. Small—typically 0.03 –0.3 kg s21, represented by transmissivity by absorbing radiation. However, for the pur-
0.1 kg s21; poses of this paper, a revised ‘effective’ fraction of heat
. Medium—typically 0.3–3 kg s21, represented by 1 kg s21; radiated is defined as F0, accounting for the reduced trans-
. Large—typically 3–30 kg s21, represented by 10 kg s21; missivity in the area being deluged, (but not including the
. Major failure—over 30 kg s21. transmissivity of the atmosphere between the fire and the
receiver which is outside the deluged area). The incident radi-
For each size category, the information is provided to allow ation at a distance, s (m), from the fire can be estimated using
assessment of the fire size, the incident radiation field to per-
sons and objects outside the fire, the thermal loading to _
1000 tF 0 MH
objects engulfed by the fire, the smoke hazard presented Is ¼ (11)
4 ps2
by the fire, the likely effects of active water deluge and the
effects of confinement. (It should be noted that for leak This ‘point source’ model can be useful for estimating the
sizes of 10 kg s21 and above, the fire is large compared distance to a given radiation level. In the case of personnel,
with the average offshore module and it may be necessary it is important to consider both the level of radiation and the
to consider the fire to be ‘confined’.) duration of exposure. The thermal dose is often defined
as the product of the incident radiation and the duration
(kJ m22). However, the effect on people correlates better
with a ‘thermal load’, J defined by
Incident Radiation to the Surroundings ð
J ¼ Is(4=3) dt (12)
An obvious hazard presented by a fire is the thermal radi-
ation to the surroundings, in particular to personnel during
escape and evacuation. The incident thermal radiation, I, to Threshold values of dosage associated with serious injury
a person or object from a fire can be described in terms of or death can be found in the literature (Lees, 2005; UKOOA,
the flame emissive power, E, and the view factor, V, of the 2006). However, as a guide, a radiation level of about 5 kW
flame from the position of the receiver as m22 can be tolerated for about one minute and consequently
represents a level from which it is reasonable to assume that
escape, without significant injury, would usually be possible
I ¼ VE t (kW m2 ) (9) for an average employee wearing typical protective clothing.
However, the point source model is not suitable for estimat-
The view factor is a function of the flame shape. Conse- ing incident radiation to locations close to (certainly within
quently, most integral or empirical mathematical models of one flame length) of the flame, where it may be significantly
fires will assume some kind of simplified flame shape which in error. In the near field, a mathematical model which defines
is then used to calculate the view factor. The flame average a realistic flame shape (and ideally a variation of surface
surface emissive power is also a function of the flame emissive power over the flame to allow for smoke obscura-
shape. Therefore, average surface emissive powers used tion) should be used. Typical values of F and F0 are provided
by the model will not necessarily be the same as those in the tabulated guidance for the different sizes and types of
measured during an actual fire. jet fire so that estimates of the far field incident radiation
In the far field, (typically more than two flame lengths away) hazard can be made using the point source model equations
the flame shape is not critical, so a simplified approach can (10) and (11) given above.
be taken using the ‘point source’ model, whereby the difficul-
ties of defining a flame shape and associated average sur-
Thermal Loading to Engulfed Objects
face emissive power can be avoided. In this approach, the
fraction of the heat of combustion of the fuel radiated to the The thermal load per unit area, ql, to an object engulfed by
surroundings, F, is defined in terms of the flame area, Af, fire will be a combination of radiation from the flame (qrl) and
the flame surface emissive power, E, and the net rate of convection from the hot combustion products (qcl) passing

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B3): 207– 220
HYDROCARBON JET FIRE HAZARDS IN THE OIL AND GAS INDUSTRY 217

over the object surface. Hence the thermal load can be writ- provided in the tabulated guidance for the jet fire types and
ten as sizes.
However, researchers often quote heat fluxes measured
ql ¼ qrl þ qcl ¼ 1f sTf4 þ tf sTa4 þ h(Tf  Ts ) during experiments using calorimeters (for total heat flux)
and radiometers (for radiative flux). These instruments are
However, not all the thermal loading is necessarily designed to have a surface emissivity close to 1 and are
absorbed by the surface, some may be reflected back and maintained at a low temperature throughout the experiments.
the surface will also lose heat by radiation. Hence the total Hence the fluxes measured and reported for calorimeters are
absorbed load per unit area (qa) is given by given by equation (13) with 1s ¼ 1 and Ts  333 K and can be
regarded as a conservative estimate of the total heat flux
qa ¼ 1f as sTf4  1s af sTs4 þ as tf sTa4 initially absorbed by an engulfed object. Radiometers are
designed and calibrated to measure 1f sTf4 . At an early
 1s tf sTs4 þ h(Tf  Ts )
stage, whilst Ts is low, the flux measured by a radiometer is
approximately (qra =1s ) and so, if 1s is taken as 1, it provides
Assuming that both the flame and surface can be con-
a conservative estimate of the radiative flux absorbed by an
sidered as diffuse grey bodies, then as ¼ 1s, af ¼ 1f and
object. Subtracting the measured radiative flux from the
tf ¼ 1 2 1f. Furthermore, the term involving Ta is small and
measured total heat flux enables the initial convective flux
can be neglected, giving
absorbed by the object to be determined. Using an estimated
or measured value for Tf enables h to be determined at the
qa (kW m2 ) ¼ qra þ qca
measurement location. In this way, experimentally measured
¼ 1s s(1f Tf4  Ts4 ) þ h(Tf  Ts ) (13) flux levels can be used to derive input data for transient heat-
up calculations. Consequently, typical values of the total,
radiative and convective fluxes (as defined in the above
When considering an actual object engulfed by a fire, the
with 1s ¼ 1 and Ts  333 K) are also provided in the tabulated
flame temperature exposed to different parts of the object
guidance for the range of fire types and sizes. These rep-
surface may vary; similarly the velocities in the flame (and
resent initial heat fluxes to the fire engulfed object (that is,
hence convective heat transfer coefficient) may vary.
when the object surface is still cold and the temperature
Hence, to determine the total load absorbed by an object
difference is greatest). A simplified method for heat-up calcu-
the above equation should be summed over the area of the
lations can then be used (such as that given in UKOOA,
object.
2006; Energy Institute, 2003) to determine the temperature
Also, as the object engulfed in the flame heats up, the
of the engulfed object with time.
absorbed load will reduce. This is particularly the case with
the convective load which reduces linearly with increasing
object temperature. Therefore, for an accurate transient cal-
Tabulated Guidance
culation of the temperature rise of an object, the parameters
Tf, 1f and h are required, together with the emissivity of the Table 1 provides guidance values for gas jet fires and
surface (which itself may change as the surface heats up). Table 2 for two-phase jet fires. As noted above, the maximum
Where possible, typical values of these parameters are heat fluxes to engulfed objects have been found to occur

Table 1. Guidance values for gas jet fires.


21
Size (kg s ) 0.1 1.0 10 .30 Effect of confinement

Flame length (m) 5 15 40 65 Affected by enclosure shape


Fraction of heat radiated, F 0.05 0.08 0.13 0.13 and openings
CO level (% v/v) and smoke CO , 0.1 CO , 0.1 CO , 0.1 CO , 0.1 Increased CO up to about 5%
concentration (g m23) Soot 0.01 Soot 0.01 Soot 0.01 Soot 0.01 v/v at a vent prior to external
flaming, but after external
flaming ,0.5% v/v exiting
the flame. Soot levels
depend on equivalence ratio
from about 0.1 g m23 at
f ¼ 1.3 to 2.5 g m3 at f ¼ 2
Initial total heat flux (kW m22) 180 250 300 350 Increased heat loadings up to
Initial radiative flux (kW m22) 80 130 180 230 400 kW m22 (280 kW m22
Initial convective flux (kW m22) 100 120 120 120 radiative, 120 kW m22
Flame temperature, Tf (K) 1560 1560 1560 1560 convective, Tf ¼ 1600 K,
Flame emissivity, 1f 0.25 0.4 0.55 0.7 1f ¼ 0.75, h ¼ 0.09 kW m22
Convective heat transfer 0.08 0.095 0.095 0.095 K21)
coefficient, h (kW m22 K21)
Effect of deluge No effect on heat loadings to engulfed objects. Risk of extinguishment and
In far field, take F0 ¼ 0.8 F for one row of water sprays, F0 ¼ 0.7 F explosion hazard if deluge
for two rows and F0 ¼ 0.5 F for .2 rows at 12 l m22 min21. activated when enclosure is
May improve combustion efficiency and reduce CO levels already hot and fire is well
within flame established

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B3): 207 –220
218 LOWESMITH et al.

Table 2. Guidance values for two-phase jet fires.

Size (kg s21) 0.1 1.0 10 .30 Effect of confinement

Flame length (m) 5 13 35 60 Affected by enclosure shape


Fraction of heat radiated, F See note below table and openings
CO level (% v/v) and smoke CO , 0.1 CO , 0.1 CO , 0.1 CO , 0.1 Increased CO up to about 5%
concentration (g m23) Soot 0.01 Soot 0.01 Soot 0.01 Soot 0.01 v/v at a vent prior to external
flaming, but after external
flaming ,0.5% v/v exiting
the flame. Soot levels
depend on equivalence ratio
from about 0.1 g m23 at
f ¼ 1.3 to 2.5 g m23 at
f¼2
Initial total heat flux (kW m22) 200 300 350 400 Increased heat fluxes, take
Initial radiative flux (kW m22) 100 180 230 280 values as per 30 kg s21
Initial convective flux (kW m22) 100 120 120 120 two-phase jet fire
Flame temperature (K) 1560 1560 1560 1560
Flame emissivity, 1f 0.3 0.55 0.7 0.85
Convective heat transfer 0.08 0.095 0.095 0.095
coefficient, h (kW m22 K21)
Effect of deluge Some benefit to engulfed objects but temperature may still rise Risk of extinguishment and
although at a slower rate. Combined area and dedicated deluge potential formation of pool
may prevent temperature rise if effectively applied.
Take F 0 as per Table 1.

Note: For fraction of heat radiated, Fm, of mixture involving x% liquid by mass: use Fm ¼ (x=100):(FL  FG ) þ FG where FG is the fraction of heat
radiated for natural gas as given in Table 1 and FL is the fraction of heat radiated for the liquid fuel involved. Take FL ¼ 0.24 for C3; 0.32 for C4,
0.45 for C6– C25 (including condensate and diesel); and 0.5 for crude oil.

when the fuel mixture is about 30% gas and 70% liquid by these guidance values with the data measured during the
mass. Consequently, the values shown in Table 2 correspond tests for maximum total heat flux and the proportion of radia-
to this worst case condition. tive and convective fluxes measured at locations on the front,
For flashing liquid fires (such as propane or butane) a top and rear surface of the fire engulfed pipe. Clearly, the
lower flame temperature of about 1300 K is likely with an simplified guidance values take no account of the spatial vari-
emissivity of 1, giving a radiative flux of about 160 kW m22. ation of heat loading over the surface of a fire engulfed object.
A convective heat transfer coefficient of about 0.08 kW m22 For a conservative assessment of the response of the pipe
K21 is suggested giving a convective flux of about 80 kW to the heat loading, the temperature of the engulfed pipe with
m22 and a total flux of about 240 kW m22. time can be calculated for a series of time steps (Dt) as
follows:
At time ti ¼ ti1 þ Dt(i  1) the thermal load absorbed by
Validation Exercise the object, qi , is given by
Consider a ‘live’ crude jet fire at 5 kg s21, which is 20%
gas, 80% oil, released at 20 bar. Three experiments with qi ¼ 1s s(1f Tf4  Ti4 ) þ h(Tf  Ti ) (i  0) (14)
these test conditions were conducted during the BFETS
Phase 2 project (SCI, 1998; Advantica, 1997a), two of
which impacted onto a pipe (Tests 9 and 11) and the third
being a free flame (Test 3).
Using Table 2, this would fall into the ‘large’ release cat-
egory of 10 kg s21 giving a flame length of 35 m which
gives a conservative assessment compared to the measured
length of 29 m for the free flame and 21 m and 27 m for the
impacting flames.
Using the formula for fraction of heat radiated below Table 2,
with FG ¼ 0.13 and FL ¼ 0.5, Fm ¼ (80=100):(0:5  0:13)
þ0:13 ¼ 0:426, Figure 14 presents incident radiation
measured during the jet fires with calculations made using
the point source model Is ¼ (1000tF MH _ m )=(4ps2 ), taking
t ¼ 0.8 and Hm ¼ 42 MJ kg21.
Table 2 suggests a maximum total heat flux of 350 kW m22
whereas during the two experiments where the flame
impacted a steel pipe, the maximum heat fluxes measured
were 348 and 370 kW m22 . Table 2 suggests that the radia-
tive flux accounted for 65% of the total flux (230 kW m22) with
34% being convective (120 kW m22). Table 3 compares Figure 14. Incident radiation from ‘live’ crude oil jet fire.

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B3): 207– 220
HYDROCARBON JET FIRE HAZARDS IN THE OIL AND GAS INDUSTRY 219

Table 3. Comparison of guidance values for heat flux with large scale data.

Percentage radiative Percentage convective

Maximum
heat flux Front Top Rear Avg Front Top Rear Avg

Guidance values 350 66% 34%

Data Test 9a 370 41% 61% 68% 57% 59% 39% 32% 43%
Data Test 11a 348 43% 86% 100% 76% 57% 14% 0% 24%
a
Test 9 impacted the pipe at 9 m and Test 11 at 15 m from the release point.

on this recent large scale experimental data, this paper has


reviewed jet fire hazards in relation to offshore situations
and provided updated guidance on the hazards they pose.
Tabulated data are provided which can be used as input for
mathematical models. Additionally, simple calculation
methods are given whereby an initial conservative assess-
ment of jet fire hazards can be readily undertaken.

NOMENCLATURE
a constant
Figure 15. Heat up of engulfed object in ‘live’ crude jet fire. Ad area of release hole, m2
Af area of flame, m2
Aw deluged area, m2
Ḃ flowrate of air entrained for combustion, kg s21
and the temperature of the object increases to Tiþ1 ¼ C specific heat capacity of steel, kJ kg21 K21
Cd coefficient of discharge of orifice
Ti þ DTi (i  0) where
Cw water curtain thickness, m
d diameter of release hole, mm
Dtqi E flame emissive power, kW m22
DTi ¼ (15)
Crx F fraction of heat radiated by a flame
H net calorific value, MJ kg21
Figure 15 shows a comparison of the calculated tempera- h convective heat transfer coefficient, kW m22 K21
I incident radiation, kW m22
ture rise of the pipe with data measured by thermocouples
J thermal load to personnel, (kW m22)4/3 s
embedded within the pipe wall at three locations during the Is incident radiation at a distance s, kW m22
two impacting fire experiments. The three locations were: L jet fire length, m
close to the point of impact of the fire (front of pipe); 908 Ṁ mass flow rate, kg s21
from this location on top of the pipe; and 1808 from the m molecular weight, kg
N number of nozzles over a deluged area
point of impact on the rear of the pipe. The data used for P absolute pressure, MPa
the calculations was: C ¼ 520 J kg21 K21, r ¼ 7850 kg Q net rate of heat released by combustion (Q ¼ MH),_ MW
m23, s ¼ 5.6697 W m22 K24, x ¼ 0.0125 m, 1s ¼ 0.7, qa total thermal load absorbed by an engulfed object, kW m22
Dt ¼ 1 s, together with Tf ¼ 1560 K, 1f ¼ 0.7 and h ¼ 95 W qi total thermal load absorbed by an engulfed object at time ti ,
kW m22
m22 K21 from Table 2.
ql total thermal load to an engulfed object, kW m22
qra radiative load absorbed by an engulfed object, kW m22
qca convective load absorbed by an engulfed object, kW m22
qrl radiative load to an engulfed object, kW m22
CONCLUSIONS qcl convective load to an engulfed object, kW m22
Jet fires pose a serious hazard to oil and gas installations R Universal gas constant, J mol21 K21
r mass ratio of air to fuel for stoichiometric combustion
with the potential for escalation to a major incident. Hence it is RI reduction in incident radiation,%
important for operators to have a good understanding of fire s distance, m
hazards posed by their installation and to be able to quantify Ta ambient temperature, K
them as part of their Safety Case. Following the Piper Alpha Tf flame temperature, K
disaster, the Interim Guidance Notes provided guidance on Tg temperature of gas release, K
Ti surface temperature at time ti , K
how to assess fire hazards based on the information avail- Ts surface temperature, K
able at that time. However, many areas of uncertainty were ti time at step i of iterative calculation, s
identified where no data was available, especially in relation Uw water droplet velocity, m s21
to two-phase fires, confinement and deluge. V view factor of flame from position of receiver
V̇w water flowrate per nozzle, l min21
Since that time a considerable body of experimental Wf water volume fraction
research of large scale jet fires has been undertaken provid- x thickness of steel, m
ing an increased understanding of jet fire behaviour. Based Z compressibility factor

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B3): 207 –220
220 LOWESMITH et al.

Greek Symbols Gosse, A.J. and Pritchard, M.J., 1995, Large scale jet fire impaction
af flame absorptivity onto a flat surface, Proceedings of the IGRC, Cannes, France.
as surface absorptivity Gosse, A.J. and Hankinson, G., 2001, Use of water deluge to mini-
1f flame emissivity mise hazards of oil and gas fires offshore, Proceedings of the
1s surface emissivity IGRC, Amsterdam, The Netherlands.
g ratio of specific heats Hankinson, G., Lowesmith, B.J., Genillon, P. and Hamaide, G., 2000,
f equivalence ratio Experimental studies of releases of high pressure gas from punc-
r density of steel, kg m23 tures and rips in above-ground pipework, International Pipeline
s Stefan–Boltzmann constant, kW m22 K24 Conference, Calgary, Canada.
t atmospheric transmissivity Hankinson, G. and Lowesmith, B.J., 2004, Effectiveness of area and
tf flame transmissivity dedicated water deluge in protecting objects impacted by crude
oil/gas jet fires on offshore installations, J Loss Prevention in the
Process Industries, 17: 119–125.
Hankinson, G., Lowesmith, B.J., Evans, J.A. and Shirvill L.C., 2007,
REFERENCES Jet fires involving releases of crude oil, gas and water, IChemE
Process Safety and Environmental Protection, 85.
Advantica, 1997a, BFETS Phase 2: Horizontal jet fires of oil and gas, Lees, 2005, Lees’ Loss Prevention in the Process Industries, Vol. 1
unpublished Advantica report. (Elsevier Butterworth-Heinemann Ltd, London, UK).
Advantica, 1997b, Large scale natural gas jet fires impacting on flat Pritchard, M.J. and Cowley, L.T., 1991, Thermal impact on structures
surfaces, unpublished Advantica report. from large scale jet fires, Safety Developments in the Offshore
Advantica, 1997c, JIP: The effectiveness of water area deluge in miti- Industry, Inst of Mech Eng, Glasgow.
gating the effect of fires, unpublished Advantica report. Roberts, T.A., 2004, Effectiveness of an enhanced deluge system to
Advantica, 2000a, JIP: A programme of large scale experiments to protect LPG tanks and sensitivity to blocked nozzles and delayed
study the effectiveness of water deluge in mitigating potential off- deluge initiation, J Loss Prevention in the Process Industries, 17:
shore jet and pool fires, unpublished Advantica report. 151– 158.
Advantica, 2000b, JIP: Large scale experiments to study jet fires of Steel Construction Institute (SCI), 1992, Interim guidance notes for
oil/gas/water mixtures, unpublished Advantica report. the design and protection of topside structures against explosion
Bennett, J.F., Cowley, L.T., Davenport, J.N. and Rowson, J.J., 1991, and fire.
Large scale natural gas and LPG jet fires, final report to EC, Shell Steel Construction Institute (SCI), 1998, JIP on blast and fire engineer-
Report TNER.91.022. ing for topside structures Phase 2, final summary report, SCI 253.
Birch, A.D., Brown, D.R., Cook, D.K. and Hargrave, G.K., 1988, Sekulin, A.J. and Acton, M.R., 1995, Large scale experiments to
Flame stability in underexpanded natural gas jets, Combustion study horizontal jet fires of mixtures of natural gas and butane,
Science and Technology, 58: 267– 280. final report to the EC, Advantica report GRC R 0367.
Chamberlain, G.A., 1994, An experimental study of large scale com- Shell, 1992, Large scale butane jet fires data, unpublished Shell
partment fires, Trans I Chem E, Part B, 72: 211 –219. reports TNER.92.012 to TNER.92.053.
Chamberlain, G.A. and Brightwell, H.M., 1994, Large scale compart- Shirvill, L.C. and White, G.C., 1994, Effectiveness of deluge systems
ment fires, experimental data, HSE report OTO 94 011-024. in protecting plant and equipment impacted by high velocity natural
Cook, D.K., Fairweather, M., Hammonds, J. and Hughes, D.J., 1987, gas jet fires, International Symposium on Heat and Mass Transfer
Size and radiative characteristics of natural gas flares. Part I— in the Chemical Process Industry, Rome, Italy.
Field scale experiments, Chem Eng Res Des, 65: 310– 317. Shirvill, L.C., 2004, Efficacy of water spray protection against pro-
Cowley, L.T. and Johnson, A.D., 1992, Oil and Gas Fires: Character- pane and butane jet fires impinging on LPG storage tanks,
istics and Impact. JIP BFETS Phase 1, published by HSE as OTI J Loss Prevention in the Process Industries, 17: 111 –118.
92 596. UKOOA, 2006, Fire and explosion guidance. Part 2: Avoidance and
Cullen, W.D., 1990, The Public Enquiry into the Piper Alpha Disaster mitigation of fires, UK Offshore Operations Association, final draft
(Department of Energy, HMSO, London, UK). available from www.fireandblast.com\lgn-update.
Davenport, J.N., 1994a, Large scale natural gas/butane mixed fuel Wickens, M.J. and Lowesmith, B.J., 1993, Fire and explosion
jet fires, final report to EC, Shell report TNER.94.030. hazards—large scale experiments to assess and mitigate their
Davenport, N., 1994b, Large scale natural gas/kerosene mixed fuel effects. Presented at Inst Gas Engineers, Eastern Section.
jet fires, final report to the API, Shell report TNER.94.061.
Drysdale, 1985, An Introduction to Fire Dynamics (John Wiley &
Sons, Chichester, UK). The manuscript was received 29 June 2006 and accepted for
Energy Institute, 2003, Guidelines for the Design and Protection of publication after revision 6 October 2006. The paper was published
Pressure Systems to Withstand Severe Fires (Energy Institute, online ahead of print 20 March 2007.
London, UK).

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B3): 207– 220

You might also like