Anda di halaman 1dari 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 2

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

Abstract: High pressure jet res 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 incident 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 re hazards. However, many areas of uncertainty were identied where no data was available. Areas of particular concern identied in the Interim Guidance Notes were two-phase jet res, the effect of connement on jet res 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 reassesses jet re hazards in an offshore environment and provides updated guidance on the hazards they pose, including tabulated data and simple calculation techniques for predicting jet re hazards. Keywords: jet res; two-phase jet res; offshore re hazards; water deluge; thermal load.

BACKGROUND
The oil and gas industry produces, processes and conveys a wide range of hydrocarbon fuels from natural gas to heavy crude oil in large quantities and often at high pressures. An ignited accidental release of pressurized fuel can present a serious jet re hazard to industry personnel and to the installation itself potentially leading to escalation of the incident due to further inventory loss or structural collapse. For example, the riser failure which occurred during the Piper Alpha Incident in 1988 (Cullen, 1990), contributed to the collapse of the platform and a major loss of life. Offshore installations present particular problems in relation to the protection of personnel from hazardous events. Space restrictions lead to closely packed process pipework and hence the likelihood of escalation following a relatively small re is increased compared to an onshore facility. The close proximity of personnel and the difculties of escape are an important consideration. Hence it is important for operators to have a good understanding of the re hazards posed by their installation and to be able to demonstrate that the risk is managed effectively as part of their Safety Case.
207

Correspondence to: Dr B.J. Lowesmith, Department of Chemical Engineering, Loughborough University, Loughborough, Leicestershine, LE11 3TU, UK. E-mail: b.j.lowesmith@ lboro.ac.uk

DOI: 10.1205/psep06038 09575820/07/ $30.00 0.00 Process Safety and Environmental Protection Trans IChemE, Part B, May 2007 # 2007 Institution of Chemical Engineers

Prior to the Piper Alpha incident, there was little public domain information concerning the quantication of re hazards which could be accessed by operators and the incident led to the initiation of a major Joint Industry Project (JIP) coordinated by the Steel Construction Institute (SCI) entitled Blast and Fire Engineering Project for Topside Structures (BFETS) in 1990. Phase 1 reviewed the current understanding of re and explosion hazards offshore, identied areas requiring further research and provided Interim Guidance Notes (IGNs) for use by the offshore industry (Cowley and Johnson, 1992; SCI, 1992). In particular, the IGNs included guidance on the kinds of res likely to occur offshore and, where possible, guidance on their quantication, for example, in terms of re size and thermal loading. However, the guidance recognized many areas of uncertainty where no reliable data could be provided. This led directly to the initiation of Phase 2 of the project where major research projects were undertaken to obtain full scale data on the hazards presented by offshore res and explosions to address the gaps in understanding identied during Phase 1. As part of Phase 2, jet res involving oil and gas mixtures were undertaken by Advantica Limited
Vol 85 (B3) 207220

208

LOWESMITH et al.
water deluge systems on re behaviour and consequences are also summarized. Guidance is then presented for a simplied approach to quantifying jet re hazards (in terms of values for ame size, temperature, heat loads, and so on) for the range of re sizes likely to be considered in a quantied risk assessment (QRA). These typical values for jet re hazards are based on the experimental data reviewed here supplemented with predictions by validated models (developed by Shell and Advantica) and can be used to assess the potential hazard to personnel and the likely affect to re impacted obstacles. The approach taken in dening jet re hazards is similar to, but more detailed than, that presented within the Energy Institute guidance on severe res (Energy Institute, 2003). As the focus is offshore operations, the fuels for which guidance is required are principally pressurized gas, oil fractions and their mixtures, sometimes including water. The guidance does not extend to activities involving liqueed natural gas (LNG). Experimental studies generally used processed oil fractions such as diesel or kerosene and these are considered to be representative of the behaviour of light crude oils. Some experimental studies related to LPG (propane and butane) and, although not generally experienced offshore except as part of the composition, where appropriate the results of these studies have been included.

(formerly British Gas Research and Development) and conned condensate jet res were studied by SINTEF in Norway (SCI, 1998; Advantica, 1997a). Advantica were well qualied to undertake this work since from the early 1980s Advantica had been studying jet re hazards associated with natural gas, principally related to gas transmission pipelines. Recognizing the importance of studying these hazards at a representative scale, large and full scale experiments were undertaken at their Spadeadam Test Site in Cumbria to determine the re characteristics (Wickens and Lowesmith, 1993; Cook et al., 1987; Hankinson et al., 2000), although much remains unpublished. Other studies of jet res (conducted jointly with Shell Global Solutions) included other fuels such as propane and butane of particular relevance to onshore storage facilities (Bennett et al., 1991; Shell, 1992; Davenport, 1994a; Sekulin and Acton, 1995). After Phase 2, Advantica continued work on gas jet re hazards (Gosse and Pritchard, 1995; Advantica, 1997b) and Shell undertook large scale experiments studying gas/ kerosene jet res (Davenport, 1994b). Attention was also focused on re mitigation using active water deluge. Advantica initiated two JIPs to study the effectiveness of water deluge to mitigate jet and pool re hazards (Advantica, 1997c, 2000a), whilst Advantica and Shell jointly conducted a JIP to study jet res involving oil/gas/water mixtures (Advantica, 2000b). Only limited information from these JIPs has been published thus far (Gosse and Hankinson, 2001; Hankinson and Lowesmith, 2004; Hankinson et al., forthcoming). Experiments studying dedicated vessel deluge with jet res have also been conducted at Spadeadam and elsewhere (Shirvill and White, 1994; Shirvill, 2004; Roberts, 2004). The effect of connement (with and without deluge) on re behaviour was studied during the BFETS JIP experiments conducted at SINTEF and subsequently in further studies by Shell (Chamberlain, 1994; Chamberlain and Brightwell, 1994). It was always the intention of the BFETS JIP that the IGNs would be updated after Phase 2 and this was completed initially in relation to explosion hazards. After some delay, the updating of the IGNs for re hazards is now in progress with the support of the UK Offshore Operators Association (UKOOA, 2006).

NATURE AND CHARACTERISTICS OF JET FIRES


Jet res can be produced following the pressurized release of a variety of fuel types. The simplest being a pressurized gas giving rise to a gas jet re. A pressurized liquid/gas mixture (such as live crude or gas dissolved in a liquid) will give rise to a two-phase jet re. The gas stream atomizes the liquid into droplets which are then evaporated by radiation from the ame. However, a pressurized release of a liquid can also give rise to a jet re in which two-phase behaviour is observed if the liquid is able to vaporize quickly. This is most likely to occur when a liquid is released from containment at a temperature above its boiling point at ambient conditions whereupon ash evaporation occurs (e.g., propane, butane) and a ashing liquid jet re results. Non-volatile liquids (e.g., kerosene, diesel, or stabilized crude) are unlikely to be able to sustain a two-phase jet re, unless permanently piloted by an adjacent re; 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 re hazards, information on the nature of re hazards likely to be experienced offshore and guidance on their quantication is being reviewed taking into account the wealth of large scale experimental data that has been generated during and since Phase 2 of the BFETS JIP, as described above. This paper presents the results of this review in relation to jet re hazards which forms the basis of the guidance relating to jet res which is to be included within the updated industry guidance notes (UKOOA, 2006). Based on data from the large scale experiments outlined above including much data which has not been previously published, the nature and characteristics of hydrocarbon jet res, especially in relation to offshore oil and gas installations, are described. The effects of connement and active

Flame Stability
Whether or not a stable jet re will arise following the releases of a pressurized hydrocarbon will depend principally upon the nature of the fuel, the size of the hole from which the release occurs and the geometry of the surroundings. In the case of natural gas, it has been found that for free jets (not impacting) some combinations of hole size and pressure cannot produce stable ames (Birch et al., 1988). Figure 1 shows that for hole sizes under 30 mm diameter, there is a lower bound pressure which vertical high pressure releases must exceed to produce stable jet res. In practice this means that most small leaks will be inherently unstable and will not support a ame without some form of ame stabilization, such as the presence of another re 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 re. By contrast, the generally lower exit velocities from ashing liquid releases lead to ashing liquid jet res with shorter ame lift-offs and proportionately shorter and more buoyant ames overall. These lower velocities also result in res that are more wind affected whilst the higher hydrocarbon content of these fuels increases the ame luminosity. However, releases involving gas dissolved in, or mixed with, a liquid can result in a two-phase jet re which combines the worst aspects of both the gas jet re and the ashing liquid jet re, that is, high velocities and high ame luminosity. Apart from providing ame stabilization, impact onto an obstacle may modify the shape of a jet re. Objects which are smaller than the ame half-width at the point of impact are unlikely to modify the shape or length of the ame to any great extent. However, impact onto a large vessel may signicantly shorten the jet re, and impact onto a wall or roof could transform the jet into a radial wall jet where the location and direction of the re 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 res.

Combustion Characteristics and Smoke


provide a permanent pilot or stabilization as a result of impact onto an object such as pipework, vessels or the surrounding structure. Figure 1 also includes data from horizontal free jet res with and without general area deluge at two different owrates (Advantica, 2000a) from which it can be seen that the unstable region is larger for horizontal jets and that deluge further increases ame instability. However, in the highly congested environment offshore, impact within a short distance is very likely, and hence small leaks are most likely to stabilize on the nearest point of impact. In the case of high pressure releases of natural gas, the mixing and combustion is relatively efcient resulting in little soot (carbon) formation except for extremely large release rates. Hence little or no smoke is produced by natural gas jet res (typically ,0.01 g m23). CO concentrations in the region of 5 7% v/v have been measured within a jet re itself but this is expected to drop to less than 0.1% v/v by the end of the ame. Considerably more soot is produced in jet res involving higher hydrocarbons, although there is no available experimental data quantifying the difference. Limited measurements in the smoke one ame length downstream of the end of a live crude jet re determined a percentage obscuration of typically 10% over a 200 mm path length (Advantica, 2000a). This corresponds to a visibility distance of about 5 m (Drysdale, 1985).

Flame Shape and Appearance


A jet re is a turbulent diffusion ame produced by the combustion of a continuous release of fuel. Except in the case of extreme connement, which might give rise to extinguishment of gas jet res or liquid drop-out with two-phase jet res, the combustion rate will be directly related to the mass release rate of the fuel. In the offshore context, operating pressures are high so the ow of an accidental pressurized gas release into the atmosphere will be choked having a velocity on release equal to the local speed of sound in the uid (called sonic releases). A factor which is often overlooked is the noise produced by sonic gaseous releases. This is usually high pitched and so loud that it may prevent effective radio communication between personnel. As a result emergency actions could be hampered. For pressurized gaseous releases, following an expansion region downstream of the release point, the ame itself commences in a region of sub-sonic velocities as a blue relatively non-luminous ame. Further air entrainment and expansion of the jet then occurs producing the main body of the gas jet re as turbulent and yellow. In the absence of impact onto an object, these res are characteristically long and thin and highly directional. The high velocities within the released gas mean that they are relatively unaffected by

Jet Fire Length


Jet re size is primarily related to the mass release rate. For gaseous releases this, in turn, is related to the size of the leak (hole diameter) and the pressure (which may vary with time as a result of emergency blowdown). For nongaseous fuels, the liquid content results in relatively higher release rates for a given aperture and pressure compared to gaseous releases and, when the release is two-phase (such as may arise from a relatively long pipe connected to a storage vessel containing a liquid above its boiling point), estimating the release rate is non-trivial. This is because it is necessary to know what proportion of the uid is in each phase and this proportion may be changing along the length of the pipe due to heat transfer through the pipe wall and pressure changes. Figure 2 shows jet re lengths for a range of fuels plotted against the power of the ame in megawatts, Q ( mass release rate x net caloric value). The gure includes a correlation for jet re length, L, based on the majority of the natural gas data and all the higher hydrocarbon data, which is L 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 ame length.

For high pressure gaseous releases, the mass release rate through an orice, M, can be related to the area of the hole, Ad, and pressure, P, using v u   g1 ! u g1 gm 2 6t _ M Cd Ad P10 RTg Z g 1

(2)

For natural gas, this can be approximated for circular holes of diameter, d (mm), by _ M d 2 P 103 (3)

Substituting into equation (1), the jet re length for natural gas can be approximated by L 0:933d 0:746 P0:373 (4)

Flame Radiative Emissions


As noted above, the combustion process within a natural gas jet re is relatively efcient and produces little soot (carbon). Consequently, these ames are not as luminous as higher hydrocarbon ames. Radiation emissions from natural gas jet res arise mostly from water vapour and carbon dioxide, except for very large releases where soot production starts to enhance the process. The long thin shape may also result in a ame path which may not be optically thick. The net result is that the radiative heat transfer to the surroundings is lower than for comparable higher hydrocarbon jet res. This is reected in the fraction of heat

radiated, F, for such res as can be seen in Figure 3. (F is dened as (energy released as radiation from the ame surface/net energy of combustion)). As can be seen, F increases with carbon number, reecting the increased radiative emissions from soot within the higher hydrocarbon jet res. For gas liquid fuel mixtures, interpolation based on the percentage liquid within the mixture provides a reasonable estimate of F as can be seen in Figure 4 for butane/natural gas mixtures. A special case of interest at some installations is live crude that includes a signicant quantity of water. Experiments (Hankinson et al., 2007) have shown that mixtures with a water cut [dened as (mass of water/mass of fuel) 100%) of up to 125% remain ammable, although not necessarily capable of supporting a stable ame in the absence of some other supporting mechanism. The inclusion of water also slightly increases ame length and ame buoyancy, and signicantly reduces the amount of smoke produced. For water cuts less than 50% there is little impact on the fraction of heat radiated but for higher water cuts the fraction of heat radiated is reduced (Figure 5).

Heat Loads and Flame Temperatures


The thermal load to an engulfed object in a jet re will be a combination of radiative load and convective load from the hot combustion products passing over the surface. Clearly the total heat ux which is imparted to an engulfed object will vary over the surface of the object. In addition, the relative proportions of convective and radiative heat ux 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 be experienced close to the point of impact of a ame where the highest velocities occur, whereas the highest radiative heat load will be experienced where the more radiative part of the ame (usually towards the end of the ame) is viewed by the object. As the more radiative part of the ame is closer to the tail, this can result in the highest overall heat uxes being experienced on the rear surface of an engulfed object which may seem counter-intuitive. Figure 6 shows total heat uxes experienced by a horizontal

cylinder (pipe) impacted by a horizontal high pressure gas jet re (Pritchard and Cowley, 1991). The cylindrical surface is presented at by cutting along the rear. When the pipe was located at 21 m (towards the end of the ame) the maximum heat uxes were experienced at the point of impact on the front of the pipe [Figure 6(a)]. At 15 m [Figure 6(b)] the heat loads were relatively uniform around the pipe, but at 9 m [Figure 6(c)] the heat loads were greatest to the rear of the pipe due to radiation from the tail of the ame.

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 ux 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 transfer from higher hydrocarbon jet res is generally greater than that from natural gas ames and the generally lower velocities arising from ashing liquid releases (such as propane or butane) result in a lower convective ux to engulfed objects. For the rear surface of an engulfed object, Figure 7 shows that the fraction of the heat ux which is radiative increases from about 0.5 for natural gas to about 0.8 for fuels containing a large proportion of higher hydrocarbons (data from Bennett et al., 1991; Davenport, 1994a, b; Advantica, 1997a). In the case of a pressurized gasliquid mixture (such as live crude), the high gas velocities may still occur and result in a high convective contribution, whilst the higher hydrocarbon content maintains a high radiative contribution; making these type of jet res a worst case in terms of total heat ux to engulfed obstacles. Experimental work (Davenport, 1994a, b; Advantica, 1997a) suggests that the maximum combined uxes occur for gas liquid mixtures which are about 60 80% by mass of liquid (Figure 8). This

phenomenon can be explained by considering the maximum time averaged ame temperatures measured in over 180 large scale gas jet and ashing liquid jet res (Bennett et al., 1991; Sekulin and Acton, 1995; Advantica, 1997b) presented in Figure 9. As can be seen, the maximum ame temperatures are higher for the gas jet re (on average 12808C and up to 15008C), compared to a ashing liquid re which has a lower ame temperature (typically 10508C). In the case of two-phase jet res, the ame temperature is dominated by the gas content but the ame emissivity will also be enhanced by the higher hydrocarbon content leading to overall higher radiative uxes for such mixtures. Figure 8 includes data with mass owrate between 2.5 and 5 kg s21 and shows that for a ashing liquid fuel the maximum heat ux is generally around 200 kW m22 whereas for gas only (0% liquid), the maximum heat ux is typically 250 275 kW m22. In experiments involving natural gas at higher owrates (up to 10 kg s21) maximum heat uxes up

Figure 7. Radiative heat ux as a fraction of total heat ux to rear of engulfed object.

Figure 8. Maximum heat uxes to an engulfed object for gas liquid 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

Figure 9. Maximum ame temperatures.

to 320 kW m22 were measured. This was due to a higher radiative component arising from increased ame thickness and soot production within the ame. Broadly speaking, neglecting spatial variations, for a given location of an object within a ame (as a proportion of ame length), the convective component for a gas jet re is more or less constant with increasing size of release, whereas the radiative component increases with release rate as the ame becomes optically thick and more smoky. Hence the relative proportion of convective to radiative ux varies with re size. For fuels containing water, water cuts under 50% result in no signicant reduction in heat uxes to engulfed objects (,10%). However, over 50% the ames are signicantly less radiative, and the overall heat ux to an obstacle can be reduced by 40% or more (Advantica, 2000b; Hankinson et al., 2007) (Figure 10).

Water Deluge of Jet Fires


The most usual forms of active water deluge are water curtains (used to protect an escape corridor), general area deluge and dedicated vessel deluge. The success or otherwise of active water deluge in mitigating the effect of jet res depends upon the nature of the jet re (gas, liquid or two-phase) and objective to be achieved (reduction of incident thermal radiation or protection of engulfed objects). As noted above, the activation of general area deluge can adversely affect the stability of high pressure gas jet res,

Figure 10. Effect of water cut in the fuel on maximum heat uxes to an engulfed object.

particularly if the re 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 providing adequate ame stabilization. Although there is some evidence that the deluge increases combustion efciency resulting in lower CO and increased CO2 levels within the ame (Advantica, 2000a), in most other respects deluge has little effect on the size, shape and thermal characteristics of a high pressure gas jet re. 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 lm over the vessel in the presence of the high velocity jet, and so dry patches form where the temperature rise is undiminished by the action of deluge (Shirvill and White, 1994). Compared to the situation with a gas jet re, the use of dedicated vessel deluge to protect a vessel against a ashing liquid jet re (e.g., propane, butane) can be more effective. The water interacts with the ame to some extent; reducing the ame luminosity and the amount of smoke produced. Nevertheless, at typical application rates (10 15 l m22 min21) it cannot be relied upon to maintain a water lm over the vessel and hence to prevent vessel temperature rise in areas where dry patches form, although the rate of rise may be expected to reduce to 20 70% of the rate without deluge for a propane jet re (Shirvill, 2004). However, in tests with an increased water application of 40 l m22 min21, a 2 tonne LPG tank was effectively protected when subjected to a 2 kg s21 ashing propane jet re (Roberts, 2004). For two-phase jet res of live crude, dedicated deluge (at 10 l m22 min21) offered limited protection and no reduction in the rate of temperature rise in the area where the re impacted the obstacle (Advantica, 2000a; Hankinson and Lowesmith, 2004). This work also showed that using area deluge alone at the standard rate of 12 l m22 min21 is unlikely to modify the ame behaviour although there is some evidence that a higher deluge rate (24 l m22 min21) can result in water interaction with the ame, resulting in a shorter ame and some reduction in heat uxes to certain areas of an engulfed object, notably the front (where ame impact occurs) and top areas. Since dedicated vessel deluge is more effective at reducing the radiative heat uxes in the region to the rear of the vessel, the combination of area deluge (at the higher rate) and dedicated vessel deluge can be effective in reducing overall heat uxes to a vessel such that the temperature rise is halted or at least the rate of temperature 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 re impacting onto a vessel (Hankinson and Lowesmith, 2004) are shown in Figure 11. Measurements in the combustion products downstream of these live crude jet res determined an obscuration factor due to smoke of typically 10% over a 200 mm path length without deluge (visibility distance of about 5 m) and no signicant 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 visibility distance of between 1 m and 2 m. The major benet of area deluge with all kinds of jet res 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 re. (a) Top of vessel; (b) front; (c) rear; (d) bottom.

the surroundings, which protects adjacent plant and in particular, aids escape by personnel. Similarly water curtains can be used to protect escape corridors. Figure 12 presents large scale data (Advantica, 1997c; Advantica, 2000a) on the reduction of incident radiation due to water deluge by MV57 type nozzles spaced 2.84 m apart. For this arrangement, operating at 12 l m22 min21, incident radiation levels can be reduced by about 20% for a single row of nozzles, 3040% for two rows and 4060% for more than two rows (general area deluge).

Increased deluge rates can further reduce incident radiation levels: 60 70% at 18 l m22 min21; 80 90% at 24 l m22 min21 for general area deluge. Nozzles producing smaller droplet sizes can have an enhanced mitigation effect, but there is an increased risk that the droplets will be blown away by the wind (Advantica, 2000a). The mitigation of incident radiation is due to attenuation of radiation by the water droplets in the atmosphere, that is, effectively reducing the atmospheric transmissivity, t. The reduction in incident radiation, Is, with distance, s, due to this attenuation is characteristically expressed by an equation of the form Is 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 Vw litres per minute and producing droplets with a velocity Uw, the water volume fraction, Wf, is given by Wf _ N:V W 1 AW :UW 60 000 (6)

Figure 12. Reduction in incident radiation by water curtains of MV57 nozzles.

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 radiation as a function of average water pathlength for a range 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

Figure 13. Effectiveness of water deluge in reducing incident radiation.

droplets with a Sauter mean diameter of 300 micron; HV60 producing droplets in the range 670 750 micron depending on the operating pressure; MV57 producing droplets 640870 micron; and a prototype Large Droplet Nozzle (LDN) producing droplets about 1890 micron. The best-t equation of the appropriate form tted to the data is y 100 (1 2 e 21005x). Therefore, the reduction in incident radiation, RI (%), for a curtain of thickness Cw is RI 100 (1 e1005Wf Cw ) (7)

Confined Jet Fires


The behaviour of a jet re within a conned or partially conned area will depend upon the degree of connement and the direction of the jet relative to the ventilation opening. It should be noted that certain ventilation patterns could lead to ame instability and extinguishment. The amount of air required for complete stoichiometric combustion of methane is about 17.2 times the fuel owrate on a mass basis. For higher hydrocarbons this gure is about 15. The equivalence ratio, f, for a conned re in terms of the mass burning rate of the fuel and the mass owrate of air entrained (M and B) is dened as

fr

_ M _ B

(8)

where r is the mass ratio of air to fuel required for stoichiometric burning (that is r  15). If f , 1 then the re is well ventilated (also termed fuel controlled) and if f . 1 the re is under-ventilated (or ventilation controlled). If ventilation is plentiful (f , 1) or the jet is directed out of a conned region through a vent then there may be little difference in jet re characteristics compared to an unconned re. However, if the release rate is large relative to the size of the connement or the ventilation openings are small then the re may not be able to entrain enough air for complete combustion inside the compartment (f . 1). This is likely to result in increased levels of incomplete combustion products such as CO, increased levels of smoke (soot) and increased ame temperatures, particularly in regions close to the ceiling of a compartment where hot combustion products may be

trapped and recirculate. This leads to increased heat uxes to objects and surfaces compared to an unconned re (SCI, 1998; Chamberlain, 1994). The location where combustion occurs and the hottest parts of the ame may also shift due to the connement. In tests (Chamberlain, 1994) involving horizontal jet res 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 unconned res, the behaviour of under-ventilated conned res changes with time as the air initially available within the compartment is consumed, and this may lead to external aming after a period of time when the body of ame moves through the vent in order to nd the oxygen required for combustion. For gas jet res, 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 ame (Chamberlain, 1994). Soot production is related to the equivalence ratio f and hence the degree of ventilation and may range from about 0.1 g m23 at f 1.3 to up to 2.5 g m23 at f 2 (Chamberlain and Brightwell, 1994). The worst case condition in terms of re severity is likely to occur if the jet re is slightly under-ventilated as this leads to high heat release rates and enhanced soot production. Deluge of a conned gas jet re may lead to ame extinguishment and hence a serious explosion hazard from the continuing release. In the case of a two-phase jet re, extinguishment may result in a mist-air explosion hazard and/or the formation of a liquid pool. The likelihood of ame extinguishment is signicantly increased if the surroundings are already hot at the time the deluge is activated as the main mechanism which results in extinguishment of the jet re is inerting, that is evaporation of the water droplets leading to a mixture of gas/air/steam within the compartment which is outside the ammable limits. The water vapour may also contribute to ame instability by reducing the burning velocity. However, if the deluge is activated at an early stage, prior to the compartment walls becoming hot, then the re may not be extinguished and some benet in terms of reduced ame temperatures and wall temperatures may accrue (SCI, 1998; Chamberlain, 1994). In tests (Advantica, 2000a) involving only partial connement around the upper area of a module, during which dead spaces occurred close to the ceiling, area deluge was found to be benecial in reducing heat uxes to the ceiling surface, reducing the ame extent and the amount of smoke produced.

A SIMPLIFIED APPROACH TO ASSESSING JET FIRE HAZARDS


It is normal practice for a QRA of an oil and gas installation to consider a range of sizes of event. The critical parameter in determining re size is the mass release rate, which in turn is related to the leak size and pressure. A leak classication system which is often used, considers holes of 3, 10, 30 and 100 mm diameter which, for an operating pressure of about 100 bar, correspond to gaseous release rates of about 0.1, 1, 10 and 100 kg s21. Ideally, during a QRA, the re size and thermal loading from res should be assessed using mathematical models

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

216

LOWESMITH et al.
heat released by combustion as F EAf 1000 Q

which have been extensively validated against large scale data. A range of such models are available on a licence or consultancy basis. However, a simplied approach was proposed in the Interim Guidance Notes (SCI, 1992), whereby correlations for ame dimensions were suggested for jet and pool res together with guidance on typical heat loadings to engulfed objects. Similarly, guidance on typical heat loadings from res was provided in the Energy Institute document for severe jet and pool res (Energy Institute, 2003). In this paper, a similar approach is taken in order to provide tabulated guidance on assessing jet re hazards, based on recent knowledge and large scale experimental work. The guidance values (which are conservative) can then be used to undertake an initial simplied QRA. If necessary, problem areas can then be identied for more rigorous assessment. Guidance values for jet re hazards are provided for the following range of sizes of leak: . Smalltypically 0.03 0.3 kg s21, represented by 0.1 kg s21; . Mediumtypically 0.33 kg s21, represented by 1 kg s21; . Largetypically 330 kg s21, represented by 10 kg s21; . Major failureover 30 kg s21. For each size category, the information is provided to allow assessment of the re size, the incident radiation eld to persons and objects outside the re, the thermal loading to objects engulfed by the re, the smoke hazard presented by the re, the likely effects of active water deluge and the effects of connement. (It should be noted that for leak sizes of 10 kg s21 and above, the re is large compared with the average offshore module and it may be necessary to consider the re to be conned.)

The incident radiation received in the far eld at a distance, s (m), from the re, Is, is then expressed as Is _ 1000 tF MH 2 4 ps (10)

The atmospheric transmissivity will depend upon the prevailing atmospheric conditions (absolute humidity) and the path length, but might typically be 0.8 on a dry day for distances likely to considered. However, fog would signicantly reduce this value. The activation of water deluge on an offshore installation, producing water droplets in the air, effectively reduces the transmissivity by absorbing radiation. However, for the purposes of this paper, a revised effective fraction of heat radiated is dened as F0, accounting for the reduced transmissivity in the area being deluged, (but not including the transmissivity of the atmosphere between the re and the receiver which is outside the deluged area). The incident radiation at a distance, s (m), from the re can be estimated using Is _ 1000 tF 0 MH 4 ps2 (11)

Incident Radiation to the Surroundings


An obvious hazard presented by a re is the thermal radiation to the surroundings, in particular to personnel during escape and evacuation. The incident thermal radiation, I, to a person or object from a re can be described in terms of the ame emissive power, E, and the view factor, V, of the ame from the position of the receiver as I VE t (kW m2 ) (9)

This point source model can be useful for estimating the distance to a given radiation level. In the case of personnel, it is important to consider both the level of radiation and the duration of exposure. The thermal dose is often dened as the product of the incident radiation and the duration (kJ m22). However, the effect on people correlates better with a thermal load, J dened by (4=3) (12) J Is dt Threshold values of dosage associated with serious injury or death can be found in the literature (Lees, 2005; UKOOA, 2006). However, as a guide, a radiation level of about 5 kW m22 can be tolerated for about one minute and consequently represents a level from which it is reasonable to assume that escape, without signicant injury, would usually be possible for an average employee wearing typical protective clothing. However, the point source model is not suitable for estimating incident radiation to locations close to (certainly within one ame length) of the ame, where it may be signicantly in error. In the near eld, a mathematical model which denes a realistic ame shape (and ideally a variation of surface emissive power over the ame to allow for smoke obscuration) should be used. Typical values of F and F0 are provided in the tabulated guidance for the different sizes and types of jet re so that estimates of the far eld incident radiation hazard can be made using the point source model equations (10) and (11) given above.

The view factor is a function of the ame shape. Consequently, most integral or empirical mathematical models of res will assume some kind of simplied ame shape which is then used to calculate the view factor. The ame average surface emissive power is also a function of the ame shape. Therefore, average surface emissive powers used by the model will not necessarily be the same as those measured during an actual re. In the far eld, (typically more than two ame lengths away) the ame shape is not critical, so a simplied approach can be taken using the point source model, whereby the difculties of dening a ame shape and associated average surface emissive power can be avoided. In this approach, the fraction of the heat of combustion of the fuel radiated to the surroundings, F, is dened in terms of the ame area, Af, the ame surface emissive power, E, and the net rate of

Thermal Loading to Engulfed Objects


The thermal load per unit area, ql, to an object engulfed by re will be a combination of radiation from the ame (qrl) and 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


over the object surface. Hence the thermal load can be written as
4 ql qrl qcl 1f sTf4 tf sTa h(Tf Ts )

217

However, not all the thermal loading is necessarily absorbed by the surface, some may be reected back and the surface will also lose heat by radiation. Hence the total absorbed load per unit area (qa) is given by
4 4 qa 1f as sTf4 1s af sTs as tf sTa 4 1s tf sTs h(Tf Ts )

Assuming that both the ame and surface can be considered as diffuse grey bodies, then as 1s, af 1f and tf 1 2 1f. Furthermore, the term involving Ta is small and can be neglected, giving qa (kW m2 ) qra qca
4 1s s(1f Tf4 Ts ) h(Tf Ts )

(13)

When considering an actual object engulfed by a re, the ame temperature exposed to different parts of the object surface may vary; similarly the velocities in the ame (and hence convective heat transfer coefcient) may vary. Hence, to determine the total load absorbed by an object the above equation should be summed over the area of the object. Also, as the object engulfed in the ame heats up, the 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 calculation of the temperature rise of an object, the parameters Tf, 1f and h are required, together with the emissivity of the surface (which itself may change as the surface heats up). Where possible, typical values of these parameters are

provided in the tabulated guidance for the jet re types and sizes. However, researchers often quote heat uxes measured during experiments using calorimeters (for total heat ux) and radiometers (for radiative ux). These instruments are designed to have a surface emissivity close to 1 and are maintained at a low temperature throughout the experiments. Hence the uxes measured and reported for calorimeters are given by equation (13) with 1s 1 and Ts % 333 K and can be regarded as a conservative estimate of the total heat ux initially absorbed by an engulfed object. Radiometers are designed and calibrated to measure 1f sTf4 . At an early stage, whilst Ts is low, the ux measured by a radiometer is approximately (qra =1s ) and so, if 1s is taken as 1, it provides a conservative estimate of the radiative ux absorbed by an object. Subtracting the measured radiative ux from the measured total heat ux enables the initial convective ux absorbed by the object to be determined. Using an estimated or measured value for Tf enables h to be determined at the measurement location. In this way, experimentally measured ux levels can be used to derive input data for transient heatup calculations. Consequently, typical values of the total, radiative and convective uxes (as dened in the above with 1s 1 and Ts % 333 K) are also provided in the tabulated guidance for the range of re types and sizes. These represent initial heat uxes to the re engulfed object (that is, when the object surface is still cold and the temperature difference is greatest). A simplied method for heat-up calculations can then be used (such as that given in UKOOA, 2006; Energy Institute, 2003) to determine the temperature of the engulfed object with time.

Tabulated Guidance
Table 1 provides guidance values for gas jet res and Table 2 for two-phase jet res. As noted above, the maximum heat uxes to engulfed objects have been found to occur

Table 1. Guidance values for gas jet res. Size (kg s


21

0.1 5 0.05 CO , 0.1 Soot 0.01

1.0 15 0.08 CO , 0.1 Soot 0.01

10 40 0.13 CO , 0.1 Soot 0.01

.30 65 0.13 CO , 0.1 Soot 0.01

Effect of connement Affected by enclosure shape and openings Increased CO up to about 5% v/v at a vent prior to external aming, but after external aming ,0.5% v/v exiting the ame. Soot levels depend on equivalence ratio from about 0.1 g m23 at f 1.3 to 2.5 g m3 at f 2 Increased heat loadings up to 400 kW m22 (280 kW m22 radiative, 120 kW m22 convective, Tf 1600 K, 1f 0.75, h 0.09 kW m22 K21) Risk of extinguishment and explosion hazard if deluge activated when enclosure is already hot and re is well established

Flame length (m) Fraction of heat radiated, F CO level (% v/v) and smoke concentration (g m23)

Initial total heat ux (kW m22) Initial radiative ux (kW m22) Initial convective ux (kW m22) Flame temperature, Tf (K) Flame emissivity, 1f Convective heat transfer coefcient, h (kW m22 K21) Effect of deluge

180 80 100 1560 0.25 0.08

250 130 120 1560 0.4 0.095

300 180 120 1560 0.55 0.095

350 230 120 1560 0.7 0.095

No effect on heat loadings to engulfed objects. In far eld, take F0 0.8 F for one row of water sprays, F0 0.7 F for two rows and F0 0.5 F for .2 rows at 12 l m22 min21. May improve combustion efciency and reduce CO levels within ame

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 res.

Size (kg s21) Flame length (m) Fraction of heat radiated, F CO level (% v/v) and smoke concentration (g m23)

0.1 5 CO , 0.1 Soot 0.01

1.0

10

.30 60 CO , 0.1 Soot 0.01

Effect of connement Affected by enclosure shape and openings Increased CO up to about 5% v/v at a vent prior to external aming, but after external aming ,0.5% v/v exiting the ame. Soot levels depend on equivalence ratio from about 0.1 g m23 at f 1.3 to 2.5 g m23 at f2 Increased heat uxes, take values as per 30 kg s21 two-phase jet re

13 35 See note below table CO , 0.1 Soot 0.01 CO , 0.1 Soot 0.01

Initial total heat ux (kW m22) Initial radiative ux (kW m22) Initial convective ux (kW m22) Flame temperature (K) Flame emissivity, 1f Convective heat transfer coefcient, h (kW m22 K21) Effect of deluge

200 100 100 1560 0.3 0.08

300 180 120 1560 0.55 0.095

350 230 120 1560 0.7 0.095

400 280 120 1560 0.85 0.095

Some benet to engulfed objects but temperature may still rise although at a slower rate. Combined area and dedicated deluge may prevent temperature rise if effectively applied. Take F 0 as per Table 1.

Risk of extinguishment and potential formation of pool

Note: For fraction of heat radiated, Fm, of mixture involving x% liquid by mass: use Fm (x=100):(FL FG ) 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 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 mass. Consequently, the values shown in Table 2 correspond to this worst case condition. For ashing liquid res (such as propane or butane) a lower ame temperature of about 1300 K is likely with an emissivity of 1, giving a radiative ux of about 160 kW m22. A convective heat transfer coefcient of about 0.08 kW m22 K21 is suggested giving a convective ux of about 80 kW m22 and a total ux of about 240 kW m22.

Validation Exercise
Consider a live crude jet re at 5 kg s21, which is 20% gas, 80% oil, released at 20 bar. Three experiments with 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 ame (Test 3). Using Table 2, this would fall into the large release category of 10 kg s21 giving a ame length of 35 m which gives a conservative assessment compared to the measured length of 29 m for the free ame and 21 m and 27 m for the impacting ames. Using the formula for fraction of heat radiated below Table 2, with FG 0.13 and FL 0.5, Fm (80=100):(0:5 0:13) 0:13 0:426, Figure 14 presents incident radiation measured during the jet res with calculations made using _ the point source model Is (1000tF MHm )=(4ps2 ), taking t 0.8 and Hm 42 MJ kg21. Table 2 suggests a maximum total heat ux of 350 kW m22 whereas during the two experiments where the ame impacted a steel pipe, the maximum heat uxes measured were 348 and 370 kW m22 . Table 2 suggests that the radiative ux accounted for 65% of the total ux (230 kW m22) with 34% being convective (120 kW m22). Table 3 compares

these guidance values with the data measured during the tests for maximum total heat ux and the proportion of radiative and convective uxes measured at locations on the front, top and rear surface of the re engulfed pipe. Clearly, the simplied guidance values take no account of the spatial variation of heat loading over the surface of a re engulfed object. For a conservative assessment of the response of the pipe to the heat loading, the temperature of the engulfed pipe with time can be calculated for a series of time steps (Dt) as follows: At time ti ti1 Dt(i ! 1) the thermal load absorbed by the object, qi , is given by qi 1s s(1f Tf4 Ti4 ) h(Tf Ti ) (i ! 0) (14)

Figure 14. Incident radiation from live crude oil jet re.

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

HYDROCARBON JET FIRE HAZARDS IN THE OIL AND GAS INDUSTRY


Table 3. Comparison of guidance values for heat ux with large scale data. Percentage radiative Maximum heat ux Guidance values Data Test 9a Data Test 11a
a

219

Percentage convective

Front

Top 66%

Rear

Avg

Front

Top 34%

Rear

Avg

350 370 348 41% 43% 61% 86%

68% 100%

57% 76%

59% 57%

39% 14%

32% 0%

43% 24%

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 re 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 assessment of jet re hazards can be readily undertaken.

NOMENCLATURE
Figure 15. Heat up of engulfed object in live crude jet re. a Ad Af Aw B C Cd Cw d E F H h I J Is L M m N P Q qa qi ql qra qca qrl qcl R r RI s Ta Tf Tg Ti Ts ti Uw V Vw Wf x Z constant area of release hole, m2 area of ame, m2 deluged area, m2 owrate of air entrained for combustion, kg s21 specic heat capacity of steel, kJ kg21 K21 coefcient of discharge of orice water curtain thickness, m diameter of release hole, mm ame emissive power, kW m22 fraction of heat radiated by a ame net caloric value, MJ kg21 convective heat transfer coefcient, kW m22 K21 incident radiation, kW m22 thermal load to personnel, (kW m22)4/3 s incident radiation at a distance s, kW m22 jet re length, m mass ow rate, kg s21 molecular weight, kg number of nozzles over a deluged area absolute pressure, MPa _ net rate of heat released by combustion (Q MH), MW total thermal load absorbed by an engulfed object, kW m22 total thermal load absorbed by an engulfed object at time ti , kW m22 total thermal load to an engulfed object, kW m22 radiative load absorbed by an engulfed object, kW m22 convective load absorbed by an engulfed object, kW m22 radiative load to an engulfed object, kW m22 convective load to an engulfed object, kW m22 Universal gas constant, J mol21 K21 mass ratio of air to fuel for stoichiometric combustion reduction in incident radiation,% distance, m ambient temperature, K ame temperature, K temperature of gas release, K surface temperature at time ti , K surface temperature, K time at step i of iterative calculation, s water droplet velocity, m s21 view factor of ame from position of receiver water owrate per nozzle, l min21 water volume fraction thickness of steel, m compressibility factor

and the temperature of the object increases to Ti1 Ti DTi (i ! 0) where DTi Dtqi Crx (15)

Figure 15 shows a comparison of the calculated temperature rise of the pipe with data measured by thermocouples embedded within the pipe wall at three locations during the two impacting re experiments. The three locations were: close to the point of impact of the re (front of pipe); 908 from this location on top of the pipe; and 1808 from the point of impact on the rear of the pipe. The data used for the calculations was: C 520 J kg21 K21, r 7850 kg m23, s 5.6697 W m22 K24, x 0.0125 m, 1s 0.7, Dt 1 s, together with Tf 1560 K, 1f 0.7 and h 95 W m22 K21 from Table 2.

CONCLUSIONS
Jet res pose a serious hazard to oil and gas installations with the potential for escalation to a major incident. Hence it is important for operators to have a good understanding of re hazards posed by their installation and to be able to quantify them as part of their Safety Case. Following the Piper Alpha disaster, the Interim Guidance Notes provided guidance on how to assess re hazards based on the information available at that time. However, many areas of uncertainty were identied where no data was available, especially in relation to two-phase res, connement and deluge. Since that time a considerable body of experimental research of large scale jet res has been undertaken providing an increased understanding of jet re behaviour. Based

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

220
Greek Symbols af ame absorptivity as surface absorptivity ame emissivity 1f surface emissivity 1s g ratio of specic heats f equivalence ratio r density of steel, kg m23 s StefanBoltzmann constant, kW m22 K24 t atmospheric transmissivity tf ame transmissivity

LOWESMITH et al.
Gosse, A.J. and Pritchard, M.J., 1995, Large scale jet re impaction onto a at surface, Proceedings of the IGRC, Cannes, France. Gosse, A.J. and Hankinson, G., 2001, Use of water deluge to minimise hazards of oil and gas res offshore, Proceedings of the IGRC, Amsterdam, The Netherlands. Hankinson, G., Lowesmith, B.J., Genillon, P. and Hamaide, G., 2000, Experimental studies of releases of high pressure gas from punctures and rips in above-ground pipework, International Pipeline Conference, Calgary, Canada. Hankinson, G. and Lowesmith, B.J., 2004, Effectiveness of area and dedicated water deluge in protecting objects impacted by crude oil/gas jet res on offshore installations, J Loss Prevention in the Process Industries, 17: 119125. Hankinson, G., Lowesmith, B.J., Evans, J.A. and Shirvill L.C., 2007, Jet res involving releases of crude oil, gas and water, IChemE Process Safety and Environmental Protection, 85. Lees, 2005, Lees Loss Prevention in the Process Industries, Vol. 1 (Elsevier Butterworth-Heinemann Ltd, London, UK). Pritchard, M.J. and Cowley, L.T., 1991, Thermal impact on structures from large scale jet res, Safety Developments in the Offshore Industry, Inst of Mech Eng, Glasgow. Roberts, T.A., 2004, Effectiveness of an enhanced deluge system to protect LPG tanks and sensitivity to blocked nozzles and delayed deluge initiation, J Loss Prevention in the Process Industries, 17: 151 158. Steel Construction Institute (SCI), 1992, Interim guidance notes for the design and protection of topside structures against explosion and re. Steel Construction Institute (SCI), 1998, JIP on blast and re engineering for topside structures Phase 2, nal summary report, SCI 253. Sekulin, A.J. and Acton, M.R., 1995, Large scale experiments to study horizontal jet res of mixtures of natural gas and butane, nal report to the EC, Advantica report GRC R 0367. Shell, 1992, Large scale butane jet res data, unpublished Shell reports TNER.92.012 to TNER.92.053. Shirvill, L.C. and White, G.C., 1994, Effectiveness of deluge systems in protecting plant and equipment impacted by high velocity natural gas jet res, International Symposium on Heat and Mass Transfer in the Chemical Process Industry, Rome, Italy. Shirvill, L.C., 2004, Efcacy of water spray protection against propane and butane jet res impinging on LPG storage tanks, J Loss Prevention in the Process Industries, 17: 111 118. UKOOA, 2006, Fire and explosion guidance. Part 2: Avoidance and mitigation of res, UK Offshore Operations Association, nal draft available from www.reandblast.com\lgn-update. Wickens, M.J. and Lowesmith, B.J., 1993, Fire and explosion hazardslarge scale experiments to assess and mitigate their effects. Presented at Inst Gas Engineers, Eastern Section. The manuscript was received 29 June 2006 and accepted for publication after revision 6 October 2006. The paper was published online ahead of print 20 March 2007.

REFERENCES
Advantica, 1997a, BFETS Phase 2: Horizontal jet res of oil and gas, unpublished Advantica report. Advantica, 1997b, Large scale natural gas jet res impacting on at surfaces, unpublished Advantica report. Advantica, 1997c, JIP: The effectiveness of water area deluge in mitigating the effect of res, unpublished Advantica report. Advantica, 2000a, JIP: A programme of large scale experiments to study the effectiveness of water deluge in mitigating potential offshore jet and pool res, unpublished Advantica report. Advantica, 2000b, JIP: Large scale experiments to study jet res of oil/gas/water mixtures, unpublished Advantica report. Bennett, J.F., Cowley, L.T., Davenport, J.N. and Rowson, J.J., 1991, Large scale natural gas and LPG jet res, nal report to EC, Shell Report TNER.91.022. Birch, A.D., Brown, D.R., Cook, D.K. and Hargrave, G.K., 1988, Flame stability in underexpanded natural gas jets, Combustion Science and Technology, 58: 267 280. Chamberlain, G.A., 1994, An experimental study of large scale compartment res, Trans I Chem E, Part B, 72: 211 219. Chamberlain, G.A. and Brightwell, H.M., 1994, Large scale compartment res, experimental data, HSE report OTO 94 011-024. Cook, D.K., Fairweather, M., Hammonds, J. and Hughes, D.J., 1987, Size and radiative characteristics of natural gas ares. Part I Field scale experiments, Chem Eng Res Des, 65: 310 317. Cowley, L.T. and Johnson, A.D., 1992, Oil and Gas Fires: Characteristics and Impact. JIP BFETS Phase 1, published by HSE as OTI 92 596. Cullen, W.D., 1990, The Public Enquiry into the Piper Alpha Disaster (Department of Energy, HMSO, London, UK). Davenport, J.N., 1994a, Large scale natural gas/butane mixed fuel jet res, nal report to EC, Shell report TNER.94.030. Davenport, N., 1994b, Large scale natural gas/kerosene mixed fuel jet res, nal report to the API, Shell report TNER.94.061. Drysdale, 1985, An Introduction to Fire Dynamics (John Wiley & Sons, Chichester, UK). Energy Institute, 2003, Guidelines for the Design and Protection of Pressure Systems to Withstand Severe Fires (Energy Institute, London, UK).

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

Anda mungkin juga menyukai