You are on page 1of 8

OTC 14132

Controlling Hydrocarbon Fires in Offshore Structures


G. A. Chamberlain, Shell Global Solutions
Copyright 2002, Offshore Technology Conference
This paper was prepared for presentation at the 2002 Offshore Technology Conference held in
Houston, Texas U.S.A., 6 9 May 2002.
This paper was selected for presentation by the OTC Program Committee following review of
information contained in an abstract submitted by the author(s). Contents of the paper, as
presented, have not been reviewed by the Offsh ore Technology Conference and are subject to
correction by the author(s). The material, as presented, does not necessarily reflect any
position of the Offshore Technology Conference or its officers. Electronic reproduction,
distribution, or storage of any part of this paper for commercial purposes without the written
consent of the Offshore Technology Conference is prohibited. Permission to reproduce in print
is restricted to an abstract of not more than 300 words; illustrations may not be copied. The
abstract must contain conspicuous acknowledgment of where and by whom the paper was
presented.

Abstract
This paper reviews improvements in the understanding of fire
hazards and protection in offshore structures since the Piper
Alpha disaster in 1988. Important new knowledge on fire
development, spread and escalation have come about through
a number of industry sponsored initiatives, which have also
demonstrated how to control and mitigate fire hazards by good
design and effective protection. The most notable gains in
knowledge have been in the areas of unconfined two-phase
hydrocarbon jet fires, confined jet and pool fires (compartment
fires), and understanding the value of area deluge for fire
mitigation.
Background
The Piper Alpha disaster was a tragic milestone in influencing
the Safety Case regime [1]. Since then there have been a
number of Industry sponsored initiatives, which have
contributed to significant advances in the science and
understanding of the consequences of accidental hydrocarbon
fires in offshore structures and the engineering of safe design.
While some uncertainties remain, the application of this
knowledge, via a Hazards and Effects Management Process or
appropriate Standard (e.g. ISO 13702:1999), is normally
sufficient to improve design and operation of plant in a cost
efficient way.
A short review of the fire hazard aspects of the Piper
Alpha disaster helps to focus on credible scenarios. Next to
each event are some pertinent questions asked of the main fire
hazards and the effectiveness of possible protective barriers
that could be brought to bear in safety studies and design:
The initial blast ruptured oil lines in the adjacent
module. How severe was the resulting fire? To what
extent does the surrounding congestion and
confinement affect the severity? How toxic is the
smoke?

The Tartan riser failed on the outboard side of the


emergency shut-off valve resulting is a full-bore jet
fire. What protection should be applied to risers and
valves? Is passive fire protection (PFP) adequate?
Will PFP age with time and then be less effective
against fires? Is a sub-sea isolation valve effective
against rise fires? How close to the platform should
it be located?
The water deluge system did not operate. How
effective are water-based systems in controlling
different types of fire?
How effective are blowdown systems against
different types of fire? What strategies should be
adopted?
The Blast and Fire Engineering for Topside Structures
(BFETS) Project [2,3] started in 1990 was the first major
initiative to provide interim guidance on how to reduce the
potential for such disasters. This project also identified gaps
in knowledge and recognized the need for full-scale
experiments and further guidance, which have been addressed
in numerous follow-up studies and other Joint Industry
Projects. In this paper I review the considerable progress
made to address the questions raised above.
Fire Hazards
An accidental release of flammable hydrocarbon results in a
variety of fire consequences, dependent on the type and initial
state of the hydrocarbon. There are essentially three fluid pre release conditions determining the "source term" for the fire
events:
1. Liquid at ambient pressure, e.g. diesel, gasoline or oil tank,
2. Liquid at pressure above ambient, e.g. separator contents,
pumped crude oil, oil-well blow-out,
3. Gas or vapour above ambient pressure, e.g. gas pipeline,
vapour space in separator, gas blowout.
Complete definition of the source term then requires
calculation of the outflow rate, and the physical form and
dimensions of the release.
For example, a flammable liquid release from an
atmospheric storage tank will always result firstly in the
formation of a liquid pool, immediate ignition of which results
in a transient fireball which burns back to a spray jet fire and
burning pool.

G. A. CHAMBERLAIN

A liquid release from pressurised storage initially leads to


a spray jet. The spray would consist of both vapour and liquid
phases, in the form of droplets, if the liquid is volatile e.g.
condensate, or simply droplets for non-volatile releases e.g.
stabilised crude oil. Mechanical and thermodynamic forces
both play a role in causing the jet to break up in to droplets.
Just a few bar is required for atomisation of the jet [4]
independent of any flashing processes that may be occurring.
The effect this has on subsequent dispersion, degree of liquid
rainout and fire behaviour is critical to an assessment of the
hazardous consequences, and hence the extent and type of
protection system required. Jet and pool fires can occur on
immediate ignition after an initial transient fireball, but liquid
rainout is unlikely if a released liquid, such as LPG, readily
flashes to vapour in the atmosphere.
A pressurised gas release normally results in a
dispersing gas jet, which may contain aerosols if the gas cools
sufficiently during release to allow condensation of some or
all of the components. Entrainment of air normally is
sufficient to re-evaporate the aerosol droplets. Ignition creates
the classic jet fire, typified by refinery flares. In extraordinary
cases the release of hot gas containing heavy components
could lead to condensation and agglomeration of droplets and
a rainout pool.
Jet Fires. A jet fire is a turbulent diffusion flame resulting
from the combustion of a fuel continuously released with
some significant momentum in a particular direction. In
principle any pressurised fuel has the potential to generate a
jet fire.
Jet flame stability. Not all releases give rise to stable flames.
A correlation [5,6] derived from extensive experimental data
[7,8,9], shows that for unobstructed releases of natural gas
from orifices less than about 30mm diameter, certain drive
pressures will not sustain a stable flame. Thus accidental
damage to small-bore high pressure fittings might reasonably
be expected not to result in a stable flame in open
environments. In highly congested industrial plants however,
the likelihood of flame stabilisation by impact on adjacent
surfaces is high.
The increased burning velocity associated with higher
hydrocarbon gases results in greater stability and smaller
critical diameters. On the other hand, jet releases of liquid
kerosene up to 2.5 kg/s are unable to sustain stable flames
when released as a horizontal jet because of insufficient
droplet evaporation [10]. Similar self-extinction was evident in
the crude oil jet releases in Phase 2 of the BFETS. [11,12].
Jet flame size and external heat flux. In typical plant
environments free jet fires could be considered unlikely but, to
date, most predictive models have been developed and
validated for such fires, and it is worthwhile reviewing the
progress made. Impacting and confined fire behaviour is the

OTC 14132

subject of ongoing research, discussed later. The main hazard


from a free jet fire is thermal radiation to the surroundings. A
simple expression for the radiation flux received by an object
q is q=tVS, where t is the atmospheric transmissivity, V is the
view factor of the flame surface and S is the flame surface
emissive power. For most calculations the Shell jet flame
model [13,14] which represents the flame as a solid body - the
frustum of a cone radiating uniformly over its surface adequately predicts the incident flux for safety engineering
purposes, and is considerably more accurate than single point
source representations (e.g. in API RP521) for objects close to
the flame.
The applicability of empirical models is however limited
by the range of experimental data on which they are based.
Computational fluid dynamics (CFD) models have been
improved in recent years but require considerable computer
power, computer time and expertise. Some results however are
remarkably good [15]. Additional output from the model
includes such important information as convective and
radiative components of heat flux within the flame and
external radiation field, which have been validated by
experiment for natural gas and propane gas flames.
Jet fire internal heat flux. An accidental ignited release is
likely to impact on adjacent objects. Thus the thermal loading
is required to assess the risk of escalation by failure of
pipework, or vessels, or structure. The object surface is
subjected to convection from the high velocity hot combustion
products and to radiation emitted by the flame. These
components of total heat flux not only vary along and across
the flame but also vary in time due to the turbulent nature of
the jet and fluctuations in wind. Measurements [16] in sonic
natural gas jet flames up to 10 kg/s indicate maximum time averaged total heat fluxes to a cold surface of about 250
kW/m2 split roughly equally between convection and
radiation. Hazard analyses often make use of these data
extremes to predict the temperature rise of engulfed plant.
Good agreement with experiment is found if a gas velocity of
40 m/s and a flame kinetic temperature of 1200 o C are used in
the heat transfer calculation. In some treatments [17] the
convective heat transfer coefficient and flame radiative
temperature can be derived experimentally and combined in
physical models to predict the thermal load to fire engulfed
objects. An alternative approach is to use the direct
measurements obtained from large scale experiments taking
care
not
to
extrapolate
too
far
from
the
experimental conditions.
Shell [18] and British Gas [19] carried out many tests
with propane, butane and mixtures of natural gas and butane at
flow rates up to 28 kg/s for the single component fuels and 2.5
kg/s for the mixtures. In general, the flames were more
radiative than natural gas and produced more smoke. They
were also more buoyant and wind affected. For butane
concentrations up to 40% the flame properties were similar to

OTC 14132

CONTROLLING HYDROCARBON FIRES IN OFFSHORE STRUCTURES

those of natural gas. Davenport [10] carried out similar tests


with kerosene/natural gas jet fires. The fires were even more
radiative than with butane mixtures, but pure kerosene releases
could not sustain flames under the test conditions. Measured
time-averaged heat fluxes to a cold surface reached about 300
kW/m2 at high levels (60-80% by mass) of either butane
or kerosene.
In Phase 2 of the BFETS [3] the jet fire
characteristics of unconfined releases of crude oil were
measured. Releases of crude oil and natural gas consisted of
separate fluid streams, which combined several diameters
downstream in the jet fire. The two nozzle arrangement was
chosen to provide a convenient way of varying the gas-oil
ratio (GOR) and is close to but not strictly representative of a
release of live crude, such as could occur in a well blowout.
Some general conclusions can be drawn from these
tests. The size of crude oil jet fires was similar to other
hydrocarbon liquid and two phase jet fires. All the flames
were highly radiative, with maximum time-averaged surface
emissive powers from 203 to 409 kW/m2 . Increasing the gas
content in crude oil increased the emitted radiation. The
fraction of heat radiated in the mixed releases reached about
0.3 consistent with other 2-phase jet flames. Increasing the gas
content in crude oil increased the incident heat flux on an
impinged pipe object by increasing both the radiative and
convective components. The radiative component was,
however, always higher at the rear of the object. Spot
maximum heat fluxes were around 350 kW/m2 , similar to
other mixed releases.
Further experimental studies were carried out as a
separate Joint Industry Project [20] to quantify the hazards
posed by realistic releases of simulated live crude oil
containing gas and water. Live crude was prepared from
stabilized crude oil commercial propane and natural gas.
GORs of 500 to 1500 scf/bbl at 30 to 100 barg were studied.
Releases of about 5 kg/s were maintained by over-pressuring
the storage vessel with natural gas. The resultant horizontal jet
flames were about 20m long quite buoyant and very smoky. In
a second phase of the project water was injected into the
release stream to examine the effect on flame radiation and
stability. Small amount of water had little effect on the jet fire,
larger amounts dramatically reduced smoke production and
increasing the water content still further took the flame to the
point of extinction. This project confirmed that jet fires from
live crude releases are not more severe than other mixed
hydrocarbon releases. The flames were shorter and more
buoyant than those investigated in the BFETS Phase 2 study,
but the heat fluxes were similar.
Pool Fires. A pool fire is a turbulent diffusion fire burning
above a pool of vaporising liquid fuel where the fuel vapour
has very low initial momentum. The vaporisation of high
boiling point fuels is controlled mainly by radiative feedback
to the pool surface. The hazard effects are controlled largely
by the size of the pool and the fuel type. These determine the
mass burning rate and the size of the visible flame above the

pool. Wind tilt drags the flame beyond a confining wall


boundary and can bring the fire closer to downwind objects.
Two distinct regions are created in the flame zone - a region of
continuous burning and a smokier region where flame appears
intermittently. Empirical models [21,22,23] usually represent
the flame as a tilted cylinder which take into account these two
regions by assigning different factors for smoke obscuration.
The external radiation field is calculated using a view factor
and flame surface emissive power as before.
Relatively little information is available on the
thermal loading from pool fires to engulfed objects. However,
it is known that radiative heat transfer dominates (about 80%)
and that flame temperatures reach about 1000o C. The internal
heat flux of kerosene or JP-4 pool fires is around 150 kW/m2
but increases to about 250 kW/m2 for large LNG or LPG
pool fires.
Confined fires. The behaviour of jet and pool fires can be
significantly modified when walls and ceilings, such as the
boundaries of offshore compartments, confine the fire. In
addition, hot combustion products rapidly fill ceiling spaces
and impose fire loadings on any objects located there. If the
compartment openings are small or the release rate of fuel is
high, the fire may not be able to entrain enough air for
complete combustion inside the compartment. The fire is then
said to be ventilation controlled. As well as the hazards
already discussed, additional ones exist [24]. These include
external flaming from compartment openings, impaired
visibility along escape routes, increased CO hazard, explosion
hazard from unburnt fuel if the fire terminates due to lack of
oxygen, and possible increased thermal loading to objects due
to greater amounts of hot soot. Recent large-scale studies
[25,26] have confirmed that fire global stoichiometry si a
useful correlating parameter. Gaseous propane jet fires
burning in slightly fuel rich conditions represent a worst case
because they combine high heat release with enhanced soot
production. The heat flux was about 50% radiative in the
impingement zone of the propane jet totalling about 350
kW/m2 in near stoichiometric propane jets. The heat flux well
away from the impingement zone was nearly 100% radiative.
Pool fires, on the other hand, are more severe in overventilated conditions. When ventilation controlled the pool fire
burning rate is governed by the size of the ventilation opening
and the evaporation rate tends to drop to match the lower
burning rate. Thus the pool fire limits itself to the slightly
under-ventilated regime. Persaud et al [27] show that the effect
of insulation on the compartment walls is to raise the average
smoke temperature by 200-400o C.
The results of Phase II of the BFETS project, carried
out at SINTEF NBL, confirmed and extended these findings to
condensate jet and pool fires and larger scales [28,29]. For jet
fires it was found that there was no significant dependency of
scale. Gas temperatures were between 1100 and 1300o C.
Maximum heat fluxes were 170 kW/m2 for fuel rich jet flames

G. A. CHAMBERLAIN

and 250 kW/m2 for fuel lean flames. There was no significant
difference between propane and condensate jet fires except
within the impingement zone. The hot spot on the ceiling in
the propane gas jet fires was replaced by a cold spot for the
condensate jet fires.
A weak overall dependency on fuel air ratio was
observed for jet fires, but localised effects were significant. In
very confined regions, hot spots can occur whereby radiation
is trapped. Temperature rises can be rapid and, in these tests,
high enough to melt the steel surface of pipe obstacles. These
localised effects appear to require good local mixing of air and
fuel and radiative feedback from nearby large hot surfaces.
Although the degree of congestion was generally not typical of
offshore modules, future application of techniques, such as
computational fluid dynamics with representative combustion
chemistry, should unravel some of these complexities.
A split vent of equal area to a single vent greatly
increased the air flow to the fire, producing a chimney
effect, and simple models based on the A root H
approximation for air inflow are invalid for these geometries.
The release pressure of condensate had a large effect
on the temperature rise of the impinged object. At low
pressure, large droplets with long burn out times were
generated, giving a cold spot on ceiling. At high pressure,
atomisation occurred producing smaller droplets with short
burn out times and no cold spot.
The fire stoichiometry of burning pools was
controlled by the air inflow and not by the pool fire burning
rate, confirming earlier observations. There was no difference
in burning rate between pool fires on water or steel. Fuel
controlled pool fires at large scale burnt at the same rate as in
the open (for the condensate used this was 0.065 kg/m2 /s) and
produced rapid temperature rises (up to 1300o C), and high heat
fluxes (up to 320 kW/m2 ) in the insulated compartments.
Combustion of soot is thought to play an important role under
these conditions. If the pool fire became ventilation controlled,
the heat release rate fell, the evaporation rate of fuel dropped,
and all the available air entrained into the compartment was
consumed. The fire stoichiometry settled in the slightly fuel
rich regime and unburnt pyrolysis products burnt at the vent.
For sound practical reasons the confined fire rig was
fully insulated during all tests. Offshore modules do not have
the same degree of insulation. Insulation of the test rig results
in higher final temperatures of the boundaries of the
compartment and higher smoke temperatures. A true steady
state was not achieved in any test because the final measured
wall and ceiling temperatures were not equal, or even similar
to, the final measured smoke temperature. Nevertheless, the
walls of a non-insulated compartment with radiation loss alone
from the outer surfaces would be expected to be cooler by a
factor of about 1.2 in absolute temperature. This extra
radiation loss feeds back to the smoke temperature, which in
turn changes the stoichiometry and hence the chemistry and
the heat release within the compartment. These feedback
processes are complex, but not intractable, and it is reasonable
to assume that sufficient information was generated by these

OTC 14132

tests to enable modelling of the fire behaviour and


extrapolation to offshore conditions.
In over-ventilated compartments, in which excess
combustion air is present, the situation is further complicated
by the fact that soot oxidation only becomes important at the
higher temperatures which are readily attained in the insulated
compartment. Since soot plays an important role in radiation
heat transfer any process which changes the soot concentration
can have a profound effect on the fire dynamics. The
experimenter measures extremely high temperatures and heat
fluxes in over-ventilated pool fires, for example. The air-rich,
hot surroundings promote partial soot oxidation, thereby
releasing more heat and increasing the radiation transfer from
hot soot particles. This process appears to become important
when smoke temperatures reach about 1200-1300o C. In a noninsulated compartment the smoke temperatures may never be
high enough for soot oxidation to switch on and the
escalation to higher temperatures may not occur. The pool fire
would then be expected to burn like a pool fire in the open.
Another effect of insulation is to mask some of the
more subtle influences of stoichiometry or scale on the
temperatures reached in ventilation controlled jet fires.
Decreasing the stoichiometry by reducing the vent area at a
given scale has little effect on the smoke layer temperature in
the insulated compartment (both in practice and theoretically
over the range studied), but is expected to reduce the
temperature somewhat when the insulation is removed.
Increasing the scale at a fixed vent size will have no effect on
the smoke temperature in a fully insulated compartment, as
observed, but will lead to reduced smoke temperatures in a
non-insulated compartment, because the heat (radiation) losses
from the boundaries must increase with scale due to the larger
surface area over which losses can occur. Quantification of
these global changes is readily achieved by application of
physical models, for example [27], that properly account for
the heat and mass balances.
In summary, the general level of understanding of
compartment fire behaviour is now sufficiently good to assess
most compartment fire hazards with some confidence for
modules having simple geometries. In particular jet fire and
pool fire temperatures, smoke layer temperatures, heat fluxes
to surfaces within the module, the extent of external flaming
and internal impingement zones can be reasonably well
predicted. Estimates for CO concentrations in the smoke layer
are also available based on empirical relationships to
temperature and flame stoichiometry. Future improvement in
model development should focus on evaluating the
combustion product emissions from module vents, particularly
for CO and smoke concentrations which are both treated
simplistically in most models at present.
Control and Mitigation
Water based systems. Water spray systems are routinely
installed on offshore platforms for mitigation of fire hazards,
but their effectiveness has only been assessed in recent years
[30]. A study by Shell and British Gas [31] has shown at

OTC 14132

CONTROLLING HYDROCARBON FIRES IN OFFSHORE STRUCTURES

realistic scale that a dedicated deluge system (10.2


litres/m2/min) cannot be relied upon to prevent excessive
temperature rise when a vessel (2m diameter, 8m long) is
subjected to a high velocity gas jet fire (3 kg/s). The high
velocity jet readily penetrated the water film. High pressure
water from a monitor was however more effective if aimed
directly at the impinged area.
The same vessel was used to determine the effectiveness
of a commercial directed deluge system in protecting against
flashing liquid propane jet fires of various sizes and separation
distances [32]. The deluge system was designed to deliver
10.2 litres/m2 /min in accordance with NFPA 15 but actually
delivered 17.6 l/m2 /min, more in accordance with the design
rules in reference [33]. It was found that a typical deluge on an
LPG tank could not be relied upon to maintain a water film
over the whole tank surface in an impinging jet fire. Where the
water film breakdown results in a dry patch the tank wall can
reach temperatures above 120 o C but the rate of temperature
rise will be significantly less than the unprotected case by a
factor of 1.5 to 5.8. Releases close to the LPG tank can result
in liquid propane impinging the target vessel with local low
temperatures and icing on the vessel wall. High temperature
gradients are created to the surrounding hotter areas.
A similar experimental study [34] was carried out to test
whether the Tentative Rules, originally advocated to protect
against pool fires, would protect a 1.2m diameter, 4.5m long
LPG tank against a 2 kg/s flashing liquid propane jet fire.
Similar conclusions were drawn and some extra design
guidance has been recommended [35].
Area deluge typical of the arrangement used offshore has
been shown [5] and confirmed by the BFETS Project [28,29]
to extinguish compartment jet fires in the test conditions used.
This however may be undesirable because of the risk of
further ignition and explosion. Early application of water
deluge while the compartment was still relatively cool did not
extinguish the jet fire, and the average smoke layer
temperature fell to below 200o C within 2 minutes. The main
mechanism of extinction is thought to be one of inerting the
entrained air by steam. The extent of hot surroundings play an
important part in determining the steam loading and further
work is necessary to develop safety engineering guidance.
Pool fires do not extinguish in the same deluge conditions, but
the fires are effectively controlled and burn at reduced rates.
External flames became intermittent in the case of fuelcontrolled fires, but were otherwise extinguished and were
replaced by copious smoke and steam.
Passive fire protection systems. Passive fire protection
coatings can guarantee effective protection against offshore
fires, if they are applied correctly and tested in realistic fire
conditions [36,37]. Passive fire protection materials have
traditionally been specified on the basis of fire resistance tests
carried out in a furnace operating under time-temperature
conditions defined by a fire curve [38]. Such tests do not

replicate the balance of radiative and convective heat transfer,


high gas velocities and thermal shock that are major factors
with regard to the performance of passive fire protection in
actual fires and in particular, jet fires resulting from high
pressure gas leaks.
Following early development work by Shell Research
[36] and SINTEF NBL [39], the UK Health and Safety
Executive, in conjunction with the Norwegian Petroleum
Directorate, set up a working group to examine jet fire testing
and to work with industry to develop a standard jet fire test
procedure for passive fire protection materials. Through the
Jet Fire Test Working Group the early work evolved into a test
procedure, the Interim Jet Fire Test for Determining the
Effectiveness of Passive Fire Protection Materials [40] and
subsequent Jet Fire Resistance Test [41], which reproduces the
key conditions of a large-scale (3kg/s) natural gas jet fire using
a 0.3 kg/s propane gas jet flame. This test is now widely used
to assess PFP coatings and systems.
CFD modelling is becoming increasingly important
for predicting the heat fluxes to vessels or test specimens. A
recent example to predict the detailed structure of the flames
in the Jet Fire Test and large-scale jet fires is given in
reference [42]. The model enables the comparison of difficult to-measure properties such as gas velocities and temperatures
at any point in the flow-field and radiative and convective heat
transfer. The CFD model can then be used to optimise the
design of new jet fire tests.
Vessel blow-down. Current industry guidance, e.g. API 521,
does not consider severe fires particularly impinging jet fires
that could lead to catastrophic failure of vessels before the
inventory has been safely removed. The efficacy of blowdown systems can be placed under increased scrutiny now that
fire boundary conditions and better understanding of the fluid
thermodynamics during de-pressurisation are available. An
example of such an analysis is summarised here but there is a
clear opportunity for extending the knowledge more generally.
A major concern though is that there is no
experimental validation.
Assume an equilibrium mixture of methane, ethane,
propane and butane at 120 bar and 45o C in a vessel 5m long
by 1m diameter and designed to blow-down to 7 bar in 15
minutes. The vessel is 30% full. When engulfed by a typical
natural gas jet fire of 150 kW/m2 internal radiative flux and 40
m/s velocity gas at 1400o C, calculations using the
BLOWFIRE code, an extension of the BLOWDOWN code
[43, 44] suggest that the vessel will remain intact. If the
mixture concentration is changed such that the equilibrium
pressure is 30 bar, then the vessel is predicted to fail in 9
minutes in the same fire. Clearly a different blow-down
strategy is required in this case. Similarly, in a fully developed
compartment fire, the 30 bar vessel is predicted to fail in 6
minutes, but the 120 bar vessel again survives.

G. A. CHAMBERLAIN

Sub-sea isolation valves. The need for a sub-sea isolation


valve (SSIV) in a pipeline can be based on an analysis of the
fire hazard consequences of pipeline rupture. For example, if
fire engulfment of a critical structural member, a platform leg
for instance, is of concern and the leg is say about 35m distant
from the release point, then jet fire analysis tells us that
releases of greater than 8 kg/s should be considered.
Engulfment time contours of the type shown in Figure 1 can
be constructed for gas pipelines to aid the correct positioning
of the SSIV. If for example the critical engulfment time is
1000 seconds before loss of leg integrity then the SSIV could
be placed up to 350m along the pipeline. Note that a worst
case hole size is about 40 mm in this case. In general, hole
sizes ranging from about 30 to 50 mm appear to cover worst
case assessments for gas releases.
The minimum distance of the SSIV is governed by
sub-sea rupture of the pipeline and the creation of a gas fire on
the sea surface. Realistic assessments of the hazards posed by
sea fires can be performed but the methodology remains so far
largely unpublished.
Concluding remarks
This paper cannot do justice to the rapid advances of recent
years in the assessment of the fire hazards posed by accidental
releases of hydrocarbons. The interested reader is strongly
urged to read the references cited herein. However many
validated methods are now available to the engineer for cost
effective safe design and research continues to address
remaining areas of uncertainty, such as
chemistry of compartment fires,
water droplet/flame interactions for fire mitigation,
empirical models for the behaviour of PFP materials,
experimental validation of models for vessel failure by
fire attack,
integration of hazard consequence modelling with event
probability in a consistent and demonstrable way by the
use of computer tools.
References
1. Lord Cullen, The Public Inquiry into the Piper Alpha
Disaster, Den (HMSO), Nov. 1990.
2. Steel Construction Institute, Interim guidance notes for
the design and protection of topside structures against
explosions and fire, Report no. SCI-P-112, 1992.
3. Steel Construction Institute, Blast and fire engineering for
topside structures, Phase 2, Final Summary Report, SCI
Publication Number 253, 1998.
4. Bowen, P.J., and Shirvill, L.C, "Combustion hazards posed
by the pressurised atomisation of high-flashpoint liquids",
J.Loss Prev. Process Ind., (1994) 7, No. 3.
5. Chamberlain, G.A.: "An Experimental study of water
deluge on compartment fires", International Conference on
Modelling and Mitigating the Consequences of Accidental
Releases of Hazardous Materials, AIChE, New Orleans
(1995).

OTC 14132

6. Chamberlain, G.A.: Evaluating offshore fires and


explosions, Proceedings of the International Conference
on Changing Health and Safety Offshore - The agenda for
the next 10 years, Aberdeen (July 22-24 1998) UK Health
and Safety Executive.
7. Kalghatgi, G.T.: "Blow-out stability of gaseous jet
diffusion flames. Part 1: Still air", Comb. Sci. and Tech.,
(1981) 26, 241.
8. Birch, A.D., Brown, D.R., Cook, D.K., and Hargrave,
G.K.: "Flame stability in under expanded natural gas jets",
Comb. Sci. and Tech., (1988) 58, 267.
9. McCaffrey, B.J., and Evans, D.D.: "Very large methane jet
diffusion flames", 21st Symposium (International) on
Combustion, The Combustion Institute (1986).
10. Davenport, N: "Large scale natural gas/kerosene mixed
fuel jet fires. Final report to the American Petroleum Inst.",
(1994) Shell Research report no. TNER.94.061.
11. Acton, M.R. and Evans, J.A.: Horizontal jet fires of oil
and gas, report to the Blast and Fire Engineering Project
for Topside Structures Phase II, Steel Construction
Institute (1997).
12. Cracknell, R.F. and Chamberlain, G.A.: Interpretation of
experimental data for unconfined crude oil jet fires, report
to the Blast and Fire Engineering Project for Topside
Structures Phase II, Steel Construction Institute (1997).
13. Chamberlain, G.A.: "Developments in design methods for
predicting thermal radiation from flares",
Chem.Eng.Res.Des. (1987) 65, 299.
14. Johnson, A.D., Brightwell, H.M., and Carsley, A.J.: "A
model for predicting the thermal radiation hazard from
large scale horizontally released natural gas jet fires",
Trans. IChemE., (1994) 72, Part B.
15. Johnson, A.D., Ebbinghaus, A., Imanari, T., Lennon, S.P.,
and Marie, N.: "Large-scale free and impinging turbulent
jet flames - numerical modelling and experiments",
IChemE Symp. Series 141, Hazards XIII, UMIST,
Manchester (1997).
16. Cowley, L.T., and Pritchard, M.J.: "Large scale natural gas
and LPG jet fires and thermal impact on structures",
Gastech 90, 14th International LNG/LPG Conference,
Amsterdam (1990).
17. Cracknell, R.F., Davenport, N., and Carsley, A.J.: "A
model for heat flux on a cylindrical target due to the
impingement of a large-scale natural gas flame", 2nd
European Conf. on Major Hazards Onshore and Offshore,
IChemE, (Oct 1995).
18. Davenport, N.: "Large scale natural gas/butane mixed fuel
jet fires. Final report to the European Commission", (1994)
Shell Research Report no. TNER.94.030.
19. Sekulin, A.J. and Acton, M.R.: "Large scale experiments to
study horizontal jet fires of mixtures of natural gas and
butane", British Gas report to CEC, GRC R0367 (1995).
20. Joint Industry Project managed by BG Technology, Large
scale experiments to study jet fires of crude oil/gas
mixtures and crude oil/gas/water mixtures, work in
progress.

OTC 14132

CONTROLLING HYDROCARBON FIRES IN OFFSHORE STRUCTURES

21. Considine, M.: "Thermal radiation hazard ranges from


large hydrocarbon pool fires", Safety and Reliability
Directorate Report SRD R297, UKAEA (1984).
22. Pritchard, M.J. and Binding, T.M.: "FIRE2: A new
approach for predicting thermal radiation levels from
hydrocarbon pool fires", IChemE Symposium Series No.
130 (1992).
23. Johnson, A.D.: A model for predicting thermal radiation
hazards from large scale LNG pool fires, IChemE
Symposium Series No. 130, (1992) 507-524.
24. Wighus, R., Meland, O., and Vembe, B.: "Smoke hazard in
offshore platform fires", SINTEF Report STF25 A91007
(1991).
25. Chamberlain, G.A., "An Experimental study of large scale
compartment fires", Trans. IChemE, (1994) 72, Part B,
211.
26. Chamberlain, G.A., "The hazards posed by pool fires in
offshore platforms", Trans. IChemE., (1996) 74, Part B,
81.
27. Persaud, M.A., Chamberlain, G.A., and Cuinier, C., "A
model for predicting the hazards from large scale
compartment jet fires", IChemE Symp. Series 141,
Hazards XIII, UMIST, Manchester (1997).
28. Chamberlain, G.A., Persaud, M.A., Wighus, R., and
Drangsholt, G., Blast and Fire Engineering for Topside
Structures. Test Programme F3, confined jet and pool fires.
Final Report, Steel Construction Institute (1997).
29. Persaud, M.A., Wighus, R., and Chamberlain, G.A, Blast
and Fire Engineering for Topside Structures. Test
Programme F3, confined jet and pool fires. Interpretation
Report, Steel Construction Institute (1997).
30. Wighus, R (1991), "Active fire protection: extinction of
enclosed gas fires with water sprays", SINTEF Report
STF25 A91028.
31. Shirvill, LC, and White, GC, "Effectiveness of deluge
systems in protecting plant and equipment impacted by
high velocity natural gas jet fires", International
Symposium on Heat and Mass Transfer in Chemical
Process Industry Accidents, Rome, 1994.
32. Bennett, JF, Shirvill, LC and Pritchard, MJ, Efficacy of
water spray protection against jet fires impinging on LPG
storage tanks, HSE Contract Research Report 137/1997,
ISBN 0 7176 1395 X, 1997.

33. Fire Offices Committee, Tentative rules for medium and


high velocity water spray systems, 1979.
34. Beckett, H, Cooke, G and Roberts, TA, Water deluge
protection against jet fires: Effectiveness of industry
recommended application rate, HSL Internal Report
PS/98/09, 1998.
35. Roberts, TA, Medonos, S and Shirvill, LC, Review of the
response of pressurised process vessels to fire attack, UK
HSE Offshore Safety Report, OTO 2000 051, 2000.
36. Shirvill, LC, Performance of passive fire protection in jet
fires, Major Hazards Onshore and Offshore, I.Chem.E
Series No. 130, Manchester, 1992.
37. Hunt, RJ, Shirvill, LC, and Earles, GA, "Design and
application of passive fire protection for the deck of the
Mars tension leg platform", Offshore Technology
Conference, Houston, 1997.
38. Shirvill, LC, and White, GC, Fire testing - A review of
past, current and future methods, OMAE Conference,
Copenhagen, 1995.
39. Wighus, R, and Shirvill, LC, A test method for jet fire
exposure, 7th International Symposium on Loss
Prevention and Safety Promotion in the Process Industries,
Taormina, 1992.
40. Interim Jet Fire Test for Determining the Effectiveness of
Passive Fire Protection Materials, Offshore Technology
Report OTO 93-028, UK HSE, 1993.
41. Jet Fire Test Working Group, Jet fire resistance test of
passive fire protection materials, UK HSE, Offshore
Technology Information Report OTI 95 634, 1995.
42. Johnson, AD, Shirvill, LC, and Ungut, A, CFD
calculation of impinging gas jet flame, UK HSE, Offshore
Technology Report, OTO 1999 011, 1999.
43. Haque, MA, Richardson, SM and Saville, G, "Blowdown
of pressure vessels - I. Computer Model", Trans. IChemE,
part B, Proc. Safety Environ. Protection, Vol. 70, 3-9, 1992
44. Haque, MA, Richardson, SM, Saville, G, Chamberlain,
GA, and Shirvill, LC, "Blowdown of pressure vessels - II.
Experimental validation of computer model and case
studies", Trans. IChemE, part B, Proc. Safety Environ.
Protection, Vol. 70, 10-17, 1992.

G. A. CHAMBERLAIN

OTC 14132

Time (s) for mass flow > 8 kg/s


(34'' pipe)

H o l e d i a m e t e r / m m

2 0 0

200

1 5 0

400

1 0 0

600

400
200

800

1 0 0 0

600
800

1 4 0 0

400

5 0

1 0 0 0

1 6 0 0

1 8 0 0
1 4 0 0
1 0 0 0
800
600
400

800

600
200

1 2 0 0

400
200
0

0
1 0 0

2 0 0

3 0 0

4 0 0

5 0 0

6 0 0

Length of pipe to SSIV (inc. riser)/m

Figure 1 - Contour map of time for which release is greater than 8 kg/s as a function of hole diameter and length of closed pipe section.

You might also like