Flixborough Some Additional Lessons       Presentation by J. I. Cox (FELLOW)

(Reprinted from The Chemical Engineer, May 1976, Issue No. 309, pp.353-8)


NOTE           This paper began life as a presentation to an IChemE symposium in December 1975. It and other papers presented at the symposium then appeared in The Chemical Engineer in April/May 1976 and later was specially reprinted by virtue of its insight into the causation hypotheses. This website version - and its companion piece Flixborough Revisited
appears with passages highlighted in red if thought to be of contemporary interest and/or directly relevant to the more contentious issues at the Public Inquiry.

Introduction

The explosion at the Nypro (UK) Ltd caprolactam plant at 4.53 p.m. on 1 June 1974 took place in a large cloud of cyclohexane that escaped when a temporary pipe collapsed. The Court of Inquiry into the Flixborough Disaster decided that the defects of this plant modification, coupled with a rise in operating temperatures and pressures, was responsible. They state (Para 209) “the disaster was caused by the introduction into a well-designed and constructed plant of a modification which destroyed its integrity”.

In consequence of this technical assessment, subsequent discussion of the report has concentrated on management issues. Little of general application has emerged so far from the technical investigations, and Nypro may be the only company to appreciate the full significance of the investigations conducted during the Public Inquiry — although industry is of course well aware of the general implications.

In this presentation, lessons for plant safety are considered without regard to whether these had, or were thought to have had, direct bearing on the disaster. In adopting this approach, alternative explanations for the disaster (Fig 1) are recalled to clarify the technical issues and to emphasise the degree of probability of such phenomena occurring (again). The paper thus serves to indicate alternative views on the cause of the disaster - though the main purpose of the Nottingham IChemE Symposium was to draw attention to the technical lessons.

The investigations are considered in greater depth by Cottrell and Swann1 and Ball2 (metallurgy) and Gugan3 (combustion), whereas this paper deals with the overall implications.

1 Water Sprays

A 20 inch pipe and bellows assembly was installed as a temporary bridge between the fourth and sixth cyclohexane oxidation reactors (Fig 2) when the fifth reactor was removed at the end of March 1974 following the discovery of a ½ long crack in the stainless steel lining. The inner crack had developed from a 6 ft crack in the mild steel cladding and this, in turn, had been caused by nitrate corrosion from a temporary spray of cooling water (to which nitric acid had been added for pH control).

This spray was intended as protection against a most unlikely event. Several months earlier a small air leak had been noticed in the air supply line to the fifth reactor. Nypro’s criteria for defining a leak were more stringent than often used in the industry. It was decided that protection was needed in case the air compressor stopped unexpectedly and the block valves and compressor failed to prevent back-flow. The spray had been intended to condense any cyclohexane that might escape.

It is ironic that this ultra-cautious action by a “safety conscious” company (Paras 201, 202) was the starting point for the disaster. Users of water sprays, whether to dilute leaks, improve heat exchanger performance, or any other of the numerous reasons for their application, are warned (Para 212) that contaminated water can cause stress corrosion. Another lesson is that safety measures often involve an element of compromise between alternative hazards and that the blind application of ‘rules’ can have unexpected results.

2 Court Investigations

These dealt with the mode of failure of the 20 inch pipe assembly. The Court concluded that the assembly was subjected to conditions of pressure and temperature more severe than any which had previously prevailed” and which were sufficient to and did cause the assembly to rupture, and thus to release large quantities of cyclohexane. Such cyclohexane formed a cloud of vapour (mixed with air) which exploded” (Para 225).

Experiments with a duplicate assembly were conducted on behalf of the Court by the Safety in Mines Research Establishment (SMRE)4. They were unable to reproduce the jack-knife of the disaster (illustrated in Fig 3c), but did achieve double-bellows instability (‘squirm’ in the Court’s terminology) just above normal working conditions (Fig 3a). It was established (Para 123) “that rupture initiated by jack-knifing was unlikely to occur before relief valve pressure” and that, if a squirm did occur without bellows rupture, “the assembly with both bellows squirmed would not then rupture save at pressures above relief valve pressure”.

By elegant mechanical engineering reasoning, Professor Newland reconciled the SMRE experiments with a jack-knife collapse5. He calculated that the initial downwards movement towards the ‘squirm’ position, under only slightly different conditions to those of the SMRE simulation tests, could provide kinetic energy to deform the lower mitred bend and initiate a jack-knife. The results of his calculations6 are summarised in Table 1 and his proposed mode of failure is illustrated by Fig 3b (above).

Table 1 shows that, for a 43 inch scaffolding span (as used by SMRE), squirm would occur at 9.7 kg-cm-2  for the expected bellows stiffness of 2 800 lbf/in. As this is only 0.2 kg-cm-2 into the 9.5-10.8 kg-cm-2 range for jack-knifing, the chances of a subsequent jack-knife may be expressed roughly as 0.2/1.3 (about 15%) - an improbability confirmed by the SMRE squirm without jack-knife at 9.8 kg-cm-2.

By contrast, the 78 inch span favoured by the Court (Para 124) would allow squirm below 9 kg-cm2 with no chance of subsequent jack-knifing. So it had to be assumed (Para 125) that the Flixborough bellows were stiffer than 2800 lbf/in. The fact that they failed to squirm at 9.2 kg-cm-2 earlier the same day does not validate this assumption as the temperature was then lower than 150°C (Para 86) - probably 120°C. This reinforces the admitted “low probability” (Para 191) that a jack-knife would be caused by severe operating conditions.

TABLE 1        Internal pressures causing squirm and jack-knifing after squirm

                                                                                Internal Pressures (kg-cm-2)

Span of scaffold supports

Temperature (a)

For Squirm (Fig 3a)
Bellows axial stiffness
(lbf/in)

For Jack-Knifing after Squirm (Fig 3b) Probabilities

   

2800(b)

3300

0%

50%

100%

43"

150°C

9.7

11.3

9.5

10.1

10.8

 

160°C

9.5

11.1

9.3

9.9

10.6

78"

150°C

8.8

10.3

9.3

9.9

10.6

 

160°C

8.6

10.1

9.1

9.7

10.4

             N.B.  (a) Operating conditions, 155°C and 8.6-8.8 kg-cm-2.
                       (b) A compromise between 2860, calculated by the manufacturers, and 2752,
                       the actual value of the SMRE test bellows, made from the same material batch.

This improbability in no way invalidates the conclusion (Para 63, 209) that the 20 inch pipe assembly as installed was quite unsuitable and the recommendations with respect to non-compliance with BS3351:1971 stand. The bellows and pipe were shown to be separately suitable for the duty but the assembly was not pressure-tested in situ. Had this been done pneumatically at 155°C, the probable result would have been a squirm as in the SMRE test (Fig 3a) and the defective design would have been revealed.

In view of the dangers associated with pneumatic testing, BS3351 recommends hydraulic tests and the Court hypothesises (Para 73) that “Such a test would almost certainly have caused failure”. Whether or not this is right, it would have been relevant to comment that hydraulic testing would have been impracticable on the plant as designed and it is advisable to design provision for pressure-testing of sections of plant.

With respect to the content of BS335I 1971 it could have been mentioned that reference to “Advice of the manufacturers” is not satisfactory in a British Standard. In this case the bellows’ manufacturers supplied their clients (the plant constructors) with their ‘Designer’s Guide’ but Nypro did not receive copies until two months after the disaster. It is unrealistic to assume, as do British Standards, that users of plant equipment have ready access to manufacturer’s advice at all relevant times.

3 Plant Conditions

Operation

The disaster occurred whilst cyclohexane was recycling at essentially working pressures and temperatures awaiting a delivery of nitrogen to supplement the stocks needed for shut­downs and emergencies. Overpressure was a potential hazard during this prolonged start-up period as the pressure control valve (PCV on Fig 2) was not suitable for these conditions and was isolated by block valves. This could have been avoided by the provision of a tight shut­off control valve or a split-range controller.

However, the pressure was not uncontrolled: the inlet block valves were also closed (to prevent ingress of vapour) and the temperatures were on automatic control before the shift hand-over. In this situation a rise in pressures and temperatures required either deliberate operator intervention or equipment malfunction. Neither hypothesis was substantiated by the settings of the valves examined after the explosion and the presumption (Paras 87, 225) that the deceased allowed pressures and temperatures to reach levels “more severe” than “had previously prevailed” is not readily accepted by those who know the plant and the way it was operated. Nevertheless even the possibility that this could have occurred emphasises the need for attention to safe design for start-up as well as normal operation.

Nitrogen

During normal operation an automatic shut-down and purge of the oxidation reactors took place if the ‘Low Low Nitrogen Level’ was reached (so that oxidation stopped when nitrogen stocks were too low to cope with an emergency). If this had happened by mistake during warm-up, there might have been a pressure rise of about 0.8 kg-cm-2  in five minutes - assuming the purge was not stopped.

Although this theory was rejected (Para 88) - there was no supporting and some opposing evidence - it illustrates an important and not uncommon problem. A safety device (in this case the automatic nitrogen purge) can be a hazard in circumstances other than those for which it was designed. So provision for deactivation was allowed - an option which accords with the reality and complexity of plant operation. (In contrast, the Factory Inspectorate advocated rigid immutable written instructions for plant operators and that deactivation of such safety devices not be allowed during plant operation).

Internal explosion

King7 suggested a mechanism dependent upon plant scale - a significant temperature difference in the fourth oxidation reactor allowing a ‘cool’ slug of water at the bottom of its otherwise hot contents. At 4.53 pm on 1 June the water might have boiled unexpectedly and caused a sharp pressure rise. Although feasible, the amount of water and the delay before boiling both appear larger than possible at Flixborough. However, designers should be alive to the possibility that new phenomena become significant for large plant sizes. It is of interest also that King’s hypothesis gains credibility from the temporary absence of the stirrer from reactor 4. This illustrates how an apparently innocuous modification may have unexpected significance. But similar plants have not recorded sudden pressure rises during start-up, even when operating without agitators and after water displacements.

Other internal explosion mechanisms were suggested and rejected. All depend upon detailed chemical reactions that do not have general application and so are not discussed herein. The relevance of an internal (or an external) explosion is that it circumvents the problem that a slow rise in internal pressure was shown by the Court’s investigation (section 2) to be unlikely to cause the jack-knife failure of the disaster.

4 Zinc embrittlement

The 20 inch pipe would have jack-knifed from a fairly small external shock wave: certainly no more than 4 psi, possible as little as 0.01 psi. So an external explosion, especially if from above, could provide an explanation for the jack-knife and the presence of foreign debris (Court Plate 10 and Fig 3c), which must have been trapped in mid-flight.

The centre of the big explosion was east south east of the oxidation reactors (Fig 4), and two fan rotors had landed 60 ft and 140 ft away, roughly at right angles to the direction of the main blast. The finned tubes of the southernmost fin-fan cooler (from which these fan rotors had been ejected) had shattered into 2-3 ft lengths and, since the supporting struts were on the west, the rotor must have passed through the overhead bank of finned tubes. As the collapse of the plant structure would have blocked this flight path, a local explosion seems to have occurred before the collapse of the bank of fin-fan coolers (which presumably fell with the big explosion).

(N.B. The furthest flung Fan Rotor landed on waste ground and was not engulfed in the subsequent general fire. There were signs that the bearings’ grease had begun to melt whilst the rotor was in its normal position. In addition the rotor was covered in soot from the flash fire of the main explosion. These facts were consistent with a fire and explosion in the fin-fans prior to the main explosion.)

Mr Orbons, a DSM metallurgist, showed that the shattered tubes had suffered zinc embrittlement. The fins were galvanised mild steel but the tubes (of the southernmost cooler only) were stainless. Detailed studies of zinc embrittlement confirmed that zinc can crack stainless steel within “a fraction of a second” of the temperature reaching 800-900°C (Court Appendix II).

So it seemed possible that a relatively small fire shattered the finned tubes and released cyclohexane for an explosion in the fan housing. Nevertheless, evidence for a pre-event9 and apparent sighting of a fan rotor in flight10 and the eye-witnesses (right-hand corner, Fig 4) who reported flames from the position of the fin-fan coolers before the main blast were all discounted by (Para 164) ‘No evidence of any kind was given to suggest that this was a reasonable possibility much less a probability.”

Zinc embrittlement was not confined to the fin-fan coolers and the phenomenon assumed some importance in relation to the damage to an 8 inch line (section 7). Experiments with the galvanised wire used round pre-formed lagging confirmed (Court Appendix II) that “Zinc from a wire can cause embrittlement. The wire needs to be close to the specimen but contact is not essential.” The Court concluded nonetheless (Para 162) that they “cannot regard it (zinc embrittlement from galvanised wire) as a real possibility”.

These conclusions severely weaken the important warning and recommendation (Para 213) “that the attention of industry should be drawn to these matter”. The experiments showed that the juxtaposition of zinc and stainless steel is very dangerous11 and it is disappointing that the Court did not recommend that galvanised mild steel integral with stressed stainless steel should be prohibited, especially since they accept that “a relatively small but fierce fire can, if there is a source of zinc nearby, cause a sudden catastrophic failure”.

5 The Unconfined Vapour Explosion

Studies of the wreckage, discharge calculations and wind tunnel simulations all confirmed that the location and size of the major explosion were consistent with discharge 20-25 seconds after the collapse of the 20 inch line. Although most eyewitnesses reported flames some time before the explosion, Para 93 suggests that ignition did not occur until the cloud reached the Hydrogen Plant. During the hearings it was implied that earlier ignition might have prevented the explosion whereas a slower escape might have merely delayed it.

Fig 4 Simplified Site Plan (omitting details not referenced in paper)
[redrawn for this website in colours and in original landscape orientation]

Gugan3a suggests instead that flames were present throughout the escape but were kept at bay by the rapid expansion of the cyclohexane cloud. When the discharge slowed, flames moved inwards from several directions culminating in a highly ‘efficient’ explosion. This is consistent with the 20-25 second period expected from the discharge calculations and is substantiated by eyewitnesses12. It implies that the time of ignition was not relevant to the size of the explosion but that the rate and quantity of discharge was all-important. Moreover, it suggests that the inventory of over 200 tons of superheated cyclohexane circulating through the plant without any isolation valves was a serious design weakness: the Court’s theory does not.

Designers of plant should bear in mind that a high inventory of a critical process fluid presents a safety hazardand that measures may be needed to reduce total inventory or instal devices to limit the critical inventory at risk. Although there are practical difficulties associated with isolation valves, these are not insurmountable and it is most disturbing that the Court (Para 203) absolved the plant design from criticism on this count.

Ignition preceded the unconfined vapour explosion at Feyzin by an appreciable period, confirming that Gugan’s theory is feasible with all its implications for plant design, fire­fighting and site planning. It implies that an explosion - not a big fire - can be expected from a large and fast enough escape and suggests a design criterion for the numbers of isolating valves needed to prevent this type of explosion.

6 Partial Collapse of the 20 inch Line

The foregoing suggested that, after a small fire and an explosion in the fin-fan coolers, the 20 inch line could have jack-knifed and, following this collapse, a large escape of cyclohexane kept the already-present flames at the periphery of an expanding cloud. But the five laboratory witnesses with a clear and immediate view of the first cyclohexane escape did not report an emission from the fin-fan coolers (Fig 5): their evidence was more consistent with a low-level escape much nearer the 20 inch line.

 

Fig.5 – Laboratory viewpoints

A         Plan view showing initial positions of witnesses and various other features

This led to the hypothesis that a simple ‘squirm’ cracked the bellows, initiated a small escape (which went down into the plant and then out from under the plinths), and thus started the sequence of events. But no evidence of a preliminary crack in the bellows was found and the hypothesis did not explain the size of the initial escape or the damage to an 8 inch line nearby (Fig 5a, plan view and Court Plate 10). If correct, it would have confirmed but not added to the lessons of Section 2.

7 Damage to the 8 inch Line

    The ruptured elbow G (Fig 6) was the most obvious but not the only damage to the 8 inch line. A large (3 inch long) zinc-embrittled crack had appeared on the east face, 5 ft above the elbow, and a spatter of smaller cracks around the rupture and further downstream. The line received detailed study1 as the first reported escape was consistent with its location (Fig 5b), and no witnesses reported seeing vapour from the clearly visible nozzle of reactor 6 (Fig 5d) or movement of the 20 inch line as the escape emerged.

    Cottrell and Swann’s investigations are summarised in the Court’s Appendix II but not interpreted. These studies established that the elbow ruptured through creep failure at essentially working pressure and at a temperature of at least 900°C. As such conditions were not possible after the loss of line pressure2a, this led the writer to eventual acceptance that failure of the 8 inch line must have preceded the big explosion.

    From the standpoint of understanding what must have happened, the most helpful finding was the inferred temperature gradient of the line whilst still under pressure (Fig 6a). It was shown that the elbow reached perhaps 950°C but that the temperature then dropped away very rapidly and, only 5 ft above the elbow, probably never exceeded the temperature generally prevailing in the post-explosion fire. Such highly localised heating could not occur during a truly general fire and the only credible source for the required local frame was accepted to be a leak from the adjacent non-return valve. However, this leak would have impinged on the elbow only if re-directed axially by an essentially intact lagging box (an impossibility in the post-explosion fire).

    The hypothesis that there was a pre-explosion leak from this joint was a “theoretical possibility” (Para 154) because the non-return valve had two bolts untightened which, after investigation, it was accepted (Para 143) “were probably loose before the disaster”. Moreover, SMRE found that parts of the relevant gasket were missing (as opposed to having disintegrated).13

    So it seemed conceivable that there was a gasket blow at the non-return valve which resulted in a flame directed at elbow G and this caused the (pre-explosion) rupture. Relevant details of this, the 8 inch line hypothesis, are outlined below and the complete conjectured sequence of events appears as Fig lb.

Fig.6 – Elbow G from east

A         As found - with inferred temperatures (°C)
B        
Conjectured - early in leaking period with lagging and cladding partly consumed at elbow


Notes
(a) Zinc embrittlement requires 800-900°C for at least a fraction of a second whilst the affected surface is free from a surface oxide layer.
(b) Swelling for 5 minutes and 15 kg/cm
2 (normal working pressure) requires the temperatures shown. Longer periods or higher pressures need lower temperatures for the same swelling effect. The areas recorded as <850°C had circumferences within the manufacturing tolerance and may have been at 150°C throughout.
(c) The preferred exit of a leak from the lagging box may be presumed to be close to the loose bolts of the non-return valve (NRV).

Initial leak and ignition

The conjectured escaping fluid would have been a cyclohexane/water mixture and the aluminium cladding over the pipe insulation was not electrically bonded. Para 154 accepted “that gaskets can blow and that a flashing liquid can produce a charge of static electricity on conductors such as aluminium” and that this “could cause ignition”. The contraction and expansion of the line during the 3-day shut-down could explain the timing of the gasket blow. Lessons include:

1.     Checking systems should be absolutely dependable to ensure that flanges are fitted correctly and properly tightened.

2.     Codes of practice for piping material specifications are appropriate for such services and might usefully recommend a minimum flanges policy.

3.     There should be restrictions on the use of sandwich connections (of the type used for the non-return valve illustrated by Fig 6) in which two gasket joints are held by a single set of bolts.

4.     Thermal insulation cladding should be electrically bonded on all equipment, including piping, as is the general practice for vessels.

5.     Special care on these points is needed with hot two-phase mixtures because of the extra hazard from static electricity with oil/water mixtures and flashing liquids14

The flame from the lagging box

On initial emergence from the lagging box most of the escape would have been directed towards the elbow G, roughly as illustrated in Fig 6b. Bluff body turbulence would have made the off-port flame stable at the intrados, but this situation would not continue indefinitely. All parties were “satisfied that ... the lagging box would have been destroyed and the necessary directional flame with it” (Para 161). The difference of opinion (Paras 155-157) was over how long the directed flame would last15. The Court, though not venturing an estimate of its own, clearly thought 10 minutes excessive.

It became evident during the hearings that the concept of an off-port flame is not readily understood, and the fact that aluminium can be and is used for burner nozzles was met with frank disbelief. It would be most unfortunate if industry shared this misconception: one lesson from the investigations is that the location of joints and the design of lagging boxes should ensure that potential leaks should not be directed at highly stressed equipment and, where this cannot be avoided, not from the distance which is the optimum for flame stabilisation. In principle of course, the most important precaution is to eliminate the potential leaks but, in recognition of the impossibility of 100% security in this respect, it would be useful to leave certain joints unlagged for inspection.

The deluge system

As shown in the pre-disaster photograph 3b, the plant had water deluge protection for vessels and critical items such as level glasses—although not for the non-return valve or elbow G. Over one hundred spray nozzles at the south end of the oxidation unit could be actuated by 68°C quartz sensor bulbs on a parallel air network, but there was no evidence that these operated on June 1st.

The nearest sensor bulb visible on the photograph is over 10 ft distant and this raised doubts concerning the adequacy of plant protection currently employed in industry. Experimental work by the Fire Research Establishment16, assumed that there was an additional sensor bulb within 5 ft of the elbow, hidden from view by a water spray. (In Para 160, it is said that there was “convincing evidence” for a sensor in this position, even though it could not be found in the wreckage - as were all the others - and did not appear on installation drawings).

The FRE experiments showed that there is no certainty that the deluge would have been activated even by the assumed ‘hidden’ quartz bulb: sensors that rely on conduction and convection afford much less protection than commonly supposed. Substantially more conduction/convection sensors or radiation-activated sensors are needed. By an unfortunate oversight, the original report on these experiments, and the calculations refuted in Para 157, are not listed with the “Reports presented to the Court” (Appendix VI), giving credence to the claim (Para 160), “No-one from any sources produced, before the end of our hearings ... any calculations (to show that the deluge would not have operated”). Industry needs to be alerted to the far-reaching implications of this part of the Flixborough investigations.

The petal and 3 inch cracks

The 8 inch line was lagged with preformed rock wool sections kept in position by lengths of 14 SWG galvanised wire, wrapped around the sections and spaced 12 inch apart. The twisted ends on vertical sections were laid down the axis of the pipe as illustrated on Fig 6b and the aluminium cladding placed over all.

There was a zinc-embrittled crack on the opened out ‘petal’ of the tulip rupture, about 12 inch from the expected termination of the lagging on the horizontal section adjacent to the NRV. The 3 inch crack was about 5 ft up the vertical section above the elbow and had been nucleated from (at least) 5 separate points in an almost precise vertical line (on the east side, by the walkway from which the lagger would have worked). As these positions were consistent with zinc embrittlement from expected positions of galvanised wire (and inconsistent with a random spatter of zinc droplets), experiments with this wire were undertaken by Cottrell and Swann. These showed (Appendix II) that such wire could indeed cause zinc embrittlement.

The Court did not accept that galvanised wire provided the zinc for these cracks, apparently because they visualised total disintegration of the cladding and lagging (with vaporisation of zinc at 907°C). But experience of small-scale fires shows that preformed sections will fall after disintegration at only one or two critical points (at around 600°C). The comment (Para 162) that the Court “cannot regard (zinc embrittlement from galvanised wire) as a real possibility” is no guarantee that stainless steel pipe could not be embrittled in a small fire by galvanised wire.

    (Note. Both hypotheses accepted that the smaller zinc cracks formed after the rupture of elbow G (Fig 1). However the 20 inch hypothesis was unable to explain why these appeared only on south east faces of pipe below the 3 inch crack or how the zinc was able to penetrate the surface oxide layer formed by several minutes - at least - of a general fire. As the crack orientations were consistent with the stress-time sequence of Fig lb (the 8 inch hypothesis), the Court was obliged to attach (Para 187) “no special significance to the direction of zinc cracks” and did not proffer an explanation for the stress history of the line).2b

Creep failure

It was accepted (Para 214) that fracture through creep “can be produced in a matter of minutes by a small fierce fire” but not accepted (Para 162) that the envisaged flame from the lagging box could have raised the pipe temperature to over 900°C. Experiments by Gugan achieved red heat from flame impingement on a pipe containing flowing water but were criticised on the grounds that the flow rates used were too low and that, in practice, liquid would have swept away the vapour blanket needed for such temperatures to be attained.

Unfortunately the actual flow conditions were not known. As the plant was recirculating prior to start-up, low circulation could have been expected and, in confirmation of this, the discharge valve on the pump was found throttled (Fig 2). But the flow was level-controlled and the circulation could have been varied without adverse effects. DSM suggested that the flow could have been one-third of normal and this was used by Gugan as basis for his experiments. However, once the 3 inch crack had opened (this required no more than 800°C at the outer surface), the flow through the elbow would have been reduced by at least half that figure (to around 25 lb-s-1).

The flow would have been further reduced as a result of vaporisation of the cyclohexane/water mixture and, above 500°C, by decomposition of the cyclohexane into benzene, hydrogen and carbon. The consequence of extra pressure drop downstream from elbow G (and the throttle valve) would have been less flow and even more vaporisation. Such factors, well known to boiler engineers, show that large flows do not necessarily provide pipes with protection from high temperatures in a fire.

Stronger pipe would delay but not avoid creep failure. The more fruitful approach is to eliminate the factors that might cause such a fire. The creep failure investigations have emphasised the vulnerability of plants to small fierce fires—the main overall lesson from the 8 inch line hypothesis.

8 Lessons from a Public Inquiry

Long before the SMRE experiments were concluded, or the Cottrell/Swann investigations even begun, Nypro’s technical advisors were rebuked publicly by the Chairman of the Court because they were not prepared to express a view on the cause of the disaster. The fact that we waited until the evidence was available does not prove us right - or the Court wrong - but it does illustrate that a public inquiry is not necessarily the best way to conduct a technical investigation. The procedures encourage rigid and premature postures to be adopted and discourage investigation in depth.

Moreover, having come to a conclusion, a Court will tend to concentrate on arguments germane to its own conclusions and so undervalue other potential lessons. Paras 116-117 and 166-171 typify this approach. A cine-film taken from location 31 (Fig 4), within a minute of the explosion, shows a vertical phenomenon “said to be a near vertical turbulent jet of flame coming from the rupture in the 8 inch line” The Court agreed that if a vertical flame could come from the ruptured elbow, this “would strongly support the 8 inch hypothesis “. To test this conjecture, a ¾ inch plastic elbow was made which copied the geometry of the rupture. On discharge, this produced a substantial upwards component thus emphatically reinforcing the 8 inch hypothesis.

The report does not mention this experiment, specifically designed to give a realistic simulation of (he direction of emission from the ruptured elbow, but refers, instead, and at length, to experiments on a different topic, with a ¼ inch elbow which “did not ... attempt to reproduce the “tulip” effect in the original split “. The Court concluded this section with the stricture (Para 172), “We have dealt with this particular point in some detail for it appears to us to be a good example of the way in which the enthusiasm for the 8 inch hypothesis felt by its proponents has led them to overlook obvious dejects which in other circumstances they would not have failed to realise.”

This same point is mentioned for it is a good example of the way in which the Court’s commitment for the 20 inch hypothesis led them to present their conclusions in a way that does not help the reader to assess contrary evidence. The Court could still be right that a single unsatisfactory modification caused the disaster but this is no reason for complacency. There are many other lessons. It is to be hoped that the respect normally accorded to the findings of a Court of Inquiry will not inhibit chemical engineers in looking beyond the report in their endeavours to improve the already good safety record of the chemical industry.

Acknowledgements

Thanks are due to colleagues of L. H. Manderstam and Partners (UK) Ltd, and other co­workers on the investigations and to Nypro (UK) Ltd, for many helpful suggestions in the preparation of this paper. The views expressed are entirely my own personal responsibility.

References

1 Cottrell. A. H. and Swann, P. R.. Chem Eng. (London). April 1976. p. 266

2 BalI, J. C. Chem. Eng. (London). April 1976. p. 275

   a ibid Section 4.1

   b ibid. sections 4 & 5

3 Gugan, K., Chem. Eng. (London), May 1976, p. 341

   a ibid. section 4.2

   b ibid. Fig 2

4 Waterhouse, D. and Games, G. A. C., Construction of and tests on a reconstructed bridging pipe assembly, SMRE report to Public inquiry

5 Newland, D. E., Report on an investigation of possible causes of failure of the 20-inch by-pass assembly at Flixborough, Report to Public Inquiry

6 Appendix I and Para 124 of Court’s Report and Ref. 5

7 King, R., Process Eng., September 1975, p. 69

8 Evans, G. O. H. M. Factory Inspectorate, Private communication.

9 For example, Professor Sir Frederick Warner’s Proof of Evidence, page 41 and the preliminary Police Report, page 39.

10 The eye-witness at location 15 of Fig 4 could not see the plant and the spectacular vapour escapes (and flames) because of a control room. He reported* (Day 3, pp 81-91) an object something like a round disc spiralling in the air ... with a 6 ft wide cone of flames beneath it ... about 3 ft wide ... about 100-120 ft up immediately after the explosion (i.e. (initial noise) ... by the time 1 just turned my head around.” He estimated its flight as up from Section 27 and down to the Flaking Plant. As he was 600 ft away, this could well describe the trajectory of the furthest flung fan rotor (see Fig 4) which was “waggling about a little (as if) on a central axis” (as might be expected from the object with a thin shaft and small counter-weight shown on Fig 5).

11 Elliott, D., Zinc embrittlement of stainless steel - A postscript to Flixborough, (Unpublished). At the Nottingham symposium, Dr D Elliot reported that British Steel had conducted tests on this phenomenon and were confident that a satisfactory technical solution had been found.

12 All the more than thirty eye-witnesses with a clear view of the area reported flames prior to the major explosion (Fig 3). Moreover, whereas the earlier viewers reported flames moving outwards, others reported the opposite movement.

13 Foley, J. H. and Nicholson, C. E., Metallurgical examination of damaged pipes from section 25A, SMRE report to Public Inquiry.

14 Klinkenberg and van der Maine, Electrostatics in the Petroleum Industry, Elsevier and Guide to Fire Protection in the Chemical Industry. CIA.

15 The metallurgical evidence indicated that 4 minutes would cause creep failure at 950°C and working pressure so, allowing for heating-up, a slightly longer period would have been necessary in practice. No one was in a position to see the flame but the witness at location 25 (Fig 4) saw a “wisp of steam” 3-5 minutes before the main explosion (Para 112) rising from behind Reactor 3 (i.e. the position of the 8 inch line). He did not recall seeing this emission only a ‘few’ minutes earlier, suggesting perhaps that the time from gasket failure to creep rupture was less than, say, 10 minutes.

16 Nash, P. and Theobald, C. R., The use of automatic sprinklers as fire sensors in chemical plant, Paper to Nottingham IChemE symposium, December 1975.

* ADDITIONAL NOTE  (2005) see Flixborough FAQs for further details of this eyewitness evidence.