Sie sind auf Seite 1von 4

Rapid depressurization of pressure vessels

Afzal Haque, Stephen Richardson, Graham Saville and Geoffrey Chamberlain*


Department of Chemical Engineering, imperial college, London SW72BY, UK
Shell Expro, 1 Altens Farm Road, Aberdeen AB9 2HY, UK
Experiments were conducted on the rapid depressurization of large pressure vessels. Measure-
ments taken included the pressure, temperatures at a large number of positions both within the
fluid phase(s) and on the wall of the vessel, and composition, all as a function of time during the
blowdown process. The systems studied included subcritical and supercritical, condensing and
non-condensing. From these experiments, an understanding of the physical processes involved
during blowdown was evolved. This was incorporated into a mathematical model of blowdown,
and implemented in a computer program. The model correctly predicts all the major phenomena
observed in.the experiments, as a function of time.
(Keyw or ds: pr essur e vessel s; hazar ds; physi c al pr oc esses)
Rapid depressurization of pressure vessels, usually
called blowdown, is a hazardous operation, at least as
far as large pressure vessels are concerned. The hazard
arises because of the very low temperatures generated
within the fluid during rapid depressurization. This
inevitably leads to a reduction in the temperature of the
vessel itself and possibly to a temperature below the
ductile-brittle transition temperature of the steel from
which the vessel was fabricated.
In the release of gas from a vessel at high pressure,
the thermodynamic path taken by the gas inside the
vessel is one of constant specific entropy if the process
is adiabatic. This would result in the production of very
low temperatures (78 K) for the release of, e.g.
nitrogen at 150 bar down to ambient pressure. This
should lead to substantial heat transfer from the walls
of the containing vessel, and hence to the removal of
the adiabaticity of the process. As far as the wall is
concerned. there are a number of competing processes:
heat transfer by convection to the cold gas inside the
vessel; heat conduction through the wall of the vessel;
and heat transfer by convection between the surround-
ings and the vessel.
If the state inside the vessel, either initially or at
any stage in the blowdown, is one in which the vessel
contains e.g. liquid condensate with a vapour phase
above, there is the additional complication of heat
transfer from the vessel wall to the liquid by boiling.
This will result in the evaporation of material from the
liquid phase and its transfer into the vapour phase. All
the time this is happening, depressurization is taking
place through an exit pipe that could be removing
either the vapour phase or the liquid phase, according
to its position in the vessel. It is therefore quite difficult
Received 12 September I989
Presented at the First Int. Conf. on Loss of Containment, 12-14 September 1989.
London. UK
09504230/90/010004-0463.00
0 1990 Butterwort~ &Co. (Publishers) Lfd
4 J . Loss Prev. Process Ind., 1990, Vol3, J anuary
to predict how the system will behave.
In process plant, particularly in the petroleum
industry, there are many large vessels operating under
pressure and containing hydrocarbon mixtures. These
vessels may be horizontal or vertical, and have inlet and
exit pipes at the top, bottom or sides. Depressurization
is frequently necessary, and in an emergency it may
have to be rapid. It is under these circumstances that
the lowest wall temperatures will be observed. The
objective of this paper is to describe some of the work
performed to predict the behaviour of a vessel during
blowdown.
Experimental
The physical processes taking place during blowdown
are a complicated mix of several phenomena. To
determine the more important of these, a programme
of experimental work was carried out involving the
depressurization of three vessels ranging in diameter
from 5 to 110 cm, with length to diameter ratios from
10 to 3 respectively.
In the case of a gas-filled vessel, experiments
showed that the dominant mode of heat transfer at high
pressures was natural convection. This was most clearly
illustrated by examining the blowdown of a horizontal
vessel through an exit on the axis of the end closure of
the vessel. The vessel was 1.5 m long, with 0.27 m i.d.
and walls 2.5 cm thick: It was instrumented with a
pressure transducer to measure the internal pressure,
and = 80 bare wire thermocouples to measure the
temperature within the gas space in the vessel and the
surface temperatures of the vessel wall, both inside and
out. Pressurization was with nitrogen to 150 bar, and
depressurization was through an orifice of diameter
6.35 mm. A data logger was used to record all the
temperatures and the pressure once every 3 s. Blow-
Rapid depressurization of pressure vessels: A. Haque et al.
TOP
BOTTOM
Figure 1 Isotherms (left hand side) and schematic streamlines
induced by natural convection (right hand side) in a vertical
plane parallel to the ends of a horizontal vessel
down to atmospheric pressure was essentially complete
after = 100 s. Figure I (left hand side) shows a plot of
isotherms in a vertical plane parallel to the ends of the
horizontal vessel, obtained by interpolation and
smoothing of the raw experimental measurements. It
indicates the presence of natural convection with
streamlines roughly as shown on the right hand side of
Figure I . The isotherms are essentially the same on all
such vertical planes within the vessel, indicating that
the axial forced convecting component, due to flow of
gas towards the exit port, is dwarfed by the natural
convection circulating component. Measurements
made with this vessel in a vertical position indicate that
natural convection also dominates in this case.
Apart from the region close to the wall, the
temperature gradients within the gas space were quite
small and the band of temperatures shown in Figure 2 is
typical of those observed during blowdown of a
gas-filled vessel. Figure 2 also shows the corresponding
variation of the inner wall temperature. This too varied
little over the whole of the vessel.
Multicomponent mixtures behave in a similar
manner at high temperatures, where the mixture is
single phase gas, but show a markedly different
behavjour once condensation takes place. Figure 3
shows the bulk fluid temperature measured within the
same vertical vessel when a 70% nitrogen/30% carbon
dioxide mixture (molar composition) was blown down
from the top, with initial conditions of 150 bar, 20C.
At first, when the mixture was single phase, the
temperature fell in an essentially spatially uniform
manner within the vessel. After = 10 s, the tempera-
tures split into two bands, with those recorded in the
top part of the vessel continuing to fall and following
0
-20
Y
-?
, - 40
z
b
E
d
- 60
5
rn
. c
Q - 80
s
5
- 100
- 120
0 20 40 60 80 100
Time/s
Figure 2 Variation of temperature with time during blowdown
of nitrogen: the upper band refers to inner wall temperatures
and the lower band to gas temperatures; the solid lines indicate
predictions made using the BLOWDOWN package
20
- 20
d- ,
- i i
5
z
k
8
- 40
;
I I I I I I I I I I
0 20 40 60 80
Time/s
Figure 3 Variation of temperature with time during blowdown
of a 70% nitrogen/30% carbon dioxide mixture: the upper band
refers to bottom zone temperatures and the lower band to top
zone temperatures; the solid lines indicate predictions made
using the BLOWDOWN package
the lower band in the diagram. The temperatures
measured near the bottom of the vessel, however,
showed first a rise and then a steady fall. This
behaviour was caused by the onset of condensation. At
first, the amount of condensate formed was small and
as it reached the bottom wall of the vessel, which was
still relatively warm, it partially evaporated, driving off
the more volatile components and thereby raising the
J . Loss Prev. Process Ind., 1990, Vol3, J anuary 5
Rapid depressorization of pressure vessels: A. Haque et al.
boiling temperature. However, more and more conde-
nsate was formed in the upper part of the vessel, due to
expansion. This fell to the bottom, and, as the bottom
of the vessel itself was cooled, the evaporation rate fell
off and a pool of liquid gradually accumulated. Since
the pressure in the system was still falling, this liquid
experienced evaporative cooling. Clearly, a very com-
plex and competitive process is going on in this liquid
pool, but the experimental evidence indicates that
eventually evaporative cooling dominates the be-
haviour and the liquid temperature falls, as shown in
Figure 3.
Only a few centimetres depth of liquid were
formed during this experiment, but the evaporative
cooling was sufficient to reduce the inner wall temper-
ature at the bottom of the vessel to - 10C (at the top of
the vessel the corresponding temperature was 15C)
before the pressure reached 5 bar, at which point the
liquid layer, which was now almost pure carbon
dioxide, solidified to form carbon dioxide snow.
Similar behaviour was observed during blowdown
of natural gas/propane mixtures from the 1.1 m dia-
meter vessel. However, the much closer volatilities of
methane and propane, as compared with nitrogen and
carbon dioxide, did produce some quantitative differ-
ences. Thus, the depth of liquid condensate foimed
when depressurizing a 30 mole /o propane mixture was.
in relative terms. an order of magnitude larger. Since
propane has a triple point pressure < 1 bar. this
condensate remained liquid until atmospheric pressure
was reached. at which stage its temperature was close
to the normal boiling point of pure propane, indicating
that almost all of the volatile components had been
driven out. The inside wall temperature at the bottom
of the vessel was almost identical to that of the liquid.
The extensive series of measurements made on the
blowdown of the two larger vessels for various com-
positions provided the data base for modelling blow-
down. Most of the measurements were made on
vertical vessels blown down from the top. and the
model described applies specifically to this. However,
measurements were made on vertical vessels blown
down from the bottom and also on horizontal vessels
blown down from the end, but these are not discussed
here.
Computer model
A computer package called BLOWDOWN has been
developed to simulate rapid depressurization of vessels.
The vessel is divided into two zones: the top zone
contains all of the vapour together with any suspended
liquid-phase droplets; and the bottom zone contains all
of the liquid phase that has dropped out of the top
zone, forming a pool on the bottom of the vessel. If no
liquid is present the bottom zone is eliminated, but it
reappears if condensation occurs at any time. Each
zone is assumed to be at a spatially uniform temper-
ature and composition; all zones are assumed to be at a
spatially uniform pressure.
6 J. Loss Prev. Process Ind., 1990, Vol3, J anuary
The continuous depressurization process is re-
placed in the computer model by a series of discrete
time steps. Each time step is sub-divided into a number
of simple thermodynamic and heat transfer sub-steps
according to the following algorithm:
Select a pressure decrement.
Perform an isentropic flash on each zone.
Calculate the rate of discharge through the choke.
Calculate the duration of the time step and the
amount of fluid discharged.
Calculate the heat transfer coefficients for each
zone.
Perform energy and mass balances over the contents
of each zone and an energy balance over the vessel
wall.
If depressurization is complete, stop; otherwise,
repeat this process.
During blowdown from the top of the vessel, the rate of
discharge through a choke is calculated by requiring the
fluid to follow an isentropic path, the fluid emerging
from the choke with a speed equal to the local speed of
sound of the gas-phase component. The speed of sound
in the choke, V,, is determined by the simultaneous
solution of the equations:
Hi = H, + :VIC
s, = s,
where H denotes enthalpy, S entropy, p pressure, o
volume and T temperature; subscripts i and c refer to
conditions far upstream of the choke (where the flow is
essentially stagnant) and in the choke, repectively; and
C, and C,, denote specific heats at constant pressure
and at constant volume, respectively.
A number of possible situations could arise in
practice, since the fluid approaching a choke could be
one phase (vapour) or two phase (liquid and vapour)
and the fluid in the choke could be in a metastable state
or in thermodynamic and phase equilibrium. (The fluid
approaching the choke is assumed to be in equilib-
rium). Any combination of these phase states can be
handled, but as a result of comparison with the
experimental measurements, the preferred method is
to use one phase or two phase, as appropriate for the
fluid approaching the choke, and equilibrium for that
within the choke. In all cases, once the state within the
choke is evaluated, the mass flow rate through it
follows from a knowledge of the density and speed of
sound in this state, together with the cross-sectional
area of the orifice.
Heat transfer in the top zone allows for forced
convection I and natural convection; in the bottom
zone, it allows for nucleate3 and film4 boiling; outside
the vessel, it allows for natural convection2. Energy
balances over the contents and wall of the vessel are
always carried out using standard numerical methodsj
and solved by a matrix inversion methodh. A mass
balance over the contents is required only if liquid is
present, since mass transfer between zones only occurs
as a result of evaporation from the liquid pool in the
bottom zone, and sedimentation from the top zone.
Figures 2 and 3 show predicted and measured
temperatures. It can be seen from Figure 2 that
BLOWDOWN successfully predicts the temperature of
the gas and of the inner wall. Similarly, Figure 3 shows
that it successfully predicts the temperatures in the top
and bottom zones when condensation occurs. The
package for depressurization has also been validated
for natural gas and for natural gas containing up to 30%
propane.
Conclusion
The experiments undertaken and the model developed
have given a good understanding of the physical
processes occuring during blowdown, even for multi-
component multiphase systems. This work is now being
extended to cover: free water, which often forms
Rapid depressurization of pressure vessels: A. Haque et al.
during blowdown, especially offshore; pipe lines, over
which there is a significant pressure drop during
blowdown; and arbitrary combinations of vessels and
pipe lines, to simulate real systems more effectively.
Acknowledgements
This work was sponsored by Shell UK Exploration &
Production, under contract number C/22225.
References
I Chcmicol Engineers Handbook (Eds. R. H. Perry and C.H.
Chilton). 5th edition. McGraw-Hill. New York. USA. IY73. pp.
10.12- 10. IS
2 Chemical Engineers Handbook (Eds. R. H. Perry and C.H.
Chilton). Sth edition. McGraw-Hill. New York, USA, 1973, pp.
10. IO- 10. 12
3 Rosenhow. W. M. Trutts ASME 1952.74. Y6Y
4 Jordan-. D. P. Advunces Heut Trunsfcr 196X. 5. 55
5 Incropera. F. P. and De Witt. D. P. in Fundamentals of Heat &
Mass Transfer. 2nd edition. Wiley. New York. USA. IYRS.
chapter5
h Thomas. L. H. C~mnt Pure Marh 1954.7. 195
J . Loss Prev. Process Ind., 1990, Vol3, J anuary 7

Das könnte Ihnen auch gefallen