Blowdown
Blowdown
Blowdown
PII: S0957-5820(19)30262-9
DOI: https://doi.org/10.1016/j.psep.2019.10.035
Reference: PSEP 1972
Please cite this article as: Shafiq U, Shariff AM, Babar M, Azeem B, Ali A, Bustam MA, A
review on modeling and simulation of blowdown from pressurized vessels and pipelines,
Process Safety and Environmental Protection (2019),
doi: https://doi.org/10.1016/j.psep.2019.10.035
This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.
of
1
CO₂ Research Centre (CO2RES), Universiti Teknologi PETRONAS, 32610 Bandar Seri
ro
Iskandar, Perak, Malaysia
2 -p
Department of Chemical and Materials Engineering, University of Jeddah, Jeddah, Kingdom
of Saudi Arabia
re
lP
*
Corresponding Author. Tel.: +60 5 3687530, (Azmi Shariff)
na
Graphical abstract
ur
Jo
Page | 1
of
ro
-p
re
Abstract
lP
when there is an escape of flammable gaseous mixture that can cause potential fire or
na
explosion. One of the scenarios that causes such accidents is the blowdown process.
important to design optimally to make sure that blowdown valve is according to the
requirements. For the safe use of a pressure relief system, some of the parameters are critical,
Jo
for example, selection of construction material, sizing of relief valves, temperature, and
pressure, etc. There is no literature currently available that discusses all the mathematical
models or simulation tools for optimum design of the blowdown process. This subject matters
because the available models or tools cover different aspects of blowdown process. A
meticulous review is required to present the applications of these models and tools based on
Page | 2
the accidental scenarios. Therefore, this paper critically reviews the models and tools that are
developed purposely to calculate optimum blowdown parameters based on fluid and vessel
conditions. Recommendations are given for the development of new simulation tool to
modeling
Nomenclature
of
BP British Petroleum MSM: Marginal Stability Two-Fluid Model
ro
CCS: Carbon Capture and Sequestration PR: Peng & Robinson
CFD:
EoS:
Computational Fluid Dynamics
Equation of State
PRV:
QRA:
-p Pressure Relief Valve
Equilibrium
Page | 3
MOC: Method of Characteristics
1 Introduction
Air-tight pressurized vessels and pipelines are widely used to transport or hold gases and
process liquids or fluids in petrochemical plants, refineries, and process industries. These
vessels are generally subjected to high-pressure loading and unidentical external or internal
hazardous and dangerous because of characteristic operational pressure of the vessel [2]. The
of
rupture of pipeline and subsequent eruption of fire broke down the platform into seabed
resulting in a loss of 167 lives during the Piper Alpha incident [3]. Many other tragedies have
ro
been witnessed that caused a significant number of fatalities due to pipeline and vessel
-p
rupture [4, 5]. After the Piper Alpha tragedy, the offshore industry carried out a great deal of
research work to understand the characteristics of hydrocarbon fires and explosions [6, 7].
re
The results indicate that a pressure above 4 bar developed in typical topsides of the vessels is
enough to inflict the structural damage such as weld failure, local instability, or extensive
lP
inelasticity.
na
The National Board of Boiler and Pressure Vessel Inspectors recorded a 24 % increase in
the number of accidents involving pressure vessels during 1999-2000. Steam heating, power
boilers, and unfired pressure vessels were the major sources of such accidents [8]. From 1992
ur
to 2001, a total of 23,338 accidents related to pressure vessels were recorded [8-10].
Jo
Similarly, according to Swedish Nuclear Power Inspectorate [11], more than 66 % (2476) of
the pipeline failure incidents recorded during 1994-1999 were triggered due to leakage or
puncture. Common reasons for pressure vessel failures are illustrated in Figure 1.
Page | 4
4% 8%
Design 2%
Misapplication 23%
Installation
Maintenance
2%
Repair
Operation
61%
External
of
In oil and gas platforms or offshore operations, overpressure of pressurized vessels also
ro
arises during emergency situations due to fire or valves' malfunction. These industries usually
work with CO₂, natural gas, and fossil fuels [12, 13]. Most of the accidents at offshore oil and
-p
gas industries, nuclear power plants, and chemical plants cause the outflow of radioactive,
re
toxic, explosive, or flammable materials. Particularly in the oil and gas industry, failure or
rupture of pressurized vessels poses higher risk due to the possible escape of flammable
lP
gaseous mixture in a fire event. Therefore, process vessels are depressurized for the
17]. Blowdown is a typical way of minimizing the failure hazard of vessels when an
emergency arises in a process industry. However, for the safe use of a pressure relief system,
ur
some of the parameters are crucial; for example, temperature, pressure, selection of
construction material, and sizing of relief valves. Prediction of precise minimum vessel wall
Jo
temperature during depressurization process can affect the selection of construction material,
reduce over-design, and subsequently lower the project cost. Similarly, over-design can also
Page | 5
of
Figure 2: Important parameters for safe blowdown process
ro
In recent years, safety and process engineers have encountered several safety-related
issues while dealing with high-pressure blowdown. The key principle of sustainable
-p
depressurization process design is to decrease the pressure and discharge the inventory safely
re
in the minimum possible time [18]. The discharge rate of compressible fluids generally
depends upon pressure, pressure drop, temperature, friction factor, Reynolds number,
lP
pipeline diameter, roughness, length, and gas properties [19]. However, the rapid expansion
of gas and generation of vapors can expose the pressure vessel to pressurized thermal shock
na
during an emergency blowdown process. This shock initiates due to integrated stress from
rapid transitions in pressure and temperature that result in the non-uniform distribution of
ur
temperature in vessel walls and, therefore, lead to the differential contraction and expansion.
Pressurized thermal shocks can also cause embrittlement of vessel walls and subsequent
Jo
“failure by fatigue” due to very low temperature generated inside the vessel because of Joule-
design a sustainable blowdown process, there is dire need of a technique to predict the mass
efflux. Moreover, to calculate optimum blowdown time, a precise balance between maximum
acceptable depressurization time [20] and minimum fluid and wall temperature that may be
Page | 6
safely considered, is required. Some of the important factors to be considered before
of
ro
Figure 3: Factors to be considered before designing a sustainable pressure relief system
-p
Conventionally, several numerical models and simulation tools are reported in literature
re
for the prediction of maximum possible blowdown time and fluid and vessel conditions. The
numerical models have been developed by commonly adopting the Homogenous Equilibrium
lP
Model (HEM) assumption such that each phase is considered to be at the phase equilibrium
and thermal equilibrium. Some of the researchers also coded simulation tools based on their
na
numerical models. However, the available mathematical models or simulation tools cover
the existing correlations, numerical models, and computational tools that have capability to
Jo
calculate optimum blowdown parameters for pressure vessels and pipelines, is presented in
this paper to summarize existing work and set out most overlooked aspects.
In case of blowdown of vessels, a hazard may emerge due to very low temperature gained
by the material inside the process vessel. This can potentially lead to rupture even with a
Page | 7
small amount of force if temperature falls below the ductile-brittle transition temperature of
vessel's material of construction [21]. If free water is present inside vessel, it can result in
growth of hydrates or formation of liquid condensate that can get carried-over into the flare
or vent system. If large droplets of liquid condensate or continuous liquid stream gets into the
flare, they will start falling from the top covering entire flare into flames. This condition is
extremely dangerous as this falling and burning liquid may cause a major fire incident in
vicinity of the plant [22]. To date, several numerical models and simulation tools have been
developed that can simulate different blowdown scenarios of vessels based on vessel sizing,
of
blowdown conditions, and compositions etc. A timeline of these simulation tools or
ro
numerical models is illustrated in Figure 4 followed by a brief description (Some of the
authors didn't name their simulation tools; therefore, we have mentioned the name of the
Figure 4: Timeline for blowdown simulation tools and numerical models for vessels
ur
2.1 DIERS
A consortium of several companies founded DIERS program in 1976 to design pressure
Jo
relief system for runaway reactions and develop the additionally required technology. DIERS
tool, and designed a bench-scale prototype apparatus. The aim was to predict two-phase flow
venting and analyze the possible applicability of various two-phase (vapor-liquid) flashing
flow methods for sizing of relief systems. In this model, it is considered that boiling takes
Page | 8
place through the entire volume of liquid rather than solely at the surface. Each bubble
occupies volume and displaces the liquid surface upward. Individual bubbles are able to rise
(slip) through the liquid with a velocity that depends on buoyancy and surface tension,
whereas they are retarded by viscosity and foamy character of the fluid [23]. DIERS program
evolved over time; the initial phase included study of vapor-liquid phase dynamics, second
phase involved the experimental investigations for small and large scale integral blowdown
and vented runaway reaction, and the final phase included the development of a coded
computational package along with a bench scale prototype experimental apparatus [24].
of
2.2 BLOWDOWN
ro
A brief description of prediction and experimentation of fluid and vessel behavior during
blowdown process was presented by Haque et al. in 1990 [25]. Experiments were performed
-p
on blowdown of pressurized vessels containing pure N₂, 70 % N₂–30 % CO₂, and natural
of vessels. The developed BLOWDOWN package has the ability to measure the temperature,
For the experimental investigation, a 6.325 mm orifice was used to blowdown the mixture
from a 0.086 m³ vessel with 25 mm of wall thickness. Pure N₂ took almost 100 s to reach the
ur
atmospheric pressure from 15 MPa. Whereas, the vessel containing 70 % N₂–30 % CO₂
binary mixture reached the atmospheric pressure in 60 s. The dynamic change in gas and wall
Jo
determination, R2, and geometric mean bias of the model data are calculated. For this model,
geometric mean bias of different models is presented in Figure 33 and Figure 34 respectively.
Page | 9
295
(a)
275
Temperature/K
255
235
215
195
175
0 10 20 30 40 50 60 70 80 90 100
Time/s
0 0
(b)
of
-20 -5
Wall temperature/K
Gas temperature/K
-40 -10
ro
-60
-80
-100
-120
-p -30
re
0 20 40 60 80 100
Time/s
lP
Figure 5: Dynamic temperature changes during blowdown for (a) 70 % N₂–30 % CO₂
mixture and (b) pure N₂ [25]
Subsequently, an extension in the simulation tool was made that can also be used to
na
simulate the blowdown of pipeline, particularly a long subsea pipeline. Extension of this
work by Haque et al. in 1992 covers simulation study of free water (produced during
ur
depressurization) for the arbitrary combinations of vessels and pipelines [26]. This extended
version of BLOWDOWN tool was validated again in 1992 by Haque et al. while supported
Jo
by several case studies [27, 28]. In the first case study; 53804 m long, 16.5″ subsea gas line
containing 51 MMSCF methane at 283 K was blown down from a pressure of 12 MPa. In the
second case study, full-bore depressurization of 2000 m long pipeline containing 1975 tonnes
of methane condensate was carried out. The initial pressure inside the pipeline was 20 MPa
Page | 10
Another comparison of BLOWDOWN predictions and experimental data was made
during LPG test performed by British Petroleum (BP) and Shell on the Isle of Grain in 1985
BLOWDOWN package are also available in literature [32]. In all the experimental studies,
occurring and an ability to make predictions with adequate certainty. However, use of
extended corresponding states principle introduces uncertainties that make the simulation
computationally difficult. Moreover, the limitations of a cubic equation of state (EoS) near or
of
at the critical region are well documented but their effect on the accuracy of blowdown
ro
simulation is still unknown. Further improvement and simulation of practical problems imply
the assumptions of homogeneity when the flow is two-phase and the quasi-steadiness could
be neglected.
-p
re
2.3 FRICRUP
A simple mechanistic model named FRICRUP was coded by Norris et al. in FORTRAN
lP
program to predict single and multi-phase flow behavior during blowdown of vessels or
pipelines [33]. HEM and thermodynamic equilibrium model (TEM) assumptions along with
na
no relative velocities between liquid and vapor phases were taken into account. Moreover,
FRICRUP tool was validated with the available data for gaseous mixture of CO₂, carbonated
ur
water, and air. The comparison shows that a reasonable mechanistic model has been
developed that can predict single and multi-phase blowdown from the vessel as well as the
Jo
pipe. Regardless of its complexities, the mechanistic model indicates disagreement with
multi-phase flow possibly due to TEM assumptions. A year later, Norris et al. conducted
another experimental study using HC gases [34]. The results were similar to the ones
achieved while using the non-HC gases because the basic mechanical model assumptions
Page | 11
2.4 BLOWSIM
BLOWSIM model [14, 35], developed by Mahgerefteh and Wong is based on three
cubic-EoSs including Peng and Robinson (PR) [36], Soave-Redlich-Kwong (SRK) [37], and
pressure, mass flow rate, and fluid and wall temperature, all as a function of time.
1. Heat transfer between fluid phases and their corresponding sections of the vessel wall.
of
2. Non-equilibrium effect between phases.
ro
4. Interphase fluxes because of condensation and evaporation.
-p
The performance of the featured model was assessed by the correlation of predicted data
for liquid and vapor phases as well as unwetted and wetted walls. The achieved results and
lP
their comparison with different EoSs as well as the experimental data are illustrated in
Figure 6 and 7. BLOWSIM's predictions are quite insensitive to the Cubic-EoS used. In
na
general, it can be inferred that choice of the Cubic-EoS has a negligible effect on predicted
results. Moreover, both numerical models are capable of reasonably predicting the dynamic
ur
pressure variations. The R-square noted for TCC (TWU) model for temperature predictions is
≈ 0.992.
Jo
Page | 12
12
Exp.
BLOWDOWN
10
SRK
PR
8
Pressure/MPa
TCC
0
0 200 400 600 800 1000 1200 1400
of
Time/s
Figure 6: Correlation between predicted and experimental pressure-time profiles [14]
ro
295
Exp.
290 BLOWDOWN
Bulk vapor temperature/K
285
280
-p SRK
PR
TWU
275
re
270
265
lP
260
255
250
245
na
240
0 300 600 900 1200 1500
Time/s
ur
Figure 7: Correlation between predicted and experimental vapor temperature profile [14]
2.5 PHAST
Jo
The Unified Dispersion Model (UDM) developed by Woodward and Cook in the early
90's was later implemented into DNV software package, termed as PHAST in version 6.0
[39-41]. PHAST is a helpful tool to study the initial release of mixture as well as the far-field
dispersion. Previously, PHAST was limited only to vapor and liquid phase release of the
chemicals. However, the upgraded version allows to study the occurrence of fluid to solid
transition during release of CO₂. PHAST was validated against several experimental studies
Page | 13
for unpressurized release of CO₂ [39, 42]. Later, Witlox et al. performed some more
experimental studies for the validation against pressurized release of CO₂ [43]. An overview
of the validation of different models including flammable effect, dispersion and pool
experimental database for validation of the above models and scenarios [44].
2.6 P. S. Cumber
A numerical model was coded by Cumber for the prediction of vent sizing of pressurized
vessel during blowdown batch process [45]. The pressure vessel containing gas mixture is
of
considered as a single unit with thermodynamic equilibrium assumptions. This model has a
ro
vast range of venting models implemented for a two-phase system. However, selection of the
most suitable model for a specific case study necessitates the user to have a comprehensive
-p
information on the scope of application of the respective model. For this particular case,
predicted blowdown data was plotted against the experimental data as shown in Figure 8 and
re
9. The trend indicates that predicted gas temperature in the initial phases of depressurization
lP
temperature with an R-square ≈ 0.445 because of the adiabatic vessel wall assumption as
na
shown in Figure 9. The significant difference is with respect to the phase temperature
measurement that shows that thermodynamic equilibrium assumption is not effective for this
ur
case. However, the measured pressure of vessel was precisely predicted. For the vessels
depressurizing over long time-scales, the temperature of the vessel was under-predicted after
Jo
first few minutes as a consequence of assuming no heat transfer through the vessel walls.
This did not have a substantial effect on the simulated mass flow rate though. Another
discrepancy with the application of this model is the robustness that needs to be addressed
[46].
Page | 14
14
Pre.
12 Exp.
10
Pressure/MPa
0
0 200 400 600 800 1000 1200 1400
Time/s
of
Figure 8: Dynamic pressure profile during blowdown of LPG [45]
ro
310
Pre.
300
Exp. LIQUID
290
280
-p Exp. GAS
Temperature/K
270
re
260
250
lP
240
230
220
0 200 400 600 800 1000 1200 1400
na
Time/s
2.7 A. Fredenhagen
Previously accomplished research work on depressurization of pure CO₂ leads towards
Jo
successful modeling of the pressure transitions and axial void profile during the expansion
pressurized vessel containing CO₂ with impurities (N₂) was performed by Fredenhagen and
Eggers in 2001 to study the effect of impurities [51]. In this study, the axial void profiles,
axial temperature profiles, and pressure and mass were measured. Top vented cylindrical
Page | 15
vessel with 0.05 m³ volume, 0.242 m ID, and a 4:5 height to diameter ratio was used. The
venting line was linked to a quick opening ball valve with a cross-sectional area of 50 mm².
PR EoS was used for the calculation of phase equilibrium. For the prediction of this process,
a simulation tool was also coded based on the energy and mass balance and a drift flux
model. The dynamic pressure and temperature profiles predicted by featured simulation tool
and the experimental results are illustrated in Figure 10 and 11. The computational model
predicted reasonable results of pressure transition throughout the process. The overall R-
square for temperature predictions is ≈ 0.955, however, the temperature predictions at low
of
pressure were not much precise. Consequently, an improvement needs to be made for more
ro
precise modeling of thermodynamic behavior of the mixture.
14
Pre.
12
10
-p Exp.
re
8
Pressure/MPa
6
lP
2
na
0
0 10 20 30 40 50 60 70 80
Time/s
ur
Figure 10: Calculated and measured pressure transients for 17 mm² orifice [51]
Jo
Page | 16
300
Pre.
290 Exp.
280
Temperature/K
270
260
250
240
230
0 20 40 60 80 100 120 140
of
Time/s
ro
2.8 H. Mahgerefteh
Most comprehensive blowdown model reported in literature was BLOWDOWN, later on
-p
extended as BLOWSIM as discussed earlier. Although validated extensively with the
re
experimental results, the prediction is still limited to depressurization under ambient
temperature. This is a serious drawback since the blowdown situations may involve a fire
lP
case most of the time. To overcome this issue in 2002, Mahgerefteh et al. described the
following the depressurization of vessel under fire [52]. The model is able to predict the tri-
axial transient pressure and thermal stress produced in both unwetted and wetted wall
ur
sections of the vessel. The fluid/wall heat transfer was obtained using the following equation.
Where ℎ is the heat transfer coefficient, 𝑇𝑤′ (𝑎, 𝑡) refers to the unknown wall temperature
at time 𝑡, 𝑇𝑍 is the liquid or vapor temperature, 𝐴 is the heat transfer area and 𝑛 stands for
power index. A hypothetical case study for the depressurization of a vessel containing higher-
HC was performed. The vessel was supposed to be at the initial pressure and temperature of
11.6 MPa and 293 K respectively. Two types of blowdown scenarios were simulated: (1) the
Page | 17
vessel under ambient conditions and (2) the vessel under fire radiating a heat flux of 90
kW/m². The predicted pressure and temperature profiles for a vessel under ambient and fire
conditions are illustrated in Figure 12. The predicted results show that thermal radiation
stress and triaxial thermal profiles indicates that rupture occurs in dry section of vessel
because of the combined effect of resident total stress and thermally induced mechanical
weakening of vessel walls. The featured numerical model serves as a powerful tool for the
calculation of minimum required blowdown time to eradicate the risk of vessel puncture
of
under fire.
16 650
ro
Pressure (FC) FC = Fire Condition (Temperature)
14 Pressure (AC) AC = Ambient Condition (Temperature) 600
Vapor temperature (FC)
12 550
Vapor temperature (AC)
-p
Temperature/K
Pressure/MPa
2 250
0 200
0 200 400 600 800 1000 1200 1400
na
Time/s
Figure 12: Temperature and pressure histories under ambient and fire conditions [52]
ur
2.9 SDM
Over the years, the researchers' primary focus has been on the performance of pressure
Jo
relief valve (PRV) using dynamic modeling approach. Unfortunately, most of the processes
were simulated for dynamic properties of PRV at the fixed inlet scenarios and not during the
reclosing process [53]. To predict the dynamic properties of PRV during reclosing process, a
simplified dynamic model (SDM) was presented by Song et al. [54]. When the disc moves,
PRV behaves as one degree-of-freedom system and motion of the moving part (stem and
Page | 18
disc) was resolved based on force balance when open. The motion of value was simulated
based on second-order differential equation deduced from Newton's second law (Eq. 2).
𝑚ẍ + 𝑐ẋ + 𝐹𝑠 = 𝐹𝑓 − 𝐹𝑔 + 𝑓𝑐 (2)
Where 𝑚, ẍ, and ẋ represent the mass of the moving parts, acceleration, and velocity of
the disc part in the moving direction respectively. 𝑐 represents the damping coefficient and
𝐹𝑠 stands for the spring force acting on the disc that was calculated using 𝑘(𝑥0 + 𝑥(𝑡)). 𝐹𝑓 ,
𝐹𝑔 , and 𝑓𝑐 are the flow force acting on the disc part, the gravity of disc, and the coulomb
of
motion was calculated using Eq. 3.
ro
𝑚ẍ/𝜓 + 𝑐ẋ/𝜓 + 𝑘(𝑥0 + 𝑥)/𝜓 = (𝜋/4). 𝐶𝑓 . 𝜙 (3)
Where 𝑘 is the spring stiffness. 𝜓 and 𝜙 are the reference force and pressure difference
-p
ratio. The dynamic numerical model was based on the major principles of the steady
re
computational fluid dynamics (CFD) analysis and rigid-body motion of the PRV. Afterwards,
the results from a case study revealed that SDM predicted the blowdown of conventional
lP
PRV reliably as shown in Figure 13. In addition, the effect of reclosing/opening process and
spring stiffness can also be easily studied using this dynamic numerical model.
na
5.25
4.75
4.25
ur
2.75
Jo
2.25
Popping
1.75
1.25
0.75
0.25
-0.25
17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34
Time/s
Figure 13: Comparison of experimental results and SDM [54]
Page | 19
2.10 J. Zhang
To Simulate the blowdown process for the supercritical water-cooled reactor (SCWR), a
transient analysis code using a comprehensive numerical model was developed by Zhang et
al. in 2013 [55]. Three different phase separation models were implemented for the
was used to calculate subcooled, supercritical, and overheated fluid by neglecting phase
separation effect. Secondly, assuming that bubble rising and blowdown periods are
comparable and ignoring neither discharge rate nor the bubble rising velocity, bubble rising
of
model describes this process properly. Thirdly, a complete separation model is considered
suitable for conditions when bubble rising velocity is faster as compared to the discharge
ro
speed and when the droplets falling velocity is higher than discharge speed for condensation
-p
because there is enough time for vapor dome development in both the cases.
Supercritical CO₂ and H₂O were blown down initially from 313 K temperature and 25
re
MPa pressure for experimental investigation and verification of the developed code. The
lP
mm² orifice for depressurization. The experimental study and predicted pressure drop results
with different models as shown in Figure 14. The R-square for pressure predictions is ≈ 0.377
na
by complete separation model. The consequence of the initial condition on pressure drop was
considered for different regions divided by the relationship between corresponding pseudo-
ur
critical and initial temperature. Furthermore, it was concluded that both void fraction and
Jo
blowdown speed increase with the decrease in initial pressure and an increase of initial
Page | 20
20
Initial Conditions: Exp.
18
P₀ = 20 MPa Homogeneous model
16 Bubble rising model
T₀ = 313 K
14 A₀ = 17 mm² Complete separation model
Pressure/MPa
12
10
8
6
4
2
0
0 20 40 60 80 100 120 140
of
Time/s
Figure 14: Dynamic pressure comparison of different models [55]
ro
2.11 VBsim
To predict the depressurization behavior of vessel containing two-phase (V-L) HC
-p
mixture with non-ideal behavior, Alessandro et al. described the modeling and experimental
validation of the new unsteady model, VBsim (Vessel Blowdown Simulator) [56]. The
re
featured model accounts for non-equilibrium effects between constituent fluid phases such as
lP
temperature difference between phases. The numerical model evaluates dynamic fluid
temperature, pressure, and composition using EoS for non-ideal gas. SRK and PR EoS with
na
Van der Waals mixing rules have been implemented. Vapor phase discharge (𝜓𝐺,𝑜𝑢𝑡 ) was
estimated using equations of motion for compressible fluid through the orifice (Eq. 4 and 5).
ur
𝜓𝑖𝑠 = { 𝛾/(𝛾−1)
(5)
𝑆𝐺,𝑜𝑟 ɼ1/𝛾 √2𝛾/(𝛾 − 1)𝑝𝜌𝐺 (1 − ɼ(𝛾−1)/𝛾 ) 𝑝/𝑝𝑒𝑥𝑡 < (2/(𝛾 + 1))
𝐺
Where 𝑀𝑊 is the vapor phase molar weight and 𝐶𝐷𝐺 is fixed to a default value of 0.84. 𝜌𝐺
is the vapor density, 𝑆𝐺,𝑜𝑟 is the orifice area, 𝑝 is the vessel pressure, 𝑝𝑒𝑥𝑡 is the atmospheric
pressure, ɼ is the ratio of 𝑝 & 𝑝𝑒𝑥𝑡 , and 𝛾 is the isentropic coefficient of the real gas. The
Page | 21
equation for an incompressible fluid through an orifice was used for the calculation of liquid
Where 𝑆𝐿,𝑜𝑟 , 𝜌𝐿 and 𝐻𝐿 are the orifice area, liquid density, and liquid static head
𝐿
respectively. The 𝑀𝑊 stands for liquid phase molar weight. The liquid discharge coefficient,
𝐶𝐷𝐿 , is fixed to a default value of 0.61. The liquid and the vapor daughter phase compositions
of
𝑥𝑖 = 𝑧𝑖 /(1 − 𝛽 + 𝛽𝐾𝑖 ’) (8)
ro
𝑦𝑖 = 𝐾𝑖 𝑧𝑖 /(1 − 𝛽 + 𝛽𝐾𝑖 ) (9)
Where 𝐾𝑖 is the initial equilibrium coefficient and the overall vapor fraction, 𝛽, is
-p
calculated by solving the Rachford–Rice equation. For experimental validation of the
re
developed model, the predicted results were plotted against the experimental data from
Haque et al. (PR and SRK) and HSE (SRK) [26, 28, 57]. A comparison with the previously
lP
defined blowdown model (BLOWDOWN) prediction was also made as shown in Figure 15.
Considering the overall performance, VBsim produced reasonable results with R-square ≈
na
0.682 for temperature predictions and less CPU time requirements. However, the impact of
the model could be significantly improved if compared with further experimental data.
ur
Jo
Page | 22
320
Vbsim
310
BLOWDOWN
300 Exp. UP
290 Exp. DOWN
Temperature/K
280
270
260
250
240
230
220
0 200 400 600 800 1000 1200
Time/s
of
Figure 15: VBsim and BLOWDOWN predictions vs experimental data (S9) [56]
ro
2.12 A. Park
Previously, Haque et al. and Mahgerefteh et al. developed models to study heat transfer
-p
between the mixture and the vessel walls. However, the correlation used and the value of the
heat transfer coefficient were unclear. Therefore, Park et al. (2018) presented an elaborated
re
heat transfer model including combined convection (forced and natural), multi-layer transient
conduction, and the nucleate boiling using PR EoS [58]. Mass and energy conservations for
lP
the fluid were calculated using equations presented by Speranza and Terenzi [59]. The model
consists of sub-algorithms of flash calculation, discharge rate calculation, and heat transfer
na
rate calculation in order to estimate the realistic expansion path. The predicted results were
compared with the measured and simulated data from literature including renowned
ur
simulation tools such as BLOWDOWN, BLOWSIM, VBsim, BLOW, and commercial tools
Jo
such as VessFire 1.2 and Aspen HYSYS v9. The model reasonably predicted the vapor
temperature within the experimental range while BLOWDOWN, VBsim, and HYSYS
predicted 2–10 K higher temperature. Similarly, the model also precisely predicted the wall
temperature within the experimental range while VessFire v1.2 and BLOW predicted almost
2–8 K higher wall temperature in contact with vapor, and VBsim predicted 10 K higher wall
Page | 23
300
Vapor
Inner Wall Temperature/K
290
280
270
Exp. Liquid
260
Park et al
BLOWDOWN
250 HYSYS v9®
BLOW
240
0 200 400 600 800 1000 1200
of
Time/s
Figure 16: Vessel wall temperature histories and comparison with other simulation tools and
numerical models [58]
ro
3 Design of Blowdown System for Pressure Pipelines
-p
The blowdown of a pipeline and a vessel differs from each other due to a considerable
pressure difference within the vessel but not within the pipeline. For blowdown of a vessel,
re
pressure drop is across the orifice, whereas for a pipeline blowdown, pressure drop is along
lP
the line and across the orifice if it is small enough. In case of blowdown of pipelines, the
hazard emerges not only following the low temperature generated in the pipeline walls but
na
also due to the high efflux rates and large total efflux that arise when a massive inventory in a
typical line is depressurized [22, 60]. To avoid accidents caused by these issues, several
numerical models have been developed. Later, some of them have also been coded as
ur
computerised program usually known as simulation tools. A timeline for the available
Jo
simulation tools and numerical models capable to simulate the depressurization process is
illustrated in Figure 18 followed by a brief description (Some of the authors didn't name their
simulation tools; therefore, we have mentioned the names of those authors instead of the
simulation tool).
Page | 24
1995 1999 2011 2013 2014
of
3.1 META
Marginal stability two-fluid model (MSM) based on extended Geurst's variational
ro
principle to generalize multi-component two-phase dispersion was presented by Chen et al. in
-p
1995 [61-64]. Thermodynamic equilibrium assumptions were made to develop the equations
profound, the flow was considered to be marginally stable to achieve hyperbolicity of the
lP
equations.
In the later volume of publication, the simplified numerical method developed by Chen et
na
al. in 1993 [65] was modified and applied to the MSM for multi-component mixtures [66].
Subsequently, MSM and HEM were incorporated on the basis of simplified numerical
ur
phase Analyser) was coded. A comparison of different case studies was made between HEM
and the MSM with experimental data for single or binary component two-phase blowdown. A
100 m long horizontal commercial steel pipeline with 0.15 m inner diameter (ID) containing
LPG (95 mole % propane and 5 mole % butane) was blown down. The testing temperature
and pressure were kept between 288–293 K (ambient temperature) and 800–2100 kPa
respectively. The bore size was varied from 0.05 m to full-bore in diameter and controlled
Page | 25
with the circular orifice plate. Figure 18 and 19 show the history of the pressure and
temperature variations with respect to time at the closed ends. An R-square ≈ 0.957 was
1200
Exp.
1000 META-HEM
META-MSM CS
800
Pressure/kPa
600
400
of
200
ro
0
0 5 10 15 20 25
Time/s -p
Figure 18: Dynamic pressure changes' histories at the closed end [65]
re
300
Exp.
290 META-HEM
lP
280 META-MSM CS
Temperature/K
270
260
na
250
240
ur
230
220
0 5 10 15 20 25
Jo
Time/s
Figure 19: Dynamic temperature changes' histories at the closed end [65]
The comparison of HEM predictions and the measured data has been exceptionally
reasonable. All the predicted temperature, pressure, inventory, and void fraction transitions
correspond with the measured data. The only peculiarity was in case of the void fraction that
Page | 26
was over-predicted after 7 s at the closed end. The most significant findings are presented as
follows.
1. The simplified numerical method predicted reasonable results for blowdown through
2. In case of a full-bore short pipeline blowdown (< 10 m), both mechanical and thermal
of
important for the transient flow calculations.
ro
3.2 Fairuzov
Previously, most of the work encompassed the modeling of depressurization from vessels
-p
or relatively short pipelines [67-69]. Therefore, a numerical simulation model for the
re
depressurization of relatively large pipeline transporting multi-component flashing mixtures
was coded by Fairuzov in 1998 [70]. The effect of heat transfer between fluid flow and
lP
pipeline wall during blowdown phenomenon was also described. Heat transfer model, the
hydrodynamic model, and break-flow model were the building blocks of the featured model.
na
No experimental study was conducted, however, the simulation tool was validated against the
already available experimental data from BP and Shell on the Isle of Grain [30]. The
ur
calculations were carried out on an HP Apollo Series 700 working station using a double-
Fairuzov proposed a higher amount of heat transfer amongst fluid and pipe walls during
the depressurization process, hence, the adiabatic assumption for simulation became invalid.
Therefore, effect of thermal capacitance was integrated into the model using a different
method in the development of energy conversion equation. The study unveiled considerable
Page | 27
Consequently, thermal capacitance should not be ignored especially in case of the flashing
fluids. The dynamic pressure and temperature profiles are reported in Figure 20 and 21
respectively. Figure 20 illustrates that the model slightly over-predicted depressurization rate
at the closed end of pipe. However, when the fluid flow was considered to be adiabatic, the
temperature at the open end was significantly predicted with R-square ≈ 0.914.
1200
Closed end (Exp.)
Closed end (Pre.)
1000
Open end (Exp.)
Open end (Pre.)
800
of
Pressure/kPa
600
ro
400
200 -p
0
0 5 10 15 20 25
re
Time/s
Figure 20: Dynamic pressure transitions at the open and closed ends of the line [70]
lP
300
290
na
280
270
Temperature/K
260
ur
250
Closed end (Exp.)
240
Closed end (Pre.)
Jo
Page | 28
3.3 CNGS-MOC
An efficient mathematical model (CNGS-MOC) for the simulation of full-bore rupture
(FBR), based on the method of characteristics (MOC) was developed by Mahgerefteh et al. in
1999 [71]. The MOC was implemented to simulate FBR or depressurization of a long pipe
containing two-phase or condensable HC mixtures. This method was used to develop a more
suitable method than the Finite Difference Method (FDM) and Finite Element Method (FEM)
because both had troubles in handling the choking scenarios at the ruptured end. Moreover,
the MOC is more precise than FDM and FEM because it can resolve the blocked flow
of
intrinsically using Mach line characteristics. The MOC significantly reduces number of
iterations involved while solving the system of simultaneous equations. The use of curved
ro
characteristics in conjunction with fast-mathematical algorithm and compound nested grid
-p
system leads to a dramatic reduction in CPU time in addition to improved accuracy. The
model was extensively validated using field data obtained during Isle of Grain [29] as well as
re
Piper Alpha tragedy [72]. The simulation studies were performed on the basis of HEM
assumptions. In Figure 22, the predicted data with CNGS-MOC and field data are correlated
lP
with those based on other simulation tools including BLOWDOWN [25], META-HEM [64,
66, 72], MSM-CS [64, 66, 72] as well as PLAC [73]. Both META-HEM and CNGS-MOC
na
produced reasonably precise predictions while other models' predictions were relatively poor.
The CNGS-MOC predicted the physical process with R-square ≈ 0.850 for pressure
ur
predictions.
Jo
Page | 29
1200
Exp.
CNGS-MOC
1000
META-HEM
MSM-CS
800
Pressure/kPa
BLOWDOWN
PLAC
600
400
200
0
0 5 10 15 20 25
of
Time/s
Figure 22: Dynamic pressure histories at the open end for the P42 test LPG [71]
ro
3.4 A. Oke
Regardless of the fact that most common pipeline tragedies include rupture, majority of
-p
the reported models so far are based on 1-D axial flow. Others just treat the pipeline as a
vessel discharging through an orifice. Therefore, Oke et al. developed a model that can
re
predict the transient release rate through the punctured plane based on MOC [74]. The model
lP
accounts for axial and radial flow and real fluid behavior as well as the rupture locality with
respect to pipeline length. The PR EoS was employed to get suitable phase and
na
In case of 1-D unsteady flow, assuming phase and thermodynamic equilibrium among
phases, the energy, momentum, and continuity conservation equations for a fluid in a rigidly
Jo
clamped pipeline were taken from literature [77]. The continuity equation was reformulated
as follows.
Where ℎ and 𝜑 represent enthalpy and thermodynamic function respectively. 𝜌 and 𝑢 are
density and velocity as a function of time (𝑡) and distance (𝑥) respectively. Using MOC, the
Page | 30
conservation of continuity, energy, and momentum equations can be substituted with three
compatibility equations along with their corresponding characteristic lines. The isolated flow
was modeled based on termination of the inlet feed upon rupture. In case of non-isolated
flow, inlet feed is supposed to be stopped after a prescribed period of time. Figure 23 is
schematic illustration of the fluid flow analysis following rupture and emergency isolation
Valve End
of
ro
Pump End
Control
Volume
-p
Figure 23: Rupture plane and fluid flow study following line puncture [74]
In Figure 23, C stands for match line, C0 stands for path line, and J represents solution
re
joints. The simulation of pressure and velocity profiles specifies three separate fluid flow
regimes predominant within the pipe that manage the dynamic transitions of the release rate.
lP
For validation of the numerical model, the simulated outcomes were plotted against the data
from experiments performed by BP and Shell on the Isle of Grain [30]. The results ended up
na
with a reasonable prediction of the physical process with R-square ≈ 0.942 for temperature at
Page | 31
1200
Exp. (closed end)
1000 Pre. (closed end)
Exp. (opened end)
800
Pressure/kPa
600
400
200
0
0 5 10 15 20 25
of
Time/s
Figure 24: Pressure transitions at closed and opened end of P40 (LPG) test [74]
ro
300
290
280
-p
Temperature/K
270
re
260
250 Exp. (closed end)
lP
220
0 5 10 15 20 25
Time/s
Figure 25: Temperature transitions at closed and opened end of P40 (LPG) test [74]
ur
3.5 VPM
Until 2011, no numerical model was developed to analyze the blowdown conditions for a
Jo
vent pipeline. Therefore, Rajiwate developed a vent pipe model (VPM) to investigate the
behavior of compressible gas in a vent pipeline using thermodynamic and fluid dynamic
approach [78]. Integrated with the GERG EoS, the use of REFPROP makes a highly adequate
Page | 32
Temperature, pressure, and density relations in terms of stagnation properties were
defined previously [81-83]. Mass efflux prediction for pipe is the central step for the
development of blowdown model for vent system in terms of static pressure and static
𝐺 = 𝑃𝑀√𝛾/𝑍𝑅𝑇 (11)
The above equation for mass flux in terms of stagnation properties was represented as:
Where 𝑇𝑜 , 𝑃𝑜 are the stagnation temperature and pressure respectively, 𝛾 is the specific
of
heat ratio, and 𝑍, 𝑅 and 𝑀 are the compressibility factor, universal gas constant, and match
ro
number respectively. For a given downstream and stagnation pressure, geometry, and
initial pressure and temperature of 200-400 kPa and ≈ 19 °C respectively. The experimental
lP
data were plotted only for a steady-state period that showed a good agreement with the
predicted values.
na
3.6 OLGA
Most of the reported data on experimental study of CO₂ blowdown [49, 85-88] describes
ur
findings from small-scale laboratory setups of flow loops and vessels. Thus, it is of great
interest for carbon capture and sequestration (CCS) community to collect field data from
Jo
existing large-scale CO₂ facilities. Therefore, a large-scale CO₂ pipeline blowdown was
carried out by Clausen et al. in 2012 while pointing the possible limitations of existing
dynamic multi-phase flow simulator (OLGA) [89]. An onshore buried 50 km long 24″ ID
pipeline containing 93000 tonnes of CO₂ (>99 % Pure) was initially depressurized from
supercritical conditions. The initially reported temperature and pressure were 304 K and 8.1
Page | 33
MPa respectively. The blowdown process was modeled using OLGA V5.3.2, a multi-phase
thermo-hydraulic simulation tool developed by SPT Group [90]. The new coded module of
OLGA for CO₂ single component was used that utilizes Span-Wagner EoS for pure CO₂ [91,
92].
The total blowdown time was noted to be nearly 10.50 h while OLGA predicted duration
of 10 h and 25 min. Both simulation and experimental data showed that the blowdown stayed
well above the CO₂ triple point, and no traces of solidification were found as shown in Figure
26. However, the minimum simulated temperature was almost 17 K lower than the minimum
of
measured temperature. The deviation between simulated and measured blowdown path
ro
reveals inability of the simulation tool to handle CO₂ along with impurities. Even a very
9
8
CO₂ VLE
Exp.
-p
re
7 Pre.
Phase Envelop
6
lP
Pressure/MPa
5
4
3
na
2
1
0
250 260 270 280 290 300 310 320
ur
Temperature/K
Figure 26: Measured upstream pressure and temperature [89]
Jo
To further study the effect of impurities on CO₂ blowdown and reliability of the OLGA
tool, Huh et al. conducted experimentation in 2014 with 2–8 % N₂ addition by mass fraction
[93]. The experimental apparatus comprises of the pressurization part, liquefying part, mixing
zone, and the test section. Total length of the testing apparatus was 51.96 m with 3.86 mm
ID. The temperature and pressure measurements were observed at four different positions as
Page | 34
illustrated in Figure 27 and 28. However, the simulation tool did not predict temperature drop
precisely. The results show that the numerical simulator predicted better pressure drop at
higher concentration of impurities. Furthermore, the R-square value at initial portion (7.93 m)
of pipeline is ≈ 0.697. While, at last portion (51.85 m), the temperature predictions are much
precise with R-square ≈ 0.859. It can be inferred that the numerical simulator requires further
10
7.93 m
9
23.13 m
8
of
38.32 m
7 51.85 m
Pressure/MPa
6 7.93 m
ro
5 23.13 m
4 38.32 m
51.85 m
3
2
1
-p
re
0
-5 0 5 10 15 20 25
Time/s
lP
Figure 27: Pressure transition of simulated and measured data for 92 % CO₂–8 % N₂
blowdown [93]
296
na
294
292
ur
290 7.93 m
Temperature/K
23.13 m
288 38.32 m
Jo
286 51.85 m
7.93 m
284 23.13 m
282 38.32 m
51.85 m
280
-10 -5 0 5 10 15 20 25 30 35 40
Time/s
Figure 28: Temperature changes of simulated and measured data for 92 % CO₂–8 % N₂
blowdown [93]
Page | 35
3.7 S. Brown
The commonly reported vessel/pipeline outflow numerical models were based on HEM
assumptions where the fluid phase was considered to be at the mechanical and thermal
phenomena, that is, delayed bubble formation was ignored. In 2013, Brown et al. extended
the prior work by establishing and validating homogeneous relaxation model (HRM) for
depressurization of dense phase CO₂ pipeline [94]. Initially, the phases were assumed to be at
mechanical and thermal equilibrium, hence, the mass, momentum, and energy conservation
of
were presented by Eq. 13, 14, and 15 respectively.
ro
𝜕𝜌𝑢/𝜕𝑡 + (𝜕𝜌𝑢2 + 𝑃)/𝜕𝑧 = −𝑓𝑤 𝜌𝑢2 /𝐷𝑝 (Momentum Balance) (14)
Where 𝜌, 𝐷𝑝 , 𝑃 , 𝑢, and 𝑓𝑤 are the mixture density, pipeline diameter, pressure, velocity,
re
and fanning friction factor respectively. All of these factors were evaluated using Chen's
lP
correlation (1979) [95] as function of space (𝑧) and time (𝑡). 𝐸 stands for the energy of the
mixture. For the non-equilibrium V-L transition, HRM was used with the mechanical
equilibrium assumption such that no phase slip was maintained. The non-equilibrium V-L
na
transition equation with thermodynamic equilibrium for the vapor mass fraction is presented
in Eq. 16.
ur
Where 𝑥 represents the dynamic vapor quality and 𝜃 stands for relaxation time accounting
for the phase change transition delay. Angielczyk et al. proposed an empirically determined
𝜃 = 2.15 × 10−7 (𝑥(𝜌/𝜌𝑠𝑣 )−0.54 ((𝑃𝑠 (𝑇𝑖𝑛 ) − 𝑃)/(𝑃𝑐 − 𝑃𝑠 (𝑇𝑖𝑛 )))−1.76 (17)
Page | 36
Where 𝑃𝑐 , 𝜌𝑠𝑣 and 𝑃𝑠 are critical pressure, saturated vapour density at given pressure, and
reliability, the hypothetical shock tube test was conducted initially for both single (Liquid)
and two-phase CO₂. Table I presents the initial conditions and pipeline characteristics for the
COOLTRANS FBR dense-phase CO₂ shock tube test. Moreover, CO₂PipeHaz CO₂ pipeline
Table I: Predominant conditions and pipeline characteristics for the COOLTRANS shock
tube test
of
Pipeline characteristics Initial conditions
Pipeline length 144 m Fluid temperature 278.35 K
ro
Internal diameter 150 mm Fluid pressure 15.335 MPa
temperature. In general, HRM produced plausibly good estimations of the experimental data
lP
especially during the initial period of blowdown as shown in Figure 29. R-square for pressure
predictions is ≈ 0.924 that indicates reliability of the model. It was observed that FBR clearly
na
indicates fully dispersed flow and, hence, the phase-slip can't be ignored.
ur
Jo
Page | 37
8
Exp.
7
HRM
6
Pressure/MPa
5
4
3
2
1
0
0 0.5 1 1.5 2 2.5 3
of
Time/s
Figure 29: Pressure-time profile at 0.1 m from release end of the CO₂PipeHaz
ro
FBR test [94]
3.8 S. Martynov
-p
Since most of the prior reported blowdown simulation models have been developed for
V-L phase depressurization, they were not appropriate for simulating solid CO₂ release.
re
Martynov et al. in 2013 extended HEM for the development of outflow model accounting for
lP
CO₂ solidification during blowdown [98]. A set of equations describing the HEM flow
including mass conversion, and momentum and energy balance equations were used. For
mass conservation, Eq. 13 was used. Eq. 14 and 15 were modified to Eq. 18 and 19 to
na
Page | 38
Where 𝑅, 𝑇, 𝑝, and 𝑣 are the universal gas constant, temperature, pressure, and specific
volume respectively. 𝑎 and 𝑏 are the empirical parameters accounting for the intermolecular
attraction forces and molecular volume respectively. The instantaneous heat flux at the pipe
wall was calculated based on pipeline wall temperature transitions with time using Eq. 21.
𝑎𝑣𝑒𝑟𝑎𝑔𝑒 𝑎𝑣𝑒𝑟𝑎𝑔𝑒
𝑞𝑤 = 𝜌𝑤 𝐶𝑝,𝑤 𝛿𝑤 ∗ ((𝑑𝑇𝑤 )/𝑑𝑡) (21)
𝑎𝑣𝑒𝑟𝑎𝑔𝑒
Where 𝑇𝑤 represents the average pipeline wall temperature. 𝜌𝑤 , 𝐶𝑝,𝑤 , and 𝛿𝑤 stand
for density, pipeline heat capacity, and wall thickness of pipeline respectively.
of
The featured numerical model was validated against experiments conducted by Dalian
University of Technology (DUT) under CO₂PipeHaz project. The model precisely predicts
ro
the pressure transitions in the pipeline for initially supercritical fluid (Test 1 and 2). The
-p
initial experimental conditions for all the case studies are presented in Table II.
In case of an initially two-phase fluid (Test 3), the given HEM model is not much reliable
because of R-square ≈ 0.661 for pressure predictions as shown in Figure 30. Moreover, the
ur
release model also predicted the dry ice formation and explained the transition phenomena
Page | 39
6
Pre. (heating)
5 Pre. (adiabatic)
Exp.
4
Pressure/MPa
0
0 50 100 150 200 250 300 350
of
Time/s
Figure 30: Time variation of the pressure measured in test 3 [98]
ro
3.9 M. Drescher
HEM was validated mostly against the experimental data from CO₂ pressure release.
-p
However, Dresher et al. (2014) executed experiments with different CO₂–N₂ compositions to
study the potency of the modified HEM for a multi-component system [100]. For simulation
re
study, PR EoS was used and experiments were performed on a 140 m long horizontal tube
lP
with 10 mm ID. Initially, the binary mixture in the supercritical region was at 12 MPa
pressure and a temperature of approximately 293 K. The numerical results for all three fluid
na
compositions regarding the relative and absolute value match the experimental results very
well as illustrated in Figure 31 and 32. With the increase in composition, the model
predictions become more precise as maximum R-square ≈ 0.919 for temperature prediction is
ur
noted in case of 90 % composition. Despite few complexities in the HEM employed, it still
Jo
contains several simplifications in view of the comparatively reasonable results. Further work
could involve the advanced heat-transfer models considering the effect of flow regimes or
boiling.
Page | 40
300
70 % HEM
295 70 % Exp.
290 80 % HEM
80 % Exp.
285 90 % HEM
Temperature/K
280 90 % Exp.
275
270
265
260
255
0 5 10 15 20 25 30
of
Time/s
Figure 31: Comparison of numerical and experimental (PT-40, TT-40) dynamic
ro
temperature changes for different compositions at 50 m [100]
12
10
-p 70 % HEM
70 % Exp.
80 % HEM
re
8 80 % Exp.
Pressure/MPa
90 % HEM
6 90 % Exp.
lP
2
na
0
0 5 10 15 20 25 30
Time/s
ur
Figure 32: Comparison of numerical and experimental (PT-40, TT-40) dynamic pressure
changes for different compositions at 50 m [100]
Jo
4 Major Findings
Table III presents a matrix of modeling outlooks presented in literature that could aid in a
holistic simulation of blowdown use in the industrial sector. Based on the literature reviewed,
Page | 41
Commonly, the models focused on pressure and mixture temperature transitions with
Limited studies on vessel wall temperature and void fraction are also available,
assumptions.
Mostly, the models were validated against previously published experimental data,
of
Table III: Summary of modeling approaches
ro
Title Pressure Temperature Flowrate Void
Mixture Wall Fraction
BLOWDOWN ✓ ✓ ✓ ✓ ✓
BLOWSIM
P. S. Cumber
✓
✓
✓
✓
-p
✓
✗
✓
✓
✗
✓
A. Fredenhagen ✓ ✓ ✗ ✓ ✓
re
SDM ✓ ✗ ✗ ✗ ✗
J. Zhang ✓ ✓ ✗ ✓ ✓
VBsim ✓ ✓ ✓ ✓ ✗
lP
A. Park ✓ ✓ ✓ ✓ ✗
META ✓ ✓ ✗ ✓ ✓
Fairuzov ✓ ✓ ✗ ✓ ✓
CNGS-MOC ✓ ✓ ✗ ✓ ✗
na
A. Oke ✓ ✓ ✓ ✓ ✗
OLGA ✓ ✓ ✗ ✓ ✗
S. Brown ✓ ✓ ✓ ✓ ✗
ur
S. Martynov ✓ ✓ ✓ ✓ ✗
M. Drescher ✓ ✓ ✗ ✓ ✗
Jo
Some of the numerical models or simulation tools produced incredibly precise predictions
of the physical phenomenon occurring during the blowdown process. The BLOWSIM
produced most precise results with R-square ≈ 0.9925 followed by BLOWDOWN with R-
square ≈ 0.977. BLOWSIM, META, OLGA etc. also produced the accurate results before the
process, however, they lack the ability to simulate some of the significant parameters. On the
Page | 42
other hand, BLOWDOWN has the ability to predict most of the important parameters
accurately. Cumber and J. Zhang produced poor results with R-square lower than 0.4. A bar
chart (Figure 33) illustrates R-square value for most of the reviewed simulation tools or
numerical models. For the quantitative comparison of reviewed models, the geometric mean
bias is taken into account (Figure 34). A perfect model is said to have a geometric mean bias
equal to 1.0. A geometric mean bias less than 1.0 refers to underprediction and above 1.0
of
0.8
ro
0.6
R-square
0.4
-p
A. Fredenhagen
BLOWDOWN
P. S. Cumber
CNGS-MOC
S. Martynov
M. Drescher
BLOWSIM
re
S. Brown
Fairuzov
J. Zhang
0.2
A. Park
A. Oke
VBsim
META
OLGA
lP
Page | 43
1.8
1.6
Brown
Zhang
1.4
BLOW CNGS
1.2 DOWN META
MOC OLGA
Cumber
1
BLOW Freden Park Fairuzov A. oke
0.8 SIM hagen
0.6
0.4 Martynov
of
Figure 34: MG bias values for different models
ro
5 Summary
Different simulation tools and numerical models for the calculation of optimum
-p
parameters during blowdown process are reviewed. Table IV summarizes the operating
re
conditions, vessel dimensions, characteristics, and key challenges. Several novel emerging
blowdown models have been reviewed and a comparative study has been presented.
lP
na
ur
Jo
Page | 44
Table IV: Comparative analysis of different simulation tools and numerical models.
f
oo
Name/Title Year Composition Condition Vessel/Pipeline Orifice Characteristics Key Challenges Ref
Volume Size
(mm)
Models + Tools
N₂, 70 % N₂ – - -
pr
BLOWDOWN 1990 150 bar & 1.5 m long vessel 6.325 Good understanding of the physical No work in cryogenic conditions [25,
(1990) 30 % CO₂, 20 °C with 10.6″ processes occurring or emergency 26, 28]
Model + Tool NG/C₃ & HC diameter - Ability to make predictions with an - Use of the extended principle of
acceptable uncertainty corresponding states for
e-
- Based on a three fluids model (vapor, generating the pertinent fluid
liquid and free water) thermophysical properties
- A simulation study is restricted
to depressurization under
Pr
ambient conditions only
- Not fully documented
1991 Case 1: C₁–C₆ 117 bar & 40 km long 0.4191 - Good understanding of the physical - Homogeneity assumption when [27]
+ 4.03 % N₂ 10 °C pipeline processes occurring the flow is two-phase
- Ability to make predictions with an - Quasi-steadiness assumption
acceptable uncertainty - No work on cryogenic conditions
Case 2: C₁– l
200 bar & 20000 m long 0.4064 - Suitable for planned blowdown
na
C₅, C₁₀, C₂₀ + 40 °C pipeline
3 % H₂O
1996 LPG (95 % C₃ 11.3-22.5 100 m long 35-154 - Predictions have been shown to be in - No work on Natural gas [29]
– 5 % C₄) bar & pipeline with 2-6″ at least adequate, and often good, - No work in cryogenic conditions
ur
Page | 45
Fairuzov 1998 LPG (95 % 11.25 bar 100 m long Full-bore - Multi-component, two-phase pipe - Thermal equilibrium between the [70]
f
(Model + Tool) propane & 5 & 19.9 °C pipeline with 150 rupture flow model fluid and pipe wall was not
oo
% butane mm internal - Effect of heat transfer is discussed reached
diameter - For non-adiabatic fluid flow - Rigorous conjugated heat
- Well prediction of experimental data transfer model should be used for
transient heat transfer description
BLOWSIM 1999 64 % C₁, 6 % 116 atm & N/A 10 mm - Multi-component, the multi-phase - Limited data on experimental [14]
pr
(Model + Tool) C₂, 28 % n-C₃ 293 K pipe flow model validation
& 2 % n-C₄ - BLOWSIM performs better than - No work on Natural gas
BLOWDOWN - Thermal and mechanical
- The cubic equation of state has equilibrium assumption [94]
e-
positive effects on data - No work in cryogenic conditions
or emergency situation
CNGS-MOC 1999 LPG (95 % 11.25 bar 100 m long FBR - Reasonably accurate predictions - No Experimental Work [71]
(Model + Tool) propane & 5 & 19.9 °C pipeline with 150 - Less CPU time - Less data on experimental
Pr
% butane mm internal - improving accuracy validation
diameter
P. S. Cumber 2001 LPG (C₁ 66.5 120 bar & 3.24 m height 10 mm - Precise prediction of pressure and - No experimental study conducted [45]
(Model + Tool) %, C₂ 3.5 % 25 °C with 0.54 m ID mass flow rate - Imprecise temperature prediction
& C₃ 30 % - Less computational time - No heat transfer through vessel
wall assumption
A. 2001 N₂ – CO₂ l
25-15 bar 0.05 m³ vessel 29-17 - A good prediction of the process - Tool commercially not available [51]
na
Fredenhagen & 298 K with 0.242 m mm² - Applicable for Multi-component - Unable to predict equipment
(Model + Tool) diameter process conditions
- Limited data on experimental
validation
- No data on multi-phase process
ur
H. 2002 64 % C₁, 6 % 116 bar & 3.24 m vessel 10 mm - Prediction of thermodynamic - No experimentation validation [52]
Mahgerefteh C₂, 28 % n–C₃ 293 K with 1.13 m properties under fire and ambient - No simulation tools developed
(Model) & 2 % n–C₄ diameter and 59 conditions
mm wall - For multi-component and multi-phase
Jo
thickness process
- Vessel conditions calculations
A. Oke 2004 LPG 21.6 bar & 100 m long with 150 - The model accounts for axial and - No experimental study
(Model) 20 °C 0.154 m ID radial flow, real fluid behavior as well performed
as the locality of puncture with respect - No validation for natural gas
to the length of pipe mixture and cryogenic conditions
Page | 46
VPM 2011 78.12 % N₂, 200-400 120 m long N/A - For multi-component process - It can only predict steady state [78]
f
(Model) 20.96 % O₂ & kPa and ≈ pipeline - Ability to calculate pipe wall conditions
oo
0.92 % Ar 19 °C temperature - It is not suitable for the multi-
phase process
OLGA 2012 CO₂ 81 bar & 50 km long, 24″ 8″ - Experimental data from large-scale - Non-reliable predications of [89]
(1991) 31 °C pipeline CO₂ unit experimental data
Tool - Initially supercritical conditions - No work on multi-component &
pr
- Multi-phase process multi-phase process
2014 CO₂, 2–8 % 85b bar & 51.96 m tube with N/A - Studied Effect of impurities on CO₂ - No study on the effect of initial [93]
N₂ 20 °C 3.86 mm ID blowdown conditions
- Good predictions of initial pressure
e-
drop
J. Zhang 2013 CO₂ & Water 25 MPa & 4 m height & 2 m 0.008 m² - A good prediction of the process - Limited data on experimental [55]
380 °C diameter - Supercritical conditions validation
- Multi-component and multi-phase - No work on natural gas
Pr
process - Only work on supercritical
conditions
S. Brown 2013 CO₂ 15.335 144 m long FBR - Model predictions produced a - Lack of sufficient and reliable [94]
(Model) MPa & pipeline with 150 reasonable agreement experimental data
278.35 K mm diameter - Multi-phase process - No work on the multi-component
- Delayed phase transition has an process
l
7 MPa & 37 m long insignificant effect on the - No work on natural gas and
na
298.35 K pipeline with 40 depressurization rate of pipe cryogenic conditions or
mm ID emergency situation
S. Martynov 2013 CO₂ 53-86 bar 256 m long 50 mm - Good prediction for initially - Unable to predict initially two- [98]
(Model) & 34-39 pipeline with 233 supercritical fluid phase fluid
°C mm diameter - Prediction of dry ice formation - No work on cryogenic or
ur
emergency conditions
- Experimental study performed
only for single component
M. Drescher 2014 CO₂ + 10–30 120 bar & 140 m long 9.5 mm - Good predication of experimental - Require advancement of heat- [100]
Jo
Page | 47
- Consider non-equilibrium effects
f
between constituent fluid phases
oo
A. Park 2017 HC Mixture 293-303 K 3.24m length with 10 mm - Two-phase multi-component system - No experimental work performed [58]
& 117.5- 1.13m ID prediction. No study in the cryogenic region.
120 bar - Can predict wall temperature
contracting liquid and vapor.
- More precise predictions than
pr
previously available models
e-
l Pr
na
ur
Jo
Page | 48
6 Conclusion
In this paper, the available models for the optimum blowdown parameters' calculation are
existing literature are also highlighted. The available models are capable of solving the
blowdown scenarios, however, there are still some limitations listed as follows.
Only a few lab-scale experimental studies are reported for the validation of
models. The reliability of models will be improved if also validated against the
of
large scale experimental facilities.
More extensive experimental studies on natural gas and other gaseous mixtures
ro
could help to improve the validity of models.
-p
Limited studies are reported for the blowdown of pressure vessels under fire
conditions.
re
Experimental studies for blowdown from the cryogenic conditions could also be
considered. It would help to study the blowdown process under worst case
lP
scenarios especially when a phase change is involved that can lead to the
formation of solids.
na
A new numerical model can also be developed to overcome the problems related
considered.
In addressing these research areas, industry should be empowered to make effective and
sustainable retrofit decisions while simultaneously reducing hazards and improving the
safety. This paper has focused on solutions for industrial hazards related to depressurization.
Page | 49
However, the outcome of the identified research areas has wide-ranging applications with
techniques that could be applied to solve many other system simulation challenges.
Acknowledgement
of
The authors would like to extend their most profound gratitude to the CO₂ Research Centre
ro
this Review article.
-p
re
lP
na
ur
Jo
Page | 50
References
[1] Barma M, Saidur R, Rahman S, Allouhi A, Akash B, Sait SM. A review on boilers energy
use, energy savings, and emissions reductions. Renew Sustain Energy Rev. 2017;79:970-83.
[3] Cullen W. The Public Inquiry into the Piper Alpha Disaster.(Department of Energy,
of
[4] Bond J. Institute of Chemical Engineers accidents database. Rugby. UK: Institute of
ro
Chemical Engineers. 2002.
[5] Fletcher S. US Senate ready to act on pipeline safety as public attention sharpens. Oil Gas
-p
J. 2001;99:58-60.
re
[6] Selby C, Burgan B. Blast and fire engineering for topside structures-phase 2: final
[7] Tolloczko J. Interim guidance notes for the design and protection of topside structures
http://bulktransporter.com. 2001.
[9] The National Board of Boiler and Pressure Vessel Inspectors, National Board Bulletin
Jo
Page | 51
[11] Lydell BO. Pipe failure probability—the Thomas paper revisited. Reliab Eng Syst Safe.
2000;68:207-17.
[12] Talbi B. CO₂ emissions reduction in road transport sector in Tunisia. Renew Sustain
[13] Milano J, Ong HC, Masjuki H, Chong W, Lam MK, Loh PK, et al. Microalgae biofuels
as an alternative to fossil fuel for power generation. Renew Sustain Energy Rev.
2016;58:180-97.
of
[14] Mahgerefteh H, Wong SM. A numerical blowdown simulation incorporating cubic
ro
[15] Moss DR. Pressure vessel design manual: Elsevier; 2004.
-p
[16] Cui H, Wang W, Li A, Li M, Xu S, Liu H. Failure analysis of the brittle fracture of a
thick-walled 20 steel pipe in an ammonia synthesis unit. Eng Fail Anal. 2010;17:1359-76.
re
[17] Onyebuchi VE, Kolios A, Hanak DP, Biliyok C, Manovic V. A systematic review of key
lP
challenges of CO₂ transport via pipelines. Renew Sustain Energy Rev. 2017.
[18] Gradle R. Design of Gas Pipeline Blowdowns. Enrgy Proced Cdn. 1984:15-20.
na
[19] Ouyang L-b, Aziz K. Steady-state gas flow in pipes. J Petrol Sci Eng. 1996;14:137-58.
[21] Byrnes W, Reid R, Ruccia F. Rapid depressurization of gas storage cylinder. Ind Eng
Jo
[23] Simon LL, Introvigne M, Fischer U, Hungerbühler K. Batch reactor optimization under
Page | 52
[24] Fisher H, Forrest H, Grossel SS, Huff J, Muller A, Noronha J, et al. Emergency relief
system design using DIERS technology: The Design Institute for Emergency Relief Systems
of
[27] Richardson S, Saville G. Blowdown of pipelines. Offshore Europe: Society of
ro
[28] Haque M, Richardson S, Saville G, Chamberlain G, Shirvill L. Blowdown of pressure
-p
vessels. II. Experimental validation of computer model and case studies. Trans IChemE, Part
1996;74:235-44.
[30] Cowley L, Tam V. Consequences of Pressurized LPG Releases: The Isle of Grain Full
na
Scale Experiments, submitted 13th Int. Conf LNG/LPG Conference and Exhibition, Kuala
Lumpur, Malaysia1988.
ur
[31] Tam V, Higgins R. Simple transient release rate models for releases of pressurised liquid
[32] Saville G, Richardson S, Barker P. Leakage in ethylene pipelines. Process Saf Environ.
2004;82:61-8.
[33] Norris III H, Puls R. Single-phase or multiphase blowdown of vessels or pipelines. SPE
Page | 53
[34] Norris III H. Hydrocarbon blowdown from vessels and pipelines. SPE Annual
[35] Wong MA. Development of a mathematical model for blowdown of vessels containing
[36] Peng D-Y, Robinson DB. A new two-constant equation of state. Ind Eng Chem Fund.
1976;15:59-64.
of
Chem Eng Sci. 1972;27:1197-203.
[38] Twu CH, Coon JE, Cunningham JR. A new cubic equation of state. Fluid Phase
ro
Equilibra. 1992;75:65-79.
-p
[39] Witlox H, Holt A. A unified model for jet, heavy and passive dispersion including
[40] Woodward JL, Cook J, Papadourakis A. Modeling and validation of a dispersing aerosol
[41] Woodward JL, Papadourakis A. Reassessment and reevaluation of rainout and drop size
[42] Witlox H, Holt A. “Validation of the unified dispersion model against Kit Fox field
[43] Witlox HW, Harper M, Oke A, Stene J. Phast validation of discharge and atmospheric
dispersion for pressurised carbon dioxide releases. J Loss Prevent Proc. 2014;30:243-55.
Page | 54
[44] Witlox HW, Fernandez M, Harper M, Oke A, Stene J, Xu Y. Verification and validation
of Phast consequence models for accidental releases of toxic or flammable chemicals to the
[45] Cumber PS. Predicting outflow from high pressure vessels. Process Saf Environ.
2001;79:13-22.
[46] Skouloudis AN. Benchmark exercises on the emergency venting of vessels. J Loss
of
[47] Eggers R, Green V. Pressure discharge from a pressure vessel filled with CO₂. J Loss
ro
[48] Gebbeken B, Eggers R. Thermohydraulic phenomena during vessel release of initially
-p
supercritical carbon dioxide. International symposium two-phase ow modelling and
experimentation1995.
re
[49] Gebbeken B, Eggers R. Blowdown of carbon dioxide from initially supercritical
lP
Carbon Dioxide. Proc of the 4 World Conference on Experimental Heat Transfer, Fluid
[51] Fredenhagen A, Eggers R. High pressure release of binary mixtures of CO₂ and N₂.
[52] Mahgerefteh H, Falope GB, Oke AO. Modeling blowdown of cylindrical vessels under
[53] API Recommended Practice 520 Part 1—Sizing and Selction, sixth ed. 1993.
Page | 55
[54] Song X-G, Park Y-C, Park J-H. Blowdown prediction of a conventional pressure relief
2013;57:279-88.
blowdown of pressure vessels containing two-phase hydrocarbons mixtures with the partial
of
[57] Roberts T, Medonos S, Shirvill L. Review of the response of pressurised process vessels
ro
and equipment to fire attack. 2000.
-p
[58] Park A, Ko Y, Ryu S, Lim Y. Numerical modeling of rapid depressurization of a
[59] Speranza A, Terenzi A. Blowdown of hydrocarbons pressure vessel with partial phase
separation. Applied and Industrial Mathematics in Italy: World Scientific; 2005. p. 508-19.
na
[60] Jarrah AA. Modeling and simulation of rapid depressurizations of hydrocarbon gas
mixture flowing at unsteady high velocity in piping transport system. International Journal of
ur
Page | 56
[64] Chen J, Richardson S, Saville G. Modelling of two-phase blowdown from pipelines—i.
[65] Chen J, Richardson S, Saville G. A simplified numerical method for transient two-phase
A simplified numerical method for multi-component mixtures. Chem Eng Sci. 1995;50:2173-
87.
of
[67] Lahey RT, Moody F. The Thermal Hydraulics of a Boiling Water Nuclear Reactor
ro
[68] Reibold W, Reocreux M, Jones O. Blowdown phase. Nuclear Reactor Safety Heat
Transfer. 1981:325-78. -p
[69] Jackson J, Liles D, Ransom V, Ybarrondo L. LWR system safety analysis. Nuclear
re
Reactor Safety Heat Transfer. 1981;12:415.
lP
[70] Fairuzov YV. Blowdown of pipelines carrying flashing liquids. AIChE journal.
1998;44:245-54.
na
[71] Mahgerefteh H, Saha P, Economou IG. Fast numerical simulation for full bore rupture of
[74] Oke A, Mahgerefteh H, Economou I, Rykov Y. A transient outflow model for pipeline
Page | 57
[75] DeReuck K, Craven R, Elhassan A. Transport Properties of Fluids: Their Correlation,
Prediction and Estimation, J. Millat, JH Dymond, and CA Nieto de Castro, eds., IUPAC.
[76] Assael MJ, Trusler JM, Tsolakis TF. Thermophysical properties of fluids: an
of
[78] Rajiwate FLH. Investigation of compressible fluid behaviour in a vent pipe during
ro
[79] Lemmon DEW, D. M. L. Huber, and D. M. O. McLinden. REFPROP Manual. Boulder:
[81] Bansal RK. Fluid mechanics and hydraulic machines. 9 ed ed. New Delhi: Laxmi
[82] Saad MA. Compressible Fluid Flow. 2 ed. ed. New Jersey: Prentice Hall; 1993.
[83] Shapiro AH. The dynamics and thermodynamics of compressible fluid flow (vol2).
ur
1954.
Jo
[84] Parker G. ‘Pop’safety valves: a compressible flow analysis. International journal of heat
[85] de Koeijer G, Borch JH, Jakobsenb J, Drescher M. Experiments and modeling of two-
phase transient flow during CO₂ pipeline depressurization. Enrgy Proced. 2009;1:1683-9.
Page | 58
[86] Clausen S, Munkejord ST. Depressurization of CO₂–a numerical benchmark study.
[88] Cho M-I, Huh C, Jung J-Y, Kang S-G. Experimental study of N₂ impurity effect on the
of
pipeline. Enrgy Proced. 2012;23:256-65.
[90] Bendiksen KH, Maines D, Moe R, Nuland S. The dynamic two-fluid model OLGA:
ro
Theory and application. SPE Prod Eng. 1991;6:171-80.
-p
[91] Span R, Wagner W. A new equation of state for carbon dioxide covering the fluid region
from the triple‐point temperature to 1100 K at pressures up to 800 MPa. J Phys Chem Ref
re
Data. 1996;25:1509-96.
lP
mixtures: Models for transient simulation. Int J Greenh Gas Con. 2013;15:174-85.
na
[93] Huh C, Cho M-I, Hong S, Kang S-G. Effect of impurities on depressurization of CO₂
for the full bore rupture of dense phase CO₂ pipelines. Int J Greenh Gas Con. 2013;17:349-
56.
[95] Chen NH. An explicit equation for friction factor in pipe. Ind Eng Chem Fund.
1979;18:296-7.
Page | 59
[96] Angielczyk W, Bartosiewicz Y, Butrymowicz D, Seynhaeve J-M. 1-D modeling of
supersonic carbon dioxide two-phase flow through ejector motive nozzle. 2010.
CO₂pipehaz: quantitative hazard assessment for next generation CO₂ pipelines. Institute of
phase releases of carbon dioxide from high-pressure pipelines. Process Saf Environ.
of
2014;92:36-46.
ro
carbon dioxide solid‐vapor equilibrium. Greenh : Sci Tech. 2013;3:136-47.
-p
[100] Drescher M, Varholm K, Munkejord ST, Hammer M, Held R, de Koeijer G.
[101] Hanna S, Chang J, Strimaitis D. Hazardous gas model evaluation with field
Page | 60