You are on page 1of 21

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/337603699

A review on modeling and simulation of blowdown from pressurized vessels


and pipelines

Article  in  Process Safety and Environmental Protection · November 2019


DOI: 10.1016/j.psep.2019.10.035

CITATIONS READS

12 1,610

6 authors, including:

Umar Shafiq Azmi Shariff


Institute of Engineering Fertilizer Research Universiti Teknologi PETRONAS
30 PUBLICATIONS   122 CITATIONS    127 PUBLICATIONS   1,765 CITATIONS   

SEE PROFILE SEE PROFILE

Muhammad Babar Mohamad azmi Bustam


Universiti Teknologi PETRONAS Universiti Teknologi PETRONAS
29 PUBLICATIONS   289 CITATIONS    319 PUBLICATIONS   5,695 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Ionic liquid for biomedical applications View project

biological activities and phytochemical analysis of potentially active compounds in cranberry View project

All content following this page was uploaded by Muhammad Babar on 10 February 2020.

The user has requested enhancement of the downloaded file.


Process Safety and Environmental Protection 133 (2020) 104–123

Contents lists available at ScienceDirect

Process Safety and Environmental Protection


journal homepage: www.elsevier.com/locate/psep

A review on modeling and simulation of blowdown from pressurized


vessels and pipelines
Umar Shafiq a , Azmi M. Shariff a,∗ , Muhammad Babar a , Babar Azeem a , Abulhassan Ali b ,
Mohamad Azmi Bustam a
a
CO2 Research Centre (CO2RES), Universiti Teknologi PETRONAS, 32610 Bandar Seri Iskandar, Perak, Malaysia
b
Department of Chemical and Materials Engineering, University of Jeddah, Jeddah, Saudi Arabia

a r t i c l e i n f o a b s t r a c t

Article history: In process industry, failure or rupture of pressurized vessel is very dangerous especially when there is
Received 14 February 2019 an escape of flammable gaseous mixture that can cause potential fire or explosion. One of the scenarios
Received in revised form that causes such accidents is the blowdown process. Therefore, it becomes crucial to control blowdown
24 September 2019
process to prevent such accidents. It is important to design optimally to make sure that blowdown valve
Accepted 31 October 2019
is according to the requirements. For the safe use of a pressure relief system, some of the parameters are
Available online 13 November 2019
critical, 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
Keywords:
Blowdown
tools for optimum design of the blowdown process. This subject matters because the available models
High-pressure vessels or tools cover different aspects of blowdown process. A meticulous review is required to present the
Hazards applications of these models and tools based on the accidental scenarios. Therefore, this paper critically
Depressurization reviews the models and tools that are developed purposely to calculate optimum blowdown parameters
Simulation tools based on fluid and vessel conditions. Recommendations are given for the development of new simulation
Modeling tool to simulate phase change conditions especially when solid formation is involved.
© 2019 Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
1.1. Concerns and considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
2. Design of blowdown system for pressure vessels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
2.1. DIERS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
2.2. BLOWDOWN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
2.3. FRICRUP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
2.4. BLOWSIM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
2.5. PHAST . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
2.6. P. S. Cumber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
2.7. A. Fredenhagen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
2.8. H. Mahgerefteh . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
2.9. SDM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
2.10. J. Zhang . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
2.11. VBsim . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
2.12. A. Park . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
3. Design of blowdown system for pressure pipelines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
3.1. META . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
3.2. Fairuzov . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

∗ Corresponding author.
E-mail address: azmish@utp.edu.my (A.M. Shariff).

https://doi.org/10.1016/j.psep.2019.10.035
0957-5820/© 2019 Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123 105

3.3. CNGS-MOC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113


3.4. A. Oke . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
3.5. VPM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
3.6. OLGA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
3.7. S. Brown . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
3.8. S. Martynov . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
3.9. M. Drescher . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
4. Major findings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
5. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
6. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

Nomenclature

BP British Petroleum
CCS Carbon Capture and Sequestration
CFD Computational Fluid Dynamics
EoS Equation of State
FBR Full Bore Rupture
FDM Finite Difference Method
FEM Finite Element Method
HC Hydrocarbon
HEM Homogenous Equilibrium Model
HRM Homogenous Relaxation Model Fig. 1. Common reasons for pressure vessel failures.
ID Inner Diameter
MAD Maximum Absolute Deviation
is enough to inflict the structural damage such as weld failure, local
META Multi-Component Equilibrium Two-Phase Analyser
instability, or extensive inelasticity.
MOC Method of Characteristics
The National Board of Boiler and Pressure Vessel Inspectors
MSM Marginal Stability Two-Fluid Model
recorded a 24 % increase in the number of accidents involving pres-
PR Peng & Robinson
sure vessels during 1999–2000. Steam heating, power boilers, and
PRV Pressure Relief Valve
unfired pressure vessels were the major sources of such accidents
QRA Quantitative Risk Assessment
(Transporter B. Boiler, 2001). From 1992–2001, a total of 23,338
SCWR Supercritical Water-Cooled Reactor
accidents related to pressure vessels were recorded (Transporter
SDM Simplified Dynamic Model
B. Boiler, 2001; The National Board of Boiler and Pressure Vessel
SRK Soave-Redlich-Kwong
Inspectors, 2002; AirConditioning, 2002). Similarly, according to
TEM Thermodynamic Equilibrium Model
Swedish Nuclear Power Inspectorate (Lydell, 2000), more than 66 %
VBsim Vessel Blowdown Simulator
(2476) of the pipeline failure incidents recorded during 1994–1999
V-L Vapor-Liquid
were triggered due to leakage or puncture. Common reasons for
VPM Vent Pipe Model
pressure vessel failures are illustrated in Fig. 1.
V-S Vapor-Solid
In oil and gas platforms or offshore operations, overpressure of
VLE Vapor Liquid Equilibrium
pressurized vessels also arises during emergency situations due to
fire or valves’ malfunction. These industries usually work with CO2,
natural gas, and fossil fuels (Talbi, 2017; Milano et al., 2016). Most
1. Introduction of the accidents at offshore oil and gas industries, nuclear power
plants, and chemical plants cause the outflow of radioactive, toxic,
Air-tight pressurized vessels and pipelines are widely used to explosive, or flammable materials. Particularly in the oil and gas
transport or hold gases and process liquids or fluids in petro- facilities, failure or rupture of pressurized vessels poses higher risk
chemical plants, refineries, and process industries. These vessels due to the possible escape of flammable gaseous mixture in a fire
are generally subjected to high-pressure loading and unidentical event. Therefore, process vessels are depressurized for the trans-
external or internal pressures in comparison to the ambient pres- portation of flammable or environmentally toxic gaseous mixture
sure (Barma et al., 2017). Consequently, it becomes potentially to a safer place (Mahgerefteh and Wong, 1999; Moss, 2004; Cui
hazardous and dangerous because of characteristic operational et al., 2010; Onyebuchi et al., 2017). Blowdown is a typical way of
pressure of the vessel (Ladokun et al., 2010). The rupture of pipeline minimizing the failure hazard of vessels when an emergency arises
and subsequent eruption of fire broke down the platform into in a process facilities. However, for the safe use of a pressure relief
seabed resulting in a loss of 167 lives during the Piper Alpha inci- system, some of the parameters are crucial; for example, tempera-
dent (Cullen, 1990). Many other tragedies have been witnessed that ture, pressure, selection of construction material, and sizing of relief
caused a significant number of fatalities due to pipeline and vessel valves. Prediction of precise minimum vessel wall temperature dur-
rupture (Bond, 2002; Fletcher, 2001). After the Piper Alpha tragedy, ing depressurization process can affect the selection of construction
the offshore industry carried out a great deal of research work to material, reduce over-design, and subsequently lower the project
understand the characteristics of hydrocarbon fires and explosions cost. Similarly, over-design can also be reduced by precise predic-
(Selby and Burgan, 1998; Tolloczko, 1992). The results indicate that tion of maximum flow rate during blowdown. Critical parameters
a pressure above 4 bar developed in typical topsides of the vessels for a safe blowdown process are portrayed in Fig. 2.
106 U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123

sure vessels and pipelines, is presented in this paper to summarize


existing work and set out most overlooked aspects.

2. Design of blowdown system for pressure vessels

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 small amount of
force if temperature falls below the ductile-brittle transition tem-
perature of vessel’s material of construction (Byrnes et al., 1964).
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 conden-
sate 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 (Richardson and Saville,
1993). To date, several numerical models and simulation tools have
been developed that can simulate different blowdown scenarios
of vessels based on vessel sizing, blowdown conditions, and com-
positions etc. A timeline of these simulation tools or numerical
models is illustrated in Fig. 4 followed by a brief description (Some
of the authors didn’t name their simulation tools; therefore, we
Fig. 2. Important parameters for safe blowdown process.
have mentioned the name of the authors instead of the simulation
tool).

2.1. DIERS
1.1. Concerns and considerations
A consortium of several companies founded DIERS program
In recent years, safety and process engineers have encountered in 1976 to design pressure relief system for runaway reactions
several safety-related issues while dealing with high-pressure and develop the additionally required technology. DIERS research
blowdown. The key principle of sustainable depressurization pro- program performed several experimental studies, coded a com-
cess design is to decrease the pressure and discharge the inventory prehensive simulation tool, and designed a bench-scale prototype
safely in the minimum possible time (Gradle, 1984). The discharge apparatus. The aim was to predict two-phase flow venting and ana-
rate of compressible fluids generally depends upon pressure, pres- lyze the possible applicability of various two-phase (vapor-liquid)
sure drop, temperature, friction factor, Reynolds number, pipeline flashing flow methods for sizing of relief systems. In this model, it
diameter, roughness, length, and gas properties (L-b and Aziz, is considered that boiling takes place through the entire volume of
1996). However, the rapid expansion of gas and generation of liquid rather than solely at the surface. Each bubble occupies vol-
vapors can expose the pressure vessel to pressurized thermal shock ume and displaces the liquid surface upward. Individual bubbles are
during an emergency blowdown process. This shock initiates due able to rise (slip) through the liquid with a velocity that depends
to integrated stress from rapid transitions in pressure and temper- on buoyancy and surface tension, whereas they are retarded by vis-
ature that result in the non-uniform distribution of temperature cosity and foamy character of the fluid (Simon et al., 2008). DIERS
in vessel walls and, therefore, lead to the differential contraction program evolved over time; the initial phase included study of
and expansion. Pressurized thermal shocks can also cause embrit- vapor-liquid phase dynamics, second phase involved the experi-
tlement of vessel walls and subsequent “failure by fatigue” due to mental investigations for small and large scale integral blowdown
very low temperature generated inside the vessel because of Joule- and vented runaway reaction, and the final phase included the
Thomson effect. Therefore, to predict the effects of vessel/pipeline development of a coded computational package along with a bench
rupture or leak, or to design a sustainable blowdown process, there scale prototype experimental apparatus (Fisher et al., 2010).
is dire need of a technique to predict the mass efflux. Moreover,
to calculate optimum blowdown time, a precise balance between 2.2. BLOWDOWN
maximum acceptable depressurization time (Api, 1990) and min-
imum fluid and wall temperature that may be safely considered, A brief description of prediction and experimentation of fluid
is required. Some of the important factors to be considered before and vessel behavior during blowdown process was presented by
designing the blowdown system are illustrated in Fig. 3. Haque et al. in 1990 (Haque et al., 1990). Experiments were per-
Conventionally, several numerical models and simulation tools formed on blowdown of pressurized vessels containing pure N2, 70
are reported in literature for the prediction of maximum possi- % N2–30 % CO2, and natural gas/hydrocarbon mixture. Based on the
ble blowdown time and fluid and vessel conditions. The numerical experiments performed and assessment of available data, a com-
models have been developed by commonly adopting the Homoge- putational tool called BLOWDOWN was coded to simulate rapid
nous Equilibrium Model (HEM) assumption such that each phase is depressurization of vessels. The developed BLOWDOWN package
considered to be at the phase equilibrium and thermal equilibrium. has the ability to measure the temperature, pressure, and compo-
Some of the researchers also coded simulation tools based on their sition as a function of time.
numerical models. However, the available mathematical models or For the experimental investigation, a 6.325 mm orifice was used
simulation tools cover different aspects of blowdown or depres- to blowdown the mixture from a 0.086 m3 vessel with 25 mm of
surization process. Therefore, a literature review of the existing wall thickness. Pure N2 took almost 100 s to reach the atmospheric
correlations, numerical models, and computational tools that have pressure from 15 MPa. Whereas, the vessel containing 70 % N2–30
capability to calculate optimum blowdown parameters for pres- % CO2 binary mixture reached the atmospheric pressure in 60 s. The
U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123 107

Fig. 3. Factors to be considered before designing a sustainable pressure relief system.

Fig. 4. Timeline for blowdown simulation tools and numerical models for vessels.

dynamic change in gas and wall temperature is depicted in Fig. 5.


To analyze the predicted results, the coefficient of determination,
R2 , and geometric mean bias of the model data are calculated. For
this model, the R2 for temperature prediction is ≈ 0.977. A compar-
ative illustration of R2 values and geometric mean bias of different
models is presented in Figs. 33 and 34 respectively.
Subsequently, an extension in the simulation tool was made that
can also be used to 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 depres-
surization) for the arbitrary combinations of vessels and pipelines
(Haque et al., 1992a). This extended version of BLOWDOWN tool
was validated again in 1992 by Haque et al. while supported by
several case studies (Richardson and Saville, 1991; Haque et al.,
1992b). In the first case study; 53,804 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 depressur-
ization of 2000 m long pipeline containing 1975 tonnes of methane
condensate was carried out. The initial pressure inside the pipeline
was 20 MPa and initial temperature was around 313 K.
Another comparison of BLOWDOWN predictions and exper-
imental data was made during LPG test performed by British
Petroleum (BP) and Shell on the Isle of Grain in 1985 (Richardson
and Saville, 1996; Cowley and Tam, 1988; Tam and Higgins, 1990). Fig. 5. Dynamic temperature changes during blowdown for (a) 70 % N2–30 % CO2
Further experimental investigations on the prediction of blow- mixture and (b) pure N2 (Haque et al., 1990).
108 U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123

down parameters using BLOWDOWN package are also available


in literature (Saville et al., 2004). In all the experimental studies,
BLOWDOWN package provides a reasonable understanding of the
physical processes occurring and an ability to make predictions
with adequate certainty. However, use of extended corresponding
states principle introduces uncertainties that make the simula-
tion computationally difficult. Moreover, the limitations of a cubic
equation of state (EoS) near or at the critical region are well doc-
umented but their effect on the accuracy of blowdown 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.

Fig. 6. Correlation between predicted and experimental pressure-time profiles


2.3. FRICRUP
(Mahgerefteh and Wong, 1999).

A simple mechanistic model named FRICRUP was coded by Nor-


ris et al. in FORTRAN program to predict single and multi-phase
flow behavior during blowdown of vessels or pipelines (Norris and
Puls, 1993). HEM and thermodynamic equilibrium model (TEM)
assumptions along with 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 water, and air. The comparison shows that a reason-
able mechanistic model has been developed that can predict single
and multi-phase blowdown from the vessel as well as the pipe.
Regardless of its complexities, the mechanistic model indicates
disagreement with multi-phase flow possibly due to TEM assump-
tions. A year later, Norris et al. conducted another experimental
study using HC gases (Norris, 1994). The results were similar to
the ones achieved while using the non-HC gases because the basic Fig. 7. Correlation between predicted and experimental vapor temperature profile
mechanical model assumptions remained the same. (Mahgerefteh and Wong, 1999).

2.5. PHAST
2.4. BLOWSIM
The Unified Dispersion Model (UDM) developed by Woodward
BLOWSIM model (Mahgerefteh and Wong, 1999; Wong, 1998), and Cook in the early 90’s was later implemented into DNV soft-
developed by Mahgerefteh and Wong is based on three cubic- ware package, termed as PHAST in version 6.0 (Witlox and Holt,
EoSs including Peng and Robinson (PR) (Peng and Robinson, 1976), 1999; Woodward et al., 1995; Woodward and Papadourakis, 1995).
Soave-Redlich-Kwong (SRK) (Soave, 1972), and newly developed PHAST is a helpful tool to study the initial release of mixture as
TCC Cubic-EoS (Twu et al., 1992). BLOWSIM (BLOWdown- well as the far-field dispersion. Previously, PHAST was limited only
SIMulation) requires a minimum number of input parameters. It is to vapor and liquid phase release of the chemicals. However, the
computationally efficient and can predict pressure, mass flow rate, upgraded version allows to study the occurrence of fluid to solid
and fluid and wall temperature, all as a function of time. BLOWSIM transition during release of CO2. PHAST was validated against sev-
model considers the following: eral experimental studies for unpressurized release of CO2 (Witlox
and Holt, 1999, 2001). Later, Witlox et al. performed some more
experimental studies for the validation against pressurized release
1 Heat transfer between fluid phases and their corresponding sec-
of CO2 (Witlox et al., 2014). An overview of the validation of
tions of the vessel wall.
different models including flammable effect, dispersion and pool
2 Non-equilibrium effect between phases.
spreading/evaporation, and discharge models was provided along
3 The effect of sonic flow at the orifice.
with extensive experimental database for validation of the above
4 Interphase fluxes because of condensation and evaporation.
models and scenarios (Witlox et al., 2018).

The performance of the featured model was assessed by the 2.6. P. S. Cumber
correlation of predicted data generated from BLOWDOWN and
experimental results obtained from blowdown of condensable gas. A numerical model was coded by Cumber for the prediction
Typical output results include dynamic temperature and pressure of vent sizing of pressurized vessel during blowdown batch pro-
profiles for liquid and vapor phases as well as unwetted and wet- cess (Cumber, 2001). The pressure vessel containing gas mixture
ted walls. The achieved results and their comparison with different is considered as a single unit with thermodynamic equilibrium
EoSs as well as the experimental data are illustrated in Figs. 6 and 7. assumptions. This model has a vast range of venting models imple-
BLOWSIM’s predictions are quite insensitive to the Cubic-EoS used. mented for a two-phase system. However, selection of the most
In general, it can be inferred that choice of the Cubic-EoS has a negli- suitable model for a specific case study necessitates the user to
gible effect on predicted results. Moreover, both numerical models have comprehensive information on the scope of application of the
are capable of reasonably predicting the dynamic pressure varia- respective model. For this particular case, predicted blowdown data
tions. The R-square noted for TCC (TWU) model for temperature was plotted against the experimental data as shown in Figs. 8 and 9.
predictions is ≈ 0.992. The trend indicates that predicted gas temperature in the initial
U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123 109

Fig. 8. Dynamic pressure profile during blowdown of LPG (Cumber, 2001). Fig. 10. Calculated and measured pressure transients for 17 mm2 orifice
(Fredenhagen and Eggers, 2001).

Fig. 9. Dynamic temperature profile during blowdown of LPG (Cumber, 2001).


Fig. 11. Calculated and measured temperature transients (Fredenhagen and Eggers,
2001).
phases of depressurization is in agreement with the experimental
data. However, ultimately it poorly predicts temperature with an
dictions is ≈ 0.955, however, the temperature predictions at low
R-square ≈ 0.445 because of the adiabatic vessel wall assumption
pressure were not much precise. Consequently, an improvement
as shown in Fig. 9. The significant difference is with respect to the
needs to be made for more precise modeling of thermodynamic
phase temperature measurement that shows that thermodynamic
behavior of the mixture.
equilibrium assumption is not effective for this case. However, the
measured pressure of vessel was precisely predicted. For the ves-
2.8. H. Mahgerefteh
sel depressurizing over long time-scale, the temperature of the
vessel was under-predicted after first few minutes as a conse-
Most comprehensive blowdown model reported in literature
quence of assuming no heat transfer through the vessel walls. This
was BLOWDOWN, later on extended as BLOWSIM as discussed ear-
did not have a substantial effect on the simulated mass flow rate
lier. Although validated extensively with the experimental results,
though. Another discrepancy with the application of this model is
the prediction is still limited to depressurization under ambient
the robustness that needs to be addressed (Skouloudis, 1992).
temperature. This is a serious drawback since the blowdown situ-
ations may involve a fire case most of the time. To overcome this
2.7. A. Fredenhagen
issue in 2002, Mahgerefteh et al. described the establishment of a
numerical model for quantitative risk assessment (QRA) of rupture
Previously accomplished research work on depressurization of
following the depressurization of vessel under fire (Mahgerefteh
pure CO2 leads towards successful modeling of the pressure tran-
et al., 2002). The model is able to predict the tri-axial transient
sitions and axial void profile during the expansion process (Eggers
pressure and thermal stress produced in both unwetted and wetted
and Green, 1990; Gebbeken and Eggers, 1995, 1996; Gebbeken and
wall sections of the vessel. The fluid/wall heat transfer was obtained
Eggers, 1997). However, an experimental investigation on blow-
using the following equation.
down of the top vented pressurized vessel containing CO2 with
 n  
impurities (N2) was performed by Fredenhagen and Eggers in 2001 ’
Q = h Tw (a, t) ’
Tw (a, t) − TZ At (1)
to study the effect of impurities (Fredenhagen and Eggers, 2001).

In this study, the axial void profiles, axial temperature profiles, and Where h is the heat transfer coefficient, Tw (a, t) refers to the
pressure and mass were measured. Top vented cylindrical vessel unknown wall temperature at time t, TZ is the liquid or vapor
with 0.05 m3 volume, 0.242 m ID, and a 4:5 height to diameter temperature, A is the heat transfer area and n stands for power
ratio was used. The venting line was linked to a quick opening ball index. A hypothetical case study for the depressurization of a ves-
valve with a cross-sectional area of 50 mm2 . PR EoS was used for sel containing higher-HC was performed. The vessel was supposed
the calculation of phase equilibrium. For the prediction of this pro- to be at the initial pressure and temperature of 11.6 MPa and 293 K
cess, a simulation tool was also coded based on the energy and respectively. Two types of blowdown scenarios were simulated:
mass balance and a drift flux model. The dynamic pressure and (1) the vessel under ambient conditions and (2) the vessel under
temperature profiles predicted by featured simulation tool and the fire radiating a heat flux of 90 kW/m2 . The predicted pressure and
experimental results are illustrated in Figs. 10 and 11. The compu- temperature profiles for a vessel under ambient and fire condi-
tational model predicted reasonable results of pressure transition tions are illustrated in Fig. 12. The predicted results show that
throughout the process. The overall R-square for temperature pre- thermal radiation compensates the temperature drop to a large
110 U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123

Fig. 12. Temperature and pressure histories under ambient and fire conditions
(Mahgerefteh et al., 2002).
Fig. 13. Comparison of experimental results and SDM (Song et al., 2013).

extent. A comprehensive analysis of pressure 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 fea-
tured numerical model serves as a powerful tool for the calculation
of minimum required blowdown time to eradicate the risk of vessel
puncture under fire.

2.9. SDM

Over the years, the researchers’ primary focus has been on the
performance of pressure relief valve (PRV) using dynamic model-
ing approach. Unfortunately, most of the processes were simulated
for dynamic properties of PRV at the fixed inlet scenarios and not
during the reclosing process (API, 1993). To predict the dynamic Fig. 14. Dynamic pressure comparison of different models (Zhang et al., 2014).
properties of PRV during reclosing process, a simplified dynamic
model (SDM) was presented by Song et al. (Song et al., 2013). When calculate subcooled, supercritical, and overheated fluid by neglect-
the disc moves, PRV behaves as one degree-of-freedom system and ing phase separation effect. Secondly, assuming that bubble rising
motion of the moving part (stem and disc) was resolved based on and blowdown periods are comparable and ignoring neither dis-
force balance when open. The motion of value was simulated based charge rate nor the bubble rising velocity, bubble rising model
on second-order differential equation deduced from Newton’s sec- describes this process properly. Thirdly, a complete separation
ond law (Eq. 2). model is considered suitable for conditions when bubble rising
mẍ + c ẋ + Fs = Ff − Fg + fc (2) velocity is faster as compared to the discharge speed and when the
droplets falling velocity is higher than discharge speed for conden-
Where m, ẍ, and ẋ represent the mass of the moving parts, accel- sation because there is enough time for vapor dome development
eration, and velocity of the disc part in the moving direction in both the cases.
respectively. c represents the damping coefficient and Fs stands Supercritical CO2 and H2O were blown down initially from
for the spring force acting on the disc that was calculated using 313 K temperature and 25 MPa pressure for experimental inves-
k(x0 + x (t)). Ff , Fg , and fc are the flow force acting on the disc part, tigation and verification of the developed code. The experiments
the gravity of disc, and the coulomb friction parameters respec- were performed on a cylindrical vessel with a 4 m height, 2 m diam-
tively. One degree-of-freedom system based on the equation of eter and a 17 mm2 orifice for depressurization. The experimental
motion was calculated using Eq. 3. study and predicted pressure drop results with different models
as shown in Fig. 14. The R-square for pressure predictions is ≈
mẍ/ + c ẋ/ + k(x0 + x)/ = (/4).Cf . (3)
0.377 by complete separation model. The consequence of the ini-
Where k is the spring stiffness. and  are the reference force and tial condition on pressure drop was considered for different regions
pressure difference ratio. The dynamic numerical model was based divided by the relationship between corresponding pseudo-critical
on the major principles of the steady computational fluid dynamics and initial temperature. Furthermore, it was concluded that both
(CFD) analysis and rigid-body motion of the PRV. Afterwards, the void fraction and blowdown speed increase with the decrease in
results from a case study revealed that SDM predicted the blow- initial pressure and an increase of initial temperature, yet vice versa
down of conventional PRV reliably as shown in Fig. 13. In addition, for the fluid inventory.
the effect of reclosing/opening process and spring stiffness can also
be easily studied using this dynamic numerical model. 2.11. VBsim

2.10. J. Zhang To predict the depressurization behavior of vessel containing


two-phase (V-L) HC mixture with non-ideal behavior, Alessan-
To Simulate the blowdown process for the supercritical dro et al. described the modeling and experimental validation
water-cooled reactor (SCWR), a transient analysis code using a of the new unsteady model, VBsim (Vessel Blowdown Simula-
comprehensive numerical model was developed by Zhang et al. in tor) (D’Alessandro et al., 2015). The featured model accounts for
2013 (Zhang et al., 2014). Three different phase separation mod- non-equilibrium effects between constituent fluid phases such
els were implemented for the calculation of stagnation enthalpy of as temperature difference between phases. The numerical model
two-phase fluid. Firstly, a homogenous mixture model was used to evaluates dynamic fluid temperature, pressure, and composition
U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123 111

Fig. 15. VBsim and BLOWDOWN predictions vs experimental data (S9)


Fig. 16. Vessel wall temperature histories and comparison with other simulation
(D’Alessandro et al., 2015).
tools and numerical models (Park et al., 2018).

using EoS for non-ideal gas. SRK and PR EoS with Van der Waals mix-
ing rules have been implemented. Vapor phase discharge ( G, out ) nucleate boiling using PR EoS (Park et al., 2018). Mass and energy
was estimated using equations of motion for compressible fluid conservations for the fluid were calculated using equations pre-
through the orifice (Eqs. 4 and 5). sented by Speranza and Terenzi Speranza and Terenzi (2005). The
model consists of sub-algorithms of flash calculation, discharge rate
G, out = CDG G
is /MW (4)
calculation, and heat transfer rate calculation in order to estimate
⎧  (+1)/2(−1)  /(−1)
⎨ SG, or √pG 2/( + 1) p/pext ≥ 2/( + 1)
the realistic expansion path. The predicted results were compared
with the measured and simulated data from literature including
⎩ SG, or 1/ 2/( − 1)pG 1 − (−1)/   /(−1) (5)
is =
p/pext < 2/( + 1)
renowned simulation tools such as BLOWDOWN, BLOWSIM, VBsim,
BLOW, and commercial tools such as VessFire 1.2 and Aspen HYSYS
Where MW G is the vapor phase molar weight and C G is fixed to
D
v9. The model reasonably predicted the vapor temperature within
a default value of 0.84. G is the vapor density, SG, or is the orifice the experimental range while BLOWDOWN, VBsim, and HYSYS
area, p is the vessel pressure, pext is the atmospheric pressure, predicted 2–10 K higher temperature. Similarly, the model also
is the ratio of p & pext , and  is the isentropic coefficient of the real precisely predicted the wall temperature within the experimental
gas. The equation for an incompressible fluid through an orifice was range while VessFire v1.2 and BLOW predicted almost 2–8 K higher
used for the calculation of liquid phase discharge rate (Eqs. 6 and wall temperature in contact with vapor, and VBsim predicted 10 K
7). higher wall temperature in contact with liquid as shown in Fig. 16).

L, out = CDL L L
/MW (6)
  3. Design of blowdown system for pressure pipelines
L
= SL, or L 2 (p − pext )/L + gHL (7)
The blowdown of a pipeline and a vessel differs from each other
Where SL, or , L and HL are the orifice area, liquid density, and liq- due to a considerable pressure difference within the vessel but not
uid static head respectively. The MW L stands for liquid phase molar
within the pipeline. For blowdown of a vessel, pressure drop is
weight. The liquid discharge coefficient, CDL , is fixed to a default across the orifice, whereas for a pipeline blowdown, pressure drop
value of 0.61. The liquid and the vapor daughter phase compositions is along the line and across the orifice if it is small enough. In case
were calculated by xi and yi respectively. of blowdown of pipelines, the hazard emerges not only following
the low temperature generated in the pipeline walls but also due to
xi = zi /(1 − ˇ + ˇKi ’ ) (8)
the high efflux rates and large total efflux that arise when a massive
yi = Ki zi /(1 − ˇ + ˇKi ) (9) inventory in a typical line is depressurized (Richardson and Saville,
1993; Jarrah, 2016). To avoid accidents caused by these issues, sev-
Where Ki is the initial equilibrium coefficient and the overall vapor
eral numerical models have been developed. Later, some of them
fraction, ˇ, is calculated by solving the Rachford–Rice equation.
have also been coded as computerised program usually known as
For experimental validation of the developed model, the predicted
simulation tools. A timeline for the available simulation tools and
results were plotted against the experimental data from Haque et
numerical models capable to simulate the depressurization pro-
al. (PR and SRK) and HSE (SRK) (Haque et al., 1992a, b; Roberts
cess is illustrated in Fig. 17 followed by a brief description (Some of
et al., 2000). A comparison with the previously defined blowdown
the authors didn’t name their simulation tools; therefore, we have
model (BLOWDOWN) prediction was also made as shown in Fig. 15.
mentioned the names of those authors instead of the simulation
Considering the overall performance, VBsim produced reasonable
tool).
results with R-square ≈ 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 3.1. META
data.
Marginal stability two-fluid model (MSM) based on extended
2.12. A. Park Geurst’s variational principle to generalize multi-component two-
phase dispersion was presented by Chen et al. in 1995 (Geurst,
Previously, Haque et al. and Mahgerefteh et al. developed mod- 1985a, b; Geurst, 1986; Chen et al., 1995a). Thermodynamic equi-
els to study heat transfer between the mixture and the vessel walls. librium assumptions were made to develop the equations of motion
However, the correlation used and the value of the heat transfer and energy conversion for compressible single or multi-component
coefficient were unclear. Therefore, Park et al. (2018) presented vapor-liquid (V-L) mixture. To make the mathematical simulation
an elaborated heat transfer model including combined convec- of physical process possible and profound, the flow was considered
tion (forced and natural), multi-layer transient conduction, and the to be marginally stable to achieve hyperbolicity of the equations.
112 U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123

Fig. 17. Timeline for blowdown numerical models.

Fig. 18. Dynamic pressure changes’ histories at the closed end (Chen et al., 1993). Fig. 19. Dynamic temperature changes’ histories at the closed end (Chen et al.,
1993).

In the later volume of publication, the simplified numerical


method developed by Chen et al. in 1993 (Chen et al., 1993) ever, for a long pipeline blowdown, thermal non-equilibrium is
was modified and applied to the MSM for multi-component mix- found to be insignificant.
tures (Chen et al., 1995b). Subsequently, MSM and HEM were 3 Precise predictions of the phase equilibrium and thermodynamic
incorporated on the basis of simplified numerical method, and a behavior are important for the transient flow calculations.
computational package termed as META (Multi-component Equi-
librium Two-phase Analyser) was coded. A comparison of different 3.2. Fairuzov
case studies was made between HEM and the MSM with experi-
mental data for single or binary component two-phase blowdown. Previously, most of the work encompassed the modeling of
A 100 m long horizontal commercial steel pipeline with 0.15 m depressurization from vessels or relatively short pipelines (Lahey
inner diameter (ID) containing LPG (95 mol % propane and 5 mol and Moody, 1977; Reibold et al., 1981; Jackson et al., 1981). There-
% butane) was blown down. The testing temperature and pres- fore, a numerical simulation model for the depressurization of
sure were kept between 288–293 K (ambient temperature) and relatively large pipeline transporting multi-component flashing
800–2100 kPa respectively. The bore size was varied from 0.05 m to mixtures was coded by Fairuzov in 1998 (Fairuzov, 1998). The effect
full-bore in diameter and controlled with the circular orifice plate. of heat transfer between fluid flow and pipeline wall during blow-
Figs. 18 and 19 show the history of the pressure and temperature down phenomenon was also described. Heat transfer model, the
variations with respect to time at the closed ends. An R-square hydrodynamic model, and break-flow model were the building
≈ 0.957 was found in case of META-MSM CS for the temperature blocks of the featured model. No experimental study was con-
predictions. ducted, however, the simulation tool was validated against the
The comparison of HEM predictions and the measured data has already available experimental data from BP and Shell on the Isle of
been exceptionally reasonable. All the predicted temperature, pres- Grain (Cowley and Tam, 1988). The calculations were carried out on
sure, inventory, and void fraction transitions correspond with the an HP Apollo Series 700 working station using a double-precision
measured data. The only peculiarity was in case of the void frac- FORTRAN 77 code.
tion that was over-predicted after 7 s at the closed end. The most Fairuzov proposed a higher amount of heat transfer amongst
significant findings are presented as follows. fluid and pipe walls during the depressurization process, hence,
the adiabatic assumption for simulation became invalid. Therefore,
1 The simplified numerical method predicted reasonable results effect of thermal capacitance was integrated into the model using
for blowdown through experimental validations as shown in a different method in the development of energy conversion equa-
Figs. 18 and 19. tion. The study unveiled considerable effect of thermal capacitance
2 In case of a full-bore short pipeline blowdown (< 10 m), both of pipeline on the behavior of two-phase fluid flow. Consequently,
mechanical and thermal non-equilibriums are significant. How- thermal capacitance should not be ignored especially in case of the
U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123 113

Fig. 20. Dynamic pressure transitions at the open and closed ends of the line Fig. 22. Dynamic pressure histories at the open end for the P42 test LPG
(Fairuzov, 1998). (Mahgerefteh et al., 1999).

poor. The CNGS-MOC predicted the physical process with R-square


≈ 0.850 for pressure predictions.

3.4. A. Oke

Regardless of the fact that most common pipeline tragedies


include rupture, majority of the reported models so far are based
on 1-D axial flow. Others just treat the pipeline as a vessel discharg-
ing through an orifice. Therefore, Oke et al. developed a model that
can predict the transient release rate through the punctured plane
based on MOC (Oke et al., 2003). The model 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
Fig. 21. Dynamic temperature transitions at the open and closed ends of the line suitable phase and thermodynamic equilibrium data that is partic-
(Fairuzov, 1998).
ularly appropriate for high-pressure HC mixtures (DeReuck et al.,
1996; Assael et al., 1996).
flashing fluids. The dynamic pressure and temperature profiles are In case of 1-D unsteady flow, assuming phase and thermody-
reported in Figs. 20 and 21 respectively. Fig. 20 illustrates that the namic equilibrium among phases, the energy, momentum, and
model slightly over-predicted depressurization rate at the closed continuity conservation equations for a fluid in a rigidly clamped
end of pipe. However, when the fluid flow was considered to be adi- pipeline were taken from literature (Veersteg and Malalasekera,
abatic, the temperature at the open end was significantly predicted 1995). The continuity equation was reformulated as follows.
with R-square ≈ 0.914.
[T + ϕ] dP/dt − ϕ*dh/dt + 2 a2 T *∂u/∂x = 0 (10)

3.3. CNGS-MOC Where h and ϕ represent enthalpy and thermodynamic function


respectively.  and u are density and velocity as a function of time
An efficient mathematical model (CNGS-MOC) for the sim- (t) and distance (x) respectively. Using MOC, the conservation of
ulation of full-bore rupture (FBR), based on the method of continuity, energy, and momentum equations can be substituted
characteristics (MOC) was developed by Mahgerefteh et al. in 1999 with three compatibility equations along with their correspond-
(Mahgerefteh et al., 1999). The MOC was implemented to simu- ing characteristic lines. The isolated flow was modeled based on
late FBR or depressurization of a long pipe containing two-phase or termination of the inlet feed upon rupture. In case of non-isolated
condensable HC mixtures. This method was used to develop a more flow, inlet feed is supposed to be stopped after a prescribed period
suitable method than the Finite Difference Method (FDM) and Finite of time. Fig. 23 is schematic illustration of the fluid flow analysis
Element Method (FEM) because both had troubles in handling the following rupture and emergency isolation with various boundary
choking scenarios at the ruptured end. Moreover, the MOC is more conditions from B1 -B6 .
precise than FDM and FEM because it can resolve the blocked flow In Fig. 23, C stands for match line, C0 stands for path line, and
intrinsically using Mach line characteristics. The MOC significantly J represents solution joints. The simulation of pressure and veloc-
reduces number of iterations involved while solving the system of ity profiles specifies three separate fluid flow regimes predominant
simultaneous equations. The use of curved characteristics in con- within the pipe that manage the dynamic transitions of the release
junction with fast-mathematical algorithm and compound nested rate. For validation of the numerical model, the simulated outcomes
grid system leads to a dramatic reduction in CPU time in addition were plotted against the data from experiments performed by BP
to improved accuracy. The model was extensively validated using and Shell on the Isle of Grain (Cowley and Tam, 1988). Pressure
field data obtained during Isle of Grain (Richardson and Saville, and temperature transitions at closed and opened end are shown
1996) as well as Piper Alpha tragedy (Chen, 1993). The simulation in Figs. 24 and 25 respectively. The results ended up with a reason-
studies were performed on the basis of HEM assumptions. In Fig. 22, able prediction of the physical process with R-square ≈ 0.942 for
the predicted data with CNGS-MOC and field data are correlated temperature at the closed end as shown in Fig. 25.
with those based on other simulation tools including BLOWDOWN
(Haque et al., 1990), META-HEM (Chen et al., 1995a, b; Chen, 1993), 3.5. VPM
MSM-CS (Chen et al., 1995a, b; Chen, 1993) as well as PLAC (Hall
et al., 1993). Both META-HEM and CNGS-MOC produced reasonably Until 2011, no numerical model was developed to analyze
precise predictions while other models’ predictions were relatively the blowdown conditions for a vent pipeline. Therefore, Rajiwate
114 U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123

Fig. 23. Rupture plane and fluid flow study following line puncture (Oke et al., 2003).

Fig. 24. Pressure transitions at closed and opened end of P40 (LPG) test (Oke et al., Fig. 26. Measured upstream pressure and temperature (Clausen et al., 2012).
2003).

geometry, and assumed friction factor, Eq. 11 and 12 define the


flow (Parker, 1985).
The predicted results were then validated with the experimental
study performed on a 12 m long stainless-steel pipeline. A total of 9
experimental studies were conducted with an initial pressure and
temperature of 200–400 kPa and ≈ 19 ◦ C respectively. The experi-
mental data were plotted only for a steady-state period that showed
a good agreement with the predicted values.

3.6. OLGA

Most of the reported data on experimental study of CO2 blow-


Fig. 25. Temperature transitions at closed and opened end of P40 (LPG) test (Oke
down (Gebbeken and Eggers, 1996; de Koeijer et al., 2009; Clausen
et al., 2003). and Munkejord, 2012; de Koeijer et al., 2011; Cho et al., 2013)
describes findings from small-scale laboratory setups of flow loops
and vessels. Thus, it is of great interest for carbon capture and
developed a vent pipe model (VPM) to investigate the behavior of sequestration (CCS) community to collect field data from exist-
compressible gas in a vent pipeline using thermodynamic and fluid ing large-scale CO2 facilities. Therefore, a large-scale CO2 pipeline
dynamic approach (Rajiwate, 2011). Integrated with the GERG EoS, blowdown was carried out by Clausen et al. in 2012 while pointing
the use of REFPROP makes a highly adequate prediction with the the possible limitations of existing dynamic multi-phase flow sim-
VPM (Lemmon et al., 2009; Kunz et al., 2007). ulator (OLGA) (Clausen et al., 2012). An onshore buried 50 km long
Temperature, pressure, and density relations in terms of stagna- 24 ID pipeline containing 93,000 tonnes of CO2 (>99 % Pure) was
tion properties were defined previously (Bansal, 2005; Saad, 1993; initially depressurized from supercritical conditions. The initially
Shapiro, 1954). Mass efflux prediction for pipe is the central step reported temperature and pressure were 304 K and 8.1 MPa respec-
for the development of blowdown model for vent system in terms tively. The blowdown process was modeled using OLGA V5.3.2, a
of static pressure and static temperature that is expressed as: multi-phase thermo-hydraulic simulation tool developed by SPT

Group (Bendiksen et al., 1991). The new coded module of OLGA for
G = PM /ZRT (11)
CO2 single component was used that utilizes Span-Wagner EoS for
The above equation for mass flux in terms of stagnation proper- pure CO2 (Span and Wagner, 1996; Aursand et al., 2013).
ties was represented as: 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


 (+1)/2(−1)
G = Po / To * /ZRM/ 1 + ( − 1)/2 *M 2 (12) experimental data showed that the blowdown stayed well above
the CO2 triple point, and no traces of solidification were found as
Where To , Po are the stagnation temperature and pressure respec- shown in Fig. 26. However, the minimum simulated temperature
tively,  is the specific heat ratio, and Z, R and M are the was almost 17 K lower than the minimum measured temperature.
compressibility factor, universal gas constant and match number The deviation between simulated and measured blowdown path
respectively. For a given downstream and stagnation pressure, reveals inability of the simulation tool to handle CO2 along with
U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123 115

Table 1
Predominant conditions and pipeline characteristics for the COOLTRANS shock tube
test.

Pipeline characteristics Initial conditions

Pipeline length 144 m Fluid temperature 278.35 K


Internal diameter 150 mm Fluid pressure 15.335 MPa
Pipeline wall roughness 0.05 mm Ambient pressure 0.101 MPa

Fig. 27. Pressure transition of simulated and measured data for 92 % CO2–8 % N2
blowdown (Huh et al., 2014).

Fig. 29. Pressure-time profile at 0.1 m from release end of the CO2PipeHaz FBR test
(Brown et al., 2013).

∂E/∂t + ∂u(E + P)/∂z = −u(fw u2 /Dp )(Energy Balance) (15)

Where , Dp , P, u, and fw are the mixture density, pipeline diame-


ter, pressure, velocity, and fanning friction factor respectively. All of
these factors were evaluated using Chen’s correlation (1979) (Chen,
Fig. 28. Temperature changes of simulated and measured data for 92 % CO2–8 % N2 1979) as function of space (z) and time (t). E stands for the energy of
blowdown (Huh et al., 2014). the mixture. For the non-equilibrium V-L transition, HRM was used
with the mechanical equilibrium assumption such that no phase
impurities. Even a very small amount of impurities will disturb the slip was maintained. The non-equilibrium V-L transition equation
blowdown path. with thermodynamic equilibrium for the vapor mass fraction is
To further study the effect of impurities on CO2 blowdown and presented in Eq. 16.
reliability of the OLGA tool, Huh et al. conducted experimentation
∂x/∂t + u(∂x/∂z) = (xeq − x)/ (16)
in 2014 with 2–8 % N2 addition by mass fraction (Huh et al., 2014).
The experimental apparatus comprises of the pressurization part, Where x represents the dynamic vapor quality and stands for
liquefying part, mixing zone, and the test section. Total length of relaxation time accounting for the phase change transition delay.
the testing apparatus was 51.96 m with 3.86 mm ID. The temper- Angielczyk et al. proposed an empirically determined correlation
ature and pressure measurements were observed at four different for the relaxation time as given in Eq. 17 (Angielczyk et al., 2010).
positions as illustrated in Figs. 27 and 28. However, the simula-  −0.54  −1.76
tion tool did not predict temperature drop precisely. The results = 2.15 × 10−7 x(/sv (Ps (Tin ) − P)/(Pc − Ps (Tin )) (17)
show that the numerical simulator predicted better pressure drop Where Pc , sv and Ps are critical pressure, saturated vapour density
at higher concentration of impurities. Furthermore, the R-square at given pressure, and saturation pressure at given temperature
value at initial portion (7.93 m) of pipeline is ≈ 0.697. While, at respectively. To validate numerical flow model’s reliability, the
last portion (51.85 m), the temperature predictions are much pre- hypothetical shock tube test was conducted initially for both single
cise with R-square ≈ 0.859. It can be inferred that the numerical (Liquid) and two-phase CO2. Table 1 presents the initial conditions
simulator requires further improvements in phase change model. and pipeline characteristics for the COOLTRANS FBR dense-phase
CO2 shock tube test. Moreover, CO2PipeHaz CO2 pipeline rup-
3.7. S. Brown ture experiment conditions were used as an experimental study
(Mahgerefteh et al., 2011).
The commonly reported vessel/pipeline outflow numerical A 37 m long with 40 mm ID instrumented pipeline carrying pure
models were based on HEM assumptions where the fluid phase was CO2 was blown out using an explosive charge. Pipeline’s initial con-
considered to be at the mechanical and thermal equilibrium. Con- ditions were 7 MPa pressure and 298.35 K temperature. In general,
sequently, there was no-equilibrium V-L transition and phase-slip HRM produced plausibly good estimations of the experimental data
phenomena, that is, delayed bubble formation was ignored. In 2013, especially during the initial period of blowdown as shown in Fig. 29.
Brown et al. extended the prior work by establishing and validat- R-square for pressure predictions is ≈ 0.924 that indicates reliabil-
ing homogeneous relaxation model (HRM) for depressurization of ity of the model. It was observed that FBR clearly indicates fully
dense phase CO2 pipeline (Brown et al., 2013). Initially, the phases dispersed flow and, hence, the phase-slip can’t be ignored.
were assumed to be at mechanical and thermal equilibrium, hence,
the mass, momentum, and energy conservation were presented by 3.8. S. Martynov
Eqs. 13–15 respectively.

∂/∂t + ∂u/∂z = 0(Conservation of Mass) (13) Since most of the prior reported blowdown simulation models
have been developed for V-L phase depressurization, they were not
∂u/∂t + (∂u2 + P)/∂z = −fw u2 /Dp (14) appropriate for simulating solid CO2 release. Martynov et al. in 2013
116 U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123

Table 2
Initial experimental conditions of the release.

Initial fluid condition Test 1 Test 2 Test 3

Pressure/MPa 8.0 8.6 5.3


Temperature/K 307 312 290
Inventory/tonne 6.5 4 6.5

Fig. 31. Comparison of numerical and experimental (PT-40, TT-40) dynamic tem-
perature changes for different compositions at 50 m (Drescher et al., 2014).

Fig. 30. Time variation of the pressure measured in test 3 (Martynov et al., 2014).

extended HEM for the development of outflow model accounting


for CO2 solidification during blowdown (Martynov et al., 2014). A
set of equations describing the HEM flow including mass conver-
sion, and momentum and energy balance equations were used. For
mass conservation, Eq. 13 was used. Eqs. 14 and 15 were modified
to Eqs. 18 and 19 to calculate the momentum and energy balance
respectively.
Fig. 32. Comparison of numerical and experimental (PT-40, TT-40) dynamic pres-
∂u/∂t + (∂u2 + P)/∂z = −2f w u2 /D(Momentum Balance) (18) sure changes for different compositions at 50 m (Drescher et al., 2014).

∂E/∂t + ∂u(E + p)/∂z = 4qw /D − 2f w u3 /D(Energy Balance) (19)


To calculate the properties of vapor-solid (V-S) and vapor-liquid experiments with different CO2–N2 compositions to study the
(V-L) equilibrium mixtures formed during the depressurization of potency of the modified HEM for a multi-component system
vessel/pipeline containing CO2, previously developed PR EoS by (Drescher et al., 2014). For simulation study, PR EoS was used and
Martynov et al. (Martynov et al., 2013) was extended and applied. experiments were performed on a 140 m long horizontal tube with
10 mm ID. Initially, the binary mixture in the supercritical region
p = RT/(v − b) − a/(v (v + b) + (b (v − b)) (20)
was at 12 MPa pressure and a temperature of approximately 293 K.
Where R, T , p, and v are the universal gas constant, temperature, The numerical results for all three fluid compositions regarding the
pressure, and specific volume respectively. a and b are the empiri- relative and absolute value match the experimental results very
cal parameters accounting for the intermolecular attraction forces well as illustrated in Figs. 31 and 32. With the increase in compo-
and molecular volume respectively. The instantaneous heat flux at sition, the model predictions become more precise as maximum
the pipe wall was calculated based on pipeline wall temperature R-square ≈ 0.919 for temperature prediction is noted in case of 90
transitions with time using Eq. 21. % composition. Despite few complexities in the HEM employed, it
average average still contains several simplifications in view of the comparatively
qw = w Cp,w ıw *((dTw )/dt) (21) reasonable results. Further work could involve the advanced heat-
average transfer models considering the effect of flow regimes or boiling.
Where Tw represents the average pipeline wall temperature.
w , Cp,w , and ıw stand for density, pipeline heat capacity, and wall
thickness of pipeline respectively.
4. Major findings
The featured numerical model was validated against experi-
ments conducted by Dalian University of Technology (DUT) under
Table 3 presents a matrix of modeling outlooks presented in
CO2PipeHaz project. The model precisely predicts the pressure
literature that could aid in a holistic simulation of blowdown use
transitions in the pipeline for initially supercritical fluid (Test 1 and
in the industrial sector. Based on the literature reviewed, Table 3
2). The initial experimental conditions for all the case studies are
illustrates the followings:
presented in Table 2.
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 pres- • Commonly, the models focused on pressure and mixture temper-
sure predictions as shown in Fig. 30. Moreover, the release model ature transitions with time to understand the blowdown process.
also predicted the dry ice formation and explained the transition • Limited studies on vessel wall temperature and void fraction
phenomena from V-L to V-S equilibrium at a triple point of CO2. are also available, however, further research is required when
applied to multi-phase processes.
3.9. M. Drescher • HEM is extensively used to solve the blowdown process with
thermodynamic assumptions.
HEM was validated mostly against the experimental data from • Mostly, the models were validated against previously published
CO2 pressure release. However, Dresher et al. (2014) executed experimental data, especially for pure or impure CO2.
U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123 117

Table 3
Summary of modeling approaches.

Temperature
Title Pressure Flowrate Void Fraction
Mixture Wall
√ √ √ √ √
BLOWDOWN
√ √ √ √
BLOWSIM ×
√ √ √ √
P. S. Cumber ×
√ √ √ √
A. Fredenhagen ×

SDM × × × ×
√ √ √ √
J. Zhang ×
√ √ √ √
VBsim ×
√ √ √ √
A. Park ×
√ √ √ √
META ×
√ √ √ √
Fairuzov ×
√ √ √
CNGS-MOC × ×
√ √ √ √
A. Oke ×
√ √ √
OLGA × ×
√ √ √ √
S. Brown ×
√ √ √ √
S. Martynov ×
√ √ √
M. Drescher × ×

Fig. 33. Comparative illustration of R2 values for different models.

Fig. 34. MG bias values for different models.

Some of the numerical models or simulation tools produced perfect model is said to have a geometric mean bias equal to 1.0.
incredibly precise predictions of the physical phenomenon occur- A geometric mean bias less than 1.0 refers to underprediction and
ring during the blowdown process. The BLOWSIM produced most above 1.0 refers to overprediction (Hanna et al., 1993).
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 abil- 5. Summary
ity to simulate some of the significant parameters. On the other
hand, BLOWDOWN has the ability to predict most of the impor- Different simulation tools and numerical models for the cal-
tant parameters accurately. Cumber and J. Zhang produced poor culation of optimum parameters during blowdown process are
results with R-square lower than 0.4. A bar chart (Fig. 33) illus- reviewed. Table 4 summarizes the operating conditions, vessel
trates R-square value for most of the reviewed simulation tools dimensions, characteristics, and key challenges. Several novel
or numerical models. For the quantitative comparison of reviewed emerging blowdown models have been reviewed and a compar-
models, the geometric mean bias is taken into account (Fig. 34). A ative study has been presented.
118
Table 4
Comparative analysis of different simulation tools and numerical models.

Name/Title Year Composition Condition Vessel/Pipeline Volume Orifice Size (mm) Characteristics Key Challenges Ref

Models + Tools
BLOWDOWN N2, 70 % N2 – 30 % CO2, 1.5 m long vessel with - Good understanding - No work in cryogenic (Haque et al., 1990;
1990 150 bar & 20 ◦ C 6.325 of the physical conditions or
(1990) NG/C3 & HC 10.6 diameter Haque et al., 1992a, b)
processes occurring emergency
- Ability to make - Use of the extended

U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123
predictions with an principle of
acceptable corresponding states
uncertainty for generating the
- Based on a three pertinent fluid
fluids model (vapor, thermophysical
liquid and free properties
water) - A simulation study is
restricted to
depressurization
under ambient
conditions only
- Not fully
documented

Case 1: C1 –C6 + 4.03 % 117 bar & 10 ◦ C 40 km long pipeline 0.4191 - Good understanding - Homogeneity (Richardson and
Model + Tool 1991 of the physical assumption when
N2 Saville, 1991)
Case 2: C1 –C5, C1 O, 200 bar & 40 ◦ C 20000 m long pipeline 0.4064 processes occurring the flow is
C2O + 3 % H2O - Ability to make two-phase
predictions with an - Quasi-steadiness
acceptable assumption
uncertainty - No work on
- Suitable for planned cryogenic conditions
blowdown

1996 LPG (95 % C3 – 5 % C4) 11.3-22.5 bar & 100 m long pipeline 35-154 - Predictions have - No work on Natural (Richardson and
13.3-20 ◦ C with 2-6 diameter been shown to be in gas Saville, 1996)
at least adequate, - No work in cryogenic
and often good, conditions or
agreement with the emergency situation
Isle of Grain
measurements

META (1994) LPG (95 % propane & 5 - Well prediction of - More efficient
1995 8 to 21 bar & 15 to 20 ◦ C 100 m in length, 0.15 m 0.05 m thermodynamic
(Chen et al., 1995a, b)
Model + Tool % butane experimental data
measurement
methods needed
FRICRUP (1993-94) 1993 Air, CO2, carbonated 69-138 bar & 37.8 ◦ C 25.5 km long pipeline 1.5 to 12.7 for a bottle - Predict P, T, time and - Did not predict very (Norris and Puls, 1993;
Model + Tool 1994 water, CH4 and HHC (Vessel) and 40 bar & with 7.992 diameter & 84 for Pipeline mass flowrate. well for multi-phase Norris, 1994)
15.5 ◦ C (pipeline). - For single and flow
multi-phase - Can predict
- For Vessel/Pipeline
multi-component Conditions
- No work on Natural
gas
Table 4 (Continued)

Name/Title Year Composition Condition Vessel/Pipeline Volume Orifice Size (mm) Characteristics Key Challenges Ref

Fairuzov 1998 LPG (95 % propane & 5 11.25 bar & 19.9 ◦ C 100 m long pipeline Full-bore rupture - Multi-component, - Thermal equilibrium (Fairuzov, 1998)
(Model + Tool) % butane with 150 mm internal two-phase pipe flow between the fluid
diameter model and pipe wall was
- Effect of heat not reached
transfer is discussed - Rigorous conjugated
- For non-adiabatic heat transfer model

U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123
fluid flow should be used for
- Well prediction of transient heat
experimental data transfer description

BLOWSIM 1999 64 % C1 , 6 % C2, 28 % 116 atm & 293 K N/A 10 mm - Multi-component, - Limited data on (Mahgerefteh and
(Model + Tool) n-C3 & 2 % n-C4 the multi-phase pipe experimental Wong, 1999)
flow model validation
- BLOWSIM performs - No work on Natural
better than gas
BLOWDOWN - Thermal and
- The cubic equation mechanical
of state has positive equilibrium
effects on data assumption (Brown
et al., 2013)
- No work in cryogenic
conditions or
emergency situation

CNGS-MOC 1999 LPG (95 % propane & 5 11.25 bar & 19.9 ◦ C 100 m long pipeline FBR - Reasonably accurate - No Experimental (Mahgerefteh et al.,
(Model + Tool) % butane with 150 mm internal predictions Work 1999)
diameter - Less CPU time - Less data on
- improving accuracy experimental
validation

P. S. Cumber 2001 LPG (C1 66.5 %, C2 3.5 % 120 bar & 25 ◦ C 3.24 m height with 10 mm - Precise prediction of - No experimental (Cumber, 2001)
(Model + Tool) & C3 30 % 0.54 m ID pressure and mass study conducted
flow rate - Imprecise
- Less computational temperature
time prediction
- No heat transfer
through vessel wall
assumption

A. Fredenhagen 2001 N2 – CO2 25-15 bar & 298 K 0.05 m3 vessel with 29-17 mm2 - A good prediction of - Tool commercially (Fredenhagen and
(Model + Tool) 0.242 m diameter the process not available Eggers, 2001)
- Applicable for - Unable to predict
Multi-component equipment
process conditions
- Limited data on
experimental
validation
- No data on
multi-phase process

119
120
Table 4 (Continued)

Name/Title Year Composition Condition Vessel/Pipeline Volume Orifice Size (mm) Characteristics Key Challenges Ref

H. Mahgerefteh 2002 64 % C1 , 6 % C2, 28 % 116 bar & 293 K 3.24 m vessel with 10 mm - Prediction of - No experimentation (Mahgerefteh et al.,
(Model) n–C3 & 2 % n–C4 1.13 m diameter and thermodynamic validation 2002)
59 mm wall thickness properties under fire - No simulation tools
and ambient developed
conditions
- For

U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123
multi-component
and multi-phase
process
- Vessel conditions
calculations

A. Oke (Model) 2004 LPG 21.6 bar & 20 ◦ C 100 m long with 150 - The model accounts - No experimental
0.154 m ID for axial and radial study performed
flow, real fluid - No validation for
behavior as well as natural gas mixture
the locality of and cryogenic
puncture with conditions
respect to the length
of pipe

VPM (Model) 2011 78.12 % N2, 20.96 % O2 200-400 kPa and ≈ 120 m long pipeline N/A - For - It can only predict (Rajiwate, 2011)
& 0.92 % Ar 19 ◦ C multi-component steady state
process conditions
- Ability to calculate - It is not suitable for
pipe wall the multi-phase
temperature process

OLGA (1991) 50 km long, 24 - Experimental data - Non-reliable


2012 CO2 81 bar & 31 ◦ C 8 from large-scale CO2 predications of (Clausen et al., 2012)
pipeline
unit experimental data
- Initially supercritical - No work on
conditions multi-component &
- Multi-phase process multi-phase process
- No study on the
Tool
2014 CO2, 2–8 % N2 85b bar & 20 ◦ C 51.96 m tube with N/A - Studied Effect of effect of initial (Huh et al., 2014)
3.86 mm ID impurities on CO2 conditions
blowdown
- Good predictions of
initial pressure drop

J. Zhang 2013 CO2 & Water 25 MPa & 380 ◦ C 4 m height & 2 m 0.008 m2 - A good prediction of - Limited data on (Zhang et al., 2014)
diameter the process experimental
- Supercritical validation
conditions - No work on natural
- Multi-component gas
and multi-phase - Only work on
process supercritical
conditions
Table 4 (Continued)

Name/Title Year Composition Condition Vessel/Pipeline Volume Orifice Size (mm) Characteristics Key Challenges Ref

15.335 MPa & 278.35 K 144 m long pipeline - Model predictions - Lack of sufficient and
S. Brown (Model) 2013 CO2 FBR (Brown et al., 2013)
with 150 mm diameter produced a reliable
reasonable experimental data
agreement - No work on the
- Multi-phase process multi-component
- Delayed phase process
transition has an - No work on natural
insignificant effect gas and cryogenic

U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123
on the conditions or
7 MPa & 298.35 K 37 m long pipeline depressurization rate emergency situation
with 40 mm ID of pipe
S. Martynov 2013 CO2 53-86 bar & 34-39 ◦ C 256 m long pipeline 50 mm - Good prediction for - Unable to predict (Martynov et al., 2014)
(Model) with 233 mm diameter initially supercritical initially two-phase
fluid fluid
- Prediction of dry ice - No work on
formation cryogenic or
emergency
conditions
- Experimental study
performed only for
single component

M. Drescher 2014 CO2 + 10–30 % N2 120 bar & 20 140 m long pipeline 9.5 mm - Good predication of - Require (Drescher et al., 2014)
with 10 mm ID experimental results advancement of
- Experimental heat-transfer models
validation of HEM - Temperature
with binary prediction was less
components accurate

VBsim 2015 N2 CO2–HC (up to C1 O) 290-323 K & 118.5-22 8.429 m length with 0.635-1.4 cm - Good predication of - No experimental (D’Alessandro et al.,
& HC (up to C3) bar 1.97 m diameter experimental results study performed. 2015)
- For - No study in the
multi-component cryogenic region.
and multi-phase
process
- Consider
non-equilibrium
effects between
constituent fluid
phases

293-303 K & 3.24 m length with - Two-phase - No experimental


A. Park 2017 HC Mixture 10 mm (Park et al., 2018)
117.5-120 bar 1.13 m ID multi-component work performed No
system prediction. study in the
- Can predict wall cryogenic region.
temperature
contracting liquid
and vapor.
- More precise
predictions than
previously available
models

121
122 U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123

6. Conclusion Byrnes, W., Reid, R., Ruccia, F., 1964. Rapid depressurization of gas storage cylinder.
Ind. Eng. Chem. Process. Des. Dev. 3, 206–209.
Chen, J.-R., 1993. Modelling of Transient Flow in Pipeline Blovvdown Problems.
In this paper, the available models for the optimum blow- University of London.
down parameters’ calculation are discussed. The challenges for Chen, N.H., 1979. An explicit equation for friction factor in pipe. Ind. Eng. Chem.
depressurization in process industry through a review of existing Fund. 18, 296–297.
Chen, J., Richardson, S., Saville, G., 1993. A simplified numerical method for transient
literature are also highlighted. The available models are capable two-phase pipe flow. Trans IChemE 71, 304–306.
of solving the blowdown scenarios, however, there are still some Chen, J., Richardson, S., Saville, G., 1995a. Modelling of two-phase blowdown from
limitations listed as follows. pipelines—i. A hyperbolic model based on variational principles. Chem. Eng. Sci.
50, 695–713.
Chen, J., Richardson, S., Saville, G., 1995b. Modelling of two-phase blowdown from
• Only a few lab-scale experimental studies are reported for the pipelines—II. A simplified numerical method for multi-component mixtures.
Chem. Eng. Sci. 50, 2173–2187.
validation of models. The reliability of models will be improved
Cho, M.-I., Huh, C., Jung, J.-Y., Kang, S.-G., 2013. Experimental study of N2 impu-
if also validated against the large scale experimental facilities. rity effect on the steady and unsteady CO2 pipeline flow. Enrgy Proced. 37,
• More extensive experimental studies on natural gas and other 3039–3046.
gaseous mixtures could help to improve the validity of models. Clausen, S., Munkejord, S.T., 2012. Depressurization of CO2–a numerical benchmark
study. Energy Procedia 23, 266–273.
• Limited studies are reported for the blowdown of pressure vessels Clausen, S., Oosterkamp, A., Strøm, K.L., 2012. Depressurization of a 50 km long 24
under fire conditions. inches CO2 pipeline. Enrgy Proced. 23, 256–265.
• Experimental studies for blowdown from the cryogenic con- Cowley, L., Tam, V., 1988. Consequences of pressurized LPG releases: the Isle of
grain full scale experiments. In: Submitted 13th Int. Conf LNG/LPG Conference
ditions could also be considered. It would help to study the and Exhibition, Kuala Lumpur, Malaysia.
blowdown process under worst case scenarios especially when a Cui, H., Wang, W., Li, A., Li, M., Xu, S., Liu, H., 2010. Failure analysis of the brittle
phase change is involved that can lead to the formation of solids. fracture of a thick-walled 20 steel pipe in an ammonia synthesis unit. Eng. Fail.
• A new numerical model can also be developed to overcome the Anal. 17, 1359–1376.
Cullen, W., 1990. The Public Inquiry Into the Piper alpha Disaster. Department of
problems related to dry ice formation during the blowdown pro- Energy, HMSO, London, UK.
cess. Cumber, P.S., 2001. Predicting outflow from high pressure vessels. Process Saf. Env-
• Studies on blowdown of vessels/pipelines through multiple iron. 79, 13–22.
D’Alessandro, V., Giacchetta, G., Leporini, M., Marchetti, B., Terenzi, A., 2015.
chokes needs to be considered. Modelling blowdown of pressure vessels containing two-phase hydrocarbons
• To avoid solidification during blowdown, the effect of additives mixtures with the partial phase equilibrium approach. Chem. Eng. Sci. 126,
719–729.
can be studied.
de Koeijer, G., Borch, J.H., Jakobsenb, J., Drescher, M., 2009. Experiments and mod-
eling of two-phase transient flow during CO2 pipeline depressurization. Energy
In addressing these research areas, industry should be empow- Proced. 1, 1683–1689.
de Koeijer, G., Borch, J.H., Drescher, M., Li, H., Wilhelmsen, Ø, Jakobsen, J., 2011.
ered to make effective and sustainable retrofit decisions while CO2 transport–Depressurization, heat transfer and impurities. Energy Procedia
simultaneously reducing hazards and improving the safety. This 4, 3008–3015.
paper has focused on solutions for industrial hazards related to DeReuck, K., Craven, R., Elhassan, A., 1996. In: Millat, J., Dymond, J.H., Nieto de Cas-
tro, C.A. (Eds.), Transport Properties of Fluids: Their Correlation, Prediction and
depressurization. However, the outcome of the identified research Estimation. IUPAC. Cambridge Univ. Press, Cambridge.
areas has wide-ranging applications with techniques that could be Drescher, M., Varholm, K., Munkejord, S.T., Hammer, M., Held, R., de Koeijer, G., 2014.
applied to solve many other system simulation challenges. Experiments and modelling of two-phase transient flow during pipeline depres-
surization of CO2 with various N2 compositions. Energy Proced. 63, 2448–2457.
Eggers, R., Green, V., 1990. Pressure discharge from a pressure vessel filled with CO2 .
Acknowledgements J. Loss Prev. Process Ind. 3, 59–63.
Fairuzov, Y.V., 1998. Blowdown of pipelines carrying flashing liquids. Aiche J. 44,
245–254.
The authors would like to extend their most profound grati- Fisher, H., Forrest, H., Grossel, S.S., Huff, J., Muller, A., Noronha, J., et al., 2010. Emer-
tude to the CO2 Research Centre (CO2RES), Universiti Teknologi gency Relief System Design Using DIERS Technology: The Design Institute for
Emergency Relief Systems (DIERS) Project Manual. John Wiley & Sons.
PETRONAS (UTP), Malaysia, for the accomplishment of this Review Fletcher, S., 2001. US Senate ready to act on pipeline safety as public attention
article. sharpens. Oil Gas J. 99, 58–60.
Fredenhagen, A., Eggers, R., 2001. High pressure release of binary mixtures of
CO2 and N2. Experimental investigation and simulation. Chem. Eng. Sci. 56,
References 4879–4885.
Gebbeken, B., Eggers, R., 1995. Thermohydraulic phenomena during vessel release of
Hanna, S., Chang, J., Strimaitis, D., 1993. Hazardous gas model evaluation with field initially supercritical carbon dioxide. International Symposium Two-Phase Ow
observations. Atmos. Environ. Part A Gen. Top. 27, 2265–2285. Modelling and Experimentation.
AirConditioning, H., 2002. Refridgeration-The News. Boiler Accident Report : To Err Gebbeken, B., Eggers, R., 1996. Blowdown of carbon dioxide from initially supercrit-
Is Human, November 8th. ical conditions. J. Loss Prev. Process Ind. 9, 285–293.
Angielczyk, W., Bartosiewicz, Y., Butrymowicz, D., Seynhaeve, J.-M., 2010. 1-D Mod- Gebbeken, B., Eggers, R., 1997. Investigations on vessel blowdown and flashing of
eling of Supersonic Carbon Dioxide Two-Phase Flow through Ejector Motive pressurized carbon dioxide. Proc of the 4 World Conference on Experimental
Nozzle. Heat Transfer, Fluid Mechanics and Thermodynamics (ExHFT4), 1709–1716.
Api, R., 1990. Recommended Practice., pp. 521. Geurst, J., 1986. Variational principles and two-fluid hydrodynamics of bubbly liq-
API, 1993. Recommended Practice 520 Part 1—Sizing and Selction, sixth ed. uid/gas mixtures. Physica A 135, 455–486.
Assael, M.J., Trusler, J.M., Tsolakis, T.F., 1996. Thermophysical Properties of Fluids: Geurst, J., 1985a. Virtual mass in two-phase bubbly flow. Physica A 129, 233–261.
An Introduction to Their Prediction. World Scientific. Geurst, J., 1985b. Two-fluid hydrodynamics of bubbly liquid/vapour mixture includ-
Aursand, P., Hammer, M., Munkejord, S.T., Wilhelmsen, Ø., 2013. Pipeline transport ing phase change. Philips J. Res. 40, 352–374.
of CO2 mixtures: models for transient simulation. Int. J. Greenh. Gases Control Gradle, R., 1984. Design of gas pipeline blowdowns. Energy Proc. Cdn., 15–20.
15, 174–185. Hall, A., Butcher, G., Teh, C., 1993. Transient simulation of two-phase hydrocarbon
Bansal, R.K., 2005. Fluid Mechanics and Hydraulic Machines, 9 ed. Laxmi Publications flows in pipelines. European Two-Phase Flow Group Meeting (Hannover), paper,
(P) Ltd, New Delhi, ed. p. I4.
Barma, M., Saidur, R., Rahman, S., Allouhi, A., Akash, B., Sait, S.M., 2017. A review on Haque, A., Richardson, S., Saville, G., Chamberlain, G., 1990. Rapid depressurization
boilers energy use, energy savings, and emissions reductions. Renew. Sustain. of pressure vessels. J. Loss Prev. Process Ind. 3, 4–7.
Energy Rev. 79, 970–983. Haque, M., Richardson, S., Saville, G., 1992a. Blowdown of pressure vessels. I. Com-
Bendiksen, K.H., Maines, D., Moe, R., Nuland, S., 1991. The dynamic two-fluid model puter model. Trans IChemE, Part B, Process Saf. Environ. Prot. 70, 3–9.
OLGA: theory and application. SPE Prod Eng. 6, 171–180. Haque, M., Richardson, S., Saville, G., Chamberlain, G., Shirvill, L., 1992b. Blowdown of
Bond, J., 2002. Institute of Chemical Engineers Accidents Database. Institute of Chem- pressure vessels. II. Experimental validation of computer model and case studies.
ical Engineers., Rugby. UK. Trans IChemE, Part B, Process Saf. Environ. Prot. 70, 10–17.
Brown, S., Martynov, S., Mahgerefteh, H., Proust, C., 2013. A homogeneous relaxation Huh, C., Cho, M.-I., Hong, S., Kang, S.-G., 2014. Effect of impurities on depressurization
flow model for the full bore rupture of dense phase CO2 pipelines. Int. J. Greenh. of CO2 pipeline transport. Energy Procedia 63, 2583–2588.
Gases Control 17, 349–356.
U. Shafiq, A.M. Shariff, M. Babar et al. / Process Safety and Environmental Protection 133 (2020) 104–123 123

Jackson, J., Liles, D., Ransom, V., Ybarrondo, L., 1981. LWR system safety analysis. Richardson, S., Saville, G., 1991. Blowdown of Pipelines. Offshore Europe: Society of
Nuclear Reactor Saf. Heat Transf. 12, 415. Petroleum Engineers.
Jarrah, A.A., 2016. Modeling and simulation of rapid depressurizations of hydrocar- Richardson, S., Saville, G., 1996. Blowdown of LPG pipelines. Process Saf Environ. 74,
bon gas mixture flowing at unsteady high velocity in piping transport system. 235–244.
Int. J. Ind. Eng. (SSRG-IJIE)., 3. Roberts, T., Medonos, S., Shirvill, L., 2000. Review of the Response of Pressurised
Kunz, O., Klimeck, R., Wagner, W., Jaeschke, M., 2007. The GERG-2004 Wide-range Process Vessels and Equipment to Fire Attack.
Equation of State for Natural Gases and Other Mixtures. Fortschr. Ber VDI, VDI- Saad, M.A., 1993. Compressible Fluid Flow, 2 ed. Prentice Hall, New Jersey, ed.
Verlag, Düsseldorf. Saville, G., Richardson, S., Barker, P., 2004. Leakage in ethylene pipelines. Process Saf.
Ladokun, T., Nabhani, F., Zarei, S., 2010. Accidents in Pressure Vessels: Hazard Aware- Environ. 82, 61–68.
ness. International Association of Engineers. Selby, C., Burgan, B., 1998. Blast and Fire Engineering for Topside Structures-phase
Lahey, R.T., Moody, F., 1977. The Thermal Hydraulics of a Boiling Water Nuclear 2: Final Summary Report. SCi publication.
Reactor (Monograph Series on Nuclear Science and Technology). Shapiro, A.H., 1954. The Dynamics and Thermodynamics of Compressible Fluid Flow,
L-b, Ouyang, Aziz, K., 1996. Steady-state gas flow in pipes. J. Petrol Sci. Eng. 14, vol2.
137–158. Simon, L.L., Introvigne, M., Fischer, U., Hungerbühler, K., 2008. Batch reactor opti-
Lemmon, D.E.W., Huber, D.M.L., McLinden, D.M.O., 2009. REFPROP Manual. U S mization under liquid swelling safety constraint. Chem. Eng. Sci. 63, 770–781.
Department of Commerce, Boulder. Skouloudis, A.N., 1992. Benchmark exercises on the emergency venting of vessels.
Lydell, B.O., 2000. Pipe failure probability—the Thomas paper revisited. Reliab. Eng. J. Loss Prev. Process Ind. 5, 89–103.
Syst. Saf. 68, 207–217. Soave, G., 1972. Equilibrium constants from a modified Redlich-Kwong equation of
Mahgerefteh, H., Wong, S.M., 1999. A numerical blowdown simulation incorporating state. Chem. Eng. Sci. 27, 1197–1203.
cubic equations of state. Comput. Chem. Eng. 23, 1309–1317. Song, X.-G., Park, Y.-C., Park, J.-H., 2013. Blowdown prediction of a conventional
Mahgerefteh, H., Falope, G.B., Oke, A.O., 2002. Modeling blowdown of cylindrical pressure relief valve with a simplified dynamic model. Math. Comput. Model.
vessels under fire attack. AIChE J. 48, 401–410. 57, 279–288.
Mahgerefteh, H., Saha, P., Economou, I.G., 1999. Fast numerical simulation for full Span, R., Wagner, W., 1996. A new equation of state for carbon dioxide covering the
bore rupture of pressurized pipelines. Aiche J. 45, 1191–1201. fluid region from the triple-point temperature to 1100 K at pressures up to 800
Mahgerefteh, H., Fairweather, M., Falle, S., Melheim, J., Ichard, M., Storvik, MPa. J. Phys. Chem. Ref. Data 25, 1509–1596.
I., et al., 2011. CO2Pipehaz: Quantitative Hazard Assessment for Next Speranza, A., Terenzi, A., 2005. Blowdown of Hydrocarbons Pressure Vessel with Par-
Generation CO2 Pipelines. Institute of Chemical Engineers, Rugby, tial Phase Separation. Applied and Industrial Mathematics in. World Scientific,
pp. 606–610. Italy, pp. 508–519.
Martynov, S., Brown, S., Mahgerefteh, H., Sundara, V., Chen, S., Zhang, Y., 2014. Talbi, B., 2017. CO2 emissions reduction in road transport sector in Tunisia. Renew.
Modelling three-phase releases of carbon dioxide from high-pressure pipelines. Sustain. Energy Rev. 69, 232–238.
Process Saf. Environ. 92, 36–46. Tam, V., Higgins, R., 1990. Simple transient release rate models for releases of pres-
Martynov, S., Brown, S., Mahgerefteh, H., 2013. An extended Peng-Robinson equation surised liquid petroleum gas from pipelines. J. Hazard. Mater. 25, 193–203.
of state for carbon dioxide solid-vapor equilibrium. Greenhouse: Sci. Technol. 3, The National Board of Boiler and Pressure Vessel Inspectors, URL: 2002. National
136–147. Board Bulletin. http://www.nationalboard.org.
Milano, J., Ong, H.C., Masjuki, H., Chong, W., Lam, M.K., Loh, P.K., et al., 2016. Microal- Tolloczko, J., 1992. Interim Guidance Notes for the Design and Protection of Topside
gae biofuels as an alternative to fossil fuel for power generation. Renew. Sustain. Structures Against Explosion and Fire: Steel Construction Institute.
Energy Rev. 58, 180–197. Transporter B. Boiler, URL 2001. Pressure Vessel Accidents Increase. http://
Moss, D.R., 2004. Pressure Vessel Design Manual. Elsevier. bulktransporter.com.
Norris III, H., 1994. Hydrocarbon blowdown from vessels and pipelines. SPE Annual Twu, C.H., Coon, J.E., Cunningham, J.R., 1992. A new cubic equation of state. Fluid
Technical Conference and Exhibition: Society of Petroleum Engineers. Phase Equilibra 75, 65–79.
Norris III, H., Puls, R., 1993. Single-phase or multiphase blowdown of vessels or Veersteg, K., Malalasekera, W., 1995. An introduction to computational fluid dynam-
pipelines. SPE Annual Technical Conference and Exhibition: Society of Petroleum ics: the finite volume method. Pearson Schweiz Ag. 20, 400.
Engineers. Witlox, H., Holt, A., 1999. A unified model for jet, heavy and passive dispersion
Oke, A., Mahgerefteh, H., Economou, I., Rykov, Y., 2003. A transient outflow model including droplet rainout and re-evaporation. CCPS International Conference &
for pipeline puncture. Chem. Eng. Sci. 58, 4591–4604. Workshop on Modeling the Consequences of Accidental Releases of Hazardous
Onyebuchi, V.E., Kolios, A., Hanak, D.P., Biliyok, C., Manovic, V., 2017. A systematic Materials, Sept.
review of key challenges of CO2 transport via pipelines. Renew. Sustain. Energy Witlox, H., Holt, A., Contract 44003900 for Exxon Mobil 2001. Validation of the
Rev. Unified Dispersion Model against Kit Fox Field Data.
Park, A., Ko, Y., Ryu, S., Lim, Y., 2018. Numerical modeling of rapid depressurization Witlox, H.W., Harper, M., Oke, A., Stene, J., 2014. Phast validation of discharge and
of a pressure vessel containing two-phase hydrocarbon mixture. Process Saf. atmospheric dispersion for pressurised carbon dioxide releases. J. Loss Prev.
Environ. 113, 343–356. Process Ind. 30, 243–255.
Parker, G., 1985. ‘Pop’safety valves: a compressible flow analysis. Int. J. Heat Fluid Witlox, H.W., Fernandez, M., Harper, M., Oke, A., Stene, J., Xu, Y., 2018. Verification
Flow 6, 279–283. and validation of Phast consequence models for accidental releases of toxic or
Peng, D.-Y., Robinson, D.B., 1976. A new two-constant equation of state. Ind. Eng. flammable chemicals to the atmosphere. J. Loss Prev. Process Ind. 55, 457–470.
Chem. Fundam. 15, 59–64. Wong, M.A., 1998. Development of a Mathematical Model for Blowdown of Vessels
Rajiwate, F.L.H., 2011. Investigation of Compressible Fluid Behaviour in a Vent Pipe Containing Multi-component Hydrocarbon Mixtures. University of London.
During Blowdown. Curtin University. Woodward, J.L., Papadourakis, A., 1995. Reassessment and reevaluation of rainout
Reibold, W., Reocreux, M., Jones, O., 1981. Blowdown phase. Nuclear Reactor Safety and drop size correlation for an aerosol jet. J. Hazard. Mater. 44, 209–230.
Heat Transfer., 325–378. Woodward, J.L., Cook, J., Papadourakis, A., 1995. Modeling and validation of a dis-
Richardson, S., Saville, G., 1993. Blowdown of vessels and pipelines. Institution persing aerosol jet. J. Hazard. Mater. 44, 185–207.
of Chemical Engineers Symposium Series: Hemsphere Publishing Corporation, Zhang, J., Zhu, D., Tian, W., Qiu, S., Su, G., Zhang, D., 2014. Depressurization study of
195-. supercritical fluid blowdown from simple vessel. Ann. Nucl. Energy 66, 94–103.

View publication stats

You might also like