MODELING OF HYDROGEN EXPLOSION ON A PRESSURE SWING ADSORPTION FACILITY

Similar documents
Simulations of hydrogen releases from high pressure storage systems

SIMULATIONS OF HYDROGEN RELEASES FROM A STORAGE TANKS: DISPERSION AND CONSEQUENCES OF IGNITION

CFD SIMULATIONS OF VENTILATION EFFECT ON HYDROGEN RELEASE BEHAVIOR AND COMBUSTION IN AN UNDERGROUND MINING ENVIRONMENT

Hydrogen and fuel cell stationary applications: key findings of modelling and experimental work in the HYPER project

Comparison of Large-Scale Vented Deflagration Tests to CFD Simulations for Partially Congested Enclosures

PRAGMATIC ASSESSMENT OF EXPLOSION RISKS TO THE CONTROL ROOM BUILDING OF A VINYL CHLORIDE PLANT

Engineering Models for Vented Lean Hydrogen Deflagrations

Practical Modelling & Hazard Assessment of LPG & LNG Spills

Modelling Hazardous Consequences of a Shale Gas Well Blowout

Annexure 2: Rapid Risk Assesment RISK ANALYSIS

Deborah Houssin-Agbomson, Jean-Yves Letellier, Philippe Renault, Simon Jallais

HYDROGEN - METHANE MIXTURES : DISPERSION AND STRATIFICATION STUDIES

CFD Modelling of Blast Waves from BLEVEs

Simulation and modelling of water spray in the 3D explosion simulation program FLACS. Elin Kristin Dale

Fire and Safety for Offshore drilling and production Ajey Walavalkar ANSYS Inc.

Measurement and simulation of the flow field around a triangular lattice meteorological mast

APPENDIX A: SENSITIVITY TESTS INPUTS AND SETUP TABLES

ACCIDENTAL HYDROGEN RELEASE IN GC-LABORATORY; A CASE STUDY

SUMMARY OF THE EXPERIMENTAL STUDIES OF COLD HELIUM PROPAGATION ALONG A SCALE MODEL OF THE LHC TUNNEL

Experimental determination of deflagration explosion characteristics of methane air mixture and their verification by advanced numerical simulation

I.CHEM.E. SYMPOSIUM SERIES NO. 97 BUOYANCY-DRIVEN NATURAL VENTILATION OP ENCLOSED SPACES

Direct Numerical Simulation on Hydrogen Fuel Jetting from High Pressure Tank

REQUIREMENTS FOR VALIDATION OF MATHEMATICAL MODELS IN SAFETY CASES

S.A. Klein and G.F. Nellis Cambridge University Press, 2011

HYDROGEN RISK ANALYSIS FOR A GENERIC NUCLEAR CONTAINMENT VENTILATION SYSTEM

Politecnico di Torino. Porto Institutional Repository

Investigation on Divergent Exit Curvature Effect on Nozzle Pressure Ratio of Supersonic Convergent Divergent Nozzle

Study on Intensity of Blast Wave Generated from Vessel Bursting by Gas Explosion

A METHOD FOR ASSESSING THE CONSEQUENCES OF SMALL LEAKS IN ENCLOSURES

MODELLING OF FUME EXTRACTORS C. R.

Internal Explosion Methodologies

DEFINING HAZARDOUS ZONES ELECTRICAL CLASSIFICATION DISTANCES

DETERMINATION OF HAZARDOUS ZONES FOR A GENERIC HYDROGEN STATION A CASE STUDY

SPONTANEOUS IGNITION OF PRESSURIZED HYDROGEN RELEASE

USE OF THE EXCEEDANCE CURVE APPROACH IN OCCUPIED BUILDING RISK ASSESSMENT

ANALYSIS OF HEAT TRANSFER THROUGH EXTERNAL FINS USING CFD TOOL

Study on wind turbine arrangement for offshore wind farms

CFD SIMULATIONS OF GAS DISPERSION IN VENTILATED ROOMS

The Influence Of Vessel Volume And Equivalence Ratio In Vented Gas Explosions

DETERMINATION OF HAZARDOUS ZONES FOR A GENERIC HYDROGEN STATION A CASE STUDY

The Comparison of CFD with a Traditional Method Used in an Incident Investigation Case Happened in Taiwan

Experimental Study of Hydrogen Release Accidents in a Vehicle Garage

EFFECTIVENESS EVALUATION OF FACILITIES PROTECTING FROM HYDROGEN-AIR EXPLOSION OVERPRESSURE

Tube rupture in a natural gas heater

Blast Damage Consideratons for Horizontal Pressure Vessel and Potential for Domino Effects

RESIDENTIAL WATER DISTRIBUTION

Investigation of Suction Process of Scroll Compressors

VENTED CONFINED EXPLOSIONS INVOLVING METHANE/HYDROGEN MIXTURES

PERFORMANCE OF A FLAPPED DUCT EXHAUSTING INTO A COMPRESSIBLE EXTERNAL FLOW

Characteristics of Decompression Tank Internally Pressurized With Water Using OpenFOAM Syamsuri 1, a

Identification and Screening of Scenarios for LOPA. Ken First Dow Chemical Company Midland, MI

VENTED EXPLOSION OF HYDROGEN / AIR MIXTURE: AN INTER COMPARISION BENCHMARK EXERCISE

EXPERIMENTAL STUDY OF HOT INERT GAS JET IGNITION OF HYDROGEN-OXYGEN MIXTURE

Tutorial. BOSfluids. Relief valve

Single Phase Pressure Drop and Flow Distribution in Brazed Plate Heat Exchangers

COMPUTATIONAL FLOW MODEL OF WESTFALL'S LEADING TAB FLOW CONDITIONER AGM-09-R-08 Rev. B. By Kimbal A. Hall, PE

Improving Accuracy of Frequency Estimation of Major Vapor Cloud Explosions for Evaluating Control Room Location through Quantitative Risk Assessment

DEVELOPMENT OF HIGH ALTITUDE TEST FACILITY FOR COLD JET SIMULATION

MODELING AND SIMULATION OF VALVE COEFFICIENTS AND CAVITATION CHARACTERISTICS IN A BALL VALVE

New Findings for the Application of Systems for Explosion- Isolation with Explosion Venting

CFD Modelling of LPG dispersion

AIRFLOW GENERATION IN A TUNNEL USING A SACCARDO VENTILATION SYSTEM AGAINST THE BUOYANCY EFFECT PRODUCED BY A FIRE

SHAPA TECHNICAL PAPER 10 (Revised) SIZING OF EXPLOSION RELIEF VENTS DAVID BURN FIKE UK LTD

Experimental Analysis on Vortex Tube Refrigerator Using Different Conical Valve Angles

SPONTANEOUS IGNITION PROCESSES DUE TO HIGH-PRESSURE HYDROGEN RELEASE IN AIR

SAFETY OF LABORATORIES FOR NEW HYDROGEN TECHNIQUES

Vented gas and dust explosions

EXPERIMENTAL STUDY ON THE DISCHARGE CHARACTERISTICS OF SLUICE FOR TIDAL POWER PLANT

A STUDY OF THE LOSSES AND INTERACTIONS BETWEEN ONE OR MORE BOW THRUSTERS AND A CATAMARAN HULL

HYDROGEN RELEASES IGNITED IN A SIMULATED VEHICLE REFUELLING ENVIRONMENT

ON THE EFFECT OF LIFT FORCES IN BUBBLE PLUMES

An experimental study of internal wave generation through evanescent regions

Efficiency Improvement of Rotary Compressor by Improving the Discharge path through Simulation

Uncertainty in the analysis of the risk of BLEVE Fireball in process plants and in transportation

A study of heat transfer effects on air pollution dispersion in street canyons by numerical simulations

Investigation of Cargo Tank Vent Fires on the GP3 FPSO, Part 2: Analysis of Vapour Dispersion

A Computational Assessment of Gas Jets in a Bubbly Co-Flow 1

Abstract. Geometry Obstructions Ventilation Wind conditions

CFD Based Approach for VCE Risk Assessment

Transient Analyses In Relief Systems

Basis of Structural Design

Evaluation of Northwest Citizen Science Initiative (NWCSI) March 7,2015 Report Summary

OPTIMIZATION OF INERT GAS FLOW INSIDE LASER POWDER BED FUSION CHAMBER WITH COMPUTATIONAL FLUID DYNAMICS. Abstract. Introduction

1. Air is blown through a pipe AB at a rate of 15 litre per minute. The cross-sectional area of broad

OPTIMIZING THE LENGTH OF AIR SUPPLY DUCT IN CROSS CONNECTIONS OF GOTTHARD BASE TUNNEL. Rehan Yousaf 1, Oliver Scherer 1

NEAR FIELD EFFECTS OF SMALL SCALE WATER BLEVE

Surrounding buildings and wind pressure distribution on a high rise building

Prediction of gas explosion overpressure interacting with structures for blast-resistant design

A Numerical Simulation of Fluid-Structure Interaction for Refrigerator Compressors Suction and Exhaust System Performance Analysis

Influence of rounding corners on unsteady flow and heat transfer around a square cylinder

EXPLOSION VENTING OF BUCKET ELEVATORS

This is a repository copy of The effect of vent size and congestion in large-scale vented natural gas/air explosions.

3D Turbulence at the Offshore Wind Farm Egmond aan Zee J.W. Wagenaar P.J. Eecen

Behaviour of a highly pressurised tank of GH2 submitted to a thermal or mechanical impact

WALL BOILING MODELING EXTENSION TOWARDS CRITICAL HEAT FLUX. ABSTRACT

Numerical simulation and analysis of aerodynamic drag on a subsonic train in evacuated tube transportation

The effect of back spin on a table tennis ball moving in a viscous fluid.

A Study on the Effects of Wind on the Drift Loss of a Cooling Tower

THE EXTERNAL EXPLOSION CHARACTERISTICS OF VENTED DUST EXPLOSIONS.

Rotary air valves used for material feed and explosion protection are required to meet the criteria of NFPA 69 (2014)

Transcription:

MODELING OF HYDROGEN EXPLOSION ON A PRESSURE SWING ADSORPTION FACILITY B. Angers 1, A. Hourri 1, P. Benard 1, E. Demaël 2, S. Ruban 2, S. Jallais 2 1 Institut de recherche sur l hydrogène, Université du Québec à Trois-Rivières, Québec, Canada, Benjamin.Angers@uqtr.ca, Ahmed.Hourri@uqtr.ca, Pierre.Benard@uqtr.ca 2 Air Liquide, Centre de Recherche Claude-Delorme, 7835 Jouy en Josas, France ABSTRACT This article presents a numerical study of the consequences of an hydrogen release from a pressure swing adsorption installation operating at 3 barg. The simulations are performed using FLACS/Hydrogen from GexCon. The impact of obstruction, partial confinement, leak orientation and wind on the explosive cloud formation (size and explosive mass) and on explosion consequences is investigated. In parallel, overpressures resulting from ignition are calculated as a function of the time to ignition. 1. INTRODUCTION The overpressures associated with an hydrogen cloud ignition represent a significant hazard for large scale hydrogen production facilities. In the event of an unintentional ignition, the presence of many obstacles and the high degree of confinement inherent to these facilities greatly increase the risk of higher flame velocities which in turn generate overpressure waves. The study of these hazards is of significant interest to hydrogen suppliers in order to establish and validate mitigation procedures [1,2]. In this context, we investigate the impact of confinement and turbulence inducing mechanisms, such as wind and obstacles, on the overpressure generated during the combustion of a hydrogen gas cloud arising from a Pressure Swing Adsorption (PSA) facility. Dispersion simulations results for a jet release from three 3.4 barg reservoirs using the commercial FLACS/Hydrogen program are presented. The simulation scenarios considered are two hydrogen leaks from orifices with diameters 2 mm and 35 mm in a still environment (without wind) and in windy environment where the wind velocities are 3 m/s and 5 m/s. For each scenario, two hydrogen jet directions have been considered. In the first case, the jet was oriented such that it impinges on the ground with an angle of 45 and in the second case, the jet was directed horizontally. Critical profiles of hydrogen concentration in air are presented. We also present multiple scenarios where the jet is ignited close to the ground. The overpressure resulting from the cloud ignition is estimated as a function of the leak direction and time to ignition. 2. MODELLING SCENARIO DESCRIPTION As shown in Figure 1, the PSA installation consisted of 12 linked pressurized pipes. The average distance (pitch) between adjacent pipes is presented in Table 1. The pitch was measured center to center minus the radius of adjacent pipes. Only pipes with a radius of more than.168 m were considered for the pitch calculation. The volume blockage ratio (VBR), area blockage ratio (ABR) and the piping surface area per volume for different regions of the installation are given in Table 2. The VBR and ABR are respectively defined as the percentage of volume and surface blocked by objects in a given region. The piping surface area per volume given by FLACS cofile utility is defined as the objects surface area divided by the volume of the region considered. Figure 2 and Figure 3 show volumes and areas used for the VBR, ABR and piping surface area calculation.

Figure 1. Pressure swing adsorption installation (PSA) viewed from different angles. Table 1. The average distance (pitch) between pipes inside the central region of the PSA in each direction Direction X Y Z Pitch (m).52.37.65 Figure 2. Each box shows the volume used to calculate the VBR of the central region on the left, the central wide region in the middle and the reservoir region on the right. Figure 3. The left, middle and right picture show the area used to calculate the central, the central wide and the reservoir ABR respectively.

Table 2. Volume blockage ratio (VBR), area blockage ratio (ABR) and piping surface area Region Central Central wide Reservoir Total volume (m3) 43.8 741.1 4572.6 PSA volume (m3) 19.2 33.6 873.3 VBR (%) 4.8 4.5 19.1 Total area (m2) 13.1 24. 148. PSA area (m2) 4.8 6.3 68.2 ABR (%) 36.5 26.2 46.1 Piping surface area (m2) 93.9 116.6 784.1 Piping surface area per volume (m2/m3).233.157.171 The leak was assumed to originate from a broken branch connection at one end of the system. The position of the leak is shown in Figure 4. Three of the 43.3 m 3 reservoirs contain hydrogen at a pressure of 3.4 barg and at a temperature of 45 C. The leak diameters considered, i.e. 2 mm and 35 mm, correspond to a rupture at a branch connection of 3/4" and 1"1/2 respectively. As a conservative approximation, the pressure drop through the pipes from the reservoirs to the leak was not taken into account. Figure 4. A branch connection was selected as leak origin. The leak is located.5 m above the ground. For clarity, the following nomenclature was used to define the different scenarios: 35NW45 C2 means that the leak diameter is 35 mm, there is No Wind, the jet direction is set 45 toward the ground, it is a Combustion simulation and the time to ignition is set at 2 sec after the leak onset. The jet cloud ignition was assumed to occur on the ground inside the 3% concentration envelop. For most simulations, the cloud was ignited 2 seconds and 2 seconds after the leak onset. At 2 seconds, the cloud is small but contains a higher concentration of hydrogen, while at 2 seconds, the cloud is well developed but more diluted and it covers most of the PSA. In one case (35NW45 ), the time to ignition was varied between.5 second and 6 seconds. For the scenarios with wind, the atmospheric stability Pasquill class F was chosen for 3 m/s wind, and class D for 5 m/s wind. In these cases, the reference height was set at 1 m and the ground roughness at.3 m (open flat terrain; grass, few isolated obstacles). The wind direction was the same as that of the jet. FLACS uses a rectilinear grid. In the case of jet dispersion, a zone made of cubic cells was defined at the expanded jet location. Following Gexcon good practice recommendations [3], the cell size of the initial cubic zone at the expanded jet location was determined by the effective leak area A leak so that A cv > A leak >.9A cv where A cv is the area of the control volume. The leak originates from one cell in the initial cubic zone. From that initial zone, the grid was stretched to a coarser cubic grid encompassing the PSA. The size of the central region cells encompassing the PSA was set at.5 m for

both dispersion and combustion grids. It was then stretched again outside the PSA to a coarser rectangular grid. One grid was used for all combustion simulations. Combustion grid dependency studies were done for the windless case scenario with 35 mm leak diameter directed 45 toward the ground at 2 seconds, by setting the size of the central region cells encompassing the PSA to.25 m. In this case, the difference in the maximum overpressure does not exceed 18% and -4% in the domain and on the monitor points when respectively compared with the overpressures obtained with.5 m cells. On the other hand, with an ignition time of 2 seconds the maximum overpressure was lowered with.25 m cells by -29% and -34% in the domain and on the monitor points when respectively compared with the overpressures obtained with.5 m cells. Gexcon recommend that the reactive flammable cloud be resolved with a minimum of 1 grid cells in the direction where the cloud meets confinement. Due to our choice of using the same grid for all the combustion simulations, this rule was broken in the cases where the ignition time was lower. For each scenario, the flow is choked at the jet exit. The jet outlet conditions, i.e. the leak rate, temperature, effective leak area, velocity and the turbulence parameters (i.e. turbulence intensity and turbulent length scale) for the flow, are calculated using an imbedded jet program in FLACS. FLACS can also calculate the time dependent leak and turbulence parameters for continuous jet releases in the case of high pressure vessel depressurization. The estimate assumes isentropic flow conditions through the nozzle, followed by a single normal shock (whose properties are calculated using the Rankine- Hugoniot relations), which is subsequently followed by expansion into ambient air [4]. For the 2 mm and 3 mm leak diameter, the initial mass flow rate was.52 kg/s and 1.539 kg/s respectively. A discharge coefficient of.85 was used. Overpressure measurements were taken at regular intervals along and perpendicular to the jet direction. Monitoring points were positioned on the ground and 1. m above the ground as shown on Figure 5. Figure 5. Monitoring points position

3. MODELING RESULTS Figure 6 shows the transversal cut going vertically along the axis of the jet through the hydrogen flammable contour (at the lower flammability limit (LFL): 4% molar fraction H 2 in air) at 2 seconds and 2 seconds after the leak onset for the dispersion scenarios where d = 35 mm. Simulation time: 2 seconds Simulation time: 2 seconds 35NW45 35NWHor 35W3F45 35W3FHor 35W5D45 35W5DHor Figure 6. Hydrogen molar concentration envelop profile (up to 3% in red) of the lower flammability limit contour (4% in blue) along the jet direction (Y = m) for a leak diameter of 35 mm. As shown on Figure 6 for similar cases, the wind does not influence significantly the shape of higher concentrations contours (red zones near the ground). On the other hand, the lower concentrations contours are greatly affected by the wind. They extend horizontally instead of dispersing upward when wind is present. Figure 7 shows the mass histogram taken from the combustion grid for the 35NW45 scenario at 2 sec and 2 sec. Flammable mass at 2 sec release time is lower (near 3 kg) than at 2sec (near 12kg).

However, the flammable cloud is more concentrated in hydrogen at 2 sec compared to 2 sec and is then more reactive. Figure 8 presents the maximum overpressures measured during a series of combustion simulations based on the same scenario (35NW45 ) where the time to ignition was varied from.5 sec to 6 sec. As shown, these overpressures were measured at the monitoring points as well as in the entire domain. Figure 8 also shows the hydrogen mass at stoechiometric concentration, right before the cloud was ignited. It should be noted that we could not extract from FLACS the hydrogen mass at stoechiometric concentration. We can only extract mass values from concentration intervals. However, an approximate of the stoechiometric mass of hydrogen extracted from the interval 28-32% (vol.) was plotted in Figure 8 as a function of time to ignition. Overpressure and hydrogen mass at stoechiometric concentration peaks are observed at 2 sec and 2 sec as illustrated in Figure 8. Simulation time: 2 seconds Simulation time: 2 seconds % of total mass 2 18 16 14 12 1 8 6 4 2 9.34 13.15 15.69 12.95 17.54 15.3 7.24 3.68 1.6 1.62.64.18.19.21.21.21..26 % of total mass 6 5 4 3 2 1 [4; 8[ [8; 12[ [12; 16[ [16; 2[ [2; 24[ [24; 28[ [28; 32[ [32; 36[ [36; 4[ [4; 44[ [44; 48[ [48; 52[ [52; 56[ [56; 6[ [6; 64[ [64; 68[ [68; 72[ [72; 76] 54.85 [4; 8[ [8; 12[ [12; 16[ [16; 2[ [2; 24[ [24; 28[ [28; 32[ [32; 36[ [36; 4[ [4; 44[ [44; 48[ [48; 52[ [52; 56[ [56; 6[ [6; 64[ [64; 68[ [68; 72[ [72; 76] 16.96 11. 7.93 4.45 2.38 1.4.45.38.28.3.4.4.9..4.5. Hydrogen mole fraction interval (% (vol)) Hydrogen mole fraction interval (% (vol)) Figure 7. Mass histogram from combustion grid, prior to ignition for the 35NW45 scenario. Maximum OP on Monitor points Maximum OP in Domain Hydrogen mass.35.35 Maximum overpressure (barg).3.25.2.15.1.5..3.25.2.15.1.5. Hydrogen mass at stoichiometric concentration (kg) 1 2 3 4 5 6 Time to ignition (sec) Figure 8. Maximum overpressure and hydrogen mass at stoechiometric concentration (between 28-32% (vol.)) as a function of time to ignition for the 35NW45 scenario.

The relation between time to ignition and maximum overpressure was thoroughly studied [5]. As shown on Figure 8, the highest overpressure was measured with a time to ignition of 2 sec in agreement with the experimental findings of Takeno et. al. [6] who concluded that short time to ignition resulted in higher overpressure. Royle et. al. [7] observed the same phenomena in their experiment. As shown on Figure 7, while the total amount of flammable fuel is lower at 2 sec (2.98 kg) than at 2 sec (14.19 kg), the hydrogen mass at stoechiometric concentration is higher at 2 sec (.21 kg) than at 2 sec (.15 kg). Note also on Figure 8 that the peaks in hydrogen mass at 2 sec and at 2 sec follow the same trend as the maximum overpressures. Figure 9 shows the flammable mass of hydrogen at two concentrations intervals: 4-75% and 11-75%. The 11 75% interval is useful because it correspond to the hydrogen concentrations for which a strong flame acceleration is observed by Dorofeev [8]. For both intervals, the flammable mass reaches a maximum at 2 sec. Flammable mass between 4-75% (vol.) Flammable mass between 11-75% (vol.) 16 14 Flammable mass of hydrogen (kg) 12 1 8 6 4 2 1 2 3 4 5 6 Time to ignition (sec) Figure 9. Flammable mass of hydrogen between mole fraction intervals of 4-75% (vol.) and 11-75% (vol.), right before the cloud was ignited, as a function of time to ignition for the 35NW45 scenario. The maximum overpressures for all combustion simulations are presented in Figure 1 and Figure 11. As shown on Figure 1, maximum overpressures were 5 to 6 times more powerful for horizontal jets than for the 45 jets. We see the same trend on Figure 11 for maximum overpressures measured on monitor points, although the ratio is only 2. As shown on Figure 6, in the horizontal case, most of the stoechiometric concentration contours are near the ground under the PSA. The most reactive part of the cloud is outside the congested area where most of the pipes accelerating the flame are. This is particularly true for jets with d = 2 mm and for jets oriented 45 toward the ground. On the other hand, the flammable concentration envelop encompasses the congested area much more for the horizontal jet than for the 45 jets.

Monitor Points 2.5 2.5 Domain Maximum Overpressure (bar) 2 1.5 1.5 Maximum Overpressure (bar) 2 1.5 1.5 45, no wind horizontal, no wind 45, wind 3F horizontal, wind 3F 45, wind 5D horizontal, wind 5D 5 1 15 2 5 1 15 2 Time to Ignition (sec) Time to Ignition (sec) Figure 1. Maximum overpressure as a function of time to ignition measured on all monitor points and in the domain for simulations with d = 35 mm..25 Monitor Points.25 Domain Maximum Overpressure (bar).2.15.1.5 Maximum Overpressure (bar).2.15.1.5 45, no wind horizontal, no wind 45, wind 3F horizontal, wind 3F 45, wind 5D horizontal, wind 5D 5 1 15 2 5 1 15 2 Time to Ignition (sec) Time to Ignition (sec) Figure 11. Maximum overpressure as a function of time to ignition measured on all monitor points and in the domain for simulations with d = 2 mm. Overpressure extents at 5 mbar, 14 mbar and 2 mbar are given in Figure 12. Isocontours on the XZ view plane (along jet axis) and XY view plane (1 m above ground) for the scenario (35NW45 ) at 5 mbar are shown on Figure 13, at various times after ignition.

Time to ignition = 2 sec Time to ignition = 2 sec Maximum Distance from Release Point (m) 45 4 35 3 25 2 15 1 5 5 1 15 2 Overpressure (mbar) Maximum Distance from Release Point (m) 45 4 35 3 25 2 15 1 5 5 1 15 2 Overpressure (mbar) 35NW45 35NWHor 35W3F45 35W3FHor 35W5D45 35W5DHor 2NW45 2NWHor 2W3F45 2W3FHor 2W5D45 2W5DHor Figure 12. Maximum travel distance of 5 mbar, 14 mbar and 2 mbar overpressure fronts measured from the origin of the leak at (,, ). XZ Plane (Y= m) XY Plane (Z =.5 m) 35NW45 C2 t =.128 sec 35NW45 C2 t =.171 sec Figure 13. 5 mbar isocontours at various times after ignition on the XZ plane (along jet axis) and XY plane (1 m above ground) for the 35NW45 scenario. For the 45 angle release with the 35 mm hole, even if the cloud is more reactive at 2 sec, the safety distances are lower than in the 2 sec case. At the opposite, in the horizontal release, these safety distances become larger at 2 sec compared to 2 sec case due to the entire flame propagation in the congested part of the PSA.

4. CONCLUSION We presented results of finite volume simulations for a time-dependent hydrogen release from a broken branch connection.5 m above the ground, originating from three 43.3 m 3 reservoirs at 3.4 barg of a pressure swing adsorption installation. An overpressure of.33 bar was obtained with a leak diameter of 35 mm oriented 45 toward the ground, with no wind. For all scenarios, a maximum overpressure of 2.5 bar was obtained with a leak diameter of 35 mm, oriented horizontally, with no wind, when ignited 2 sec after the leak onset. For the same scenario, a 5 mbar pressure wave travelled up to 4 m in the direction of the leak and 26 m perpendicular to it. The effects of a 3 m/s and 5 m/s wind were considered. In the presence of a 3m/sec Pasquill class F wind, a maximum overpressure of 1.7 bar was obtained with a leak diameter of 35 mm, oriented horizontally and ignited 2 sec after the onset of the leak. In this case, the 5 mbar pressure wave travelled 35 m along the jet axis and 24 m perpendicular to the jet. Two orifice diameters were considered in this study, namely 2 mm and 35 mm. An average maximum overpressure of.2 bar was obtained for most scenarios with d = 2 mm. On the other hand, with d = 35 mm, the maximum overpressure varied from.28 bar up to 2.5 bar depending on the jet orientation, wind strength and ignition time. The consequences of ignition shortly after the start of the release were also presented. The time of ignition was varied from.5 sec to 6 sec for one case. Short time to ignition usually results in higher overpressure, based on the amount of hydrogen at stoechiometric concentration. This was mainly noticeable for the simulations with d = 35 mm, less with d = 2 mm. ACKNOWLEDGEMENTS We gratefully acknowledge the support of Air Liquide. We thank Olav Roald Hansen of GexCon for useful discussions. We also thank Fernando Gomez, Tommy Henly from IRH for their help with the geometry importation and the programming of data analysis tools. REFERENCES 1. B. Angers, A. Hourri, P. Bénard, P. Tessier and J. Perrin, Simulation of Hydrogen Releases from a High Pressure Reservoir : Dispersion and Consequences of Ignition. Proceedings of the International Conference on Hydrogen Safety 25, Pisa, Italy. 2. Hourri, A., Angers, B., Bénard, P., Surface effects on flammable extent of hydrogen and methane jets, International Journal of Hydrogen Energy, 29; 34: 1569-1577. 3. FLACS V9 User s guide, Gexcon, 29. 4. Houf, W., Winters, W., Evans, G., Approximate Inflow Conditions for Subsonic Navier-Stokes Computations of Under-Expanded Jets, Unpublished manuscript. 5. Angers, B., Hourri, A., Bénard, P., Hydrogen dispersion and ignition from high pressure storage systems, IRH, Report, July 27. 6. Takeno K., Okabayashi K., Ishinose T., Koushi A., Nokana T., Hashiguchi K. and Chitose K.: Phenomena of dispersion and explosion of high pressurized hydrogen, Proceedings of the International Conference on Safety 25, Pisa, Italy. 7. M. Royle, D.B. Willoughby, Consequences of catastrophic releases of ignited and unignited hydrogen jet releases, International Journal of Hydrogen Energy, 211; 36: 2688-2692. 8. S. B. Dorofeev Flame acceleration and explosion safety applications, Proceedings of the Combustion Institute, Volume 33, Issue 2, 211, Pages 2161-2175.