High pressure jet

Last updated
Contour animation of the mole fraction variation (from 0.025 to 0.05) of a Natural Gas jet as it impinges a steel tank done through CFD. CFD simulation of a high pressure NG jet impinging a steel tank.gif
Contour animation of the mole fraction variation (from 0.025 to 0.05) of a Natural Gas jet as it impinges a steel tank done through CFD.

A high pressure jet is a stream of pressurized fluid that is released from an environment at a significantly higher pressure than ambient pressure from a nozzle or orifice, due to operational or accidental release. [1] In the field of safety engineering, the release of toxic and flammable gases has been the subject of many R&D studies because of the major risk that they pose to the health and safety of workers, equipment and environment. [2] Intentional or accidental release may occur in an industrial settings like natural gas processing plants, oil refineries and hydrogen storage facilities. [2]

Contents

A main focus during a risk assessment process is the estimation of the gas cloud extension and dissipation, important parameters that allow to evaluate and establish safety limits that must be respected in order to minimize the possible damage after a high pressure release. [3]


Mechanism and structure of a gaseous jet

Subsonic and sonic flow

When a pressurized gas is released, the velocity of the flow will heavily depend on the pressure difference between stagnant pressure and downstream pressure. By assuming an isentropic expansion of an ideal gas from its stagnant conditions (P0 , meaning the velocity of the gas is zero) to downstream conditions (P1, positioned at the exit plane of the nozzle or orifice), the subsonic flow rate of the source term is given by Ramskill's formulation: [4]

The flow rate of an ideal gas can be represented by the graphed line. As the pressure ratio decreases and the critical value is approached, the flow changes from non-choked to choked, setting an upper limit to the velocity of the gas to the speed of sound of the medium. Gas flow rate variation as the pressure ratio decreases.png
The flow rate of an ideal gas can be represented by the graphed line. As the pressure ratio decreases and the critical value is approached, the flow changes from non-choked to choked, setting an upper limit to the velocity of the gas to the speed of sound of the medium.

As the ratio between downstream condition pressure and stagnant condition pressure decreases, the flow rate of the ideal gas will increase. This behavior will continue until a critical value is reached (in air, P1/P0 is roughly 0.528, [5] dependent on the heat capacity ratio, γ), changing the condition of the jet from a non-choked flow to a choked flow. This will lead to the a newly defined expression for the aforementioned pressure ratio and, sub-sequentially, the flow rate equation.

The critical value for the pressure ratio is defined as:

This newly defined ratio can then be used to determine the flow rate for a sonic choked flow:

The flow rate equation for a choked flow will have a fixed velocity, which is the speed of sound of the medium, where the Mach number is equals to 1:

Where: 
Q= mass flow rate, kg/s
Vc= speed of sound of the medium, m/s
CD= discharge coefficient, dimensionless (its value is chosen relative to the orifice shape)
Ao= orifice/nozzle exit plane area, m2
γ= heat capacity ratio of the gas
ρ0= gas density at P0 and T0 , kg/m3
ρ1= gas density at P1 and T1 , kg/m3
T0= Temperature in stagnant conditions, K
T1= Temperature in downstream conditions, K
P0= absolute pressure in stagnant conditions, Pa
P1= absolute pressure in downstream conditions, Pa
M= the ideal gas molecular weight, kg/kmol  
R= the Universal Gas Law Constant = 8.3145 J/(mol·K)

It is important to note that if P1 keeps on decreasing, no flow rate change will occur if the ratio is already below the critical value, unless P0 also changes (also assuming that the orifice/nozzle exit area and upstream temperature stay the same).

Under-expanded jet structure

An under-expanded jet is one that manifests when the pressure at downstream conditions (at the end of a nozzle or orifice) is greater that the pressure of the environment where the gas is being released in. It is said to be under-expanded since the gas will expand, trying to reach the same pressure of its surroundings. When under-expanded, the jet will have characteristics of a compressible flow, a condition in which pressure variations are significant enough to have a strong effect on the velocity (where it can exceed the speed of sound of the gas), density and temperature. [6] It is important to note that as the jet expands and incorporates gases from the surrounding medium, the jet will behave more and more like an incompressible fluid, allowing for a general definition of the structure of a jet to be the following: [1]

The subdivision of a high pressure jet as it expands to the pressure of the surrounding conditions. High pressure jet zones.png
The subdivision of a high pressure jet as it expands to the pressure of the surrounding conditions.

Under-expanded jet classification

Further classification of the jet can be related to how the nearfield zone develops due to the compressible effects that govern it. [1] When the jet first exists the orifice or nozzle, it will expand very quickly, resulting in an over-expansion of the flow (which will also reduce the temperature and density of the flow as quickly as it depressurized). Gases that have expanded to a pressure lower than the one of the surrounding fluid will be compressed inwards, causing an increase in the pressure of the flow. If this re-compression leads to the fluid having a higher pressure than the surrounding fluid, another expansion will happen. This process will repeat until the pressure difference between ambient pressure and jet pressure is null (or close to null). [7] Compression and expansion are accomplished through a series of shock waves, formed as a result of Prandlt-Meyer expansion and compression waves. [8]

Development of the aforementioned shock waves will be related to the difference in pressure between the stagnant conditions or downstream conditions and the ambient conditions (η0 = P0/Pamb and ηe = P1/Pamb, respectively), as well as the mach number (Ma = V/Vc, where V is the velocity of the flow and Vc is the speed of sound of the medium). With varying pressure ratios, under-expanded jets can be classified as: [1]

A CFD simulation of a natural gas extremely under-expanded jet. Sections where the Mach Number is much higher than 1 are points in which the velocity of the fluid is much higher than the speed of sound, an effect that manifest due to the sudden depressurization of the gas. P[?] is ambient pressure of 101325 Pa Extremely over-expanded jet simulated through CFD.png
A CFD simulation of a natural gas extremely under-expanded jet. Sections where the Mach Number is much higher than 1 are points in which the velocity of the fluid is much higher than the speed of sound, an effect that manifest due to the sudden depressurization of the gas. P is ambient pressure of 101325 Pa


  • Moderately under-expanded jet: Nearfield with diamond shaped structures (each structure is called a cell). A Prandlt-Meyer expansion generates oblique expansion waves that expand the fluid downstream from the exit orifice. As these waves attain constant pressure from the surrounding fluid, they are deflected back as compression waves, converging in oblique shock waves (called intercepting shock). When they meet on the axis of the jet, reflected shock waves move outwardly until they attain constant pressure from the surrounding fluid, repeating the process, and in turn, recreating the cell structure (this phenomenon occurs in air at a range of 2 ≤ η0 ≤ 4 or 1.1 ≤ ηe ≤ 3).
  • Highly under-expanded jet: Nearfield with barrel shaped structures. As the pressure ratio increases, the intercepting shock waves can't meet on the axis of the jet anymore, which forces the generation of a normal shock wave when the intercepting shock waves go beyond a certain critical angle (The normal shock wave is called Mach Disk). From the interception point of the mach disk and the intercepting shock, a residual slipstream will reflect outwardly, until it reaches constant pressure from the surrounding fluid, repeating the process, recreating the barrel shaped cell structure (this phenomenon occurs in air at a range of 5 ≤ η0 ≤ 7 or 2 ≤ ηe ≤ 4 ).
  • Extremely under-expanded jet: Nearfield with a single cell structure. When the pressure ratio goes beyond a critical value (in air at a range of η0 ≥ 7 or ηe ≥ 4 ), cell numbers within the nearfield of the jet decrease, until they all coalesce into a single cell with a single mach disk. Due to the increase in velocity and lower pressure zones around the jet, ambient fluid entrainment will increase.

Natural gas release

Amongst incidental scenarios, natural gas releases have become particularly relevant within the process industry environment. [3] With an overall composition of 94.7% of methane, [9] it is important to consider how this gas can cause incremental damage when it is released. Methane gas is a non-toxic, flammable gas, that, at higher concentrations, can behave as an asphyxiant due to oxygen displacement from the lungs. [10] The main concern with methane is related to its flammability and the potential damage that could be dealt to its surroundings if the high pressure jet were to ignite into a jet fire. [11]

Three parameters that must be considered when dealing with flammable gasses are their flash point (FP), upper flammability limit (UFL) and lower flammability limit (LFL), as they are set values for any compound at a specific pressure and temperature. If we consider the fire triangle model, to induce a combustion reaction three components are needed: a fuel, an oxidizing agent and heat. When release happens in an ambient filled with air, the oxidizing agent will be oxygen (air has a constant concentration of 21% in standard conditions). [12] At an almost pure concentration, a few centimeters from the exit plane, the concentration of natural gas is too high and oxygen too low to generate any kind of combustion reaction, but as the high pressure jet develops, the concentration of its components will dilute as air entrainment increases, allowing an enrichment of oxygen within the jet. Assuming a constant concentration for oxygen, the jet must dilute enough to enter within its flammability range; below its UFL. Within this range, a flammable mixture can be made and any source of heat can jump-start the reaction. [13]

To properly judge the damage and potential risk that the jet fire can generate, several studies regarding the maximum distance that the cloud generated by the jet can reach have been made. As dilution of the jet continues due to air entrainment in the farfield, going below its UFL, the maximum distance that the flammable mixture can reach is at the point in which the concentration of the cloud is equals to the LFL of the gas, as it is the lowest concentration allowable that permits the formation of a flammable mixture between air and natural gas at standard conditions (the LFL for natural gas is 4% [9] ). Considering a free jet at sub-critical pressure (beyond the nearfield zone), its mean volume fraction axial concentration decay of any gas released in air can be defined as follows: [14]

Where: 
= mean volume fraction axial concentration of the gas, [-]
k= experimental constant coefficient, [-]
d= orifice diameter, m
z= downstream distance, m
a= virtual origin displacement, m
ρa= air density, kg/m3
ρg= released gas density, kg/m3

Computational Fluid Dynamics

Experimental data of high pressure jets have to be limited in terms of size and complexity of the scenario due to the inherit dangers and expenses correlated to the experiment itself. Alternative methods to gather data, such as representative models, can be used in order to predict what the maximum extend of the gas cloud at its LFL concentration can reach. Simpler models like a gaussian gas dispersion model (e.g., SCREEN3 - a dispersion model) or integral model (e.g., PHAST- an integral model) can be useful to have a quick and qualitative overview on how the jet may extend. Unfortunately, their inability to properly simulate jet-obstacle interactions make them impossible to use beyond preliminary calculations. This is the reason why Computational Fluid Dynamic (CFD) simulations are generally preferred for more complex scenarios. [15]

Although there exists several approaches for CFD simulations, a common approach is the use of a finite volume method that discretizes the volume into smaller cells of varying shapes. Every single cell will represent a fluid-filled volume where the scenarios parameters will be applied. Every cell that was modeled solves a set of conservation equations of mass, momentum and energy, along with the continuity equation. Fluid-obstacle interaction is then modeled with varying algorithms based on the closure turbulent model used. [16] Depending on the number of total cells within the volume, the better the quality of the simulation, the longer the simulation time. Convergence problems can arise within the simulation as large momentum, mass and energy gradients appear in the volume. The points where these problems are expected to appear (like in the nearfield zone of the jet) need to have a higher number of cells to achieve gradual changes between one cell and another. Ideally, through CFD simulations, a simpler model can be derived which, for a specific set of scenarios, allows to have results with an accuracy and precision level similar to the CFD simulation itself. [17]

Birch's Approach

Through a set of small scale experiments at varying pressures, Birch et al. formulated an equation that allowed the estimation of a virtual surface source, considering the conservation of mass between the exit plane of the orifice and the virtual surface. [18] This approach allows to simulate a compressible, under-expanded jet as an incompressible, fully-expanded jet. As a consequence, a simpler CFD model can be simulated by using the following diameter (named pseudo-diameter) as the new exit plane: [19]

Where: 
dps= Birch's pseudo diameter, m
d= orifice diameter, m
T0= temperature in stagnant conditions, K
T2= temperature in ambient conditions, K
P0= pressure in stagnant conditions, Pa
P2= pressure in ambient conditions, Pa
Ground and obstacle interaction

In the process industry, there exist a variety of cases where a high pressure jet release incident can occur. LNG storage facilities or NG pipeline systems leakage [20] can degenerate into a jet fire and, through a domino effect, cause heavy damage to the workforce, equipment and surrounding environment. For different scenarios that may happen, safety protocols have to be engineered that aim to set minimum distances between equipment and the workforce, along with preventive systems that reduce the danger of the potential incidental scenario. The following are some of the most common scenarios that may be encountered in an industrial environment: [19] [21] [22]

Jet-tank-ground interaction will have an effect on the maximum extension of the cloud at LFL concentrations. Eddies generated after their interactions will promote gas dissipation. s represents the distance between the center of the two tanks, D is the diameter of the tanks CFD simuations of a high pressure NG jet impinging two tanks.png
Jet-tank-ground interaction will have an effect on the maximum extension of the cloud at LFL concentrations. Eddies generated after their interactions will promote gas dissipation. s represents the distance between the center of the two tanks, D is the diameter of the tanks
  • Jet-Ground interaction: this is one of the most common scenarios, where the free jet does not interact with any other obstacle apart from the ground. Although the jet can dissipate to concentrations below its LFL after roughly 16 meters with no interactions (interaction with concrete and at stagnant pressure of 65 bar, a common pressure for NG pipelines, [23] with an orifice size of 2.54 mm), as the jet gets closer to the ground and makes contact with it, a dragging effect leads to a further extension. The jet has the tendency of bending downwards if close enough to the ground because of lower pressure zones below it.
  • Jet-horizontal tank interaction: jet-tank interaction will depend of the type of material used. For cylindrical steel tanks, the distance from the exit plane will have an effect with the interaction, along with its angle with respect to the axis of the jet and the height of the release of the jet. Generally, when the jet impinges the horizontal tank along the axis of the tank, the jet is brought downwards, allowing an interaction with the ground. This will often lead to the extension of the jet at its LFL concentration with respect to its free jet (the same scenario is simulated with no other obstacles but the ground). Furthermore, crosswise extension is also expected.
  • Jet-vertical tank interaction: for a cylindrical steel tank, the distance from the exit plane will have an effect with the interaction. Generally, when the jet impinges the vertical tank along the axis of the tank, the obstacle will act as a limiting factor for the extension of the jet at its LFL concentration with respect to its free jet. The impact with the tank will generate more eddies downstream the tail of the jet and limit ground interaction (at low enough flow rates and speeds), speeding up dilution of the gas below its lower flammable limit.
  • Jet-horizontal tandem tank interaction: by adding a second horizontal tank behind the first, a shortening effect is achieved. The presence of the second obstacle leads to eddy generation after the first, promoting dissipation. Moreover, the second obstacle may have a detachment effect of the jet from the ground, as it will tend to adhere to the round surface of the second tank due to the coanda effect. Distance between the two tanks will have an effect, since, after a certain distance, the second obstacle will no longer have an effect on the cloud at LFL concentration.
  • Jet-vertical tandem tank interaction: by adding a second vertical tank behind the first, a shortening effect will generally be achieved. The presence of the second obstacle leads to eddy generation after the first, promoting dissipation. Contrary to the previous scenario, the second obstacle can allow for jet-ground interaction and elongate the jet, a quality that be worsen due to the coanda effect. Distance between the two tanks will have an effect, since, after a certain distance, the second obstacle will no longer have an effect on the cloud at LFL concentration.

See also

Related Research Articles

<span class="mw-page-title-main">Mach number</span> Ratio of speed of an object moving through fluid and local speed of sound

The Mach number, often only Mach, is a dimensionless quantity in fluid dynamics representing the ratio of flow velocity past a boundary to the local speed of sound. It is named after the Austrian physicist and philosopher Ernst Mach.

<span class="mw-page-title-main">Troposphere</span> Lowest layer of Earths atmosphere

The troposphere is the lowest layer of the atmosphere of Earth. It contains 75% of the total mass of the planetary atmosphere and 99% of the total mass of water vapor and aerosols, and is where most weather phenomena occur. From the planetary surface of the Earth, the average height of the troposphere is 18 km in the tropics; 17 km in the middle latitudes; and 6 km in the high latitudes of the polar regions in winter; thus the average height of the troposphere is 13 km.

<span class="mw-page-title-main">Bernoulli's principle</span> Principle relating to fluid dynamics

Bernoulli's principle is a key concept in fluid dynamics that relates pressure, speed and height. Bernoulli's principle states that an increase in the speed of a fluid occurs simultaneously with a decrease in static pressure or the fluid's potential energy. The principle is named after the Swiss mathematician and physicist Daniel Bernoulli, who published it in his book Hydrodynamica in 1738. Although Bernoulli deduced that pressure decreases when the flow speed increases, it was Leonhard Euler in 1752 who derived Bernoulli's equation in its usual form.

<span class="mw-page-title-main">Coandă effect</span> Tendency of a fluid jet to stay attached to a convex surface

The Coandă effect is the tendency of a fluid jet to stay attached to a convex surface. Merriam-Webster describes it as "the tendency of a jet of fluid emerging from an orifice to follow an adjacent flat or curved surface and to entrain fluid from the surroundings so that a region of lower pressure develops."

Compressible flow is the branch of fluid mechanics that deals with flows having significant changes in fluid density. While all flows are compressible, flows are usually treated as being incompressible when the Mach number is smaller than 0.3. The study of compressible flow is relevant to high-speed aircraft, jet engines, rocket motors, high-speed entry into a planetary atmosphere, gas pipelines, commercial applications such as abrasive blasting, and many other fields.

<span class="mw-page-title-main">Foam</span> Form of matter

Foams are materials formed by trapping pockets of gas in a liquid or solid.

de Laval nozzle Pinched tube generating supersonic flow

A de Laval nozzle is a tube which is pinched in the middle, making a carefully balanced, asymmetric hourglass shape. It is used to accelerate a compressible fluid to supersonic speeds in the axial (thrust) direction, by converting the thermal energy of the flow into kinetic energy. De Laval nozzles are widely used in some types of steam turbines and rocket engine nozzles. It also sees use in supersonic jet engines.

Hemorheology, also spelled haemorheology, or blood rheology, is the study of flow properties of blood and its elements of plasma and cells. Proper tissue perfusion can occur only when blood's rheological properties are within certain levels. Alterations of these properties play significant roles in disease processes. Blood viscosity is determined by plasma viscosity, hematocrit and mechanical properties of red blood cells. Red blood cells have unique mechanical behavior, which can be discussed under the terms erythrocyte deformability and erythrocyte aggregation. Because of that, blood behaves as a non-Newtonian fluid. As such, the viscosity of blood varies with shear rate. Blood becomes less viscous at high shear rates like those experienced with increased flow such as during exercise or in peak-systole. Therefore, blood is a shear-thinning fluid. Contrarily, blood viscosity increases when shear rate goes down with increased vessel diameters or with low flow, such as downstream from an obstruction or in diastole. Blood viscosity also increases with increases in red cell aggregability.

<span class="mw-page-title-main">Venturi effect</span> Reduced pressure caused by a flow restriction in a tube or pipe

The Venturi effect is the reduction in fluid pressure that results when a fluid flows through a constricted section of a pipe. The Venturi effect is named after its discoverer, the 18th-century Italian physicist Giovanni Battista Venturi.

An orifice plate is a device used for measuring flow rate, for reducing pressure or for restricting flow.

<span class="mw-page-title-main">Oblique shock</span> Shock wave that is inclined with respect to the incident upstream flow direction

An oblique shock wave is a shock wave that, unlike a normal shock, is inclined with respect to the incident upstream flow direction. It will occur when a supersonic flow encounters a corner that effectively turns the flow into itself and compresses. The upstream streamlines are uniformly deflected after the shock wave. The most common way to produce an oblique shock wave is to place a wedge into supersonic, compressible flow. Similar to a normal shock wave, the oblique shock wave consists of a very thin region across which nearly discontinuous changes in the thermodynamic properties of a gas occur. While the upstream and downstream flow directions are unchanged across a normal shock, they are different for flow across an oblique shock wave.

Choked flow is a compressible flow effect. The parameter that becomes "choked" or "limited" is the fluid velocity.

<span class="mw-page-title-main">Rocket engine nozzle</span> Type of propelling nozzle

A rocket engine nozzle is a propelling nozzle used in a rocket engine to expand and accelerate combustion products to high supersonic velocities.

<span class="mw-page-title-main">Prandtl–Meyer expansion fan</span> Phenomenon in fluid dynamics

A supersonic expansion fan, technically known as Prandtl–Meyer expansion fan, a two-dimensional simple wave, is a centered expansion process that occurs when a supersonic flow turns around a convex corner. The fan consists of an infinite number of Mach waves, diverging from a sharp corner. When a flow turns around a smooth and circular corner, these waves can be extended backwards to meet at a point.

<span class="mw-page-title-main">Spray nozzle</span> Device that facilitates dispersion of liquid into a spray

A spray nozzle or atomizer is a device that facilitates the dispersion of a liquid by the formation of a spray. The production of a spray requires the fragmentation of liquid structures, such as liquid sheets or ligaments, into droplets, often by using kinetic energy to overcome the cost of creating additional surface area. A wide variety of spray nozzles exist, that make use of one or multiple liquid breakup mechanisms, which can be divided into three categories: liquid sheet breakup, jets and capillary waves. Spray nozzles are of great importance for many applications, where the spray nozzle is designed to have the right spray characteristics.

Fanno flow is the adiabatic flow through a constant area duct where the effect of friction is considered. Compressibility effects often come into consideration, although the Fanno flow model certainly also applies to incompressible flow. For this model, the duct area remains constant, the flow is assumed to be steady and one-dimensional, and no mass is added within the duct. The Fanno flow model is considered an irreversible process due to viscous effects. The viscous friction causes the flow properties to change along the duct. The frictional effect is modeled as a shear stress at the wall acting on the fluid with uniform properties over any cross section of the duct.

<span class="mw-page-title-main">Radial turbine</span>

A radial turbine is a turbine in which the flow of the working fluid is radial to the shaft. The difference between axial and radial turbines consists in the way the fluid flows through the components. Whereas for an axial turbine the rotor is 'impacted' by the fluid flow, for a radial turbine, the flow is smoothly orientated perpendicular to the rotation axis, and it drives the turbine in the same way water drives a watermill. The result is less mechanical stress which enables a radial turbine to be simpler, more robust, and more efficient when compared to axial turbines. When it comes to high power ranges the radial turbine is no longer competitive and the efficiency becomes similar to that of the axial turbines.

Isentropic nozzle flow describes the movement of a gas or fluid through a narrowing opening without an increase or decrease in entropy.

<span class="mw-page-title-main">Non ideal compressible fluid dynamics</span>

Non ideal compressible fluid dynamics (NICFD), or non ideal gas dynamics, is a branch of fluid mechanics studying the dynamic behavior of fluids not obeying ideal-gas thermodynamics. It is for example the case of dense vapors, supercritical flows and compressible two-phase flows. With the term dense vapors, we indicate all fluids in the gaseous state characterized by thermodynamic conditions close to saturation and the critical point. Supercritical fluids feature instead values of pressure and temperature larger than their critical values, whereas two-phase flows are characterized by the simultaneous presence of both liquid and gas phases.

In gas dynamics, the Kantrowitz limit refers to a theoretical concept describing choked flow at supersonic or near-supersonic velocities. When an initially subsonic fluid flow experiences a reduction in cross-section area, the flow speeds up in order to maintain the same mass-flow rate, per the continuity equation. If a near supersonic flow experiences an area contraction, the velocity of the flow will increase until it reaches the local speed of sound, and the flow will be choked. This is the principle behind the Kantrowitz limit: it is the maximum amount of contraction a flow can experience before the flow chokes, and the flow speed can no longer be increased above this limit, independent of changes in upstream or downstream pressure.

References

  1. 1 2 3 4 Franquet, Erwin; Bruel, Pascal; Perrier, Vincent; Gibout, Stéphane (August 2015). "Free underexpanded jets in a quiescent medium: A review" . Progress in Aerospace Sciences. 77: 25–53. Bibcode:2015PrAeS..77...25F. doi:10.1016/j.paerosci.2015.06.006.
  2. 1 2 Joint Research Centre, Institute for Energy and Transport, Molkov, V., Lachance, J., Tchouvelev, A., et al., State of the art and research priorities in hydrogen safety, Publications Office, 2014, https://data.europa.eu/doi/10.2790/99638
  3. 1 2 "Lessons learned from LPG/LNG Accidents". Joint Research Centre - European Commission. Joint Research Centre the European Commission's in-house science service. Retrieved 30 June 2022.
  4. Perry, Robert H.; Green, Don W. (2008). Perry's Chemical Engineering Handbook (8th ed.). McGraw-Hill Professional Pub. p. 806. ISBN   978-0071422949.
  5. "Understanding Choked Flow Gases" (PDF). O'keefe Controls. O’Keefe Controls Co. Retrieved 1 July 2022.
  6. Bar–Meir, Genick (21 May 2007). Fundamentals of Compressible Fluid Mechanics (PDF). p. 49.
  7. Scott, Jeff. "Shock Diamonds and Mach Disks". Aerospaceweb. Retrieved 29 June 2022.
  8. Linial, Itai; Sari, Re’em (2019). "Oblique shock breakout from a uniform density medium". Physics of Fluids. Racah Institute of Physics. 31 (9): 097102. arXiv: 1908.05301 . Bibcode:2019PhFl...31i7102L. doi:10.1063/1.5100060. S2CID   199668808.
  9. 1 2 "Learn About Natural Gas". Enbridge.
  10. "Methane". Canadian Center for Occupation Health and Safety. 3 July 2022.
  11. Shu-jiao, Tong; Siu-ming, Lo; Bao-zhi, Chen; Pei-hong, Zhang (2013). "Jet Fire Consequence Evaluation on the Natural Gas Transported by Pipelines". Procedia Engineering. 52: 349–354. doi: 10.1016/j.proeng.2013.02.152 . Retrieved 30 June 2022.
  12. "Air - Composition and Molecular Weight". The Engineering Toolbox. Retrieved 30 June 2022.
  13. Drysdale, Dougal (2 February 1999). An Introduction to Fire Dynamics (3rd ed.). John Wiley & Sons Inc. ISBN   0471972916.
  14. Chen, C.J.; Rodi, W. (1980). "Vertical turbulent buoyant jets: A review of experimental data". NASA Sti/Recon Technical Report A. Oxford and New York, Pergamon Press. 80: 23073. Bibcode:1980STIA...8023073C.
  15. "advantages of computational fluid dynamics". Pretechnologies. Retrieved 30 June 2022.
  16. Menter, F. R. (1994). "Two-Equation Eddy-Viscosity Turbulence Models for Engineering Applications" (PDF). AIAA Journal. 32 (8): 1598–1605. Bibcode:1994AIAAJ..32.1598M. doi:10.2514/3.12149. S2CID   120712103.[ dead link ]
  17. Xie, Qiyuan; Lu, Qizhao; Yuan, Yanhua; Zhang, Jie; Zhou, Feng (September 2021). "Numerical study on the horizontal stretching effect of ground on high-pressure vapor jets of LNG tank leakage" . Journal of Loss Prevention in the Process Industries. 72: 104526. doi:10.1016/j.jlp.2021.104526 . Retrieved 1 July 2022.
  18. Birch, A.D.; Brown, D.R.; Dodson, M.G.; Swaffield, F. (14 March 1983). "The Structure and Concentration Decay of High Pressure Jets of Natural Gas" . Combustion Science and Technology. 36 (5–6): 249–261. doi:10.1080/00102208408923739.
  19. 1 2 Colombini, Cristian; Martani, Andrea; Renato, Rota; Busini, Valentina (2020). "Ground influence on high-pressure methane jets: Practical tools for risk assessment". Journal of Loss Prevention in the Process Industries. 67: 1–11. doi:10.1016/j.jlp.2020.104240. hdl: 11311/1152115 . S2CID   225280969.
  20. Xian, Shan; Kang, Liu; Pei-Liang, Sun (2017). "Risk Analysis on Leakage Failure of Natural Gas Pipelines by Fuzzy Bayesian Network with a Bow-Tie Model". Scientific Programming. 2017: 1–11. doi: 10.1155/2017/3639524 .
  21. Colombini, Cristian; Busini, Valentina (30 September 2019). "Obstacle Influence on High-Pressure Jets based on Computational Fluid Dynamics Simulations". Chemical Engineering Transactions. 77: 811–816. doi:10.3303/CET1977136 . Retrieved 30 June 2022.
  22. Ferrario, Fabio; Mormile, Anna; Romano, Giovanni; Tombini, Paolo; Busini, Valentina. "Studio dell'interazione tra un getto ad alta pressione e ostacoli cilindrici in serie mediante CFD". POLITesi - Archivio digitale delle tesi di laurea e di dottorato. Retrieved 30 June 2022.
  23. Micucci, Lorenzo. "Natural gas pipeline systems and operations". Gas Processing & LNG. Retrieved 30 June 2022.