Accidental release source terms

Last updated

Accidental release source terms are the mathematical equations that quantify the flow rate at which accidental releases of liquid or gaseous pollutants into the ambient environment which can occur at industrial facilities such as petroleum refineries, petrochemical plants, natural gas processing plants, oil and gas transportation pipelines, chemical plants, and many other industrial activities. Governmental regulations in many countries require that the probability of such accidental releases be analyzed and their quantitative impact upon the environment and human health be determined so that mitigating steps can be planned and implemented.

Contents

There are a number of mathematical calculation methods for determining the flow rate at which gaseous and liquid pollutants might be released from various types of accidents. Such calculational methods are referred to as source terms, and this article on accidental release source terms explains some of the calculation methods used for determining the mass flow rate at which gaseous pollutants may be accidentally released.

Accidental release of pressurized gas

When gas stored under pressure in a closed vessel is discharged to the atmosphere through a hole or other opening, the gas velocity through that opening may be choked (i.e., it has attained a maximum) or it may be non-choked.

Choked velocity, also referred to as sonic velocity, occurs when the ratio of the absolute source pressure to the absolute downstream pressure is equal to or greater than [(k + 1) / 2]k / (k − 1), where k is the specific heat ratio of the discharged gas (sometimes called the isentropic expansion factor and sometimes denoted as ).

For many gases, k ranges from about 1.09 to about 1.41, and therefore [(k + 1) / 2]k / (k − 1 ) ranges from 1.7 to about 1.9, which means that choked velocity usually occurs when the absolute source vessel pressure is at least 1.7 to 1.9 times as high as the absolute downstream ambient atmospheric pressure.

When the gas velocity is choked, the equation for the mass flow rate in SI metric units is: [1] [2] [3] [4]

or this equivalent form:

For the above equations, it is important to note that although the gas velocity reaches a maximum and becomes choked, the mass flow rate is not choked. The mass flow rate can still be increased if the source pressure is increased.

Whenever the ratio of the absolute source pressure to the absolute downstream ambient pressure is less than [ (k + 1) / 2]k / (k − 1), then the gas velocity is non-choked (i.e., sub-sonic) and the equation for mass flow rate is:

or this equivalent form:

where: 
Q= mass flow rate, kg/s
C= discharge coefficient, dimensionless (usually about 0.72)
A= discharge hole area, m2
k= cp/cv of the gas
cp= specific heat of the gas at constant pressure
cv= specific heat of the gas at constant volume
= real gas density at P and T, kg/m3
P= absolute upstream pressure, Pa
PA= absolute ambient or downstream pressure, Pa
M= the gas molecular mass, kg/kmol   (also known as the molecular weight)
R= the Universal Gas Law Constant = 8314.5 Pa·m3/(kmol·K)
T= absolute upstream gas temperature, K
Z= the gas compressibility factor at P and T, dimensionless

The above equations calculate the initial instantaneous mass flow rate for the pressure and temperature existing in the source vessel when a release first occurs. The initial instantaneous flow rate from a leak in a pressurized gas system or vessel is much higher than the average flow rate during the overall release period because the pressure and flow rate decrease with time as the system or vessel empties. Calculating the flow rate versus time since the initiation of the leak is much more complicated, but more accurate. Two equivalent methods for performing such calculations are presented and compared at. [5]

The technical literature can be very confusing because many authors fail to explain whether they are using the universal gas law constant R which applies to any ideal gas or whether they are using the gas law constant Rs which only applies to a specific individual gas. The relationship between the two constants is Rs = R/M.

Notes:

Ramskill's equation for non-choked mass flow

P.K. Ramskill's equation [6] [7] for the non-choked flow of an ideal gas is shown below as equation (1):

(1)    

The gas density, A, in Ramskill's equation is the ideal gas density at the downstream conditions of temperature and pressure and it is defined in equation (2) using the ideal gas law:

(2)    

Since the downstream temperature TA is not known, the isentropic expansion equation below [8] is used to determine TA in terms of the known upstream temperature T:

(3)    

Combining equations (2) and (3) results in equation (4) which defines A in terms of the known upstream temperature T:

(4)    

Using equation (4) with Ramskill's equation (1) to determine non-choked mass flow rates for ideal gases gives identical results to the results obtained using the non-choked flow equation presented in the previous section above.

Evaporation of non-boiling liquid pool

Three different methods of calculating the rate of evaporation from a non-boiling liquid pool are presented in this section. The results obtained by the three methods are somewhat different.

The U.S. Air Force method

The following equations are for predicting the rate at which liquid evaporates from the surface of a pool of liquid which is at or near the ambient temperature. The equations were derived from field tests performed by the U.S. Air Force with pools of liquid hydrazine. [2]

where: 
E= evaporation flux, kg/m2·min of pool surface
u= windspeed just above the liquid surface, m/s
TA= absolute ambient temperature, K
TF= pool liquid temperature correction factor, dimensionless
TP= pool liquid temperature, °C
M= pool liquid molecular weight, dimensionless
PS= pool liquid vapor pressure at ambient temperature, mmHg
PH= hydrazine vapor pressure at ambient temperature, mmHg (see equation below)

If TP = 0 °C or less, then TF = 1.0

If TP > 0 °C, then TF = 1.0 + 0.0043 TP2

where: 
= 2.7183, the base of the natural logarithm system
= natural logarithm

The U.S. EPA method

The following equations are for predicting the rate at which liquid evaporates from the surface of a pool of liquid which is at or near the ambient temperature. The equations were developed by the United States Environmental Protection Agency using units which were a mixture of metric usage and United States usage. [3] The non-metric units have been converted to metric units for this presentation.

NB, the constant used here is 0.284 from the mixed unit formula/2.205 lb/kg. The 82.05 become 1.0 = (ft/m)² × mmHg/kPa.

where: 
E= evaporation rate, kg/min
u= windspeed just above the pool liquid surface, m/s
M= pool liquid molecular weight, dimensionless
A= surface area of the pool liquid, m2
P= vapor pressure of the pool liquid at the pool temperature, kPa
T= pool liquid absolute temperature, K

The U.S. EPA also defined the pool depth as 0.01 m (i.e., 1 cm) so that the surface area of the pool liquid could be calculated as:

A = (pool volume, in m3)/(0.01)

Notes:

Stiver and Mackay's method

The following equations are for predicting the rate at which liquid evaporates from the surface of a pool of liquid which is at or near the ambient temperature. The equations were developed by Warren Stiver and Dennis Mackay of the Chemical Engineering Department at the University of Toronto. [9]

where: 
E= evaporation flux, kg/m2·s of pool surface
k= mass transfer coefficient, m/s = 0.002 u
TA= absolute ambient temperature, K
M= pool liquid molecular weight, dimensionless
P= pool liquid vapor pressure at ambient temperature, Pa
R= the universal gas law constant = 8314.5 Pa·m3/(kmol·K)
u= windspeed just above the liquid surface, m/s

Evaporation of boiling cold liquid pool

The following equation is for predicting the rate at which liquid evaporates from the surface of a pool of cold liquid (i.e., at a liquid temperature of about 0 °C or less). [2]

where: 
E= evaporation flux, (kg/min)/m2 of pool surface
B= pool liquid atmospheric boiling point, °C
M= pool liquid molecular weight, dimensionless
e= the base of the natural logarithm system = 2.7183

Adiabatic flash of liquefied gas release

Liquefied gases such as ammonia or chlorine are often stored in cylinders or vessels at ambient temperatures and pressures well above atmospheric pressure. When such a liquefied gas is released into the ambient atmosphere, the resultant reduction of pressure causes some of the liquefied gas to vaporize immediately. This is known as "adiabatic flashing" and the following equation, derived from a simple heat balance, is used to predict how much of the liquefied gas is vaporized.

where: 
X= weight percent vaporized
HsL= source liquid enthalpy at source temperature and pressure, J/kg
HaV= flashed vapor enthalpy at atmospheric boiling point and pressure, J/kg
HaL= residual liquid enthalpy at atmospheric boiling point and pressure, J/kg

If the enthalpy data required for the above equation is unavailable, then the following equation may be used.

where: 
X= weight percent vaporized
cp= source liquid specific heat, J/(kg °C)
Ts= source liquid absolute temperature, K
Tb= source liquid absolute atmospheric boiling point, K
H= source liquid heat of vaporization at atmospheric boiling point, J/kg

See also

Related Research Articles

<span class="mw-page-title-main">Equation of state</span> An equation describing the state of matter under a given set of physical conditions

In physics and chemistry, an equation of state is a thermodynamic equation relating state variables, which describe the state of matter under a given set of physical conditions, such as pressure, volume, temperature, or internal energy. Most modern equations of state are formulated in the Helmholtz free energy. Equations of state are useful in describing the properties of pure substances and mixtures in liquids, gases, and solid states as well as the state of matter in the interior of stars.

<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.

In thermodynamics, the specific heat capacity of a substance is the heat capacity of a sample of the substance divided by the mass of the sample, also sometimes referred to as massic heat capacity. Informally, it is the amount of heat that must be added to one unit of mass of the substance in order to cause an increase of one unit in temperature. The SI unit of specific heat capacity is joule per kelvin per kilogram, J⋅kg−1⋅K−1. For example, the heat required to raise the temperature of 1 kg of water by 1 K is 4184 joules, so the specific heat capacity of water is 4184 J⋅kg−1⋅K−1.

<span class="mw-page-title-main">Navier–Stokes equations</span> Equations describing the motion of viscous fluid substances

The Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances, named after French engineer and physicist Claude-Louis Navier and Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842-1850 (Stokes).

<span class="mw-page-title-main">Hydrostatic equilibrium</span> State of balance between external forces on a fluid and internal pressure gradient

In fluid mechanics, hydrostatic equilibrium is the condition of a fluid or plastic solid at rest, which occurs when external forces, such as gravity, are balanced by a pressure-gradient force. In the planetary physics of Earth, the pressure-gradient force prevents gravity from collapsing the planetary atmosphere into a thin, dense shell, whereas gravity prevents the pressure-gradient force from diffusing the atmosphere into outer space.

The Prandtl number (Pr) or Prandtl group is a dimensionless number, named after the German physicist Ludwig Prandtl, defined as the ratio of momentum diffusivity to thermal diffusivity. The Prandtl number is given as:

<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">Leidenfrost effect</span> Physical phenomenon

The Leidenfrost effect is a physical phenomenon in which a liquid, close to a surface that is significantly hotter than the liquid's boiling point, produces an insulating vapor layer that keeps the liquid from boiling rapidly. Because of this repulsive force, a droplet hovers over the surface, rather than making physical contact with it. The effect is named after the German doctor Johann Gottlob Leidenfrost, who described it in A Tract About Some Qualities of Common Water.

<span class="mw-page-title-main">Euler equations (fluid dynamics)</span> Set of quasilinear hyperbolic equations governing adiabatic and inviscid flow

In fluid dynamics, the Euler equations are a set of quasilinear partial differential equations governing adiabatic and inviscid flow. They are named after Leonhard Euler. In particular, they correspond to the Navier–Stokes equations with zero viscosity and zero thermal conductivity.

<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.

The barometric formula is a formula used to model how the pressure of the air changes with altitude.

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

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

The Ostwald–Freundlich equation governs boundaries between two phases; specifically, it relates the surface tension of the boundary to its curvature, the ambient temperature, and the vapor pressure or chemical potential in the two phases.

<span class="mw-page-title-main">Nonlinear acoustics</span>

Nonlinear acoustics (NLA) is a branch of physics and acoustics dealing with sound waves of sufficiently large amplitudes. Large amplitudes require using full systems of governing equations of fluid dynamics and elasticity. These equations are generally nonlinear, and their traditional linearization is no longer possible. The solutions of these equations show that, due to the effects of nonlinearity, sound waves are being distorted as they travel.

The intent of this article is to highlight the important points of the derivation of the Navier–Stokes equations as well as its application and formulation for different families of fluids.

Volume viscosity is a material property relevant for characterizing fluid flow. Common symbols are or . It has dimensions, and the corresponding SI unit is the pascal-second (Pa·s).

The vaporizing droplet problem is a challenging issue in fluid dynamics. It is part of many engineering situations involving the transport and computation of sprays: fuel injection, spray painting, aerosol spray, flashing releases… In most of these engineering situations there is a relative motion between the droplet and the surrounding gas. The gas flow over the droplet has many features of the gas flow over a rigid sphere: pressure gradient, viscous boundary layer, wake. In addition to these common flow features one can also mention the internal liquid circulation phenomenon driven by surface-shear forces and the boundary layer blowing effect.

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

The shear viscosity of a fluid is a material property that describes the friction between internal neighboring fluid surfaces flowing with different fluid velocities. This friction is the effect of (linear) momentum exchange caused by molecules with sufficient energy to move between these fluid sheets due to fluctuations in their motion. The viscosity is not a material constant, but a material property that depends on temperature, pressure, fluid mixture composition, local velocity variations. This functional relationship is described by a mathematical viscosity model called a constitutive equation which is usually far more complex than the defining equation of shear viscosity. One such complicating feature is the relation between the viscosity model for a pure fluid and the model for a fluid mixture which is called mixing rules. When scientists and engineers use new arguments or theories to develop a new viscosity model, instead of improving the reigning model, it may lead to the first model in a new class of models. This article will display one or two representative models for different classes of viscosity models, and these classes are:

References

  1. Perry's Chemical Engineers' Handbook , Sixth Edition, McGraw-Hill Co., 1984.
  2. 1 2 3 Handbook of Chemical Hazard Analysis Procedures, Appendix B, Federal Emergency Management Agency, U.S. Dept. of Transportation, and U.S. Environmental Protection Agency, 1989. Also provides the references below:
    – Clewell, H.J., A Simple Method For Estimating the Source Strength Of Spills Of Toxic Liquids, Energy Systems Laboratory, ESL-TR-83-03, 1983.
    – Ille, G. and Springer, C., The Evaporation And Dispersion Of Hydrazine Propellants From Ground Spill, Environmental Engineering Development Office, CEEDO 712-78-30, 1978.
    – Kahler, J.P., Curry, R.C. and Kandler, R.A.,Calculating Toxic Corridors Air Force Weather Service, AWS TR-80/003, 1980.
    Handbook of Chemical Hazard Analysis, Appendix B Scroll down to page 391 of 520 PDF pages.
  3. 1 2 "Risk Management Program Guidance For Offsite Consequence Analysis" U.S. EPA publication EPA-550-B-99-009, April 1999. (See derivations of equations D-1 and D-7 in Appendix D)
  4. "Methods For The Calculation Of Physical Effects Due To Releases Of Hazardous Substances (Liquids and Gases)", PGS2 CPR 14E, Chapter 2, The Netherlands Organization Of Applied Scientific Research, The Hague, 2005. PGS2 CPR 14E Archived 2007-08-09 at the Wayback Machine
  5. Milton R. Beychok. "Calculating Accidental Release Flow Rates From Pressurized Gas Systems". Archived from the original on 4 February 2006. Retrieved 11 August 2021.{{cite web}}: CS1 maint: unfit URL (link)
  6. CACHE Newsletter No.48, Spring 1999 Gierer, C. and Hyatt, N.,Using Source Term Analysis Software for Calculating Fluid Flow Release Rates Dyadem International Ltd.
  7. Ramskill, P.K. (1986), Discharge Rate Calculation Methods for Use In Plant Safety Assessments, Safety and Reliability Directory, United Kingdom Atomic Energy Authority
  8. Isentropic Compression or Expansion
  9. Stiver, W. and Mackay, D., A Spill Hazard Ranking System For Chemicals, Environment Canada First Technical Spills Seminar, Toronto, Canada, 1993.