Chapter 5. Hazardous Material Dispersion

The learning objectives for this chapter are to:

  1. Understand air dispersion and the parameters required to describe it.

  2. Perform calculations using the neutrally buoyant plume and puff models.

  3. Understand qualitative features of dense gas dispersion.

  4. Understand various toxic effect criteria.

  5. Understand hazardous material release prevention and mitigation.

During an incident, process equipment can release hazardous materials quickly and in significant enough quantities to spread in hazardous clouds throughout a plant site and the local community. A few examples are explosive rupture of a process vessel as a result of excessive pressure caused by a runaway reaction, rupture of a pipeline containing toxic materials at high pressure, rupture of a tank containing toxic material stored above its atmospheric boiling point, and rupture of a train or truck transportation tank following an accident.

Hazardous material dispersion includes the release of toxic and flammable materials. However, this chapter focuses solely on toxic material dispersion. Toxic materials are typically measured in parts per million (ppm) by volume levels, whereas flammable gases are in volume percent levels. The dispersion models presented in this chapter are more suitable for the ppm levels of toxic materials.

Serious accidents (such as that in Bhopal, India) emphasize the importance of planning for emergencies and for designing plants to minimize the frequency and consequences of hazardous material releases. Hazardous material release models are routinely used to estimate the consequences of a release on the plant and community environments. This information is useful for adding preventive and mitigative safeguards to the process. Thus, it is important that engineers understand this procedure.

Hazardous material release and dispersion models are an important part of the consequence analysis procedure shown in Figure 4-1. The release model represents the first three steps in the consequence modeling procedure:

  1. Identifying the release incident. Which process situations can lead to a release? This was described in Sections 4-9 and 4-10.

  2. Developing a source model to describe how materials are released and the rate of release. This was presented in Chapter 4.

  3. Estimating the downwind concentrations of the hazardous materials using the dispersion models presented in this chapter. Once the downwind concentrations are known, several criteria are available to estimate the impact or effect, as discussed in Section 5-5.

Various options are available based on the predictions of the dispersion model. This could include applying concepts of inherently safer design by reducing inventories of hazardous materials, substituting with less hazardous materials, and reducing temperatures and pressures, to name a few options. It might also entail developing an emergency response plan for the plant site and with the surrounding community. Other possibilities include developing engineering modifications to eliminate or reduce the source of the release, enclosing the potential release and adding appropriate vent scrubbers or other vapor removal equipment, and adding area monitors to detect incipient leaks and providing block valves and engineering controls to eliminate hazardous levels of spills and leaks. These options are discussed in more detail in Section 5-6 on release prevention and mitigation.

5-1 Parameters Affecting Dispersion

Dispersion models describe the airborne transport of hazardous materials away from the release point and into the plant and community. After a release, the airborne material is carried away by the wind in a characteristic plume (Figure 5-1) or a puff (Figure 5-2). The maximum concentration of material occurs at the release point (which may not be at ground level). Concentrations downwind are smaller, because of turbulent mixing and dispersion of the substance with air.

A storage vessel and plume formed by a continuous release of material are shown. A circular vessel is mounted on a surface. The continuous release of plume occurs from an exit point provided in the vessel. The plume moves toward the direction of wind and dissipates downwind by mixing with fresh air. The wind direction is denoted using a rightward arrow.
Figure 5-1 Characteristic plume formed by a continuous release of material.
Three stages of puff formed by near instantaneous release of a material are shown. The stages of the formed puff are marked as t 0, t 1, and t 2. The puff moves in the direction of wind. t 0 and t 2 are comparatively less in size from t 1.
Figure 5-2 Puff formed by near instantaneous release of material. After the initial release, the puff grows in size but eventually decreases in size.

A wide variety of parameters affect atmospheric dispersion of hazardous materials:

  • Quantity or release rate of material

  • Wind speed

  • Atmospheric stability

  • Ground conditions (buildings, water, trees)

  • Height of the release above ground level

  • Momentum and buoyancy of the material released

As the wind speed increases, the plume in Figure 5-1 becomes longer and narrower. This shape occurs because the released material is carried downwind faster but is diluted faster by a larger quantity of air and increased turbulence.

Atmospheric stability relates to vertical mixing of the air. During the day, the air temperature decreases rapidly with height, encouraging vertical motions. At night, the temperature decrease is less dramatic, resulting in less vertical motion.

Temperature profiles for day and night situations are shown in Figure 5-3. During the day, the sun heats the ground, causing the temperature to be higher on the ground with the temperature decreasing above the ground. During the night, the ground cools rapidly due to thermal radiation to space, resulting in the ground having a lower temperature than the immediate air above it. The presence of clouds will affect both the daytime and nighttime temperature profiles.

A graph depicts the variation in air temperature for day and night conditions.
Figure 5-3 Air temperature as a function of altitude for day and night conditions. The temperature gradient affects the vertical air motion. (Source: Adapted from D. Bruce Turner. Workbook of Atmospheric Dispersion Estimates (Cincinnati, OH: U.S. Department of Health, Education, and Welfare, 1970), p. 1.)

Sometimes an inversion occurs, in which the temperature increases with height, resulting in minimal vertical motion. This condition most often occurs at night.

Atmospheric stability is classified into three possible classes: unstable, neutral, and stable. In unstable atmospheric conditions, the sun heats the ground faster than the heat can be removed, so that the air temperature near the ground is higher than the air temperature at higher elevations, as might be observed in the early morning hours. The instability arises because air of lower density is below air of greater density. This influence of buoyancy enhances atmospheric mechanical turbulence. With neutral stability, the air above the ground warms and the wind speed increases, reducing the effect of solar energy input, or insolation. The air temperature difference does not influence atmospheric mechanical turbulence. In stable atmospheric conditions, the sun cannot heat the ground as fast as the ground cools, so the temperature near the ground is lower than the air temperature at higher elevations. This condition is stable because the air of higher density is below the air of lower density. The influence of buoyancy suppresses mechanical turbulence.

Ground conditions affect the mechanical mixing at the surface and the wind profile with height. Trees and buildings increase mixing, whereas lakes and open areas decrease it. Figure 5-4 shows the change in wind speed versus height for a variety of surface conditions.

A graph shows the variation in the wind gradient.
Figure 5-4 Effect of ground conditions on vertical wind gradient. (Source: Adapted from D. Bruce Turner. Workbook of Atmospheric Dispersion Estimates (Cincinnati, OH: U.S. Department of Health, Education, and Welfare, 1970), p. 2.)

The release height significantly affects ground-level concentrations. As the release height increases, ground-level concentrations are reduced because the plume must disperse a greater distance vertically prior to reaching the ground. This relationship is shown in Figure 5-5.

An industrial chimney releasing plume continuously is shown. The plume moves in the direction of wind, which is shown by an arrow. As release height increases, the distance between the plume releasing source and the place where plume touches the Earth surface increases. The increased distance leads to greater dispersion and a lower concentration at ground level.
Figure 5-5 Increased release height decreases the ground concentration.

The buoyancy and momentum of the material released change the effective height of the release. Figure 5-6 demonstrates these effects. The momentum of a high-velocity jet will carry the gas higher than the point of release, resulting in a much higher effective release height. If the gas has a density less than air, the released gas will initially be positively buoyant and will lift upward. Conversely, if the gas has a density greater than air, then the released gas will initially be negatively buoyant and will slump toward the ground. The temperature and molecular weight of the released gas determine the gas density relative to that of air (with a molecular weight of 28.97). For all gases, as the gas travels downwind and is mixed with fresh air, a point will eventually be reached where the gas has been diluted adequately to be considered neutrally buoyant. At this point the dispersion is dominated by ambient turbulence.

A vessel releasing plume is shown. The first stage of the released plume, as it comes out of the source, represents initial acceleration and dilution. Then, due to dominance of internal buoyancy, the plume starts flowing downward; again, due to dominance of ambient turbulence, it starts flowing upward. The plume flows in the direction of wind, which is shown by an arrow.
Figure 5-6 The initial acceleration and buoyancy of the released material affect the plume character. The dispersion models discussed in this chapter represent only ambient turbulence. (Source: Adapted from Steven R. Hanna and Peter J. Drivas. Guidelines for Use of Vapor Cloud Dispersion Models (New York, NY: American Institute of Chemical Engineers, 1987), p. 6.)

5-2 Neutrally Buoyant Dispersion Models

Neutrally buoyant dispersion models are used to estimate the concentrations downwind of a release in which the gas is mixed with fresh air to the point that the resulting mixture is neutrally buoyant. Thus, these models apply to gases at low concentrations, typically in the ppm range.

Two types of neutrally buoyant vapor cloud dispersion models are commonly used: the plume and the puff models. The plume model describes the steady-state concentration of material released from a continuous source, whereas the puff model describes the temporal concentration of material from a single release of a fixed amount of material. The distinction between the two models is shown graphically in Figures 5-1 and 5-2. For the plume model, a typical example is the continuous release of gases from a smokestack. A steady-state plume is formed downwind from the smokestack. For the puff model, a typical example is the sudden release of a fixed amount of material because of the rupture of a storage vessel. A large vapor cloud is formed that moves downwind away from the release point.

The puff model can be used to describe a plume; a plume is simply the continuous release of puffs. However, if steady-state plume information is all that is required, the plume model is recommended because it is easier to use. For studies involving dynamic plumes (for instance, the effect on a plume of a change in wind direction), the puff model must be used.

Consider the instantaneous release of a fixed mass of material, QmQm, into an infinite expanse of air (a ground surface will be added later). The coordinate system is fixed at the source. Assuming no reaction or molecular diffusion, the concentration C of material resulting from this release is given by the advection equation

Ct+xj(ujC)=0(5-1)Ct+xj(ujC)=0(5-1)

where uj is the velocity of the air and the subscript j represents the summation over all coordinate directions x, y, and z. If the velocity uj in Equation 5-1 is set equal to the average wind velocity and the equation is solved, we would find that the material disperses much faster than predicted. This is due to turbulence in the velocity field. If we could specify the wind velocity exactly with time and position, including the effects resulting from turbulence, Equation 5-1 would predict the correct concentration. Unfortunately, no models are currently available to adequately describe turbulence. As a result, an approximation is used. Let the velocity be represented by an average (or mean) and stochastic quantity

uj=uj+uj(5-2)uj=uj+uj(5-2)

where

uj〉 is the average velocity and

ujuj is the stochastic fluctuation resulting from turbulence.

It follows that the concentration C will also fluctuate as a result of the velocity field; so

C=C+C(5-3)C=C+C(5-3)

where

C〉 is the mean concentration and

C is the stochastic fluctuation.

Because the fluctuations in both C and uj are around the average or mean values, it follows that the average of the fluctuations is zero. Thus,

uj=0C=0(5-4)uj=0C=0(5-4)

Substituting Equations 5-2 and 5-3 into Equation 5-1 and averaging the result over time yields

Ct+xj(ujC)+xjujC=0(5-5)Ct+xj(ujC)+xjujC=0(5-5)

The terms ujC and ujC are zero when averaged ujC=ujC=0, but the turbulent flux ujC term is not necessarily zero and remains in the equation.

At this point, there are two paths forward. The first approach is to define an eddy diffusivity (with units of area/time) such that

uijC=KjCxj(5-6)

However, the eddy diffusivity changes with position, wind velocity, prevailing atmospheric conditions, and time. Also, there is no direct way to measure eddy diffusivities. While the eddy diffusivity approach is useful theoretically, it cannot be practically applied.

The second approach is to use dispersion coefficients, which are described in the next section.

5-3 Pasquill–Gifford Model

Sutton1 proposed the following definition for a dispersion coefficient:

σ2x=12C2(ut)2n(5-7)

1O. G. Sutton. Micrometeorology (New York, NY: McGraw-Hill, 1953), p. 286.

with similar expressions given for σy and σz. The parameter n is used to fit the data. The dispersion coefficients σx, σy, and σz represent the standard deviations of the concentration in the downwind, crosswind, and vertical (x, y, z) directions, respectively. Values for the dispersion coefficients are much easier to measure experimentally than eddy diffusivities since the dispersion coefficients can be determined by measuring the actual downwind concentrations.

The dispersion coefficients are a function of atmospheric conditions and the distance downwind from the release. The atmospheric conditions are classified according to six different stability classes, shown in Table 5-1. These stability classes depend on wind speed and quantity of sunlight. During the day, increased wind speed results in greater atmospheric stability, whereas at night the reverse is true. This is due to a change in vertical temperature profiles from day to night, as shown in Figure 5-3.

Table 5-1 Atmospheric Stability Classes for Use with the Pasquill–Gifford Dispersion Model a, b

 

 

 

 

Nighttime conditionsd

 

Daytime insolationc

Thin overcast

Surface wind speed (m/s)

Strong

Moderate

Slight

or >4/8 Low cloud

≤3/8 Cloudiness

<2

A

A–B

B

Fe

Fe

2–3

A–B

B

C

E

F

3–4

B

B–C

C

D f

E

4–6

C

C–D

D f

D f

D f

>6

C

D f

D f

D f

D f

Stability classes:

A, extremely unstable

B, moderately unstable

C, slightly unstable

D, neutrally stable

E, slightly stable

F, moderately stable

aF. A. Gifford. “Use of Routine Meteorological Observations for Estimating Atmospheric Dispersion.” Nuclear Safety 2, no. 4. (1961): 47.

bF. A. Gifford. “Turbulent Diffusion-Typing Schemes: A Review.” Nuclear Safety 17, no. 1 (1976): 68.

cStrong insolation corresponds to a sunny midday in midsummer in England. Slight insolation corresponds to similar conditions in midwinter.

dNight refers to the period 1 hr before sunset and 1 hr after dawn.

eThese values are filled in to complete the table.

fThe neutral category D should be used, regardless of wind speed, for overcast conditions during day or night and for any sky conditions during the hour before or after sunset or sunrise, respectively.

The dispersion coefficients σy and σz for a continuous source are given in Figures 5-7 and 5-8, with the corresponding correlations given in Table 5-2. Values for σx are not provided because it is reasonable to assume that σx = σy. The dispersion coefficients σy and σz for a puff release are given in Figure 5-9 and the equations are provided in Table 5-3. The puff dispersion coefficients are based on limited data (shown in Table 5-3) and should not be considered precise.

A graph of dispersion coefficients versus distance downwind.
Figure 5-7 Dispersion coefficients for Pasquill–Gifford plume model for rural releases.
A graph of dispersion coefficients versus distance downwind.
Figure 5-8 Dispersion coefficients for Pasquill–Gifford plume model for urban releases.
A graph of dispersion coefficients versus distance downwind.
Figure 5-9 Dispersion coefficients for Pasquill–Gifford puff model.

Table 5-2 Recommended Equations for Pasquill–Gifford Dispersion Coefficients for Plume Dispersion

Pasquill–gifford Stability class

σy(m)

σz(m)

Rural conditions

 

 

A

0.22x(1 + 0.0001x)–1/2

0.20x

B

0.16x(1 + 0.0001x)–1/2

0.12x

C

0.11x(1 + 0.0001x)–1/2

0.08x(1 + 0.0002x)–1/2

D

0.08x(1 + 0.0001x)–1/2

0.06x(1 + 0.0015x)–1/2

E

0.06x(1 + 0.0001x)–1/2

0.03x(1 + 0.0003x)–1

F

0.04x(1 + 0.0001x)–1/2

0.016x(1 + 0.0003x)–1

Urban conditions

 

 

A–B

0.32x(1 + 0.0004x)–1/2

0.24x(1 + 0.001x)+1/2

C

0.22x(1 + 0.0004x)–1/2

0.20x

D

0.16x(1 + 0.0004x)–1/2

0.14x(1 + 0.0003x)–1/2

E–F

0.11x(1 + 0.0004x)–1/2

0.08x(1 + 0.0015x)–1/2

The downwind distance x has units of meters. A–F are defined in Table 5-1.

Sources: R. F. Griffiths. “Errors in the Use of the Briggs Parameterization for Atmospheric Dispersion Coefficients.” Atmospheric Environment 28, no. 17 (1994): 2861–2865; G. A. Briggs. Diffusion Estimation for Small Emissions, Report ATDL-106 (Washington, DC: Air Resources, Atmospheric Turbulence, and Diffusion Laboratory, Environmental Research Laboratories, 1974).

Table 5-3 Recommended Equations for Pasquill–Gifford Dispersion Coefficients for Puff Dispersion

Pasquill–gifford stability class

σy(m) or σx (m)

σz(m)

A

0.18x0.92

0.60x0.75

B

0.14x0.92

0.53x0.73

C

0.10x0.92

0.34x0.71

D

0.06x0.92

0.15x0.70

E

0.04x0.92

0.10x0.65

F

0.02x0.89

0.05x0.61

The downwind distance x has units of meters. A–F are defined in Table 5-1.

Sources: R. F. Griffiths. “Errors in the Use of the Briggs Parameterization for Atmospheric Dispersion Coefficients.” Atmospheric Environment
28, no. 17 (1994): 2861–2865; G. A. Briggs. Diffusion Estimation for Small Emissions, Report ATDL-106 (Washington, DC: Air Resources, Atmospheric Turbulence, and Diffusion Laboratory, Environmental Research Laboratories, 1974).

The equations for cases 1 through 4 that follow were derived by Pasquill2 using expressions of the form of Equation 5-7. These equations along with the correlations for the dispersion coefficients are known as the Pasquill–Gifford model.

2F. Pasquill. Atmospheric Diffusion (London, UK: Van Nostrand, 1962).

The geometry for the dispersion models is shown in Figure 5-10. The immediate downwind direction is x, the crosswind direction is y and the height above ground level is z. The origin at (x, y, z) is on the ground immediately below the release point. The height of the release above ground is given by Hr.

Dispersion modeling geometry is shown. The x axis represents downwind direction, the y axis represents crosswind direction, and the z axis represents height above ground level. The origin is marked as (x, y, z) or (0, 0, 0). The release height is marked as H r. The plume moves in the direction of wind, which is shown using a rightward arrow.
Figure 5-10 Geometry for dispersion modeling.

Case 1: Puff with Instantaneous Point Source at Ground Level, Coordinates Fixed at Release Point, Constant Wind Only in x Direction with Constant Velocity u

This case is represented by the following equation. Details of the derivation of Equation 5-8 are provided in the supplemental content on the book’s website.

C(x,y,z,t)=Q*m2π3/2σxσyσzexp{12[(xutσx)2+y2σ2y+z2σ2z]}(5-8)

where Qm is the instantaneous release mass.

The time dependence in Equation 5-8 is achieved through the dispersion coefficients because their values change as the puff moves downwind. If wind is absent, u = 0, and Equation 5-8 does not produce a valid result since the puff does not move downwind at all.

The ground-level concentration is given at z = 0:

C(x,y,0,t)=Q*m2π3/2σxσyσzexp{12[(xutσx)2+y2σ2y]}(5-9)

The ground-level concentration directly downwind from the release along the x axis is given at y = z = 0:

C(x,0,0,t)=Q*m2π3/2σxσyσzexp{12(xutσx)2}(5-10)

The center of the cloud is found at coordinates (ut, 0, 0). The concentration at the center of this moving cloud is given by

C(ut,0,0,t)=Q*m2π3/2σxσyσz(5-11)

The total integrated dose Dtid received by an individual standing at fixed location (x, y, z) is the time integral of the concentration:

Dtid(x,y,z)=0C(x,y,z,t)dt(5-12)

The total integrated dose at ground level is found by integrating Equation 5-9 according to Equation 5-12. The result is

Dtid(x,y,0)=Q*mπσyσzuexp(12y2σ2y)(5-13)

The total integrated dose along the x axis on the ground directly downwind from the release is

Dtid(x,0,0)=Q*mπσyσzu(5-14)

Case 2: Plume with Continuous Steady-State Source at Ground Level and Wind Moving in x Direction at Constant Velocity u

This case is represented by the following equation:

C(x,y,z)=Qmπσyσzuexp[12(y2σ2y+z2σ2z)](5-15)

where Qm is the mass release rate (mass/time)

The ground-level concentration is given at z = 0:

C(x,y,0)=Qmπσyσzuexp[12(y2σ2y)](5-16)

The ground concentration along the centerline of the plume directly downwind is given at y = z = 0:

C(x,0,0)=Qmπσyσzu(5-17)

For continuous ground-level releases, the maximum concentration occurs at the release point.

Case 3: Plume with Continuous Steady-State Source at Height Hr above Ground Level and Wind Moving in x Direction at Constant Velocity u

The solution for this case is

C(x,y,z)=Qm2πσyσzuexp[12(yσy)2]  ×{exp[12(zHrσz)2]+exp[12(z+Hrσz)2]}(5-18)

The ground-level concentration is found by setting z = 0:

C(x,y,0)=Qmπσyσzuexp[12(yσy)212(Hrσz)2](5-19)

The ground-level centerline concentrations are found by setting y = z = 0:

C(x,0,0)=Qmπσyσzuexp[12(Hrσz)2](5-20)

The maximum ground-level concentration 〈Cmax along the x axis is found using

Cmax=2QmeπuH2r(σzσy)(5-21)

The distance downwind at which the maximum ground-level concentration occurs is found from

σz=Hr2(5-22)

The procedure for finding the maximum concentration and the downwind distance is to use Equation 5-22 to determine the downwind distance, followed by using Equation 5-21 to determine the maximum concentration.

Case 4: Puff with Instantaneous Point Source at Height Hr above Ground Level and a Coordinate System on the Ground That Moves with the Puff

For this case, the center of the puff is found at x = ut. The average concentration is given by

C(x,y,z,t)=Q*m(2π)3/2σxσyσzexp[12(yσy)2]×{exp[12(zHrσz)2]+exp[12(z+Hrσz)2]}(5-23)

At ground level, z = 0, and the concentration is computed using

C(x,y,0,t)=Q*m2π3/2σxσyσzexp[12(yσy)212(Hrσz)2](5-24)

The concentration along the ground at the centerline is given at y = z = 0:

C(x,0,0,t)=Q*m2π3/2σxσyσzexp[12(Hrσz)2](5-25)

The total integrated dose at ground level is found by using Equation 5-25 in Equation 5-12. The result is

Dtid(x,y,0)=Q*mπσyσzuexp[12(yσy)212(Hrσz)2](5-26)

Case 5: Puff with Instantaneous Point Source at Height Hr above Ground Level and a Coordinate System Fixed on the Ground at the Release Point

For this case, the result is obtained using a transformation of coordinates. If the puff moves with the wind along the x axis, the equation for this case is found by replacing the existing coordinate x by a new coordinate system, xut, that moves with the wind velocity. The result is

C(x,y,z,t)=[Puff equations with moving coordinate system (Equations 5-23 through 5-25)(5-27)×exp[12(xutσx)2]

where t is the time since the release of the puff.

Isopleths

Frequently the cloud boundary defined by a fixed concentration is required for either the plume or the puff. The line connecting points of equal concentration around the cloud boundary is called an isopleth. For a specified concentration 〈C*, the isopleth at ground level is determined by dividing the equation for the centerline concentration (Equation 5-10 for a puff and Equation 5-16 for a plume) by the equation for the general ground-level concentration (Equation 5-9 for a puff and Equation 5-16 for a plume). This equation is solved directly for y:

y=σy2ln(C(x,0,0,t)C(x,y,0,t))(5-28)

The procedure is as follows:

  1. Specify the isopleth concentration 〈C*, the wind speed u, and time t.

  2. Determine the concentrations 〈C〉 (x, 0, 0, t) at fixed points along the x axis using Equation 5-10 for a puff and Equation 5-16 for a plume. Define the boundary of the cloud along the x axis.

  3. Set 〈C〉 (x, y, 0, t) = 〈C* in Equation 5-28 and determine the values of y at each centerline point determined in step 2.

  4. Plot dots at each +y and –y at each centerline point x.

  5. Connect the dots to obtain the isopleth.

The procedure is repeated for each value of t required.

Effect of Release Momentum and Buoyancy

Figure 5-5 shows that the release characteristics of a puff or plume depend on the initial release momentum and buoyancy. The initial momentum and buoyancy change the effective height of the release. A release that occurs at ground level but in an upward spouting jet of vaporizing liquid has a greater effective height than a release without a jet. Similarly, a release of vapor at a temperature higher than the ambient air temperature will rise because of buoyancy effects, increasing the effective height of the release.

Both effects are demonstrated by the traditional smokestack release shown in Figure 5-11. The material released from the smokestack contains momentum, based on its upward velocity within the stack pipe, and it is also buoyant, because its temperature is higher than the ambient temperature. As a consequence, the material continues to rise after its release from the stack. The upward rise is slowed and eventually stopped as the released material cools and the momentum is dissipated.

A source releasing hot gases continuously is shown. The formed plume moves in the direction of wind, which is shown by an arrow. The released gases cool as they mix and dilute with cool air. As plume moves away from the source, due to neutral buoyancy, it flows slightly above the releasing source.
Figure 5-11 Smokestack plume demonstrating initial buoyant rise of hot gases.

For smokestack releases, Turner3 suggested using the empirical Holland formula to compute the additional height resulting from the buoyancy and momentum of the release:

3D. Bruce Turner. Workbook of Atmospheric Dispersion Estimates (Cincinnati, OH: U.S. Department of Health, Education, and Welfare, 1970), p. 31.

ΔHr=ˉusdˉu[1.5+2.68×103Pd(TsTaTs)](5-29)

where

ΔHr is the correction to the release height Hr,

ˉus is the stack gas exit velocity (m/s),

d is the inside stack diameter (m),

ˉu is the wind speed (m/s),

P is the atmospheric pressure (millibar),

Ts is the stack gas temperature (K), and

Ta is the air temperature (K).

For heavier-than-air vapors, if the material is released above ground level, then the material will initially fall toward the ground until it disperses enough to reduce the cloud density.

Worst-Case Dispersion Conditions

For a plume, the highest concentration is always found at the release point. If the release occurs above ground level, then the highest concentration on the ground is found at a point downwind from the release.

For a puff, the maximum concentration is always found at the puff center. For a release above ground level, the puff center will move parallel to the ground and the maximum concentration on the ground will occur directly below the puff center. For a puff isopleth, the isopleth is close to circular as it moves downwind. The diameter of the isopleth increases initially as the puff travels downwind, reaches a maximum, and then decreases in diameter.

If weather conditions are not known or are not specified, then certain assumptions can be made to result in a worst-case result—that is, the highest concentration is estimated. The weather conditions in the Pasquill–Gifford dispersion equations are included by means of the dispersion coefficients and the wind speed. By examining the Pasquill–Gifford equations, it is readily evident that the dispersion coefficients and wind speed are in the denominator. In turn, the maximum concentration is estimated by selecting the weather conditions and wind speed that result in the smallest values of the dispersion coefficients and the wind speed. By inspecting Figures 5-7 through 5-9, we can see that the smallest dispersion coefficients occur with F stability. Clearly, the wind speed cannot be zero, so a finite value must be selected. The EPA4 suggests that F stability can exist with wind speeds as low as 1.5 m/s. In contrast, other risk analysts use a wind speed of 2 m/s to be consistent with the F stability class. For daytime conditions, some risk analysts use D stability with 3 m/s wind speed. The assumptions used in the calculation must be clearly reported.

4U.S. Environmental Protection Agency. RMP Offsite Consequence Analysis Guidance (Washington, DC: U.S. Environmental Protection Agency, 1996).

The worst case on the ground is frequently of interest, given that this is where people are normally present. For the plume and puff cases, we can select the following:

Plume: Along the centerline downwind from the release on ground:

C(x,0,0)=Qmπσyσzu(5-30)

Puff: At the center of the puff with the release on the ground:

C(x,0,0)=Q*m2π3/2σxσyσzu(5-31)

The maximum concentration is obtained by:

  1. Specifying the release rate, Qm and Q*m, at the maximum value.

  2. Specifying all the dispersion coefficients, σ, at the minimum value. This implies F stability.

  3. Specifying the wind speed, u, at a minimum within the stability class. For F stability, this implies 2 m/s or the EPA-suggested value of 1.5 m/s. For daytime, some risk analysts use D stability with 3 m/s wind speed.

Limitations to Pasquill–Gifford Dispersion Modeling

Pasquill–Gifford or Gaussian dispersion applies only to neutrally buoyant dispersion of gases in which the turbulent mixing is the dominant feature of the dispersion. It is typically valid only for a distance of 0.1–10 km from the release point.

It is not always clear whether a plume or puff model should be selected based on the nature of the release: The two models may produce varying results. A release of less than 15 seconds is typically modeled as an instantaneous puff release, but what about a release of 30 seconds?

The concentrations predicted by the Gaussian models are time averages. Thus, it is possible for instantaneous local concentrations to exceed the average values predicted—a difference that might be important for emergency response. The models presented here assume a 10-minute time average. Actual instantaneous concentrations may vary by as much as a factor of 2 from the concentrations computed using Gaussian models.

Example 5-1

Assume that 10 kg/s of hydrogen sulfide (H2S) gas is released 100 m above the ground. It is a clear, sunny day, 1:00 pm, with a wind speed of 3.5 m/s. Assume rural conditions. The temperature is 30°C and the pressure is 1 atm. The molecular weight of H2S is 34.08. Determine

  1. The concentration 1 km downwind on the ground.

  2. The maximum concentration from this release and the distance downwind.

  3. The maximum discharge rate, in kg/s, to result in a maximum concentration downwind of 10 ppm.

Solution

  1. This is a plume since it is a continuous release. From Table 5-1 for a clear sunny day (strong insolation) and a wind speed of 3.5 m/s, the stability class is B. From Table 5-2 for rural releases,

    σy=0.16x(1+0.0001x)1/2=0.16(1000 m)[1+(0.0001)(1000 m)]1/2σy=153 mσz=0.12x=0.12(1000 m)=120 m

    These dispersion coefficients could also be determined using Figure 5-7.

    Since this is a plume, with a release above ground level, Case 3 and Equation 5-20 applies:

    C(x,0,0)=Qmπσyσzuexp[12(Hrσz)2]

    Substituting the numbers

    C(x,0,0)=10.0 kg/s(3.14)(153 m)(120 m)(3.5 m/s)exp[12(100 m120 m)2]=3.50×105 kg/m3=35.0 mg/m3

    Use Equation 2-7 to convert to ppm, being careful to use the correct units:

    Cppm=0.08205(TPM)(mg/m3)=0.08205[303 K(1 atm)(34.08 g/mol)](35.0 mg/m3)=25.6 ppm

  2. Use Equation 5-22 to determine the value of σz at the location of the maximum concentration.

    σz=Hr2=100 m1.414=70.7 m

    Then use the equation in Table 5-3 to determine the downwind distance.

    σz=0.12x70.7 m=0.12xx=589 m

    Use the equation in Table 5-2 (or Figure 5-7) to determine the value of σy at this location,

    σy=0.16x(1+0.0001x)1/2=0.16(589 m)[1+(0.0001)(589 m)]1/2=91.6 m

    Now use Equation 5-21 to determine the maximum concentration at this location downwind.

    Cmax=2QmeπuH2r(σzσy)=(2)(100 ;kg/s)(2.718)(3.14)(3.5 m/s)(100 m)2(70.7 m91.6 m)=5.17×104 kg/m3=517 mg/m3

    This is 378 ppm from Equation 2-7.

  3. From Equation 2-7, 10 ppm = 13.7 mg/m3 =13.7×10-6kg/m3. Substituting into Equation 5-21 and solving for Qm,

Cmax =2QmeπuH2r(σzσy)13.7×106 kg/m3=2Qm(2.71)(3.14)(3.5 m/s)(100 m)2(70.7 m91.6 m)Qm=2.64 kg/s

Example 5-2

Assume that 10 kg of hydrogen sulfide (H2S) is released instantly on the ground. The atmospheric conditions are the same as Example 5-1.

  1. What is the maximum concentration at the fence line 100 m away?

  2. If the wind is 3.5 m/s, how long does it take for the puff to reach the fence line?

Solution

  1. This is a puff since the release is instantaneous. From Table 5-1, the stability class is B. From Table 5-3 for stability class B,

    σy=σx=0.14x0.92=(0.14)(100 m)0.92=9.7 mσz=0.53x0.73=(0.53)(100 m)0.73=15.3 m

    The maximum concentration is found at the center of the puff on the ground. Equation 5-11 applies:

    C(ut, 0 , 0, t)=Q*m2π3/2σxσyσz=10 kg(1.41)(3.14)3/2(9.7 m)(9.7 m)(15.3 m)=8.85×104 kg/m3=885 mg/m3

    From Equation 2-7, this is 647 ppm.

  2. The time is found from

x=utt=xu=100 m3.5 m/s=28.6 s

This is not very much time to mount much of an emergency response.

5-4 Dense Gas Dispersion

A dense gas is defined as any gas whose density is greater than the density of the ambient air through which it is being dispersed. This result can be due to a gas with a molecular weight greater than that of air or a gas with a low temperature resulting from auto-refrigeration during release or other processes.

Following a typical puff release, a cloud having similar vertical and horizontal dimensions (near the source) may form. The dense cloud will slump toward the ground under the influence of gravity, increasing its diameter and reducing its height. Considerable initial dilution will occur because of the gravity-driven intrusion of the cloud into the ambient air. Subsequently, the cloud height will increase due to further entrainment of air across both the vertical and the horizontal interfaces. After sufficient dilution occurs, normal atmospheric turbulence predominates over gravitational forces and typical Gaussian dispersion characteristics are exhibited.

Britter and McQuaid5 developed a model of dense gas dispersion by performing a dimensional analysis and correlating existing data on this phenomenon. Their model is best suited for instantaneous or continuous ground-level releases of dense gases. The release is assumed to occur at ambient temperature and without aerosol or liquid droplet formation. Atmospheric stability was found to have little effect on the results and is not a part of the model. Most of the data came from dispersion tests in remote rural areas on mostly flat terrain, so the results are not applicable to areas where terrain effects are significant.

5R. E. Britter and J. McQuaid. Workbook on the Dispersion of Dense Gases (Sheffield, UK: Health and Safety Executive, 1988).

Dense gas dispersion is beyond the scope of this book. Readers are referred to supplemental content on this subject provided on the book’s website.

5-5 Toxic Effect Criteria

Once the dispersion calculations are completed, the question arises: What concentration is considered hazardous? Concentrations based on TLV-TWA values, discussed in Chapter 2, are overly conservative and are designed for worker exposures, not short-term exposures under emergency conditions.

One approach is to use the probit equations developed in Chapter 2. These equations are also capable of including the effects resulting from transient changes in toxic concentrations. Unfortunately, published correlations are available for only a few chemicals, and actual data show wide variations from the correlations.

One simplified approach is to specify a toxic concentration criterion above which it is assumed that individuals exposed to this value or greater will be in danger. This approach has led to many criteria promulgated by several government agencies and private associations:

  • Emergency response planning guidelines (ERPGs) for air contaminants issued by the American Industrial Hygiene Association (AIHA)

  • Immediately dangerous to life and health (IDLH) levels established by U.S. National Institute of Occupational Safety and Health (NIOSH), already presented in Chapter 2

  • Emergency exposure guidance levels (EEGLs) and short-term public emergency guidance levels (SPEGLs) issued by the National Academy of Sciences/National Research Council (NRC)

  • Acute exposure guideline levels (AEGLs) established by the U.S. Environmental Protection Agency (EPA)

  • Threshold limit values (TLVs) established by the American Conference of Governmental Industrial Hygienists (ACGIH), including short-term exposure limits (TLV-STELs) and ceiling concentrations (TLV-Cs), already presented in Chapter 2

  • Permissible exposure limits (PELs) promulgated by U.S. Occupational Safety and Health Administration (OSHA), already presented in Chapter 2

  • Toxic endpoints promulgated by the EPA as part of the Risk Management Plan (RMP) regulation

These criteria and methods are based on a combination of results from animal experiments, observations of long- and short-term human exposures, and expert judgment. The following subsections define these criteria and describe some of their features.

Emergency Response Planning Guidelines

ERPGs are prepared by an industry task force and are published by the AIHA. Three concentration ranges are provided as a consequence of exposure to a specific substance:

  1. ERPG-1 is the maximum airborne concentration below which it is believed nearly all individuals could be exposed for up to 1 hour without experiencing effects other than mild transient adverse health effects or perceiving a clearly defined objectionable odor.

  2. ERPG-2 is the maximum airborne concentration below which it is believed nearly all individuals could be exposed for up to 1 hour without experiencing or developing irreversible or other serious health effects or symptoms that could impair their abilities to take protective action.

  3. ERPG-3 is the maximum airborne concentration below which it is believed nearly all individuals could be exposed for up to 1 hour without experiencing or developing life-threatening health effects.

ERPG data are shown in Table 5-4. To date, 47 ERPGs have been developed and are being reviewed, updated, and expanded by an AIHA peer review task force. Because of the comprehensive effort to develop acute toxicity values, ERPGs are becoming an acceptable industry/government norm.

Table 5-4 Emergency Response Planning Guidelines (ERPGs), in ppm

Chemical

ERPG-1

ERPG-2

ERPG-3

Acetaldehyde

10

200

1000

Acetic acid

5

35

250

Acetic anhydride

0.5

15

100

Acrolein

0.05

0.15

1.5

Acrylic acid

1.0

50

250

Acrylonitrile

10

35

75

Allyl chloride

3

40

300

Ammonia

25

150

750

Benzene

50

150

1000

Benzyl chloride

1

10

50

Beryllium

NA

25 μg/m3

100 μg/m3

Bromine

0.1

0.5

5

1,3-Butadiene

10

200

5000

n-Butyl acrylate

0.05

25

250

n-Butyl isocyanate

0.01

0.05

1

Carbon disulfide

1

50

500

Carbon monoxide

200

350

500

Carbon tetrachloride

20

100

750

Chlorine

1

3

20

Chlorine trifluoride

0.1

1

10

Chloroacetyl chloride

0.05

0.5

10

Chloroform

NA

50

5000

Chloropicrin

0.1

0.3

1.5

Chlorosulfonic acid

2 mg/m3

10 mg/m3

30 mg/m3

Chlorotrifluoroethylene

20

100

300

Crotonaldehyde

0.2

1

3

Diborane

NA

1

3

1,2-Dichloroethane

50

200

300

Diketene

1

5

20

Dimethylamine

0.6

100

350

Dimethyldichlorosilane

2

10

75

Dimethyl disulfide

0.01

50

250

Epichlorohydrin

5

20

100

Ethanol

1800

3300

NA

Ethyl chloroformate

ID

100

200

Ethylene oxide

NA

50

500

Fluorine

0.5

5

20

Formaldehyde

1

10

25

Formic acid

3

25

250

Gasoline

200

1000

4000

Hexachlorobutadiene

1

3

10

Hexafluoroacetone

NA

1

50

Hexafluoropropylene

10

50

500

Hydrogen chloride

3

20

150

Hydrogen cyanide

NA

10

25

Hydrogen fluoride

2

20

50

Hydrogen peroxide

10

50

100

Hydrogen sulfide

0.1

30

100

Isobutyronitrile

10

50

200

2-Isocyanatoethyl methacrylate

ID

0.1

1

Lithium hydride

25 μg/m3

100 μg/m3

500 μg/m3

Methanol

200

1000

5000

Methyl chloride

NA

400

1000

Methylene diphenyl diisocynate

0.2 mg/m3

2 mg/m3

25 mg/m3

Methyl isocyanate

0.025

0.25

1.5

Methyl mercaptan

0.005

25

100

Methyltrichlorosilane

0.5

3

15

Monomethylamine

10

100

500

Nitrogen dioxide

1

15

30

Perfluoroisobutylene

NA

0.1

0.3

Phenol

10

50

200

Phosgene

NA

0.5

1.5

Phosphorus pentoxide

1 mg/m3

10 mg/m3

50 mg/m3

Propylene oxide

50

250

750

Styrene

50

250

1000

Sulfonic acid (oleum, sulfur
trioxide, and sulfuric acid)

2 mg/m3

10 mg/m3

120 mg/m3

Sulfur dioxide

0.3

3

15

Tetrafluoroethylene

200

1000

10,000

Titanium tetrachloride

5 mg/m3

20 mg/m3

100 mg/m3

Toluene

50

300

1000

Trimethylamine

0.1

100

500

Uranium hexafluoride

5 mg/m3

15 mg/m3

30 mg/m3

Vinyl acetate

5

75

500

Vinyl chloride

500

5000

20,000

Notes: NA = Not appropriate, ID = insufficient data, µg = microgram. All values are in ppm unless otherwise noted.

Source: American Industrial Hygiene Association. Emergency Response Planning Guidelines and Workplace Environmental Exposure Levels (Fairfax, VA: American Industrial Hygiene Association, 2010). www.aiha.org.

Immediately Dangerous to Life and Health

NIOSH publishes IDLH concentrations to be used as acute toxicity measures for common industrial gases. An IDLH exposure condition is defined as a condition “that poses a threat of exposure to airborne contaminants when that exposure is likely to cause death or immediate or delayed permanent adverse health effects or prevent escape from such an environment.”6

6U.S. National Institute of Occupational Safety and Health. NIOSH Pocket Guide to Chemical Hazards (Washington, DC: U.S. Department of Health and Human Services, 2005). www.cdc.gov/niosh/npg/default.html.

IDLH concentrations also take into consideration acute toxic reactions, such as severe eye irritation, that could prevent escape from the contaminated site. The IDLH concentration is considered a maximum concentration above which only a highly reliable breathing apparatus providing maximum worker protection is permitted. If IDLH concentrations are exceeded, all unprotected workers must leave the area immediately.

IDLH data are currently available for a number of chemicals (see www.cdc.gov/niosh/idlh/intridl4.html). IDLH concentrations are developed to protect healthy worker populations and are not directly suitable for the general population, which includes sensitive populations such as older, disabled, or ill individuals. For flammable vapors, the IDLH concentration is defined as one-tenth of the lower flammability limit (LFL) concentration. Note that IDLH concentrations have not been peer-reviewed and that no substantive documentation for the values exists.

Emergency Exposure Guidance Levels and Short-Term Public Emergency Guidance Levels

Since the 1940s, the NRC’s Committee on Toxicology has submitted EEGLs for chemicals of special concern to the Department of Defense. An EEGL is defined as a concentration of a gas, vapor, or aerosol that is judged acceptable and that allows exposed individuals to perform specific tasks during emergency conditions lasting from 1 to 24 hours. Exposure to concentrations at the EEGL may produce transient irritation or central nervous system effects but should not produce effects that are lasting or that would impair performance of a task.

In addition to EEGLs, the National Research Council has developed SPEGLs, defined as acceptable concentrations for exposures of members of the general public. SPEGLs are generally set at 10–50% of the EEGL and are calculated to take into account the effects of exposure on sensitive heterogeneous populations.

The advantages of using EEGLs and SPEGLs rather than IDLH concentrations are
(1) a SPEGL considers effects on sensitive populations, (2) EEGLs and SPEGLs are developed for several different exposure durations, and (3) the methods by which EEGLs and SPEGLs were developed are well documented in NRC publications. EEGL and SPEGL values are shown in Table 5-5.

Table 5-5 Emergency Exposure Guidance Levels (EEGLs) from the National Research Council, in ppm

Compound

1-hr EEGL

24-hr EEGL

Source

Acetone

8500

1000

NRC I

Acrolein

0.05

0.01

NRC I

Aluminum oxide

15 mg/m3

100

NRC IV

Ammonia

100

 

NRC VII

Arsine

1

0.1

NRC I

Benzene

50

2

NRC VI

Bromotrifluoromethane

25,000

 

NRC III

Carbon disulfide

50

 

NRC I

Carbon monoxide

400

50

NRC IV

Chlorine

3

0.5

NRC II

Chlorine trifluoride

1

 

NRC II

Chloroform

100

30

NRC I

Dichlorodifluoromethane

10,000

1000

NRC II

Dichlorofluoromethane

100

3

NRC II

Dichlorotetrafluoroethane

10,000

1000

NRC II

1,1-Dimethylhydrazine

0.24a

0.01a

NRC V

Ethanolamine

50

3

NRC II

Ethylene glycol

40

20

NRC IV

Ethylene oxide

20

1

NRC VI

Fluorine

7.5

 

NRC I

Hydrazine

0.12a

0.005a

NRC V

Hydrogen chloride

20/1a

20/1a

NRC VII

Hydrogen sulfide

 

10

NRC IV

Isopropyl alcohol

400

200

NRC II

Lithium bromide

15 mg/m3

7 mg/m3

NRC VII

Lithium chromate

100 μg/m3

50 μg/m3

NRC VIII

Mercury (vapor)

 

0.2 mg/m3

NRC I

Methane

 

5000

NRC I

Methanol

200

10

NRC IV

Methylhydrazine

0.24a

0.01a

NRC V

Nitrogen dioxide

1a

0.04a

NRC IV

Nitrous oxide

10,000

 

NRC IV

Ozone

1

0.1

NRC I

Phosgene

0.2

0.02

NRC II

Sodium hydroxide

2 mg/m3

 

NRC II

Sulfur dioxide

10

5

NRC II

Sulfuric acid

1 mg/m3

 

NRC I

Toluene

200

100

NRC VII

Trichloroethylene

200

10

NRC VIII

Trichlorofluoromethane

1500

500

NRC II

Trichlorotrifluoroethane

1500

500

NRC II

Vinylidene chloride

 

10

NRC II

Xylene

200

100

NRC II

Note: All values are in ppm unless otherwise noted.

aSPEGL value.

Acute Exposure Guideline Levels

AEGLs are developed by the U.S. EPA and are used by emergency planners and responders as guidance in dealing with accidental releases of chemicals. They are expressed as specific concentrations of airborne chemicals at which health effects may occur. These guidelines are designed to protect the elderly and children and other individuals who may be susceptible to exposures.

AEGLs are determined for five relatively short exposure periods: 10 minutes, 30 minutes, 1 hour, 4 hours, and 8 hours. There are three levels—1, 2, and 3—based on the severity of the toxic effects caused by the exposure, with level 1 being the least severe and level 3 being the most severe. The three levels are further defined as follows:

Level 1: Notable discomfort, irritation, or certain asymptomatic nonsensory effects. However, the effects are not disabling and are transient and reversible upon cessation of exposure.

Level 2: Irreversible or other serious, long-lasting adverse health effects or an impaired ability to escape.

Level 3: Life-threatening health effects or death.

Airborne concentrations below the AEGL-1 levels represent exposure levels that could produce mild and progressively increasing but transient and nondisabling odor, taste, and sensory irritation or certain asymptomatic, nonsensory effects. With increasing airborne concentrations above each AEGL level, there is a progressive increase in the likelihood of occurrence and the severity of effects.

AEGL levels represent threshold levels for the general public—including susceptible populations such as infants, children, elderly individuals, persons with asthma, and those with other illnesses. However, specific individuals subject to unique responses, could experience effects below the corresponding AEGL levels. Table 5-6 provides a selected list of AEGLs.

Table 5-6 Selected Acute Exposure Guideline Levels (AEGLs) from U.S. EPA, in ppm

Chemical name

AEGL-1

 

AEGL-2

 

AEGL-3

10 min

30 min

1 hr

4 hr

8 hr

 

10 min

30 min

4 hr

1 hr

8 hr

 

10 min

30 min

1 hr

4 hr

8 hr

Carbon tetrachloride

NR

NR

NR

NR

NR

27

18

13

7.6

5.8

700

450

340

200

150

Aniline

48

16

8.0

2.0

1.0

72

24

12

3.0

1.5

120

40

20

5.0

2.5

Chloroform

NR

NR

NR

NR

NR

120

80

64

40

29

4000

4000

3200

2000

1600

Methyl bromide

NR

NR

NR

NR

NR

940

380

210

67

67

3300

1300

740

230

130

Methyl chloride

NR

NR

NR

NR

NR

1100

1100

910

570

380

3800

3800

3000

1900

1300

Hydrogen cyanide

2.5

2.5

2.0

1.3

1.0

17

10

7.1

3.5

2.5

27

21

15

8.6

6.6

Vinyl chloride

450

310

250

140

70

2800

1600

1200

820

820

12,000

6800

4800

3400

3400

Acetonitrile

13

13

13

13

NR

80

80

50

21

14

240

240

150

64

42

Carbon disulfide

17

17

13

8.4

6.7

200

200

160

100

50

600

600

480

300

150

Ethylene oxide

NR

NR

NR

NR

NR

80

80

45

14

7.9

360

360

200

63

35

Phosgene

NR

NR

NR

NR

NR

0.60

0.60

0.30

0.08

0.04

3.6

1.5

0.75

0.20

0.09

Propylene oxide

73

73

73

73

73

440

440

290

130

86

1300

1300

870

390

260

Methyl ethyl ketone

200

200

200

200

200

4900a

3400a

2700a

1700

1700

b

b

4000a

2500a

2500a

Epichlorohydrin

1.7

1.7

1.7

1.7

1.7

53

53

24

14

6.7

570

160

72

44

20

Acrolein

0.03

0.03

0.03

0.03

0.03

0.44

0.18

0.10

0.10

0.10

6.2

2.5

1.4

0.48

0.27

Acrylonitrile

1.5

1.5

NR

NR

NR

8.6

3.2

1.7

0.48

0.26

130

50

28

9.7

5.2

Vinyl acetate

6.7

6.7

6.7

6.7

6.7

46

46

36

23

15

230

230

180

110

75

Chlorobenzene

10

10

10

10

10

430

300

150

150

150

1100

800

400

400

400

Phenol

19

19

15

9.5

6.3

29

29

23

15

12

NR

NR

NR

NR

NR

Ethyl isocyanate

NR

NR

NR

NR

NR

0.20

0.065

0.034

0.0085

0.0040

0.60

0.20

0.10

0.025

0.013

Furan

NR

NR

NR

NR

NR

12

8.5

6.8

1.7

0.85

35

24

19

4.8

2.4

Ethyleneimine

NR

NR

NR

NR

NR

33

9.8

4.6

1.0

0.47

51

19

9.9

2.8

1.5

Hydrazine

0.1

0.1

0.1

0.1

0.1

23

16

13

3.1

1.6

64

45

35

8.9

4.4

Sulfur dioxide

0.20

0.20

0.20

0.20

0.20

0.75

0.75

0.75

0.75

0.75

30

30

30

19

9.6

Hydrogen chloride

1.8

1.8

1.8

1.8

1.8

100

43

22

11

11

620

210

100

26

26

Hydrogen fluoride

1.0

1.0

1.0

1.0

1.0

95

34

24

12

12

170

62

44

22

22

Ammonia

30

30

30

30

30

220

220

160

110

110

2700

1600

1100

550

390

Bromine

0.033

0.033

0.033

0.033

0.033

0.55

0.33

0.24

0.13

0.095

19

12

8.5

4.5

3.3

Chlorine

0.50

0.50

0.50

0.50

0.50

2.8

2.8

2.0

1.0

0.71

50

28

20

10

7.1

Hydrogen sulfide

0.75

0.60

0.51

0.36

0.33

41

32

27

20

17

76

59

50

NR

NR

Phosphine

NR

NR

NR

NR

NR

4.0

4.0

2.0

0.50

0.25

7.2

7.2

3.6

0.90

0.45

Nitrogen dioxide

0.50

0.50

0.50

0.50

0.50

20

15

12

8.2

6.7

34

25

20

14

11

Notes: NR, Not reported. Values are in parts per million (ppm).

aGreater than 10% LFL.

bGreater than 50% LFL.

Source: www.epa.gov/aegl/access-acute-exposure-guideline-levels-aegls-values, accessed September 6, 2018.

Threshold Limit Values

Certain ACGIH criteria may be appropriate for use as benchmarks. The ACGIH threshold limit values—TLV-STELs and TLV-Cs—are designed to protect workers from acute effects resulting from exposure to chemicals; such effects include irritation and narcosis. These criteria are discussed in Chapter 2. The TLV criteria can be used for toxic gas dispersion but typically produce a conservative result because they are designed for worker exposures during a normal 40-hour workweek.

Permissible Exposure Limits

The PELs are promulgated by OSHA and have the force of law. These levels are similar to the ACGIH criteria for TLV-TWAs because they are also based on 8-hour time-weighted average exposures. OSHA-cited “acceptable ceiling concentrations,” “excursion limits,” or “action levels” may be appropriate for use as benchmarks.

Toxic Endpoints

The EPA has promulgated a set of toxic endpoints to be used for air dispersion modeling for toxic gas releases as part of the EPA Risk Management Plan (RMP).7 The toxic endpoint is, in order of preference, (1) the ERPG-2 or (2) the level of concern (LOC) promulgated by the Emergency Planning and Community Right-to-Know Act. The LOC is considered “the maximum concentration of an extremely hazardous substance in air that will not cause serious irreversible health effects in the general population when exposed to the substance for relatively short duration.” Toxic endpoints are provided for 74 chemicals under the RMP rule and are shown in Table 5-7.

7U.S. Environmental Protection Agency. RMP Offsite Consequence Analysis Guidance (Washington, DC: U.S. Environmental Protection Agency, 1996).

Table 5-7 Toxic Endpoints Specified by the EPA Risk Management Plan

Chemical name

Toxic endpoint (mg/L)

Chemical name

Toxic endpoint (mg/L)

Gases

 

Bromine

0.0065

Ammonia (anhydrous)

0.14

Carbon disulfide

0.16

Arsine

0.0019

Chloroform

0.49

Boron trichloride

0.010

Chloromethyl ether

0.00025

Boron trifluoride

0.028

Chloromethyl methyl ether

0.0018

Chlorine

0.0087

Crotonaldehyde

0.029

Chlorine dioxide

0.0028

Cyclohexylamine

0.16

Cyanogen chloride

0.030

Dimethyldichlorosilane

0.026

Diborane

0.0011

1,1-Dimethylhydrazine

0.012

Ethylene oxide

0.090

Epichlorohydrin

0.076

Fluorine

0.0039

Ethylenediamine

0.49

Formaldehyde (anhydrous)

0.012

Ethyleneimine

0.018

Hydrocyanic acid

0.011

Furan

0.0012

Hydrogen chloride (anhydrous)

0.030

Hydrazine

0.011

Hydrogen fluoride (anhydrous)

0.016

Iron, pentacarbonyl

0.00044

Hydrogen selenide

0.00066

Isobutyronitrile

0.14

Hydrogen sulfide

0.042

Isopropyl chloroformate

0.10

Methyl chloride

0.82

Methacrylonitrile

0.0027

Methyl mercaptan

0.049

Methyl chloroformate

0.0019

Nitric oxide

0.031

Methyl hydrazine

0.0094

Phosgene

0.00081

Methyl isocyanate

0.0012

Phosphine

0.0035

Methyl thiocyanate

0.085

Sulfur dioxide (anhydrous)

0.0078

Methyltrichlorosilane

0.018

Sulfur tetrafluoride

0.0092

Nickel carbonyl

0.00067

Liquids

 

Nitric acid (100%)

0.026

Acrolein

0.0011

Peracetic acid

0.0045

Acrylonitrile

0.076

Perchloromethylmercaptan

0.0076

Acrylyl chloride

0.00090

Phosphorus oxychloride

0.0030

Allyl alcohol

0.036

Phosphorus trichloride

0.028

Allylamine

0.0032

Piperidine

0.022

Arsenuous trichloride

0.01

Propionitrile

0.0037

Boron trifluoride

 

Propyl chloroformate

0.010

compound with

 

Propyleneimine

0.12

methyl ether (1:1)

0.023

Propylene oxide

0.59

 

 

Toluene 2,6-diisocyanate

0.0070

Sulfur trioxide

0.010

Toluene diisocyanate

 

Tetramethyllead

0.0040

(unspecified)

0.0070

Tetranitromethane

0.0040

Trimethylchlorosilane

0.050

Titanium tetrachloride

0.020

Vinyl acetate monomer

0.26

Toluene 2,4-diisocyanate

0.0070

 

 

Source: U.S. Environmental Protection Agency. RMP Offsite Consequence Analysis Guidance (Washington, DC: U.S. Environmental Protection Agency, 1996).

In general, the most directly relevant toxicologic criteria currently available, particularly for developing emergency response plans, are ERPGs, SPEGLs, and EEGLs. These guidelines were developed specifically to apply to general populations and to account for sensitive populations and scientific uncertainty in toxicological data. For incidents involving substances for which no SPEGLs or EEGLs are available, IDLH concentrations provide alternative criteria. However, because IDLH concentrations were not developed to account for sensitive populations and because they were based on a maximum 30-minute exposure period, the EPA suggests that the identification of an effect zone should be based on exposure levels of one-tenth the IDLH concentration. For example, the IDLH concentration for chlorine dioxide is 5 ppm. Effect zones resulting from the release of this gas are defined as any zone in which the concentration of chlorine dioxide is estimated to exceed 0.5 ppm. Of course, the approach is conservative and gives unrealistic results; a more realistic approach is to use a constant-dose assumption for releases less than 30 minutes using the IDLH concentration.

The use of TLV-STELs and ceiling limits may be appropriate if the objective is to identify effect zones in which the primary concerns include more transient effects, such as sensory irritation or odor perception. In general, persons located outside the zone that is defined based on these limits can be assumed to be unaffected by the release.

Craig et al.8 have provided a hierarchy of alternative concentration guidelines in the event that ERPG data are not available; this hierarchy is shown in Table 5-8. These methods may result in some inconsistencies because the different methods are based on different concepts. Good judgment should prevail.

8D. K. Craig, J. S. Davis, R. DeVore, D. J. Hansen, A. J. Petrocchi, and T. J. Powell. “Alternative Guideline Limits for Chemicals without Environmental Response Planning Guidelines.” AIHA Journal (1995): 56.

Table 5-8 Recommended Hierarchy of Alternative Concentration Guidelines

Primary guideline

Hierarchy of alternative guidelines

Source

ERPG-1

 

AIHA

 

EEGL (30-min)

NRC

 

IDLH

NIOSH

ERPG-2

 

AIHA

 

EEGL (60 min)

NRC

 

LOC

EPA/FEMA/DOT

 

PEL-C

OSHA

 

TLV-C

ACGIH

 

5 × TLV-TWA

ACGIH

ERPG-3

 

AIHA

 

PEL-STEL

OSHA

 

TLV-STEL

ACGIH

 

3 × TLV-TWA

ACGIH

AIHA: American Industrial Hygiene Association.

NIOSH: National Institute for Occupational Safety and Health.

NRC: National Research Council Committee on Toxicology.

EPA: Environmental Protection Agency.

FEMA: Federal Emergency Management Agency.

DOT: U.S. Department of Transportation.

OSHA: U.S. Occupational Safety and Health Administration.

ACGIH: American Conference of Governmental Industrial Hygienists.

Source: D. K. Craig, J. S. Davis, R. DeVore, D. J. Hansen, A. J. Petrocchi, and T. J. Powell. “Alternative Guideline Limits for Chemicals without Environmental Response Planning Guidelines.” AIHA Journal (1995): 56.

5-6 Release Prevention and Mitigation

The purpose of a hazardous material release model is to provide a tool for performing release mitigation. Release mitigation is defined as “lessening the risk of a release incident by acting on the source (at the point of release) either (1) in a preventive way by reducing the likelihood of an event that could generate a hazardous vapor cloud or (2) in a protective way by reducing the magnitude of the release and/or the exposure of local persons or property.”9

9Richard W. Prugh and Robert W. Johnson. Guidelines for Vapor Release Mitigation (New York, NY: American Institute of Chemical Engineers, 1988).

The release mitigation procedure is part of the consequence modeling procedure shown in Figure 4-1. After selection of a release incident, a source model is used to determine either the release rate or the total quantity released. This information is coupled to a dispersion model and subsequent models for fires or explosions. Finally, an effect model is used to estimate the impact of the release, which is a measure of the consequence.

Risk is composed of both likelihood and consequence. Thus, an estimate of the likelihood of a release provides only half of the total risk assessment. A particular release incident might potentially have high consequences, leading to extensive plant mitigation efforts to reduce the consequence. However, if the likelihood is low, such a massive effort might not be required. Both the consequence and the likelihood must be included to assess risk.

Table 5-9 identifies a number of measures to prevent or mitigate a release. The example problems presented in this chapter demonstrate that even a small release can result in significant downwind impact. In addition, this impact can occur minutes after the initial release, reducing the time available for emergency response. Clearly, it is better to prevent the release in the first place. Inherent safety, engineering design, and management should be the first issues considered in any release mitigation procedure.

Table 5-9 Release Prevention and Mitigation Approaches

Major area

Examples

Inherent safety

Inventory reduction: Less chemicals inventoried or less in process vessels

Chemical substitution: Substitute a less hazardous chemical for one more hazardous

Process attenuation: Use temperatures and pressures closer to ambient

Engineering design

Plant physical integrity: Use better seals or materials of construction

Process integrity: Ensure proper operating conditions and material purity

Process design features for emergency control: Emergency relief systems

Spill containment: Dikes and spill vessels

Management

Operating policies and procedures

Training for vapor-release prevention and control

Audits and inspections

Equipment testing

Maintenance program

Management of modifications and changes to prevent new hazards

Security

Early vapor detection and warning

Detection by sensors

Detection by personnel

Countermeasures

Water sprays

Water curtains

Steam curtains

Air curtains

Deliberate ignition of explosive cloud

Dilution

Foams

Emergency response

On-site communications

Emergency shutdown equipment and procedures

Site evacuation

Safe havens

Personal protective equipment

Medical treatment

On-site emergency plans, procedures, training, and drills

Source: Richard W. Prugh and Robert W. Johnson. Guidelines for Vapor Release Mitigation (New York, NY: American Institute of Chemical Engineers, 1988), p. 2.

Suggested Reading

Atmospheric Dispersion Modeling

R. E. Britter and J. McQuaid. Workbook on the Dispersion of Dense Gases (Sheffield, UK: Health and Safety Executive Report 17/1988, 1988).

Center for Chemical Process Safety. Guidelines for Consequence Analysis of Chemical Releases (New York, NY: American Institute of Chemical Engineers, 1999).

Alex De Visscher. Air Dispersion Modeling: Foundations and Applications (Hoboken, NJ: Wiley, 2013).

Guidelines for Vapor Cloud Dispersion Models, 2nd ed. (New York, NY: American Institute of Chemical Engineers, 1996).

Steven Hanna, Seshu Dharmavaram, John Zhang, Ian Sykes, Henk Witlox, Shah Khajehnajafi, and Kay Koslan. “Comparison of Six Widely-Used Dense Gas Dispersion Models for Three Recent Chlorine Railcar Accidents.” Process Safety Progress 27, no. 3 (2008): 248–259.

International Conference and Workshop on Modeling the Consequences of Accidental Releases of Hazardous Materials (New York, NY: American Institute of Chemical Engineers, 1999).

S. Mannan, ed. Lees’ Loss Prevention in the Process Industries, 4th ed. (London, UK: Butterworth-Heinemann, 2012), ch. 15.

J. McQuaid. “Trials on Dispersion of Heavy Gas Clouds.” Plant/Operations Progress 4, no. 1 (1984): 58–61.

John H. Seinfeld. Atmospheric Chemistry and Physics of Air Pollution (New York, NY: Wiley, 1986), ch. 12, 13, and 14.

D. Bruce Turner. Workbook of Atmospheric Dispersion Estimates (Cincinnati, OH: U.S. Department of Health, Education, and Welfare, 1970).

Release Mitigation

Richard W. Prugh and Robert W. Johnson. Guidelines for Vapor Release Mitigation (New York, NY: American Institute of Chemical Engineers, 1988).

Problems

5-1.

  1. What instantaneous release on the ground, in g, of acrolein will result in a downwind concentration 1500 m immediately downwind on the ground equal to 0.05 ppm? It is a bright sunny day with a wind speed of 3.5 m/s. Assume rural conditions, a temperature of 25°C, and 1 atm pressure. Where does the maximum concentration occur on the ground for this release?

  2. If the release is now 10-m above ground level, what is the concentration on the ground immediately downwind at 1500 m? Where does the maximum concentration occur on the ground for this case?

5-2.

  1. What continuous release rate on the ground, in g/s, of acrolein will result in a downwind concentration equal to 0.05 ppm located 1500 m directly downwind on the ground? It is a bright sunny day with a 3.5 m/s wind speed. Assume rural conditions, a temperature of 25°C, and 1 atm pressure. Where does the maximum concentration occur on the ground?

  2. If the release now occurs 10 m above ground level, what is the concentration 1500 m directly downwind on the ground?

  3. Where is the maximum concentration directly downwind (in m) for part (b) and what is its value (in ppm)?

5-3. Emergency response for the rupture of a chemical pipeline is being evaluated. The chemical is estimated to discharge from the pipeline at a rate of 4.53 kg/s. The concentration of concern is 100 ppm—anyone downwind who is exposed to a concentration higher than this must be evacuated. Calculate the downwind evacuation distance. The pipeline is at ground level. The molecular weight of the released material is 17.03. The sunlight is moderate and the wind speed is 3.5 m/s. Assume 1 atm, 25°C, and rural conditions.

5-4. A backyard barbeque grill contains a 20-lb tank of propane. The propane leaves the tank through a valve and regulator and is fed through a ½-in. rubber hose to a dual-valve assembly. After the valves, the propane flows through a dual set of ejectors, where it is mixed with air. The propane–air mixture then arrives at the burner assembly, where it is burned. Describe the possible propane release incidents for this equipment.

5-5. A burning dump emits an estimated 3 g/s of oxides of nitrogen. What is the average concentration of oxides of nitrogen from this source directly downwind at a distance of 3 km on an overcast night with a wind speed of 7 m/s? Assume that this dump is a point ground-level source.

5-6. You have been appointed emergency coordinator for the community of Smallville, shown in Figure 5-12. ABC Chemical Company is shown on the map. It reports the presence of the following chemicals and amounts: 100 lb of hydrogen chloride and 100 gal of sulfuric acid. You are required to develop an emergency plan for the community.

  1. Determine which chemical presents the greater hazard to the community.

  2. Assuming all of the chemical is released during a 10-min period, determine the distance downwind that must be evacuated.

  3. Identify locations that might be affected by a release incident at the plant or that might contribute to the incident because of their proximity to the plant.

  4. Determine transportation routes that will be used to transport hazardous materials into or out of the facility. Identify any high-risk intersections where accidents might occur.

  5. Determine the vulnerable zone along the transportation routes identified in part (d). Use a distance of 0.5 mi on either side of the route, unless a smaller distance is indicated by part (b).

  6. Identify any special concerns (schools, nursing homes, shopping centers, and the like) that appear in the transportation route vulnerable zone.

  7. Determine evacuation routes for the areas surrounding the plant.

  8. Determine alternative traffic routes around the potential hazard.

  9. Determine the resources required to support the needs of parts (g) and (h).

  10. Identify the means required to warn the area, and describe the content of an example warning message that could be used in an emergency at the facility.

    A figure shows a map.
    Figure 5-12 Map of Smallville.
  11. Estimate the potential number of people evacuated during an emergency. Determine how these people are to be moved and where they might be evacuated to.

  12. What other concerns might be important during a chemical emergency?

5-7. A tank has ruptured and a pool of benzene has formed. The pool is approximately rectangular with dimensions of 20 ft by 30 ft. Estimate the evaporation rate and the distance affected downwind. Define the plume boundary using 10 ppm. It is an overcast day with a 9 mph wind. The temperature is 90°F.

5-8. A tank of chlorine contains 1000 kg of chlorine at 50-bar gauge. What is the maximum hole diameter (in mm) in this tank that will result in a downwind concentration equal to the ERPG-1 at a downwind distance of 300 m? Assume 1 atm, 25°C, a molecular weight of chlorine of 70.9, and that all the liquid chlorine vaporizes.

5-9. The emergency coordinator has decided that the appropriate emergency response to the immediate release of a toxic material is to alert people to stay in their homes, with doors and windows closed, until the cloud has passed. The coordinator has also indicated that homes 4000 m downwind must not be exposed to concentrations exceeding 0.10 mg/m3 of this material for any longer than 2 min. Estimate the maximum instantaneous release of material (in kg) allowed for these specifications. Be sure to clearly state any assumptions about weather conditions, wind speed, and other conditions.

5-10. A pipeline carrying benzene has developed a large leak. Fortunately, the leak occurred in a diked area and the liquid benzene is contained within the square 50 ft × 30 ft dike. The temperature is 80°F and the ambient pressure is 1 atm. It is a cloudy night with a 5 mph wind. All areas downwind with a concentration exceeding 4 times the PEL must be evacuated.

  1. Determine the evaporation rate from the dike (in lb/s).

  2. Determine the distance downwind (in mi) that must be evacuated.

  3. Determine the maximum width of the plume (in ft) and the distance downwind (in mi) where it occurs.

5-11. Use a spreadsheet program to determine the location of a ground isopleth for a plume. The spreadsheet should have specific cell inputs for release rate (g/s), release height (m), spatial increment (m), wind speed (m/s), molecular weight of the released material, temperature (K), pressure (atm), and isopleth concentration (ppm).

The spreadsheet output should include, at each point downwind, both y and z dispersion coefficients (m), downwind centerline concentrations (ppm), and isopleth locations (m).

The spreadsheet should also have cells providing the downwind distance, the total area of the plume, and the maximum width of the plume, all based on the isopleth values.

Your submitted work should include a brief description of your method of solution, outputs from the spreadsheet, and plots of the isopleth locations.

Use the following two cases for computations, and assume worst-case stability conditions:

Case a: Release rate: 200 g/s

Release height: 0 m

Molecular weight: 100

Temperature: 298 K

Pressure: 1 atm

Isopleth concentration: 10 ppm

Case b: Same as above, but release height is 10 m above the ground. Compare the plume width, area, and downwind distance for cases a and b. Comment on the difference between the two cases.

5-12. Develop a spreadsheet to determine the isopleths for a puff at a specified time after the release of material. The spreadsheet should contain specific cells for user input of the following quantities: time after release (s), wind speed (2 m/s), total release (kg), release height (m), molecular weight of released gas, ambient temperature (K), ambient pressure (atm), and isopleth concentration (ppm).

The spreadsheet output should include, at each point downwind, the downwind location, both y and z dispersion coefficients, downwind centerline concentration, and isopleth distance off-center (+/–).

The spreadsheet output should also include a graph of the isopleth location.

For your spreadsheet construction, we suggest that you set up the cells to move with the puff center. Otherwise, you will need a large number of cells.

Use the spreadsheet for the following case:

Release mass: 0.5 kg

Release height: 0 m

Molecular weight of gas: 30

Ambient temperature: 298 K

Ambient pressure: 1 atm

Isopleth concentration: 1 ppm

Atmospheric stability: F

Run the spreadsheet for a number of different times, and plot the maximum puff width as a function of distance downwind from the release.

Answer the following questions:

  1. At what distance downwind does the puff reach its maximum width?

  2. At what distance and time does the puff dissipate?

  3. Estimate the total area swept out by the puff from initial release to dissipation.

Your submitted work should include a description of your method of solution, a complete spreadsheet output at 2000 s after release, a plot of maximum puff width as a function of downwind distance, and the calculation of the total swept area.

5-13. A fixed mass of toxic gas has been released almost instantaneously from a process unit. You have been asked to determine the percentage of fatalities expected 2000 m downwind from the release. Prepare a spreadsheet to calculate the concentration profile around the center of the puff 2000 m downwind from the release. Use the total release quantity as a parameter. Determine the percentage of fatalities at the 2000 m downwind location as a result of the passing puff. Vary the total release quantity to result in a range of fatalities from 0 to 100%. Record the results at enough points to provide an accurate plot of the percentage of fatalities versus quantity released. The release occurs at night with calm and clear conditions.

Change the concentration exponent value to 2.00 instead of 2.75 in the probit equation, and rerun your spreadsheet for a total release amount of 5 kg. How sensitive are the results to this exponent?

Hint: Assume that the puff shape and concentration profile remain essentially fixed as the puff passes.

Supplemental information:

Molecular weight of gas: 30

Temperature: 298 K

Pressure: 1 atm

Release height: 0

Wind speed: 2 m/s

Use a probit equation for fatalities of the form

Y=17.1+1.69ln(C2.75T)

where Y is the probit variable, C is the concentration in ppm, and T is the time interval in minutes.

Your submitted work must include a single output of the spreadsheet for a total release of 5 kg, including the puff concentration profile and the percent fatalities; a plot of the concentration profile for the 5 kg case versus the distance in meters from the center of the puff; a plot of the percentage of fatalities versus total quantity released; a single output of the spreadsheet for a 5 kg release with a probit exponent of 2.00; and a complete discussion of your method and your results.

Additional homework problems are available in the Pearson Instructor Resource Center.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset