Anda di halaman 1dari 5

Pakistan Institute of Engineering and Applied Sciences

Radiation Safety and Radioactive Waste Management Course


(Lecture 6)
DISPERSION MODELLING (PASQUILL CONDITIONS)
The Gaussian Plume Equation predicts a plume that is symmetrical with respect to y and with respect to x-axis.
Symmetrical does not mean that the cross-section is circular. Usually the two dispersion coefficients y and z are not
necessarily equal for all the distances x and are different in majority of the cases. Different y and z means that
spreading in vertical and horizontal direction are not equal. Most often y > z , so that a contour of constant
concentration is like an ellipse, with long axis horizontal. To use Gaussian Plume equation one must know the
appropriate values of y and z . From the previous equations 11 and 12 it is expected that they have the values like:
However, if the value of Ks in the above equation is considered, it can be seen that they are assigned an arbitrary
statistical value, independent to atmospheric behavior. It seems reasonable to assume that they would depend on wind
speed and on the degree of atmospheric turbulence, which is a function of wind speed and degree of solar heating
(insolation) etc. It is also reasonable to assume that for any given degree of insolation the value of K will be linearly
proportional to the wind speed i.e. Ky / v and Kz / v are constant. Thus according to the equations 11 and 12, which
relate statistically the dispersion coefficient with diffusion coefficient, for any specified meteorological condition, both
standard deviation y and z should increase as square root of the downwind distance x .
Experimental evidence does not agree well with this prediction. The available data have been correlated and presented
in the form of plots on log-log scale for y and z as a function of x . Experimental data show that the dispersion
coefficients increase much more rapidly with distance [Figure 11.1 and 11.2]. This means that the diffusion model for
atmospheric dispersion is not an exact description of the phenomenon. The experimental data disagree with the
diffusion theory because the equation assumed for atmospheric mixing is much too simple to account for all the
complicated things that actually go in the atmosphere, even on days with simple wind patterns, which are the only
ones on which experimental tests are attempted. Thus it can be seen that the preceding derivation shows a way to
obtain a logical material balance for dispersion of a pollutant from a point source in the atmosphere, subject to some
strong simplifying assumptions. However, the values of dispersion coefficients y and z cannot yet be computed from
theory as theory cannot accommodate meteorological features. Now if the values of dispersion coefficients given in
figure 11.1 and 11.2 may be accepted as adequate representations of experimental results, they can be used along with
the equation 14 to make prediction of concentrations downwind from a point source.
Lecture delivered by Dr. Naseem Irfan
) 12 & 11 ( .........
2
......... .......
2
Eqs
v
K x
and
v
K x
z
z
y
y

) 14 (
2
) (
2
exp
2
) (
2
exp
2
2
2
2
2
2
2
2
2
Eq
H z y H z y
v
Q
z y z y z y

1
1
]
1

,
_


+ +
1
1
]
1

,
_

+
+

This is currently the most widely used method for routine calculations of air pollutant dispersion from point sources.
The experimental data on which figures 11.1 and 11.2 are based are limited and not necessarily directly applicable to
Lecture delivered by Dr. Naseem Irfan
cities. Most of the experimental data is taken were taken for steady flow of winds over grasslands ( the Salisbury Plain
in England and the grasslands of Nebraska). It can be used for cities because nothing better is available. These plots
are based on measurements for all downwind distances less than or equal to on kilometer. The values beyond that
distance are extrapolations. However, comparison with experiments shows that advanced versions of this model
predict observed concentration fairly well. Thus It has become a standard practice, therefore, to use experimental
values of dispersion coefficients in the main dispersion equation (14) to calculate effluent concentration. The label A
to F on the lines in the figure 11.1 to 11.2 corresponds to different levels of atmospheric stability conditions which is a
direct indications that these functions depends on atmospheric conditions. Thus, z, which relates to the dispersion in
the vertical direction, can be expected to increase more rapidly with distance under unstable conditions than under
stable conditions.
Pasquill obtained a set of curves by working on experimental data, for six different atmospheric conditions. The
following relations may approximate a seventh stability condition, type G which is extremely stable:

z
(G) = 3/5
z
(F) and
y
(G) = 2/3
y
(F) -------------- Eq. (11.1 & 11.2)
As can be observed from the figures 11.1 and 11.2, the less stable conditions have higher values of both y , z than
stable conditions, at all distances from the source. It is possible to estimate the stability conditions in the lower
atmosphere by simply measuring the temperature at two or more heights on a meteorological tower. The slope of the
temperature profile can be computed by dividing the temperature difference T by the difference in height Z of the
measurements. The relationship between the Pasquill stability categories and the observed T/Z is given in the table
11.1. The United States , the Nuclear Regulatory Commission (USNRC) requires the temperature ( as well as wind
speed and direction) to be continually monitored at two points, usually at 10m and 60m, on a tower or mst near every
operating nuclear plant.
The Pasquill conditions can also be determined by monitoring the fluctuations in the angle of wind vane. On days
when the atmosphere is unstable, a wind vane tends to fluctuate more widely than on days when the atmosphere is
Lecture delivered by Dr. Naseem Irfan
stable. The correlation between the standard deviation of the angle of the vane,

, to the various Pasquill categories


is shown in table 11.1. A further elaboration for these categories classification given by Turner is given in Table
11.2. Although the instrumentation to measure directly has been installed at a number of nuclear power plants, this
method is generally considered to be less reliable- interpretations of the data is more difficult than the simple
temperature measurements just described. For instance if one wants to estimate y and z at a point 0.5 km downwind
from a pollutant source on a bright summer day with a wind speed greater than 6 m/s. The table 11.2 may be used to
conclude that for a bright summer day the incoming solar radiation is strong thus stability category will be C.
Now, using Figures 11.1 and 11.2 the values of the dispersion coefficients at a downwind distance of half a kilometer
will be 56 m and 32 m for y and z respectively. Equation 14 is correct for ground level or any elevation above it.
For large values of z , the contribution of ( z + H )
2
term becomes negligible and the result is practically identical
with the equation with no ground effect.
For conditions of y = 0 and z = 0, which corresponds to the line on the ground directly under the centerline of the
plume, the exponential term in y drops out of the equation 14. Now further multiplying both sides by v / Q gives:
The function on the right depends only on effective stack height and atmospheric stability conditions through
dispersion coefficients. Now if for any specific stability condition a downwind distance x and stability conditions are
provided then one could make up a plot of [ c v / Q ] or (concentration multiplied by a constant number) for any
effective stack height can be calculated. Turner has done this for six stability conditions one of which is shown in
figure 11.3 for a stability condition C. The figure shows the relation of effective stack height with the quantity and
location of maximum pollutant concentrations. One may conclude that as the effective stack height is increased the
quantity of maximum concentration decreases and occurs at a greater distance from the stack.
Lecture delivered by Dr. Naseem Irfan
1
]
1

,
_


z y
H
Q
v
z

2 2
1
exp
1
2
Figure 11.3 Ground level [c v / Q ] values, directly under the plume center-line, as a function of downwind
distance from the source and effective stack height H, in meters, for C stability conditions.
As the plume flows downwind, it will eventually grow until it is completely mixed below the mixing height L also
shown in the figure 11.4 As discussed in daily variation of prevalent lapse rate due to warming and cooling of earth
surface, after mid afternoon, there will exist extremely unstable conditions. This will lead to vigorous vertical mixing
from the ground up to that height and due to presence of standard lapse rate above there will be negligible vertical
mixing. The rising air columns that provide good vertical mixing induce large-scale turbulence in the atmosphere
[Figure 11.5]. This turbulence is three-dimensional, so it provides good horizontal mixing as well. Pollutants released
at ground level will be mixed almost uniformly up to mixing height, but not above it. Thus the mixing height sets the
upper limit to dispersion of atmospheric pollutants. In the morning, when inversion conditions exist quite close to the
Earth surface, the mixing height is much lower and it grows during the day. Similarly mixing heights are larger in
the summer than the winter [Table 11.2]. In order to locate these mixing height from naked eye, one may differentiate
the air clear and blue region above the mixing height and hazy and brown or gray below the mixing height. Another
feature for recognition of mixing height is the clouds themselves. The tops of clouds are not perfectly uniform, but
they are all at practically the same height, which corresponds to mixing height Figure 11.5]. Up to the mixing height
rising, unstable air brings moisture up from below to form the clouds. Above the mixing height there is no
corresponding upward flow.
Lecture delivered by Dr. Naseem Irfan

Anda mungkin juga menyukai