Вы находитесь на странице: 1из 15

Comparison of Empirically Based Calculation Methods for

Pipe Flares to Computational Fluid Dynamics


Matthew Martin
Callidus Technologies, LLC
Tulsa, OK USA
Email: mmartin@callidus.com

ABSTRACT
1. INTRODUCTION
As available computational resources increase and the use of Industrial flares are used to safely dispose of combustible fluids in
computational fluid dynamics (CFD) becomes more widely the case of required depressurizing for large-scale processes. The
accepted wider classes of industrial scale combustion systems are depressurizing requirement may be planned, as in the case of
being simulated and optimized using these resources. A large periodic maintenance, or may be unplanned as is the case when
body of empirically based work has already been assembled to equipment failure may force a halt to the process.
model the behavior of the large-scale turbulent diffusion flames
A large volume of fluid must be safely disposed in a relatively
used in flares. The work of Brzustowski and others is the de-facto
short period of time for typical refining, natural gas production
standard for flare flame behavior calculations. Performing the
and chemical processes. Combustion of such a large volume of
calculations outlined by Brzustowski still requires significantly
fluid over a short period of time produces a large amount of
less time than performing a simulation of the same conditions
power. The principle heat transfer mechanism for such industrial
using CFD, however the use of CFD may allow extrapolation to
scale flames to the far field is thermal radiation, and the ability to
conditions not within the data set or calculation methods currently
design systems for the proper disposal of the relief fluids relies on
in use.
the ability to predict radiation.
Others have recently used CFD to model flare systems with
Empirically based prediction techniques are the current standard
success, but to our knowledge no direct comparison has been
method for predicting thermal radiation from industrial flares.
made with previous empirical work. This validation is necessary;
Possibly the two most popular techniques are those outlined in the
the validation typically performed for commercially available
example calculations available in ‘ANSI/API Standard 521’ [1].
CFD codes are taken from well-controlled lab-scale experiments
The relatively easy access to these methods as well as their
and not industrial scale flows. Additionally, comparison between
apparent accuracy within a range required for engineering design
the results produced by CFD and those of previous published
seem to drive their continued popularity. It is noteworthy that
calculation methods will allow one to identify when the relatively
these are not the only available calculation techniques; a survey
fast empirically based methods are adequate and when a more
paper that gives many alternative techniques is available in ‘Heat
time consuming CFD study should be performed.
Radiation from Flares’ [2].
The predicted radiation and plume trajectory of a simple pipe flare
Computational fluid dynamics (CFD) has become increasingly
in a cross wind is compared to the empirical data and calculation
popular as a design and screening tool as the associated software
results from the “API method” of 2007 and the “Brzustowski and
has become easier to use and the cost of the typically substantial
Sommer’s” method for several different cross winds.
computing resources required has decreased. The primary appeal
Identification of differences in the results, and recommendations
for the use of CFD in engineering design is that the calculations
for the when it is appropriate to apply each of the calculation
are at least in a portion based on first principles and is therefore
methods are made.
thought to provide the ability to solve a more general class of
Keywords problems. One often over-looked aspect of CFD modeling is that
Flare, Pipe Flare, Pressure-Relieving and Depressurizing Systems, many of the sub-models required to provide closure to the
Computational Fluid Dynamics descriptive equations are empirically based, particularly those
related to turbulence. Thermal radiation also proves to be
particularly difficult to model; transmission is effectively
instantaneous compared to the flow speeds in question, and
although any individual thermal radiation model may be based on
fundamental principles, many of the fluid properties that
determine the incident radiation at a surface are not.
In addition to the accuracy of any given calculation technique, the
amount of time required to perform the calculation is critical to its
usefulness. CFD calculations typically take much longer to

1
perform when compared to the relatively simple empirically Distance required from a point source of irradiation is taken from
derived approaches presented. equation 24 of C.2.5 [1], and rearranged to solve for the radiation
intensity, as shown in Equation 1 below:
2. PRELIMINARIES AND REVIEW f r × Q flare
2.1 Flare and Flow Description Qr = (1)
The flare to be modeled consists of a derrick-supported stack to be 4×π × d 2
topped with a pipe flare tip of standard design. A tip of standard Where:
design is usually comprised of a large diameter pipe with some
method of flame retention intended for lower flow rates. At higher Qr =Radiation Intensity (Btu/hr-ft2)
flow rates, the flame characteristics near the tip are dominated by fr =Fraction of Heat Released as Radiation ()
the ‘turbulent diffusion’ between the large-scale jet of flare gas
and the atmosphere, reducing the impact of the flame retention Qflare =Heat Release from Flare (Btu/hr)
device on characteristics of the flame.
d =Distance from Flame Center (ft)
The derrick-supported stack and tip are to be 197 ft tall. The tip
exit velocity is designed to be approximately 400 ft/s. The gas to The fraction of heat release as radiation is taken to be 0.10, which
be flared is effectively 100% methane. The flow rate of gas to the is at odds with the nearest corresponding tip diameter value from
flare is 102.27 lb/s. The temperature of the flare gas is –148 F. Table 10 of [1]. This is done because in practice it has been
Figure 1 shows a schematic diagram of the flare. found, but is still highly debated, that the values proposed by the
GPSA handbook provide results that are closer to measured
2.2 ‘Clause 6’ Calculation Method values.
The equations and data already found in readily available sources 2.3 Brzustowski and Sommer’s Method from
are not replicated here; the results are presented in summary form.
The calculation method for plume trajectory and radiation taken ANSI/API Standard 521
from ‘Clause 6’ of ‘ANSI/API Standard 521’ is as follows [1]: The calculation method of Brzustowski and Sommer’s as outlined
1) Calculate the required flare diameter. in ‘ANSI/API Standard 521’is essentially the same as the ‘Clause
6’ calculation, but the determination of the flame distortion and
2) Calculate the anticipated flame length. the resulting flame center differs.
3) Calculate the flame distortion caused by wind. The method also commonly makes use of an atmospheric
absorption term, which is used to include the effect of the
4) Calculate the flare stack height.
participation of water vapor in the atmosphere on the radiation
For the purposes of this simulation, the flare tip diameter has intensity from the flare. This term is purposefully left out of these
already been fixed at 26 inches by using an equivalent inverse of calculations so that the ‘Clause 6’ method, the Brzustowski and
the equation of step 1 as presented in ‘Clause 6’. Sommer’s method and the CFD model can be compared on a
consistent basis.
The anticipated flame length, as calculated from a curve fit of
Figure 8 [1] of ‘Clause 6’ is 306 ft. 2.4 Computational Fluid Dynamics Method
The software package ‘Fluent’ was used to perform the
computational fluid dynamic calculations. From the outset, the
decision was made not to produce the ‘best’ model, but rather
what could be considered a typical model, with the following
requirements:
1) The computational grid must be of modest to small size.
a. The model must be able to be executed using
typically available hardware.
b. The software licensing cost cannot be prohibitive
for evaluating flares – consider the relative cost of
the pencil and paper required for the other methods
to any software package.
c. The model must be able to be solved in a useful
amount of time (overnight at most).

Figure 1 – Schematic diagram of the derrick-supported stack


and flare tip.

2
of combustion equipment, are too narrow and long when
compared to reality. Adjustment of the turbulent Schmidt number
might remedy this, but has not been fully investigated.
Radiation was modeled using both the ‘Discrete Ordinates’ (DO)
model and the ‘P-1’ model. The DO model is applicable over all
length scales where as the P-1 model is only applicable over a
large optical thickness [3]. However, the DO model is
significantly more computationally expensive and has a particular
disadvantage for modeling flares, which will be later shown.
Fluid properties were defined such that they varied with
temperature. No attempt was made to account for dissociation,
which might also negatively impact the radiation solution.

3. APPROACH
Figure 2 – Grid Dimensions 3.1 Comparison of Various Operating
2) Special considerations known only to those well versed in Conditions
the specifics of flare design cannot be included. The general For the purposes of evaluating the CFD model against the more
usefulness of the tool must be tested. popular calculation methods, each model will be calculated using
0, 20, and 40 mph crosswinds. A comparison of the results will be
3) Optimizations that can be made to the model which may be
made and a discussion of the potential inaccuracy in each model
realistic, but cannot be known or included in the other
will be included.
models should not be included.
The velocity ratio required for the graph provided in ‘ANSI/API
A computational grid of 199,700 cells was used to represent the
Standard 521’ for the Brzustowski and Sommer’s method does
flare stack, tip and surrounding atmosphere. The grid was
not allow for the case where the crosswind velocity is zero.
comprised entirely of unstructured hexahedral elements. Details of
Extrapolation is made to allow for the evaluation of the zero wind
the derrick were not included, with only a cylinder representing
velocity case.
the flare stack. The same grid was used for all cases, whether in
still air or with a crosswind. Figure 2 shows the grid dimensions. Flame geometry will be measured and used to calculate the flame
centers from the CFD results. Radiation will be measured from the
The fluid was modeled as an incompressible ideal gas. Preferably,
CFD model and compared to the other methods at the centerline
the ideal gas law would have been used because the tip velocity
of the flare stack along the downwind side at grade.
exceeds 0.3 Mach, but limitations to the software prevent the use
of a velocity inlet to specify the crosswind in conjunction with the
ideal gas law for fluid density. An alternative specification for the 4. RESULTS
inlet velocity could possibly be found, but in the spirit of 4.1 Flame Center Calculation
producing a ‘typical’ CFD model, the velocity inlet approach was
Flame center determination for the various empirical methods
used instead. For sonic flares it is advisable that an alternative
appears to be based largely of visual observation of the
approach be found.
geometrical center of the flame. The same method will be applied
Turbulence was modeled using the ‘Realizable k-epsilon’ (RKε) to the CFD model.
model. The other standard models can become unrealistic in high The geometrical center of the flame is not necessarily the center of
strain flows, and the RKε model is reported to be more accurate in radiation. The flame is most likely far more luminous in some
predicting the spreading rate of round jets [3]. regions when compared to others and the actual temperature of
Chemical reactions were modeled using a two-step Magnussen the flame will vary along the flame length. The fluid composition
model, specifically the ‘Eddy-Dissipation’ model available in surrounding the flame front changes along the length of the flame,
Fluent. The chemical reaction rate is limited by the mixing rate of due to back mixing of combustion products, which in turn alters
the fuel and its surroundings, which seems reasonable for a large- the local absorption coefficient and the resulting thermal
scale flare [3]. For other combustion equipment modeled by this irradiation.
author, the model is known to over-predict flame temperature in Although flame geometry is generally of great interest for any
critical regions, particularly because flame will attach where it CFD model involving combustion, there is no universally
should not solely because there is enough mixing. Over prediction accepted ‘flame’ criterion; there is no ‘flame’ function provided
of flame temperature would have a direct impact on the radiation by the CFD software in order to visualize the results. Much of the
results. For flares more complex than a pipe flare, this model may difficulty in flame visualization arises from the use of chemistry
be completely inaccurate because mixing may occur at locations models that do not include radical species which are often short
other than the diffusion boundary between the flare gas jet and the lived within the flame-front and so are more reliable indicators of
atmosphere. The ‘Mixture Fraction’ model may apply well to flame; one could consider the use of the electrical ‘flame-rod’
flares based on its requirements and assumptions about reaction detector common in process heater pilots to see the utility of
progress and mixing. It has been the experience of this author that radical detection. Other methods employed are the use of various
in general the flames produced from this model, for a wide range concentrations of CO (a partial product that is often available

3
from the simulation), temperature, oxygen, or mean mixture Comparison of Predicted Flame Geometry for Various Models
Horizontal Distance Vertical Distance to
fraction. The author has found that a threshold surface of reaction to Flame Center Flame Center From
rate seems to produce the most realistic results when compared to Model Wind Speed (MPH) From Stack (ft) Stack (ft)
visible flame, but there is a dearth of empirical data with which to 'Clause 6' 0 0 153
'Clause 6' 20 114 76
validate this approach. The use of stoichiometric mixture fraction
'Clause 6' 40 128 57
to represent the flame surface seems to be a common practice. B&S* 0 0 388
B&S 20 58 69
Mixture fraction is calculated from the CFD results in the B&S 40 65 45
following manner, as used in [4] (but not as a flame surface) and CFD 0 0 166
elsewhere: CFD 20 59 51
CFD 40 67 32

sY f − Y o + Y o0 *Extrapolating Curve Fit


z = 0 0 (2) Table 1 – Flame center predictions
sY f +Y o
4.2 Radiation Calculations Using WSGGM-
Y o
0 Domain Based Absorption Coefficients
z st = = 0.055 (3) Radiant intensity data was exported from the CFD model at grade
sY + Yo0
f
0
along the centerline of the flare stack. Data is only compared
along the downwind, highest radiation, side of the stack.
Where:
One should note that the ‘Incident Radiation’ or ‘Surface Incident
Radiation’ that is available as a field variable from the CFD
s=stoichiometric mass fraction, Yf=mass fraction of fuel software is not the radiation intensity that is of interest.
Manipulation of the field variables is required to obtain the proper
Yo=mass fraction of oxidizer, Y0f=initial mass fraction of fuel
value for post-processing.
Y00=initial mass fraction of oxidizer, z=mixture fraction,
For the Discrete Ordinates radiation model the number of ‘rays’
zst=stoichiometric mixture fraction
that are emanated from each cell center in the CFD model is a user
selectable parameter. The default number of rays is 8, which is
The estimated flame length from the CFD model by mixture known to be inadequate for flare radiation calculations. Increasing
fraction is 197 ft, which does not correspond well with the API the ray count to 18 results in radiation plots similar to those of
predicted flame length of 306 ft. Past experience is that other Figure 3, which is also inadequate for properly evaluating the
models have had better correspondence with the API predicted radiation map around the flare. Increasing the ray count to 64
flame length than has been achieved in this case. Possible reasons results in a plot that appears to be adequate, but this is also
for the relatively poor correspondence include the use of dependant on the grid that is used. Increasing the ray count of the
incompressible gas coupled with the relatively high exit velocity DO model increases the time to find a solution as well as the
or shortcomings in the mixing assumptions of the chemistry memory requirements, making a very high ray count unattractive.
model. The under-prediction of visible flame geometry by The participating properties of the fluids in the system have a
stoichiometric mixture fraction has been recurrent for the author direct impact on the predicted thermal radiation. The absorption
when investigating a large variety of combustion equipment. coefficient of the participating gases, primarily CO2 and H2O, is
Using an iso-surface of 2000 PPM CO (dry) results in much dependant on the mean beam length though the participating
closer agreement with the API predicted flame length. This iso- gases from the point of radiation to the receiver.
surface has proven to be a good approximation of visible flames
across a wide variety of combustion equipment, although
adjustment may be required depending on the fuel and other
variables. The CFD model predicts a 313-331 ft long flame,
depending on the other model parameters.
Table 1 shows the results for flame center location from the
various models. Both the Brzustowski and Sommer’s and the
CFD model compare favorably for the horizontal distance from
the flame center to the stack. The vertical distance of the flame
center from the top of the flare stack is closer in agreement for all
three methods. The figures used to derive the flame centers are
given at the end of this paper.

Figure 3 – Radiation map at grade (Btu/hr-ft2). A low ray


count used with the DO model provides inadequate results to
evaluate flare irradiation.

4
Radiation at Grade - Comparison of Single Point Methods and the DO Radiation Model

1000

900

800

700
Radiation Intensity (Btu/hr-ft2)

600

CFD Results - 20 MPH Wind


500

Clause 6 Method - 20 MPH Wind


400
B&S Method - 20 MPH Wind

300

200

100

0
0 200 400 600 800 1000 1200
Distance from Flare Stack (ft)

Figure 4 - Radiation intensity (Btu/hr-ft2) at grade along the downwind side of the flare using WSGGM-
Domain Based Absorption Coefficients.

The CFD software provides several methods for determining the intensity for the other cases. From the CFD software
absorption coefficient. One method is the ’Weighted Sum of Grey documentation:
Gases – Domain Based’ calculation. The CFD software will “In problems with localized sources of heat, the P-1 model may
automatically calculate a mean beam length for the geometry over-predict the [ir]radiative fluxes. The DO model is probably
under consideration, which will in turn be used to calculate the the best suited for computing radiation for this case, although the
absorption coefficient of the participating gases. DTRM, with a sufficiently large number of rays, is also
The Brzustowski and Sommer’s calculation method performs a acceptable” [3].
similar calculation by providing a correction factor ‘τ’, which is The known over-prediction of this model is the most likely cause
used to correct for atmospheric participation. It may therefore for the coincidental agreement in the zero-wind case; the flare
seem reasonable that the beam length from the atmospheric flame most likely qualifies as a local source.
domain used in the CFD model be included in the absorption
coefficient calculations, although in this case there is no water 4.3 Radiation Calculations Using Corrected
vapour in the atmosphere to participate in the radiation
calculations. However, one must consider that the size of the
Absorption Coefficients
Given the clear failure of WSGGM-Domain Based absorption
domain for the atmosphere surrounding the flare is essentially
coefficients to predict irradiative flux that is in line with the
arbitrary apart from flow considerations, and the domain
empirically derived calculations, other methods must be
dimensions directly affect the mean beam length.
employed. The CFD software provides a WSGGM-Cell Based
For closed domains such as process heaters, the calculation method, which calculates the mean beam length based on a
technique seems to work well and reasonable answers result. characteristic cell size from the CFD model. The software
However, for open domains such as are used for flares, it is less documentation warns that the solution will be grid dependant, and
clear how to define the proper volume of absorbing gases. this has been found to be true in practice [3]. The choice of cell
Figure 4 shows the results evaluated at 20 MPH. There is very based absorption coefficients has rarely given reasonable results.
poor agreement between the CFD model and the empirical results. The software user can choose alternate methods for absorption
Results for the other wind speeds are included at the end of this coefficient calculation of their own device, and this has been done
paper. in this case. A modified mean beam length is calculated and used
The P1 radiation model appears to give better results for the zero- as opposed to the default methods employed by the software.
wind case, but still significantly under-predicts the radiation Figure 5 shows the results for a 20 MPH crosswind using the
modified absorption coefficients. The agreement seems good

5
Figure 5 - Radiation intensity (Btu/hr-ft2) at grade along the downwind side of the flare using modified
absorption coefficients.

particularly in the far field. The empirical calculation methods are


generally more useful in the far field [1 – page 89], and the
5.2 Improvements to the CFD Model
Two improvements to the CFD model readily present themselves,
agreement with the CFD model is better in this range as well. The
but were purposefully not made. The first of these is to include a
results for the other wind speeds are included at the end of this
more realistic inlet velocity profile for the wind based on
document.
empirical correlations available in literature, as opposed to the
It is important to note that the modified absorption coefficient constant velocity profile used. This can be accomplished with
method used is not merely a scalar multiplier used to adjust the relatively little effort through the use of ‘User Defined Functions’
radiation results. In this case the modification increased the (UDF’s) linked into the CFD solution routine. A second
radiation levels over those predicted by default in the software. In improvement could be to include a sink term in the energy
other cases, the modification has reduced the predicted irradiative equation to reduce the ambient temperature as the elevation
output. increases. Both of these effects are regularly included in non-CFD
plume modeling.
5. ITEMS OF INTEREST Soot models are available within the CFD code; their
5.1 Modeling of Large Domains effectiveness has not been fully evaluated by this author, but
The mass contained in the flow issuing forth from a flare tip into a preliminary work suggests the effect is of second order compared
large domain that is typically of interest for radiation modeling is to the gray gas calculations.
small compared to the mass held in the atmospheric domain. This
makes the use of standard measures of convergence for CFD 6. CONCLUSIONS
models, particularly scaled residuals, very difficult. The CFD 6.1 Comparison of Methods
practitioner is encouraged to review the scaling procedure used The CFD simulation and the Brzustowski and Sommer’s
and determine its adequacy for problems of this nature. calculation method compare very favorably for determination of
Strong buoyant forces dominate the flow at distances greater than flame center. The ‘Clause 6’ method compares reasonably well for
several orifice diameters from the flare tip for low wind speed the vertical deflection of the flame center with the other two
cases. This can make convergence very difficult compared to methods, but the horizontal deflection is generally much greater
those with higher wind speeds. Careful selection of the pressure- for comparable wind speeds. The prediction of flame center in the
velocity coupling scheme and the use of very conservative under- CFD model is relatively immune to the choice of radiation model.
relaxation factors may aid convergence.

6
Using domain based absorption coefficient calculations, the the fraction of heat radiated from the ‘Clause 6’ and Brzustowski
radiation results for the CFD model did not compare well with the and Sommer’s methods was selected to be 0.10; the absorption
other calculation methods. Despite the large discrepancy in flame coefficient calculation method functions independently of any
center prediction, the ‘Clause 6’ method and the Brzustowski and ‘target’ irradiative fraction and still produced very similar results.
Sommer’s method compared very favorably. This is due to the CFD modeling is still a useful tool for evaluating more complex
distance of the flame center to the point of measurement; if the flare tip designs particularly if the sub-models, such as oxidation
stack were shorter the results would not be as similar. chemistry, are properly applied. Additionally, for situations where
nearby structures could potentially affect the air flow to the flare,
6.2 Recommendations for the Use of flow modeling may be the only potential method to determine the
Empirically Based Methods effect on the flare flame. CFD is also recommended if the thermal
The speed with which the empirically based methods can be radiation in the near-field of the flare is particularly critical. The
calculated makes them very attractive in most cases. The greatest CFD results predicted higher thermal radiation levels near the
shortcoming for most of the available methods is the recurrent use flare stack than the empirical methods while maintaining accuracy
of a ‘fraction of heat radiated’, which is experimentally in the far field. This suggests that CFD modeling may overcome
determined and varies both with flare design and firing conditions the shortcomings of the single point methods suggested in the
beyond a change in flow rate or gas composition. ‘ANSI/API Standard 521’ [1 –page 89].
The empirically based methods presented here do not address any
either difference in flare design or more complex firing issues. It 6.4 Concluding Remarks
is therefore recommended that these methods be used when there There are obvious improvements that can be made to the CFD
are no flow obstructions nearby the flare to alter the assumed model used in order to improve the accuracy, the first of which is
effect of the wind, and when the flare can be considered to be a to increase the cell count of the model. However, the intent of this
properly functioning standard pipe flare. modeling was to examine what reasonable CFD model
requirements could produce rather than assuming that all users
Use of these methods for more complex flare tip designs is not have access to a large compute cluster and unlimited software
recommended unless specific information for the fraction of heat licenses.
radiated is available from the manufacturer. The methods might
still be used even without manufacturer’s data to provide a worst- It appears that through proper selection of the CFD models and
case bound assuming a suitably high irradiative fraction is used. careful operation of the software, very reasonable agreement with
the empirical work of others can be achieved. Furthermore, the
Due to the relatively large amount of time and resources required single point methods presented in ‘ANSI/API Standard 521’ are
to perform CFD modeling, the empirically based methods are still very useful, far more economical, reasonably accurate, and
recommended for flares of simple design in simple flow substantially faster than computational fluid dynamics if the flow
environments. field around the flare is not highly complex.
6.3 Recommendations for the Use of
7. REFERENCES
Computational Fluid Dynamics [1] Pressure-relieving and Depressuring Systems. API Standard
The results from the CFD modeling presented here show that 521, Fifth Edition, January 2007
although there is a widespread appeal for the potential generality
of CFD solutions, the result is highly dependant on the modeling [2] Guigard, S., Kinzierski, W., Harper, N. Heat Radiation from
method selected. It is shown that reproducing the characteristic Flares. Publication Number T/537 prepared for Science and
flame shape through CFD may be relatively straightforward, but Technology Branch Environmental Sciences Division
that calculating a reasonable radiation map may be considerably Alberta Environment , May 2000
more difficult. [3] Documentation for Fluent 6.3. www.fluentusers.com , as of
Use of the default absorption coefficient calculation techniques September 2007
for the modeling of industrial scale flares is not adequate; the [4] Sommerer, Y., Galley, D., Poinsot, T., Ducruix, S., Lacas, F.,
CFD practitioner must modify the method to achieve reasonable Veynante, D. Large Eddy Simulation and Experimental
results. Study of Flashback and Blow-Off in a Lean Partially
The modification may seem equivalent to using an empirically Premixed Swirled Burner. Available through Elsevier
based ‘fraction radiated’ value as is used with the single-point Science, September 2004
methods. However, the CFD calculation has no ‘knowledge’ that

7
8. Figures Used To Calculate Flame Centers

0 MPH Wind - Y-coordinate plotted against a 2000 PPM CO (Dry) surface used to the flame height (ft). 197 feet is the flare tip exit
plane. Z-direction in the CFD model is the Y direction in the API calculations.

20 MPH Wind - X-coordinate plotted against a 2000 PPM CO (Dry) surface used to determine horizontal deflection of the flame
(ft). Zero is the flare centerline.

20 MPH Wind - Y-coordinate plotted against a 2000 PPM CO (Dry) surface used to determine vertical deflection of the flame (ft).
197 feet is the flare tip exit plane. Z-direction in the CFD model is the Y direction in the API calculations.

8
40 MPH Wind - X-coordinate plotted against a 2000 PPM CO (Dry) surface used to determine horizontal deflection of the flame
(ft). Zero is the flare centerline.

40 MPH Wind - Y-coordinate plotted against a 2000 PPM CO (Dry) surface used to determine vertical deflection of the flame (ft).
197 feet is the flare tip exit plane. Z-direction in the CFD model is the Y direction in the API calculations.

9
9. WSGGM Domain Based Absorption Coefficients, Discrete Ordinates Radiation Model Results
Radiation at Grade - Comparison of Single Point Methods and the DO Radiation Model

600

500
Radiation Intensity (Btu/hr-ft2)

400

CFD Results - 0 MPH Wind


300

Clause 6 Method - 0 MPH Wind

B&S Method - 0 MPH Wind


200

100

0
0 200 400 600 800 1000 1200
Distance from Flare Stack (ft)

Radiation at Grade - Comparison of Single Point Methods and the DO Radiation Model

1000

900

800

700
Radiation Intensity (Btu/hr-ft2)

600

CFD Results - 20 MPH Wind


500

Clause 6 Method - 20 MPH Wind


400
B&S Method - 20 MPH Wind

300

200

100

0
0 200 400 600 800 1000 1200
Distance from Flare Stack (ft)

10
Radiation at Grade - Comparison of Single Point Methods and the DO Radiation Model

1200

1000
Radiation Intensity (Btu/hr-ft2)

800

CFD Results - 40 MPH Wind


600

Clause 6 Method - 40 MPH Wind

B&S Method - 40 MPH Wind


400

200

0
0 200 400 600 800 1000 1200
Distance from Flare Stack (ft)

11
10. WSGGM Domain Based Absorption Coefficients, P1 Radiation Model Results
Radiation at Grade - Comparison of Single Point Methods and the P1 Radiation Model

600

500
Radiation Intensity (Btu/hr-ft2)

400

CFD Results - 0 MPH Wind


300

Clause 6 Method - 0 MPH Wind

B&S Method - 20 MPH Wind


200

100

0
0 200 400 600 800 1000 1200
Distance from Flare Stack (ft)

Radiation at Grade - Comparison of Single Point Methods and the P1 Radiation Model

1000

900

800

700
Radiation Intensity (Btu/hr-ft2)

600

CFD Results - 20 MPH Wind


500

Clause 6 Method - 20 MPH Wind


400
B&S Method - 20 MPH Wind

300

200

100

0
0 200 400 600 800 1000 1200
Distance from Flare Stack (ft)

12
Radiation at Grade - Comparison of Single Point Methods and the P1 Radiation Model

1200

1000
Radiation Intensity (Btu/hr-ft2)

800

CFD Results - 40 MPH Wind


600

Clause 6 Method - 40 MPH Wind

B&S Method - 40 MPH Wind


400

200

0
0 200 400 600 800 1000 1200
Distance from Flare Stack (ft)

13
11. WSGGM Domain Based Absorption Coefficients, Discrete Ordinates Radiation Model
Results with Modified Absorption Coefficient Calculation Technique
Radiation at Grade - Comparison of Single Point Methods and the DO Radiation Model, Modified
Calculation Method
600

500
Radiation Intensity (Btu/hr-ft2)

400

CFD Results - 0 MPH Wind


300

Clause 6 Method - 0 MPH Wind

B&S Method - 0 MPH Wind


200

100

0
0 200 400 600 800 1000 1200
Distance from Flare Stack (ft)

Radiation at Grade - Comparison of Single Point Methods and the DO Radiation Model, Modified
Calculation Method
1200

1000
Radiation Intensity (Btu/hr-ft2)

800

CFD Results - 20 MPH Wind


600

Clause 6 Method - 20 MPH Wind

B&S Method - 20 MPH Wind


400

200

0
0 200 400 600 800 1000 1200
Distance from Flare Stack (ft)

14
Radiation at Grade - Comparison of Single Point Methods and the DO Radiation Model, Modified
Calculation Method
1400

1200

1000
Radiation Intensity (Btu/hr-ft2)

800
CFD Results - 40 MPH Wind

Clause 6 Method - 40 MPH Wind


600

B&S Method - 40 MPH Wind

400

200

0
0 200 400 600 800 1000 1200
Distance from Flare Stack (ft)

15

Вам также может понравиться