You are on page 1of 25

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/283381222

Jet flame length and thermal radiation: Evaluation with CFD simulations

Article · January 2010

CITATIONS READS
0 404

2 authors, including:

Arnab Chakrabarty
VerdeEn Chemicals
20 PUBLICATIONS   44 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Jet Fire Modeling using CFD View project

Development of novel approaches in battery recycling using electrochemistry techniques View project

All content following this page was uploaded by Arnab Chakrabarty on 22 May 2016.

The user has requested enhancement of the downloaded file.


GCPS 2010

Jet Flame Length and Thermal Radiation: Evaluation with


CFD Simulations

Arafat Aloqaily
Baker Engineering and Risk Consultants, Inc.
11000 Richmond Avenue, Suite # 350, Houston, Texas, 77042, Phone: (281) 822-3100
Aaloqaily@BakerRisk.com

Arnab Chakrabarty
Baker Engineering and Risk Consultants, Inc.

Keywords: Horizontal Jet Flame, CFD, Computational Fluid Dynamics, PHOENICS,


Thermal Hazard, Combustion Modeling, Flame Length, Process Safety

Abstract

Thermal hazard resulting from jet flames in chemical processes can be a significant
contributor to risk onsite. It is beneficial to understand and predict its characteristics with
reasonable accuracy. A review of various models typically used in process safety
applications showed that most models were developed based on vertical jet flames, but
are applied to horizontal jet flames. This generally results in over-predicting the flame
length and distances to thermal radiation thresholds of interest in process safety
applications, because it incorrectly accounts for buoyancy by predicting longer horizontal
distances rather than upward curvature. Predictions from a range of representative
models were compared with detailed analysis of jet flame through Computational Fluid
Dynamics (CFD) simulations and also compared with an actual test. Preliminary results
show that models tested and CFD predictions are reasonably consistent for small releases.
However, for large releases, the models tested over-predict the CFD results of flame
dimensions and thermal radiation distances. This paper discusses weaknesses and
strengths of various models compared to CFD simulations. BakerRisk has a
comprehensive test program in progress to validate and refine the CFD simulation
approach and define a more representative thermal hazard prediction model.

1. Introduction

To better understand and evaluate risk to personnel working in process facilities, thermal
hazard resulting from jet flames should be accurately estimated. While some structures
on site can be good for blast and toxic hazard, they could be vulnerable to thermal hazard,
especially if the flames impinge on the structure and block evacuation/escape routes.
Unfortunately, thermal hazard characteristics are not well understood and predicted,
especially for horizontal jet flames, which is the subject of investigation in this paper.
GCPS 2010

Most available thermal hazard models have focused on predicting the vertical flame
height in still air as a unique/significant parameter to define the flame characteristics and
impact [1-3]. Based on the flame length, thermal radiation levels and distances to thermal
thresholds are then predicted. Some of these models such as those discussed by AIChE
and NFPA are simple [1, 2] and others such as the model discussed by TNO are more
comprehensive and detailed [3]. Models that focus on predicting flame length for
specific materials have also been developed, such as the propane flame investigation
presented in reference [4].

The majority of models used for predicting the flame length have been validated using
vertical flame data, as it is easy to define and measure the visual flame length. Work has
also been done for testing flame models in small-scale setups [5]. Scaling up results of
small-scale experiments to applications dealing with large-scale flames, such as process
safety applications, might not be as accurate as needed. It is also not easy to test all
models for a wide range to cover these applications due to cost and safety concerns.

The thermal modeling approach of Mudan and Croce (1988), discussed in the CCPS
Guidelines for Chemical Process Quantitative Risk Analysis, uses flame length to
evaluate thermal hazard [1]. The formula for the flame length is given by
Eq. 1 below [1]:

[Eq. 1]

Where: LFlame is the length of the visible turbulent flame (m), dj is the release diameter
(m), CT is fuel mole fraction in stoichiometric fuel-air mixture, αT is ratio of moles of
reactants/moles of products for stoichiometric mixtures, Tf is the adiabatic flame
temperature, Tj is the jet temperature, M is the molecular weight, subscripts a and f
denote air and fuel respectively.

For a given fuel such as Methane, CT, Tf, αT and Ma/Mf are fixed values. The only
parameters that could vary in this formula are the temperature of the released material Tj
and the diameter of the jet dj. Flame length predictions using Eq. 1 depend mainly on the
release size and the fuel type, and are independent of relevant parameters that can affect
flame dimensions such as mass flow rate and source pressure. A similar formula is used
for flair sizing [1, 6]. Although the Mudan and Croce approach was originally developed
for vertical flames, it was modified for horizontal flames and flames in cross wind [1].

The approach discussed by the SFPE Handbook of Fire Protection Engineering defines
two regimes for vertical flames; buoyancy and momentum controlled flames [2]. For the
buoyancy regime, the flame height is given by:

[Eq. 2-a]
GCPS 2010

Where Hc is the heat of combustion, r is the mass stoichiometric ratio of air to fuel, g is
the acceleration due to gravity, Q is the total rate of heat released, Cp is the air specific
heat, subscript ∞ refers to ambient conditions. For example, Hc and r for methane are
around 891 kJ/mol and 17.2, respectively.

For the momentum-controlled flame, the vertical jet flame height for common gases can
be approximated by:

[Eq. 2-b]

Where ρs is the density of the source gas in the discharge stream.

Eq. 2-a is valid for jet flames with low momentum, whereas Eq. 2-b is valid for jet flames
with high momentum. A momentum parameter that represents the ratio between the gas
release momentum and momentum generated by purely buoyant flame is used to
determine which formula applied in a specific case. Detailed description of this approach
is given in the SFPE Handbook of Fire Protection Engineering [2].

A comprehensive model, the geometric solid flame shape model, or Chamberlain’s model
is also used extensively in process safety applications [3]. The model is more
sophisticated than the approaches presented by CCPS/API RP 521 and the SFPE. It has
been developed for flame shape and radiation fields of flares [3]. The flame is
represented with frustum of cone radiating energy from solid body with uniform emissive
power. The flame length is affected by the release diameter and velocity, and wind speed
and direction. A detailed description of the model with illustration examples is given in
the TNO Methods for the Calculations of Physical Effects [3].

For the purposes of discussion in this paper, the approach discussed by CCPS, SFPE and
TNO will be referred to as the CCPS model, SFPE model and the Chamberlain model,
respectively. These three approaches/models discussed above, will be used to predict the
flame length for comparison with detailed CFD analysis in this paper. Thermal radiation
predictions from the Chamberlain model will be compared to the CFD predictions as
well. Microsoft Excel has been used to calculate the flame length for the models
discussed above, with the exception of the Chamberlain model, for which calculations are
performed using BakerRisk’s SafeSite3G® code.

2. Numerical Study

This section describes the CFD numerical model and modeling approach used in the
simulations presented in this paper.

The jet flame results discussed herein were simulated using the commercial
Computational Fluid Dynamics (CFD) software PHOENICS, developed by CHAM, UK
[17]. The characteristic equations defining a jet flame requires solving turbulence and
combustion phenomena simultaneously resulting in a set of complicated Differential
GCPS 2010

Algebraic Equations. This makes it impracticable to solve for both phenomena directly,
due to the complex interaction between turbulence and combustion [9]. As a result, a set
of models are used to simplify the calculations. These models are discussed in the next
section.

2.1 Numerical Model Description

PHOENICS is a commercial CFD code that can be used to simulate flow, heat, and mass
transfer problems. It contains a range of turbulence, combustion and thermal radiation
models that can be used for simulating turbulent flames. PHOENICS implements the
Finite Volume formulation with structured grids to perform the numerical calculations. It
uses interpolation between cells and cells’ surfaces to evaluate scalar and vector
quantities at cells’ centers and surfaces, respectively. A special form of SIMPLE
algorithm is used in PHOENICS to handle the coupling between continuity and
momentum equations, namely SIMPLEST algorithm [7]. Convergence can be assessed
and controlled in different ways. The residuals of all variables and interesting spot values
can be monitored for relevant variables to assess convergence. PHOENICS mainly
implements relaxation techniques to control and promote convergence. An automatic
relaxation mechanism is available and was used in most of the calculations, but
parameters can be controlled manually as well [7].

Accordingly, turbulent and combustion models were used in combination to simplify the
calculations and solve for the mixing and reaction aspects of the flow. A radiation model
was also implemented to calculate thermal radiation energy from the jet flame presented
in this paper. Brief description of each model is given below.

2.1.1 Turbulent Model

The standard κ-ε turbulent model [8] was used to solve for turbulence. It simplifies
turbulence calculations by introducing transport equations to represent the turbulent
kinetic energy, κ, and the turbulent dissipation rate, ε. This turbulent model has proven
to have accurate predictions of flows with recirculation, boundary layer, and sheer layer
type of flows [9, 10]. It is widely used in engineering applications, as it represents an
optimum model that balances the need for accuracy with the demand for computational
resources [9]. Other models might require higher computational resources for an
incremental improvement of calculations accuracy (such as Large Eddy Simulation), or
sacrifices accuracy to improve calculations efficiency (such as the one-equation turbulent
models).

It has been shown that the standard κ−ε model has very good performance in predicting
jet flow, and turbulent jet flow with combustion [9, 11]. A modified form of the κ−ε
model (RNG κ−ε model) has also been shown to perform well in supersonic combustion
applications [12].
GCPS 2010

2.1.2 Reaction (combustion) Model

The Simple Chemically-Reacting System (SCRS) model, which implies mixing-


controlled fast reaction, was chosen to simulate the combustion reaction between the fuel
and air in this paper. This model was chosen because the combustion reaction between
the fuel used in this study (methane) and air is fast, and the combustion phenomena is
controlled primarily by the mixing rate of fuel jet and surrounding air [13]. This has
proven to be a valid assumption especially for highly volatile gaseous fuel such as the
one considered in this analysis.

The SCRS model, as implemented in PHOENICS, is a simplified model that


assumes/implies the following [7]:

1. The combustion reaction is assumed to be a single irreversible step reaction,


where fuel reacts with the oxidizer to form products without intermediates.
2. The heat of combustion released from the reaction is assumed to be independent
of temperature, implying equality and constancy of specific heat capacity of both
reactants and products.
3. Prandtl (Pr) and Schmidt (Sc) numbers are assumed to be equal for fuel, oxidant
and product. Pr is defined as the dimensionless ratio of momentum diffusivity to
thermal diffusivity, while Sc is defined as the dimensionless ratio of momentum
diffusivity to mass diffusivity as shown in the following equations:

[Eq. 3]

[Eq. 4]

Where ν is kinematic viscosity (m2/s), α is thermal diffusivity (m2/s), cp is


specific heat capacity (J/kg-K), µ is dynamic viscosity (Pa.s), k is thermal
conductivity (W/m-K), ρ is density (kg/m3) and D is mass diffusivity (m2/s)

This assumption implies that the rate of mass, momentum and heat transfer is
similar, which is a valid approximation for highly turbulent flow such as the one
considered in this paper. It has been proven that this assumption simplifies the
calculations significantly [7, 14].
4. Reacting mixtures are represented by two variables; the mass fraction of unburned
fuel and the fraction of material originating from fuel per unit mass of mixture.
However, combustion products are represented with a single mass fraction, which
usually represents the combustion product of different species. This simplifies the
calculations and reduces the amount of computational resources required to solve
the problem.
5. Density is calculated using the ideal gas law.
GCPS 2010

2.1.3 Radiation Model

The IMMERSOL radiation model, as implemented by PHOENICS, was used in


combination with the combustion model to estimate the heat radiation from the jet flame.
The model was chosen for this analysis as it can handle the uncertainty in the absorption
and scattering coefficients, and is not computationally demanding [7]. For detailed
description of the SCRS and IMMERSOL models, refer to the lectures prepared by
CHAM on this subject [7].

The combination of SCRS, IMMERSOL and standard κ−ε models provide a robust and
efficient solution methodology for simulating horizontal jet flames that combines good
accuracy with reasonable demand for computational resources. The results presented in
this paper are generated using these three models, as implemented in PHOENICS.

2.2 Geometry and Boundary Conditions

This paper focuses on the simulation of horizontal non-premixed methane jet flame,
which can result from burning of methane subsequent to its accidental release from
flammable processing facilities. The formulation of the problem is given in this section.
Figure 1 illustrates the simple geometry and the simulation domain used in these
calculations. The dimensions of the domain vary from case to case for optimum run time
and grid density. In the example shown in Figure 1, the size of the domain is 200m
(length) x 100m (height) x 50m (width). For each case, the dimensions were chosen
through an iterative process to optimize the use of computational resources without
sacrificing the accuracy of the solution. This also involves incorporating appropriate
boundary conditions to be enforced at the boundaries of the simulation domain.

Figure 2 shows an example of the structured grid used for the jet flame simulation. The
grid was designed to be very fine, close to the release source, and the cells size was
increased as a function of the distance of the cell away from the release source. In this
specific example, the grid for the calculation domain as shown in Figure 1, was
90x80x50 (360,000 cells). The grid size changes with the size of the simulation domain
as explained above.

The simulations presented in this paper were all performed for methane releases, varying
the quantity of material being released in each case. This was achieved by varying the
release size and source pressure. Release sizes were varied from small to large (0.5 inch,
2 inch and 6 inch), and pressure was assessed for a range of values from low to high (250
psig – 11,000 psig). Table 1 summarizes the parameters used for the simulation of each
case presented in this work. It also shows the size of the grid and domain used in the
corresponding calculation.

The type of boundary conditions (BC) used in the simulations was identical for all cases,
as they were independent of the simulation domain or released quantity. The problem is
represented with a release of methane from a horizontal pipeline, where fixed mass flow
rate BC was set at the exit of the pipeline. The pipeline was assumed to be at a height of
GCPS 2010

1.0 m above the ground level. A fixed velocity inlet BC was defined for surrounding air,
which was assumed to flow uniformly with low turbulent intensity of 5%. The released
fuel (methane) was assumed to be flowing in the same direction as the wind. This results
in the flame being symmetrical in shape around the plane that goes through the center of
the simulation domain. This in turn implies that it is sufficient to solve the problem using
half of the domain and assuming the other half as a mirror image (i.e. applying symmetry
boundary condition across the calculation domain as shown in Figure 1). An outlet BC
was set for the three other edges of the simulation domain (marked as sky, exit and side 1
in Figure 1). Outlet BC, as defined in PHOENICS, is a fixed-pressure boundary that
allows mass to leave or enter the solution domain as needed. The ground level is defined
as a Plate BC with adiabatic conditions, thus permitting neither heat nor material to flow
through. The logarithmic law is used for the wall function, and for the purpose of this
simulation, the ground roughness was ignored.

Sky
Height, H
100 m

Air
inlet
Side 1
Exit
Length, L
200 m

Pipeline:
Source

Ground Level

5m

Exit of Release
Source

1m Pipe used to represent


Release Source

Ground Level

Figure 1: Schematic Showing Geometry and Simulation Domain


GCPS 2010

The temperature of the surrounding atmosphere was set to 25 °C, and the fuel content in
the atmosphere was set to zero. These parameters applied to the air inlet BC and all
outlet BCs as shown in Figure 1. The conditions at the exit of the release source were
assumed to be pure fuel (methane) at 25 °C, and sensitivity analysis showed that
temperature has insignificant impact on the simulation results.

Side View

Front View

Grid near
Jet Source
Top View

Figure 2: Example Showing Structured Grid used in Simulations

2.3 Approach Used To Define Input Parameters for PHOENICS

The mass flow rates of the released methane is a function of equivalent hole size (release
size), pressure and temperature. Discharge flow rates used in this study were predicted
using the Energy Balance Model, assuming the release happens from a pipeline at
constant pressure [15]. Most of the cases presented in this paper result in supersonic
conditions at the release source. PHOENICS requires very fine mesh close to the release
source (jet origin) to solve supersonic flows. This makes the calculations impracticable,
especially for applications with combustion.

For process safety application, such as the one considered in this paper, details close to
the release source may not be critical. As a result, an expansion model is used in
combination with the discharge model to predict the jet parameters at the expansion
point. Detailed description of the expansion models used in this paper can be found in
the literature [16]. Calculations presented in this paper were performed using these
models as implemented in BakerRisk’s SafeSite3G® Code.
GCPS 2010

Table 1. Matrix of Simulated Cases with Input Parameters and Numerical Characteristics
Flow Source
Case Domain size1 (m) Grid Density
Source Parameters rate Equivalent ID Iterations
Number
kg/s m L W H L W H
1 0.5 inch 500 psig 0.7 0.06 70 10 50 82 45 79 32,600
2 0.5 inch 1000 psig 1.5 0.10 70 10 50 82 45 79 60,900
3 2 inch 250 psig 5.7 0.17 200 25 100 91 52 82 60, 800
4 0.5 inch 5000 psig 7.8 0.25 150 25 100 83 45 75 63,300
5 2 inch 500 psig 11.5 0.25 200 25 100 104 41 96 80,000
6 0.5 inch 11000 psig 14.7 0.33 125 25 100 83 45 75 67,700
7 2 inch 1000 psig 23.6 0.45 200 50 100 91 61 80 50,000
8 6 inch 250 psig 52.2 0.52 150 15 100 94 43 72 100,000
9 6 inch 500 psig 105.2 0.76 200 25 100 93 44 83 100,000
10 2 inch 5000 psig 124.9 1.00 200 25 100 102 40 72 100,000
11 6 inch 1000 psig 217.0 1.16 200 25 150 102 40 80 88,900
12 2 inch 11000 psig 236.2 1.34 300 40 200 61 35 66 100,000

1
L = length, W=width, and H=height
GCPS 2010

To account for the impact of the heat generated from combustion, the expansion diameter
predicted above is adjusted to account for the sudden/fast expansion that occurs when
combustion takes place. This is done by considering the difference in density between the
hot flame and the cold jet before combustion. The equivalent diameter, dequ, is defined as:

[Eq. 5]

Where dexp is the predicted expansion diameter, ρf is the density at flame temperature,
and ρexp is the density at the expansion conditions [9, 18].

The equivalent diameter is used to replace the expansion diameter in order to maintain a
low velocity and avoid the need to solve complex supersonic flows.

This approach ignores the distance from the release point to the expansion point, and
assumes that no entrainment of surrounding air takes place between the discharge and
expansion points. It also assumes that any liquid droplets present in the jet are small
enough to follow the gas flow or evaporate quickly due to the effect of heat generated by
combustion. This assumption is valid because methane is highly volatile.

The approach described here avoids performing supersonic jet calculations, and focuses
on the overall flame characteristics, such as temperature and thermal radiation profiles, as
well as fuel concentration distribution in the flame. In process safety applications, fine
details of flow near the release source may not be of interest compared to the overall
flame characteristics.

2.3.1 Approach Validation

The approach discussed above was used to simulate the results of a methane jet flame test
conducted by BakerRisk. Methane was released from a 2 inch pipe into an area with
matrix of pipelines designed to form a congestion zone2. The wind speed at the time of
the experiment was estimated to be 5 m/s with cross wind at an angle close to 45° relative
to the released fuel direction. The source pressure was 250 psig and its temperature was
ambient.

Figure 3 shows images of the flame taken from different angles during the test. The
height of the congestion structure surrounding the flame is 6 ft, which indicates that the
combined visual dimension of the flame (horizontal length and vertical height) is around
45 ft from the point of origin (which is 10 ft upstream of congestion structure). Discharge
and expansion conditions were measured/predicted as:

1. Discharged flow rate (measured during test): around 1330 lb/min (≈10 kg/s)
2. Release size: 2 inch (≈50 mm)

2
The test was performed for a different purpose, where congestion is required. The results are used for
validation of CFD approach because the test represents conditions close to those considered in this paper.
GCPS 2010

3. Discharge pressure: 218 psig


4. Expansion diameter: 96 mm
5. Expansion temperature: 111 K
6. Expansion velocity: 782 m/s
7. Density of Methane vapor at ambient conditions: 0.67 kg/m3

The CFD approach defined above was used to simulate this test and the flame envelope
was defined as the region that contains unburned fuel. In this analysis, a mass
concentration 0.1% was assumed low enough to define this region. Figure 4 shows the
results, in the form of 3D envelope of fuel concentration >0.1% mass fraction. Numbers
on the corner of each view/picture in Figure 4 correspond to the view/picture with the
same number in Figure 3. A comparison between the two figures shows that the
predicted flame shape, height and length agree reasonably well with the test results.
Minor differences that exist between the two cases could be attributed to uncertainties in
estimating the mass flow rate during the test, the wind speed and direction relative to the
release orientation, and approximations in modeling theories. The test was performed in
open atmosphere and wind conditions were not controlled throughout the test period.
More tests are being planned and will be used to extensively verify and refine this
approach.

1 2

3 4

Figure 3: Images Showing Methane Flame from BakerRisk 2007 Test


GCPS 2010

1 2

4
3

Figure 4: CFD Predictions of Methane Flame Test Shown in Figure 3, Showing Fuel Mass
Fraction 3D Profiles

3. CFD and Modeling Results

CFD results of jet flame simulations can be extracted in several formats. Of interest to
this work, is the temperature, thermal, and fuel distribution profiles. Figure 5 and
Figure 6 show the temperature profile and fuel mass fraction distribution for Case 2 in
Table 1 (0.5-inch release size and 1000-psig source pressure). As shown in these
example figures, the jet flame expands initially in the horizontal direction. When the jet
loses its momentum and buoyancy becomes significant, the flame lifts upward resulting
in a shorter flame length (horizontal distance covered with high temperature) compared to
what could have been predicted without buoyancy effect. This is a significant
characteristic for horizontal jet flames, indicating that a flame length based on vertical
flames can over-predict the horizontal flame length.

Figure 7 and Figure 8 show the temperature and mass fraction distribution for Case 8 in
Table 1 (6-inch release size and 250-psig source pressure). The mass flow rate and
equivalent diameter for this case are larger than Case 2, and the flame as a result is larger
than the flame in Case 2 (compare Figure 5 - Figure 8).
GCPS 2010

Figure 5: Jet Flame Temperature Distribution for Case 2: 0.5 inch and 1000 psig
Methane Flame

Figure 6: Jet Flame Mass Fraction Showing Flame Envelope for Case 2 (Table 1)
GCPS 2010

Figure 7: Jet Flame Temperature Distribution for Case 8: 6 inch and 250 psig Methane
Flame (Table 1)

Figure 8: Jet Flame Mass Fraction Showing Flame Envelope for Case 8 (Table 1)
GCPS 2010

3.1 Flame Dimensions and Thermal Radiation Results

In order to accurately describe the horizontal flame, the horizontal length and vertical
height of the visual flame (average height to flame tip) were estimated from the CFD
analysis results. Table 2 summarizes the CFD results of flame dimensions for the cases
modeled in this analysis. The results correlate well with the released mass flow rate and
equivalent diameter. As the mass flow rate or the diameter increase, the flame dimensions
(height and length) increase. This is expected since the diameter and mass flow rate
affect the velocity, which in turn affects the momentum/buoyancy forces in the jet flame
leading to different flame dimensions.

The CFD predicted flame dimensions can be correlated to the released flow rate (m°, in
kg/s) and equivalent diameter (dequ, in m) as shown in Eq. 6 and Eq. 7. These empirical
formulas were developed using the data presented in Table 2, and their empirical
parameters were evaluated using the least square method.

[Eq. 6]

[Eq. 7]

Where HFlame is the flame height (m), and LFlame is the flame length (m)

Figure 9 shows a comparison between the CFD predictions of flame dimensions (length
and height) and the predictions of Eq. 6 and Eq. 7.

Figure 9: Flame Dimensions: CFD vs. Eq. 6 & Eq. 7 Predictions


GCPS 2010

Table 2. Horizontal Jet Flame Dimensions and Distances to Thermal Radiation at 3 m above Ground

Source Distance to Thermal Rad.


Source Parameter Flow rate Flame Dimensions (m)
Case Equivalent ID (kW/m2) at height 3 m
Number Pressure
Size (in) kg/s m Height Length 35.0 12.5 4.0
(psig)
1 0.5 500 0.7 0.06 2.2 16 33
2 0.5 1,000 1.5 0.10 4.0 20 16 30 46
3 2 250 5.7 0.17 6.0 30 26 47 71
4 0.5 5,000 7.8 0.25 12.5 32 36 51 73
5 2 500 11.5 0.25 10.0 38 31 53 92
6 0.5 11,000 14.7 0.33 16.0 40 40 58 75
7 2 1,000 23.6 0.45 23.0 46 45 70 97
8 6 250 52.2 0.52 23.7 63 52 105 127
9 6 500 105.2 0.76 35.0 81 67 125 159
10 2 5,000 124.9 1.00 51.0 81 98 129 161
11 6 1,000 217.0 1.16 57.7 102 84 100 177
12 2 11,000 236.2 1.34 65.0 95 53 81 171
GCPS 2010

The correlations predict the CFD flame dimensions with good accuracy. The empirical
parameters of Eq. 6 and Eq. 7 indicate that the equivalent diameter has high impact on the
flame height, and low impact on the flame length. This is clear from comparing the third
empirical coefficients in Eq. 6 and Eq. 7 (0.68 vs. -0.04 for flame height and length,
respectively).

The CFD-predicted distance in meters to thermal radiation levels in kW/m2 (assuming a


recipient 3 m above the ground) is shown in Figure 10, as a function of released mass
flow rate in kg/s, for relevant thermal radiation thresholds. The distance to thermal
radiation thresholds clearly increases with an increase in the released mass flow rate,
which is expected because of the increased dimensions of the jet flame. The data
presented in Figure 10 was implemented to develop the following correlations for thermal
radiation distances as a function of the mass flow rate:

[Eq. 8]

[Eq. 9]

[Eq. 10]

Where D is the horizontal distance (m) to thermal radiation 3 m above ground, m° is the
released mass flow rate (kg/s), H is thermal radiation level (kW/m2), a and b are
empirical constants estimated by the least square method, using data in Figure 10.

Figure 10: Distances to Thermal Radiation Levels as a Function of Released Flow Rates
GCPS 2010

Figure 11 shows a comparison between the CFD predicted distance to thermal radiation
at 3 m above the ground, and the predictions of Equations 8 – 10. The Data scatters
around the 45° line, indicating a good correlation between the CFD and correlations
predictions for the range of parameters considered in this paper.

Figure 11: Distances to Thermal Radiation Thresholds: CFD vs.


Correlations 8-10 Predictions

3.2 Chamberlain Model, CCPS and SFPE Models

Incorrectly applying the Chamberlain model to horizontal jet flames will over-predict the
jet flame dimensions compared to the CFD predictions, as shown in Figure 12. This is
expected as the model was developed for vertical flames and not intended for horizontal
flames. The CFD predictions of the horizontal flame length and flame height are lower
than the predictions of the flame length by the Chamberlain model. However, a combined
jet flame dimension such as the dimension resulting from the addition of length and
height is very similar to the Chamberlain predictions of flame length. This indicates that
both the CFD and Chamberlain models predict consistent distances to complete the
combustion of flammable material in the fuel jet. The results shown in Figure 12 also
show that the Chamberlain model and CFD analysis have consistent and similar results
for small flame dimensions (i.e. small releases), but deviate for large releases. This is due
to the buoyancy impact on horizontal flames (i.e. flame lifting effect) being more obvious
for large flames, compared to small flames where the fuel burns close to the release
source (high momentum region).
GCPS 2010

When applied to a horizontal jet flame, the Chamberlain model does not account for
flame lifting, and over-predicts the horizontal distance of visual flame (i.e. flame length)
for large releases compared to the CFD simulations. This implies that using the
Chamberlain model will predict conservative ranges of thermal radiation zones compared
to CFD analysis, which can be double the CFD results for some cases. These results do
not undermine the accuracy/value of the Chamberlain model, but it suggests that a model
that takes into account the horizontal flame lifting due to buoyancy might be less
conservative than the Chamberlain. If conservatism in the analysis is not an issue, then
the Chamberlain model is a robust model to use for thermal hazard modeling in the range
of parameters considered in this paper.

Figure 12: Comparison Between CFD and Chamberlain Model Predictions of


Horizontal Flame Dimensions

A comparison between the predictions of distances to thermal radiation thresholds using


CFD and Chamberlain model is shown in Figure 13. Chamberlain model over-predicts
the distances to thermal radiation levels of horizontal jet flame by a factor between 25%
and 100% for different thermal radiation levels. This is expected because the
Chamberlain model over-predicts the horizontal flame length, since it was developed
based on vertical flames and does not consider the horizontal flame lifting due to
buoyancy once the jet flame loses momentum. The buoyancy impact on vertical flames
could be another factor contributing to higher predictions of the Chamberlain model.
GCPS 2010

Figure 13: Comparison Between CFD and Chamberlain Predictions of Distances to


Thermal Radiation Thresholds

A comparison between the Chamberlain model and flame length predictions of the CCPS
and SFPE models (Eq. 1 and Eq. 2, respectively) are shown in Table 3 and Figure 14.
The CCPS model predictions deviate from the Chamberlain model, while the SFPE
model predictions are consistent with the Chamberlain predictions. The performance of
the CCPS model is expected, since the formula is insensitive to the source parameters,
other than the release source. Using the equivalent diameter instead of the release source
diameter in Eq. 1 would improve its sensitivity to source parameters but is expected to
over-predict the flame length compared to the CFD, Chamberlain model and the SFPE
model predictions, especially for large releases.

Since the flame length predicted with the SFPE model is consistent with the Chamberlain
model predictions, its comparison with the CFD predictions would be similar to the
Chamberlain model case. Similar conclusions can be drawn for both Chamberlain model
and the SFPE model. Both models are consistent with the CFD results of the flame
horizontal length for small releases and over-predicts the CFD results for large releases.
GCPS 2010

Table 3. Flame Length Predictions using the CCPS and SFPE Approaches

Source Parameters SFPE Formula Calculations CCPS Formula of


Modane and Croce,
Release Discharge Flow Simplified Formula, 1988
Case T. P.
Size Rate Momentum Eq. 2 Eq. 1 Predictions
# Flame
Parameter, Flame
Regime Flame
inch °C psig kg/s lb/min Rm LFlame/Dj Length LFlame/Dj
Length (m)
(m)
1 0.5 25 500 0.7 95 0.21 1095 14 200 2.54
2 0.5 25 1000 1.5 195 0.19 1538 20 200 2.54
3 2 25 250 5.7 758 0.13 786 40 200 10.16
4 0.5 25 5000 7.8 1028 0.14 Momentum 3419 43 200 2.54
5 2 25 500 11.5 1520 0.12 1095 56 200 10.16
6 0.5 25 11000 14.7 1943 0.11 5068 64 200 2.54
7 2 25 1000 23.6 3124 0.11 1538 78 200 10.16
8 6 25 250 52.1 6890 0.09 786 87 200 30.48
9 6 25 500 105.1 13885 0.08 1095 115 200 30.48
10 2 25 5000 124.7 16483 0.08 Buoyant 3419 124 200 10.16
11 6 25 1000 216.8 28654 0.07 1538 154 200 30.48
12 2 25 11000 235.8 31166 0.06 5068 160 200 10.16
GCPS 2010

Figure 14: Flame Length (m) Predictions of CCPS and SFPE Approaches vs.
Chamberlain Model Predictions

4. Summary and Conclusions

Horizontal jet flame resulting from accidental methane release was simulated numerically
using the commercial CFD analysis software, PHOENICS. Several cases representing a
wide range of methane release conditions were simulated in this paper. The flame length
and distances to thermal radiation thresholds of interest in process safety applications
were estimated using PHOENICS and compared to three common models used in process
safety calculations. These models were the Chamberlain model, CCPS and SFPE models
(Eq. 1 and Eq. 2, respectively).

The CFD results showed that the horizontal flame lifts up when the jet loses its
momentum due to buoyancy, which results from the heat generated inside the flame. The
flame was, therefore, characterized by two dimensions: the flame horizontal length from
the release source, and the flame vertical height above the ground level. Both the flame
height and length were found to be strongly dependent on the released flow rate and
equivalent diameter. Both dimensions increase with an increase in the mass flow rate
released from the source. The flame height also increases with an increase in the
equivalent source diameter, but the flame length slightly decreases with an increase in the
equivalent diameter.
GCPS 2010

Distances to thermal radiation thresholds of interest, 35.0 kW/m2, 12.5 kW/m2 and 4.0
kW/m2, were extracted from the CFD results for the cases analyzed in this paper. The
distance depends mainly on the value of the thermal radiation threshold and the released
mass flow rate. The distance increases with an increase in the mass flow rate, and
decreases with an increase in thermal radiation threshold value. Empirical formulas were
developed to correlate the flame dimensions and thermal radiation distance to the relevant
parameters, using the least square methods. The performance of these empirical formulas
was assessed and discussed in the paper.

The flame dimensions were compared to the predictions of the Chamberlain, CCPS and
SFPE models. It was found that the Chamberlain model and SFPE model generally over-
predict the flame horizontal flame length for large releases compared to the CFD
predictions. However, the predictions of these two models are consistent with the CFD
results for small releases. The combined flame dimension (length + height) based on the
CFD results was found to be consistent with the predictions of the Chamberlain model
and the SFPE model for the range of parameters considered in this paper. This indicates
that all three approaches/models predict similar distance for complete combustion of
flammable material in the fuel jet. However, the Chamberlain model and the SFPE model
do not account for flame lifting due to buoyancy, thus over-predicting the horizontal
flame dimensions, because they were originally intended for vertical flames. The CCPS
model was found to be insensitive to release parameters, causing it to be inaccurate for
most cases considered in this paper.

The Chamberlain model was used to predict the distance to thermal radiation levels.
Since the model over-predicts the flame length, it also over-predicts the distances to
thermal radiation for large releases compared to the CFD predictions. The Chamberlain
model predictions for small releases are consistent with the CFD results with small
deviations.

The work presented in this paper shows that there is potentially significant benefit to be
gained by using a horizontal flame model that appropriately takes into account the effect
of flame lifting due to buoyancy and characterizes thermal hazards in a more accurate and
defensible manner. Traditional models incorrectly apply buoyancy, which causes their
results to err in a conservative direction. Highly conservative estimates of thermal
hazards could lead to conclude incorrectly that expensive mitigation measures should be
implemented, whereas more accurate calculations may show that these actions are not
necessary. BakerRisk initiated a long-term comprehensive plan to investigate the
development of a horizontal jet flame model for process safety applications, using both
large-scale experimental tests and numerical simulations.
GCPS 2010

5. References

1. AIChE CCPS, “Guidelines for Chemical Process Quantitative Risk Analysis”,


2nd edition, New York, NY
2. NFPA, “The SFPE Handbook of Fire Protection Engineering”, 3rd edition,
Quincy, Massachusetts
3. TNO, “Methods for the Calculation of Physical Effects – due to release of
hazardous materials (liquids and gasses)”, Yellow Book, PGS 2 (CPR 14E) – Part
2, Committee for the Prevention of Disasters, 3rd edition, 2nd revised print, 2005
4. Sugawa, O., and Sakai, K., “Propane Length and Width Produced by Ejected
Propane Gas Fuel from a Pipe”, Thirteenth Meeting of the UJNR Panel on Fire
Research and Safety, Volume 1, March 1996
5. Sandia National Laboratories – Experimental Data Archive,
http://www.sandia.gov/TNF/simplejet.html (accessed on January 07, 2009)
6. American Petroleum Institute, “API RP 521: Guide for Pressure Relieving and
Depressurizing Systems”, 2nd edition, 1996, Washington
7. PHOENICS Manual and lectures, CHAM, UK, 2009
8. Launder, B.E., and Spalding, D.B., Mathematical Models of Turbulence,
Academic Press, New York, 1982
9. Aloqaily, A., “A Study of Aerodynamics in Rotary Kilns with Two Burners”,
PhD thesis, University of Toronto, 2008
10. Launder, B.E., and Spalding, D. B., “The Numerical Computation of Turbulent
Flows”, Computer Methods in Applied Mechanics and Engineering, August 1990,
pp 269 – 289
11. Zhou, X. Sun, Z., Durst, F., and Brenner, G., “Numerical Simulation of Turbulent
Jet Flow and Combustion”, Int. J. Computer and Mathematics with Applications,
38, pp 179-191, 1999
12. Mattick, S.J., and Frankel, S.H., “Numerical Modeling of Supersonic
Combustion: Validation and Vitiation Studies Using Fluent”, AIAA, 41st
AIAA/ASME/SAE/ASEE joint propulsion conference & exhibit, July 2005,
Tucson, AR
13. Spalding, D.B., ”Mixing and Chemical Reaction in Confined Turbulent Flames”,
13th International Symposium on Combustion, pp 649-657, 1971, The
Combustion Institute
14. Spalding, D.B., ”Combustion and Mass Transfer”, Pergamon Press, 1979
15. Woodward, J.L, “Validation of two models for discharge rate”, J. of Hazardous
Materials, pp 219-229, 170 (2009)
16. Woodward, J.L, “Expansion zone modeling of two-phase and gas discharges”, J.
of Hazardous Materials, 33, pp 307-318, 1993
17. Spalding, D.B., “A General Purpose Computer Program for Multidimensional
One- and Two-Phase Flow”, Math. and Comp. in Sim., 23 , pp. 267–276 , 1981
18. Thring, M.W., and Newby, M.P., “Combustion Length of Enclosed Turbulent Jet
Flames”, 4th Symp. Int. Combust., pp. 789-796, 1953

View publication stats

You might also like