You are on page 1of 9

1698 Energy & Fuels 2008, 22, 1698–1706

Computational Fluid Dynamics for Simulation of Wind-Tunnel


Experiments on Flare Combustion Systems
David Castiñeira and Thomas F. Edgar*
Department of Chemical Engineering, UniVersity of Texas at Austin, 1 UniVersity Station C0400,
Austin, Texas 78712-0231

ReceiVed September 11, 2007. ReVised Manuscript ReceiVed January 10, 2008

Flaring is used extensively in the energy and petrochemical industries to dispose of unwanted combustion
gases by burning them in an open flame. However, these units may represent an important source of gas
emissions due to inefficient operation under certain conditions such as high crosswind velocities. Several
experimental studies have previously focused on flames burning in a fixed volume by using wind tunnels. In
these experiments, the entire plume of combustion products was collected, sampled, and analyzed to calculate
the combustion efficiency. Present work simulates these wind-tunnel experiments by using the commercial
computational fluid dynamics (CFD) software package Fluent 6.2. Several three-dimensional (3D) computational
models are developed, and suitable turbulence and chemistry models are applied to simulate the complex
combustion phenomena and flame downwash. The computational work was greatly reduced by applying the
laminar flamelet model, which assumes that a turbulent flame is an ensemble of small laminar structures called
flamelets. Inefficient combustion is observed at high crosswinds, and simulation results are in very good
agreement with experimental data. These results show that CFD can successfully simulate these wind-tunnel
flare experiments. The resulting simulation models could be used to estimate the hydrocarbon emissions from
chemical and petrochemical flares at crosswind conditions, an environmental issue of great importance in air
pollution models.

1. Introduction resulting in lower combustion efficiency because less oxygen


is entrained into smaller flames, and flare combustion efficiency
Industrial flares are combustion systems designed to safely may decrease rapidly as wind speed increases from 1 to 6 m/s.
dispose of waste gases from chemical and petrochemical plants. Experimental evidence suggests that flare combustion efficien-
Theoretically, these units convert effectively all hydrocarbons cies typically may be in the range of 70% at low wind speeds
to CO2 and H2O through high temperature oxidation reactions. and could be even lower at higher wind speeds.
When operating properly, industrial flares achieve a 98–99% Unfortunately, there is still great uncertainty about flare
combustion efficiency, so only a few percent of unburned efficiency and the resulting emissions. Published research on
hydrocarbons would be released to the atmosphere under these large-scale jet diffusion flames burning in an open atmosphere
conditions. is limited.4,5 However, collecting representative samples from
Unfortunately, there are many situations in which this high flames burning in an open atmosphere is very difficult. In
combustion efficiency may be compromised. For example, addition, online measurements on these systems is complicated
steam-assisted flares are designed to provide smokeless combus- by the size and turbulence of the flames, which are typically
tion by adding steam into the combustion zone. However, some located at the tip of stacks anywhere from 10 to over 100 m
experimental studies have shown that addition of steam into tall to prevent dangerous conditions at ground level.
the flare may affect the resulting combustion efficiency.1,2 While For all these reasons, two different approaches have been
adding turbulence and promoting beneficial chemical interac- proposed to analyze the effect of crosswind velocity on industrial
tions with the carbon particles, excessive steam addition can flare performance and the resulting emissions. One of these
decrease the temperature of the flame to the point where approaches involves the experimental study of reduced scale,
inefficiency becomes a concern. turbulent diffusion flames located within a wind-tunnel facility.
Another variable that affects flame destruction efficiency is Typically, a closed-circuit wind tunnel is used so samples can
the wind speed surrounding the flare. Flames at crosswind be easily collected and analyzed.
conditions may exhibit a stripping away and dilution of part of The second approach to understand industrial flare behavior
the fuel stream before it encounters a source of ignition to begin is to use computational fluid dynamics (CFD). CFD is based
its reaction.3 Moreover, crosswind reduces the flame size, on the application of fundamental physics for the prediction of
reacting flow phenomena, and it relies on the numerical solution
* To whom correspondence should be addressed. Telephone: (512) 471-
3080. Fax: (512) 471-7060. E-mail: edgar@che.utexas.edu. (4) Kuipers, E. W.; Jarvis, B.; Bullman, S. J.; Cook, D. K.; McHugh,
(1) McDaniel, M. Flare Efficiency Study; EPA-600/2-83-052; July 1983. D. R. Combustion efficiency of natural gas flares; effect of wind speed,
(2) Pohl, J. H.; Soelberg, N. R. EValuation of the Efficiency of Industrial flow rate, and pilots; Internal report; Shell Research and Technology Thorton
Flares: H2S Gas Mixtures and Pilot Assisted Flares; EPA-600/2-86-080; and British Gas Research Centre, 1996.
September, 1986. (5) Strosher, M. InVestigations of Flare Gas Emissions in Alberta;
(3) Johnson, M. R.; Kostiuk, L. W. Combust. Flame 2000, 123, 189– Environmental Technologies, Alberta Research Council: Calgary, Alberta,
200. Canada, 1996.

10.1021/ef700545j CCC: $40.75  2008 American Chemical Society


Published on Web 04/18/2008
CFD on Flare Combustion Systems Energy & Fuels, Vol. 22, No. 3, 2008 1699

of the governing transport equations for mass, energy, species, new experiments are characterized by larger burner diameters,
and momentum. Even though research has been done on the in some cases comparable to industrial flare diameters.
subjects of CFD and industrial combustion, there is very little
discussion of industrial flares, except by Baukal et al.,6 which 2. Governing Equations
discusses CFD applications in industrial combustion.
CFD has been recently applied to simulate the effect of steam CFD relies on solving conservation or transport equations
addition and air addition for several laboratory-scale, turbulent for mass, momentum, energy, and participating species. If the
nonpremixed flames.7 Detailed finite-rate chemistry models were flow is turbulent, model equations for specific turbulent quanti-
applied to predict the species concentrations in the flame, while ties have to be solved in addition. Since even with today’s super
species mass balances were set up in order to compute the computers resolving turbulent length scales directly results in
resulting flame combustion efficiency. Simulation results showed tremendous effort, Reynolds averaged equations are typically
that incomplete combustion of hydrocarbons may occur at high applied to include the physics of turbulence. Hence, the basic
steam/fuel and air/fuel ratios up to the point where these flames model equations for a fluid in turbulent flow are the Reynolds-
become extinguished. The computational work was greatly averaged Navier–Stokes (RANS) equations. For steady state,
reduced by assuming two-dimensional (2D) axisymmetric these equations are given below:
models. Reynolds numbers of these laboratory scale flames were 2.1. Continuity Equation.
comparable to those for industrial flares. ∇(Fν) ) 0 (1)
CFD has also been used to develop 3D simulations of the
effect of crosswind on a laboratory-scale, turbulent combustion where F is the density of the fluid and νj is its ensemble-averaged
flame.8 The flame was simulated at the exit of a vertical burner velocity vector, defined on a 3D domain.
that was perpendicular to the air flow, a configuration that is 2.2. Momentum Conservation Equation.
relevant to continuous gas flaring in the atmosphere. Simulations
clearly showed that for high momentum flames moderate ∇(Fνν) ) - ∇ p + ∇ (µ(∇ν + (∇ν)T) - Fν′ν′) (2)
velocities may significantly reduce the resulting combustion where ν′ is the turbulent fluctuation of the velocity vector, µ is
efficiency. the dynamic molecular viscosity of the fluid, and F is the
Unfortunately, direct application of CFD to simulate large- pressure. The overbars denote mean values. The Reynolds
scale industrial flares is very difficult. First of all, industrial flares stresses, Fνj′νj′, are extra terms that stem from decomposing
are clearly turbulent, and direct numerical simulation of turbulent solution turbulent variables into the mean and fluctuating
flows is not possible because of the wide range of time and components; these terms must be modeled in order to close eq
length scales. Thus, some type of turbulence model must be 2. A common approach employs the Boussinesq hypothesis to
applied. Second, realistic chemical mechanisms for hydrocarbon relate the Reynolds stresses to the mean velocity gradients:

( )
combustion cannot be described by a single reaction equation.
Such models may include tens of species and hundreds of ∂νi ∂νj 2
-ν′ν′ ) υt + - (υt(∇ν) + k)δij (3)
reactions that are known in detail for only a limited number of ∂xj ∂xi 3
fuels. Hence, some chemistry simplification must be made.
where the Einstein summation notation is being used; that is,
Furthermore, it is necessary to deal with complex turbulence-
δij is the Krönecker delta, νt is the eddy kinematic viscosity,
chemistry interaction due to the sensitivity of reaction rates to
and k is the kinetic energy of turbulence, defined by
local changes.
It is very challenging to perform combustion simulations for 1
large-scale flares. The total number of grids needed to capture k ) ν′iν′i (4)
2
all the combustion details makes the computational work almost
The Boussinesq approach is used in the k-ε model, which
prohibitive for large flares. Employing three-dimensional (3D)
is the turbulence model applied in this work. The advantage of
models significantly increases the computational work. Even
this approach is the relatively low computational cost associated
so, the lack of experimental data for industrial flares makes it
with the computation of the kinematic viscosity. For the k-ε
very difficult to validate potential simulation results. Thus, CFD
model,theeddyviscosityisobtainedfromthePrandtl-Kolmogorov
is restricted to the simulation of wind-tunnel experiments in
relation:
this work, in order to compare model predictions with experi-
mental data. This allows us to validate our results by direct Cµk2
comparison with experimental data. Validated simulation models υt ) (5)
could be used to estimate the actual hydrocarbon emissions from ε
chemical and petrochemical plants. where ε is the rate of turbulent kinetic energy dissipation.
The commercial software Fluent 6.2 is used in this work. As Robustness, economy, and reasonable accuracy for a wide range
a first approach, we have selected the wind-tunnel experiments of turbulent flows explain the popularity of this model in
of low-momentum jet diffusion flames of Johnson and Kostiuk,3 industrial flow and heat transfer simulations.
where natural gas was burned at crosswind velocities ranging 2.3. Energy Conservation Equation. When considering heat
from 1.0 to 11.0 m/s in a 0.0221 m diameter burner. In addition, transfer within the fluid and/or solid regions of the problem,
a new set of wind-tunnel experiments that have been experi- Fluent also solves the energy equation. This equation is given
mentally studied at CANMET Energy Technology Centre, below in a very general form:
Ottawa, are used for simulation of larger scale flares. These

(6) Baukal, C. E.; Gershtein, V. Y.; Li, X. Computational Fluid



∂t
(FE) + ∇ (ν
b(FE + p)) ) ∇ (keff ∇ T - ∑ h bJ + (τ
i i effV
b)) + Sh
j
Dynamics in Industrial Combustion; CRC Press LLC: Boca Raton, FL, 2000.
(7) Castiñeira, D.; Edgar, T. F. Energy Fuels 2006, 20, 1044–1056. (6)
(8) Castiñeira, D.; Edgar, T. F. CFD for simulation of crosswind on the
efficiency of high momentum jet turbulent combustion flames. J. EnViron. where keff is the effective conductivity, Ji is the diffusion flux
Eng., submitted for publication. of species i, and h is enthalpy. Notice that radiation effects are
1700 Energy & Fuels, Vol. 22, No. 3, 2008 Castiñeira and Edgar

Table 1. Cases Used for Simulation and Corresponding Jet Exit


Velocity (Vj), Crosswind Velocity (U), and Experimental Fuel
Jet to Crosswind Momentum Flux Ratio (R)
case Vj (m/s) U (m/s) R
B 2.08 1.33 1.43
C 2.10 2.76 0.34
D 2.09 4.09 0.15
E 2.11 5.49 0.085
F 2.11 8.27 0.038
G 2.09 11.05 0.021

Figure 1. Schematic of a closed-loop wind tunnel facility (all dimension supplementary fans were used to ensure that the plume of
in meters) at the University of Alberta. Reprinted by permission of
Elsevier Science from Johnson and Kostiuk,3 Copyright 2000 by The
combustion products was fully mixed into the wind-tunnel air
Combustion Institute. before sampling. The wind tunnel was sufficiently large that,
during a typical 5-10 min test, the concentration of hydrocar-
bons in the tunnel remained small, and the effects of reburning
were completely negligible.
From a simulation point of view, only the test section around
the burner (e.g., the box that followed the contraction section)
is strictly relevant. The dimensions of this box are 2.44 m in
width by 1.22 m in height by 11.8 m in length. However, a
5 m long box is enough to capture the flame behavior, so
unnecessary computation work could be avoided by reducing
the simulation domain. The floor of the wind tunnel was
constructed with 19 mm thick plywood, while downstream of
the flare the tunnel was covered with 30 gauge aluminum
sheeting to protect it from possible direct flame impingement.
The walls along the test section were primarily Plexiglas. The
ceiling upstream of the flare was constructed with 19 mm thick
plywood, but downstream of the flare the ceiling was made of
19 mm thick ceramic panels that could safely resist the
Figure 2. Sketch of flow structures in a low-momentum jet diffusion accidental impingement of the flame or hot combustion products.
flame in a crosswind.
The diffusion flames were established at the exit of a 24.6
mm o.d. (22.1 mm i.d.) pipe that extended 47 cm into the wind
not considered in this work, so the corresponding term has been
tunnel. The experimental setup also included a 65% blockage
removed from eq 6.
ratio perforated plate “turbulence plug” with 3 mm diameter
2.4. Species Transport Equations. Finally, for reacting holes, which was placed inside the pipe three diameters upstream
systems the species transport equations must be solved. In of the exit. The purpose of this plug was to create velocity
general form, this equation is given by profiles similar to the turbulent pipe flow expected in full-scale
∂ industrial flares, independent of the actual flow velocity in the
(FY ) + ∇ (FνYi) ) - ∇ Ji + Ri (7) laboratory-scale flares. However, computer simulation of that
∂t i
perforated plate is very difficult to perform and may introduce
where Ri is the net rate of production of species i by chemical numerical errors in the solution process. Thus, the turbulence
reaction. Fluent applies the finite volume method to discretize of the fuel gas was adjusted in our simulation to match
and solve the governing flow equations described above. the experimental turbulence intensity measured 5 mm above
the exit plane of the burner tube (see Kostiuk et al.9).
3. Wind Tunnel Configuration For these experiments, the jet exit velocity of the fuel, Vj,
was held approximately constant at 2 m/s, and the crosswind
A set of low-momentum, natural gas diffusion flames located
speed, U, was varied from 1 to 11 m/s. The external cold-flow
in a closed-loop wind tunnel were used for simulation. These
Reynolds number (Re) ranged from 1570 to 17 270 as the
flames were studied experimentally at the University of Alberta
crosswind increased from 1 to 11 m/s. Under these conditions,
and the National Research Council of Canada. Measurement
the flow regime on the outside of the pipe flare could be
of experimental combustion efficiencies was reported by
considered to be in the regime of having a laminar boundary
Johnson and Kostiuk.3 A detailed description of the experi-
layer separation. The turbulent fluctuation in the core flow of
mental setup and results was given as a final report by
the tunnel was found to be consistently less than 0.4% except
Kostiuk et al.9 A schematic representation of the wind tunnel
at low wind speeds (<2 m/s), where the intensity rises to about
is given in Figure 1.
1.8%.
The basic information about the experimental setup is given
below; refer to refs 3 and 9 for a more detailed description. The fuel gas used was sales grade natural gas (95.2% CH4,
The experimental flames were established at the exit of a 2.1% C2H6, 1.7% N2, 0.8% CO2, and 0.2% other, by volume).
burner tube mounted vertically in the wind tunnel and perpen- In order to guarantee ignition, the experiments used a retractable
dicular to the airflow. In the vertical section downstream of the hydrogen jet diffusion flame that was previously ignited by using
flame and in the upper section of the tunnel, a series of a manual high-voltage spark system. Once the hydrogen flame
was correctly positioned, the flow of flare gas was easily ignited.
(9) Kostiuk, L.; Johnson, M.; Thomas, G. Flare research project, final After the flare was ignited, the flow of hydrogen was turned
report, University of Alberta, September, 2004. off and the ignition system was retracted. Ignition can be
CFD on Flare Combustion Systems Energy & Fuels, Vol. 22, No. 3, 2008 1701

Figure 3. Schematic representation of the middle plane used to plot results.

Figure 4. Long-exposure photographs of the experimental flame (left side) and simulation results for the temperature contours along the middle
plane (right side). Moving downward, the crosswind velocity is increased from 1.33 m/s (case B) to 11.05 m/s (case G). Reprinted by permission
of Elsevier Science from Johnson and Kostiuk,3 Copyright 2000 by The Combustion Institute.

simulated in Fluent 6.2 by patching a high-temperature region grid. The final 3D grid had 321 040 cells and 990 713 faces,
near the burner tip, which helps to switch on the combustion and it was successfully checked for skewness. The simulations
reactions in the same way as a spark helps ignites an actual were performed by using the 3D segregated solver incorporated
flame. in Fluent. Implicit formulation was chosen; only steady state
solutions were obtained. A second order upwind scheme was
4. Simulation Results used to solve all equations. The SIMPLE algorithm10 was used
For this study, the commercial software Fluent 6.2 was used (10) Patankar, S. V. Numerical Heat Transfer and Fluid Flow; McGraw
for simulation. Gambit 2.0 was used to create the mesh and the Hill: New York, 1980.
1702 Energy & Fuels, Vol. 22, No. 3, 2008 Castiñeira and Edgar

the probability density function model (composition-PDF), while


still keeping the necessary degree of accuracy. Moreover, the
most difficult part of modeling these particular flames arises
from the turbulence flow patterns rather than very complex
chemistry reactions, so the laminar flamelet model seems very
appropriate in this case. By using a detailed chemical mecha-
nism, laminar opposed-flow diffusion flamelets were calculated
in Fluent. The chemical mechanism implemented in this work
was GRI 3.0,11 which contained 325 reactions and 53 species
for natural gas combustion simulation. The laminar flamelets
were then embedded in a turbulent flame by using statistical
PDF methods.
At very low jet-to-wind momentum flux ratios, the jet fluid
issuing from the burner tip may be severely deflected by the
transverse stream. Under these conditions, a significant “down-
wash” may occur as a portion of the combusting gases verge
on being drawn into the low-pressure region on the downwind
side of the stack.12 In fact, at high crosswind, industrial flares
are observed to present a standing vortex on the downstream
side of the burner tube. Notice that flame downwash creates an
ignition source that helps to stabilize the flame, so this
phenomenon is not necessarily associated with very low
combustion efficiencies, as discussed later in this paper. A more
detailed explanation for the fuel stripping mechanism under
these conditions is given by Johnson et al.13 A sketch of the
common flow structures at low-momentum jet diffusion flames
in a crosswind is shown in Figure 2. This figure illustrates the
three regions that are typically identified in these types of
flames.14
Simulation of flame downwash is complicated by the
aerodynamic interactions of the transverse air flow, burner tube,
and deflected fuel. Moreover, flame downwash is associated with
boundary layer formation in the burner tube.15 That is why the
k-ε turbulence model by itself is not capable of predicting this
Figure 5. Simulation results for the contours of CO2 mass fraction at
the middle plane.
phenomenon. More sophisticated models such as large eddy
simulation (LES) can solve the fine details of the flow, but LES
is computationally very expensive, even on pc clusters, and it
might be infeasible for practical applications.
Fluent 6.2 offers two approaches to simulate the boundary-
layer formation attached to the wall. In one approach, the
viscosity-affected inner region is not resolved. Instead, semiem-
pirical formulas called “wall functions“ are used to bridge the
viscosity-affected region between the wall and the fully-turbulent
region. In another approach, the turbulence models are modified
to enable the viscosity-affected region to be resolved with a
mesh all the way to the wall, including the viscous sublayer.
The latter is called “enhanced wall-treatment” in Fluent, and it
is the option applied to this work. In any case, a very fine mesh
Figure 6. Simulation of the three zones for a flame under a crosswind
(case E). near the burner tube must be defined to correctly simulate
boundary layer formation. The cells attached to the burner walls
to couple pressure and velocities. Radiation effects were not in this work were 0.5 mm long with a growth rate of 1.1, which
considered. Proper boundary conditions were specified. Burner was shown to be sufficient to capture the flame downwash
walls and wind-tunnel walls were defined as “walls” in Fluent. phenomenon. The realizable k-ε model was selected.
For the gas inlet surface and the wind inlet surface, the “velocity More information about the simulation models can be found
inlet” boundary condition was applied. For the wind outlet in ref 16.
surface, the “pressure outlet” boundary condition was used, and
this pressure was 1 atm. In order to compute the combustion (11) Smith, G. P.; Golden, D. M.; Frenklach, M.; Moriarty, N. W.;
efficiency of a particular flame, a material balance was set up Eiteneer, B.; Goldenberg, M.; Bowman, C. T.; Hanson, R. K.; Song, S.;
Gardiner, W. C.; Lissianski, V. V.; Qin, Z. http://www.me.berkeley.edu/
for the flame domain to compute the mass flux of the important gri_mech/, 2005.
species. Boundary cells at the wind outlet surface were refined (12) Huang, R. F.; Chang, J. M. Combust. Flame 1994, 98, 267–272.
to correctly calculate the integral values in that zone. (13) Johnson, M. R.; Wilson, D. J.; Kostiuk, L. W. Combust. Sci.
The laminar flamelet combustion model was applied because Technol. 2001, 169, 155–174.
(14) Gollahalli, S. R.; Brzustowski, T. A.; Sullivan, H. F. Trans. Can.
it greatly reduces the computational work of other complex Soc. Mech. Eng. 1975, 3, 205–214.
models such as the eddy dissipation concept model (EDC) or (15) Huang, R. F.; Wang, S. M. Combust. Flame 1999, 117, 59–77.
CFD on Flare Combustion Systems Energy & Fuels, Vol. 22, No. 3, 2008 1703

Figure 7. Velocity vector field in the middle plane for cases B and G.

Table 2. Computed Species Mass Flows (10-5 kg/s) and Resultant Combustion Efficiencies for the Simulated Casesa
case (CH4)in (CO2)in (C2H6)in (CH4)out (CO) out (CO2)out (C2H6)out ηc (%)
B 49.947 1.343 2.063 0.00 1.130 142.66 0.00 98.5
C 49.947 1.343 2.063 0.00 0.003 145.18 0.00 100.3
D 49.947 1.343 2.063 0.00 0.002 146.72 0.00 101.4
E 49.947 1.343 2.063 0.00 0.002 141.23 0.00 97.5
F 49.947 1.343 2.063 0.00 2.541 137.94 0.00 95.3
G 49.947 1.343 2.063 0.00 7.317 132.38 0.00 91.4
a Cases C and D predict efficiencies slightly over 100% due to numerical errors.

Table 3. Specification of the Three CANMET Wind-Tunnel


Experiments18
total mass pipe diameter pipe length ηc
case run label fuel (kg/h) (in.) (in.) (%)
1 FTF030305_CE5 15 1 24 97.54
2 FTF030408_CE2 15 4 36 99.78
3 FTF030414_CE5 10 6 36 98.71

Notice that the temperature contours can be used as a good


estimator of the photographs (e.g., high temperature can be
associated with high luminosity), although they do not neces-
sarily match completely. It is worth noting that the experimental
flames, especially at low crosswinds, are slightly more vertical
than the simulated flames. This visual discrepancy is likely due
to buoyancy effects, which were not accounted for in the
Figure 8. Comparison between experimental data and simulation results simulation work.
for combustion efficiency at increasing crosswind velocities. Experi-
mental data is obtained from Johnson and Kostiuk.3 Furthermore, these simulated temperature contours represent
only the steady state solution of the problem, so in reality,
Six different cases for natural gas combustion are simulated turbulent fluctuations would eventually affect the shape and
in this section. They are listed in Table 1, as identified by position of the flames. Finally, experimental results show that,
Johnson and Kostiuk.3 This table shows for each case the jet as the crosswind velocity is increased, downwash begins to occur
exit velocity (Vj), the crosswind velocity (U), and the experi- as a portion of the combustion gases verge on being drawn into
mental fuel jet to crosswind momentum flux ratio (R), defined the low-pressure region on the leeward side of the stack. This
as phenomenon is also observed in the simulation results.
Simulation results for contours of CO2 mass fraction are given
FjetVj in Figure 5. Clearly the amount of CO2 produced in the flame
R) (8) is affected by increasing wind speed, which in turn should affect
FairU2
Table 4. Air Conditions at the CANMET Wind-Tunnel
In this paper, 2D contour plots are given at the middle plane Experiments
of the wind tunnel, which is illustrated in Figure 3.
In Figure 4 (left side), a series of long-exposure photographs case P (kPa) T (°C) wind speed (m/s) relative humidity (%)
of the overall position and size of the flame for the experiments 1 100.7 -5.8 9.34 70
are shown. On the right side of this figure, the corresponding 2 103.1 -1.8 5.27 76
3 102.3 15.1 9.97 29
simulation results for the contour plots of temperature are shown.
1704 Energy & Fuels, Vol. 22, No. 3, 2008 Castiñeira and Edgar

Figure 9. Representation of the meshed model box for case 3, top view.

For both experimental flames and simulations, the flame


structures described in Figure 2 become apparent. In fact, three
different zones are observed in Figure 6, which corresponds to
case E in Figure 4. A planar stationary vortex attached to the
burner tube defines the first zone, the long axisymmetric tail of
the flame forms the third zone, and the junction that connects
these two main parts of the flame defines the second zone. The
size of the axisymmetric tail is reduced as the crosswind velocity
increases, while the recirculation zone grows down the tube as
more flow is drawn into the low-pressure region. Ultimately, at
very high crosswinds, the main tail of the flame would be
extinguished and only the recirculating vortex would remain.
A better understanding of the downwash phenomenon can
be obtained by analyzing the simulated velocity vectors in the
middle plane. Figure 7 displays these velocity vectors, which
indicate the magnitude and direction of the flow velocity at each
point near the burner wall for cases B and G. A recirculation
Figure 10. Contours of temperature (left side) and CO2 mass fraction
zone near the burner exit is clearly observed, so the flow turns
(right side) at the middle plane for the CANMET wind-tunnel
experiments. backward in this zone. This recirculation zone enlarges as the
crosswind velocity increases from 1 to 11 m/s. Hence, at high
the resulting combustion efficiency as shown in the next section.
crosswind velocities, a large portion of the fuel is drawn
While the combustion efficiency (ηc) can be defined in a number
downward.
of ways, in this work combustion efficiencies focus on the fully
oxidized combustion products with the goal of completely Thus, the shapes and positions of the flames observed in these
oxidizing all of the fuel to CO2 (Bourguignon et al.17). Equation simulations give a qualitative estimation of the flame behavior
9 shows how (ηc) is calculated: under a crosswind. However, it is important to compute the
resulting combustion efficiencies and to compare them to the
mass flow rate of carbon in CO2 produced by flame experimental measurement. Table 2 shows the computed mass
ηc )
mass flow rate of carbon in CxHy in the fuel gas stream flows by Fluent at inlet and outlet boundaries for several species
(9) involved in the chemistry of the flame. The last column
computes the combustion efficiency based on eq 9.
(16) Castiñeira, D. Three-dimensional computational fluid dynamics A graphical comparison between the experimental and
modeling of combustion flares. Ph.D. Thesis, Department of Chemical simulated combustion efficiencies is shown in Figure 8.
Engineering at the University of Texas, Austin, TX, 2006.
(17) Bourguignon, E.; Johnson, M. R.; Kostiuk, L. W. Combust. Flame As observed in Figure 8, the experimental data and simulation
1999, 119, 319–334. results are in very good agreement, considering the complexity
CFD on Flare Combustion Systems Energy & Fuels, Vol. 22, No. 3, 2008 1705

Figure 11. Velocity vectors and recirculating region at the middle plane for the CANMET wind-tunnel simulations.

Table 5. Computed Species Mass Flows (10-5 kg/s) and Resultant Combustion Efficiency for the CANMET Wind-Tunnel Simulations
case (CH4)in (CO2)in (C2H6)in (CH4)out (CO)out (CO2)out (C2H6)out ηc experimental (%) ηc simulation (%)
1 380.229 6.782 15.696 0.002 19.183 1075.988 0.000 97.54 97.56
2 380.229 6.782 15.696 0.000 0.000 1095.620 0.000 99.78 99.35
3 253.471 4.522 10.464 0.000 0.000 726.718 0.000 98.71 98.85

of these 3D simulations. Initially, the efficiency is slightly higher N2, 0.62% CO2, and 0.02% C4H10. Table 3 describes some
because more air is entrained into the flame. However, as the physical parameters of the three flames studied in this work,
crosswind velocity grows larger than 5.49 m/s (case E), the including the case number, the run label, the total mass of fuel
combustion efficiency decreases and degrades significantly with sent to the burner, the nominal pipe diameter (Schedule 40),
further increases in the wind speed (cases F and G). Both and the pipe length. The last column lists the experimental
experiments and simulations confirm these results. Notice that combustion efficiencies measured by the authors of the
simulated combustion efficiencies exhibit the same trend as the experiment.
experimental data, with a maximum efficiency located around The variable air conditions in the wind tunnel are given in
4 m/s crosswind velocity. Mismatches between simulations and Table 4, including pressure, temperature, wind speed, and
experiments are on the order of 1-2%, so these CFD models relative humidity.
successfully simulate the wind-tunnel experiments. As observed in Table 3, the pipe dimensions for cases 2 and
A flame bent over with the windward side of the flame 3 are relatively large (the pipe diameter for case 3 is comparable
detached from the stack or a flame base trapped in the to some industrial flare diameters), which makes their computer
recirculating flow in the wake of the stack is not sufficient to simulation more challenging than that of the previous cases.
cause significant inefficiencies, which was also pointed out by The simulation procedure employed was essentially the same
Johnson and Kostiuk.3 In fact, this recirculating zone seems to as that discussed. However, larger simulation domains have to
help stabilize the flame so the resulting combustion efficiency be defined for cases 2 and 3 due to their larger scale. Hence,
is higher than expected at high crosswinds. For high momentum the length of the simulation box was set as 5.0, 6.0, and 7.0 m
flames where flame downwash does not occur, the combustion for cases 1, 2, and 3, respectively. This required a larger number
efficiency may be more sensitive to high crosswinds, as shown of cells in the domain. Moreover, the grid concentration around
by Castiñeira and Edgar.8 the burner was increased for cases 2 and 3 in order to capture
in detail the physics of these flames. A schematic representation
5. Larger-Scale Wind-Tunnel Simulation of the grid model from a top view is presented in Figure 9 for
In this section, we apply CFD for the simulation of a set of case 3.
wind-tunnel experiments that have been experimentally studied The final number of grid cells for cases 1, 2, and 3 were
at the CANMET Energy Technology Centre, Ottawa. These 255 520, 566 720, and 988 960 cells, respectively.
experiments are part of a more extensive study of flare The applied models, parameters, and procedures applied in
combustion systems under crosswind conditions in closed-loop Fluent were very similar to previous simulations. Only boundary
wind tunnels, but only three particular cases are considered here conditions differ significantly in order to fit the new physical
for simulation. conditions for fuel and air. The simulation time increased
The experimental setup was similar to the wind-tunnel significantly due to the large number of cells, especially for
experiments of Johnson and Kostiuk,3 so only relevant differ- case 3. Convergence was achieved after progressively increasing
ences will be given here. First of all, the flare test facility in the under-relaxation factors for density from 0.7 to 1.0.
this case was not closed loop, and the methodology for Figure 10 shows the contour plots for temperature (left side)
measuring efficiency was different. Second, the basic dimensions and CO2 (right side) at the middle plane of the simulation box.
for the box section (i.e., the section after the contraction of Unfortunately, photographs are not available for these experi-
Figure 1) were 8.230 m in length by 1.219 m in width by 1.82 m mental flames. A significant downwash can be observed for
in height. Natural gas was used as a fuel, and its composition
was (by volume) 95.33% CH4, 2.1% C2H6, 0.13% C3H8, 1.8% (18) Gogolek, P. Personal communication. 2005.
1706 Energy & Fuels, Vol. 22, No. 3, 2008 Castiñeira and Edgar

cases 2 and 3, which is likely due to the high crosswind velocity We also simulated larger diameter flares under a crosswind
and the larger burner surface where the flame is attached. that were part of a different set of wind-tunnel experiments. In
The velocity vectors for the three corresponding simulations this case, the computational work increased due to the larger
are displayed in Figure 11. As expected, these vectors reflect simulation domain and larger number of cells. Agreement
how the flow moves downward and becomes trapped in the between the experimental results and the simulations was very
recirculation zone. For case 3, most of the flow exiting the good in all cases.
burner moves to this zone. Computational fluid dynamics is an excellent tool to simulate
Table 5 shows the computed mass flows by Fluent at inlet flares burning in wind-tunnel facilities. Use of simulation in
and outlet boundaries for several species involved in the the study of flare units yields economic savings compared to
chemistry of the flame. The last two columns show the expensive wind-tunnel experiments. Moreover, CFD allows for
experimental combustion efficiencies (ηc) based on eq 9 and the possibility of detailed analysis of species concentration
the simulated combustion efficiencies (ηc) obtained with Fluent. profiles and turbulent flow patterns within the flame, which may
High combustion efficiencies are observed in both experi- not be available experimentally.
ments and simulations in spite of the high crosswinds. This
means that most of the fuel is burned within the recirculating Acknowledgment. The authors thank the Texas Air Research
flow in the wake of the stack. Clearly, the agreement between Center for its support of this research and Dr. P. Gogolek of the
the experimental results and the simulation data is excellent, CANMET Energy Technology Centre, Ottawa, for providing the
and this agreement is even better than in previous wind-tunnel selected data from the Flare Test Facility. They also want to thank
cases. The reason for this may be that turbulence boundary Dr. David Schowalter (Fluent) for his help and recommendations
conditions were known for these CANMET wind-tunnel experi- in implementing the software.
ments. In fact, for the wind-tunnel simulations studied in section
4, it was necessary to add artificial turbulence because of the Nomenclature
perforated plate at the experimental setup, which may introduce A ) area projections over the faces of the zone
some errors in the computations. D ) burner diameter
h ) enthalpy
6. Conclusions bJi ) diffusion flux of species i
k ) kinetic energy of turbulence
CFD analysis has been successfully applied to simulate a set Keff ) effective conductivity
of closed-loop wind-tunnel experiments in order to understand LK ) Kolmogorov length scale
industrial flare behavior and potential gas emissions from these LR ) reaction length scale
units. The laminar flamelet model was shown to be accurate mi ) mass flow rate of species i through a boundary
enough to capture the most important details of the flame. A p ) pressure
thin boundary layer was defined to simulate flame downwash, R ) momentum flux ratio
a phenomenon associated with low-momentum flares under high Ri ) net rate of production of species i by chemical reaction
crosswind. Vj ) jet exit velocity
U) crosswind velocity
Results show that high crosswind velocities affect the
νj ) ensemble-averaged velocity vector
resulting combustion efficiency (in terms of total combustion ν′ ) turbulent fluctuation of the velocity vector
to CO2). However, for low momentum flares (e.g., low gas/ x ) axial position
crosswind ratios) such as these ones, a significant amount of xi ) mass fraction of species i
fuel may be trapped in the recirculating zone near the burner χ ) species mass fraction
wall, which helps stabilize the flame and avoids fuel stripping δij ) Krönecker delta
away from the burner exit without burning. This makes the ε ) rate of turbulent kinetic energy dissipation
resulting combustion efficiencies higher than expected. For high- ηc ) combustion efficiency
momentum flares (e.g., high gas/crosswind ratios), the flame is µ ) dynamic molecular viscosity of the fluid
unlikely to attach to the stack, so lower combustion efficiencies F ) density of the fluid
νt ) eddy kinematic viscosity
are observed even at high crosswinds.8 Both experimental results
and simulations confirm this. EF700545J

You might also like