Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

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

net/publication/269714692

Computational fluid dynamics modeling of laboratory flames and an


industrial flare

Article · November 2014


DOI: 10.1080/10962247.2014.948229 · Source: PubMed

CITATIONS READS

15 3,029

6 authors, including:

Kanwar Devesh Singh Daniel Chen


Oil & Gas Lamar University
18 PUBLICATIONS 93 CITATIONS 40 PUBLICATIONS 857 CITATIONS

SEE PROFILE SEE PROFILE

Helen H. Lou
Lamar University
115 PUBLICATIONS 2,177 CITATIONS

SEE PROFILE

All content following this page was uploaded by Kanwar Devesh Singh on 03 June 2019.

The user has requested enhancement of the downloaded file.


Journal of the Air & Waste Management Association

ISSN: 1096-2247 (Print) 2162-2906 (Online) Journal homepage: https://www.tandfonline.com/loi/uawm20

Computational fluid dynamics modeling of


laboratory flames and an industrial flare

Kanwar Devesh Singh, Preeti Gangadharan, Daniel H. Chen, Helen H. Lou,


Xianchang Li & Peyton Richmond

To cite this article: Kanwar Devesh Singh, Preeti Gangadharan, Daniel H. Chen, Helen H. Lou,
Xianchang Li & Peyton Richmond (2014) Computational fluid dynamics modeling of laboratory
flames and an industrial flare, Journal of the Air & Waste Management Association, 64:11,
1328-1340, DOI: 10.1080/10962247.2014.948229

To link to this article: https://doi.org/10.1080/10962247.2014.948229

Published online: 20 Oct 2014.

Submit your article to this journal

Article views: 1139

View Crossmark data

Citing articles: 1 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=uawm20
TECHNICAL PAPER

Computational fluid dynamics modeling of laboratory flames and an


industrial flare
Kanwar Devesh Singh,1 Preeti Gangadharan,1 Daniel H. Chen,1,⁄ Helen H. Lou,1
Xianchang
1
Li,2 and Peyton Richmond1
Dan F. Smith Department of Chemical Engineering, Lamar University, Beaumont, TX, USA
2
Department of Mechanical Engineering, Lamar University, Beaumont, TX, USA
⁄Please address correspondence to: Daniel H. Chen, Dan F. Smith Department of Chemical Engineering, Lamar University, Beaumont, TX 77705,
USA; e-mail: daniel.chen@lamar.edu

A computational fluid dynamics (CFD) methodology for simulating the combustion process has been validated with experimental
results. Three different types of experimental setups were used to validate the CFD model. These setups include an industrial-scale
flare setups and two lab-scale flames. The CFD study also involved three different fuels: C3H6/CH4/Air/N2, C2H4/O2/Ar, and CH4/
Air. In the first setup, flare efficiency data from the Texas Commission on Environmental Quality (TCEQ) 2010 field tests were used
to validate the CFD model. In the second setup, a McKenna burner with flat flames was simulated. Temperature and mass fractions
of important species were compared with the experimental data. Finally, results of an experimental study done at Sandia National
Laboratories to generate a lifted jet flame were used for the purpose of validation. The reduced 50 species mechanism, LU 1.1, the
realizable k-e turbulence model, and the EDC turbulence–chemistry interaction model were used for this work. Flare efficiency, axial
profiles of temperature, and mass fractions of various intermediate species obtained in the simulation were compared with
experimental data and a good agreement between the profiles was clearly observed. In particular, the simulation match with the
TCEQ 2010 flare tests has been significantly improved (within 5% of the data) compared to the results reported by Singh et al. in
2012. Validation of the speciated flat flame data supports the view that flares can be a primary source of formaldehyde emission.

Implications: Validated computational fluid dynamics (CFD) models can be a useful tool to predict destruction and removal
efficiency (DRE) and combustion efficiency (CE) under steam/air assist conditions in the face of many other flare operating variables
such as fuel composition, exit jet velocity, and crosswind. Augmented with rigorous combustion chemistry, CFD is also a powerful
tool to predict flare emissions such as formaldehyde. In fact, this study implicates flares emissions as a primary source of
formaldehyde emissions. The rigorous CFD simulations, together with available controlled flare test data, can be fitted into
simple response surface models for quick engineering use.

Introduction revision (TCEQ, 2009). Flare efficiencies depend on many factors.


Specifically, both destruction and removal efficiency (DREs) and
The most recent Texas Air Quality Studies (TxAQS 2000 and combustion efficiency (CE) depend on the heating value of the vent
TxAQS II) revealed that air quality models (Comprehensive Air gas, vent gas species, exit velocity, steam injection, assisted air,
Quality Model with extensions [CAMx], Community Multi-Scale fuel–air mixing, and wind speed (Pohl et al., 1984; Pohl and
Air Quality [CMAQ]) often significantly underpredict observed Soelberg, 1985; TCEQ, 2000, 2007; U.S. Environmental
peak O3 with the current volatile organic compounds (VOC) emis- Protection Agency [EPA], 2009, 2012; Singh et al., 2013). As a
sion inventories in the Houston–Galveston–Brazoria area (HGB). It result, DRE and CE can drop below the 98% threshold under
is generally believed that either unidentified VOC emission inven- certain high air/steam-assisted conditions even when the flare
tory sources exist or identified sources are significantly under- operation is in compliance with 40 CFR 60.18 (U.S. Government,
reported. One potential under-reported or non-reported emission 2009) as indicated in recent TCEQ and EPA reports (Allen and
source is due to flare operations (Texas Commission on Torres, 2011; EPA, 2012). A photochemical model used by Al-
Environmental Quality [TCEQ], 2007; Environ, 2006; ZEECO, Fadhli et al. (2012) predicted an increase in ozone concentrations
2007). The more common elevated flares are also subject to cross- by more than 15 ppb at low DREs. Another issue is that the
wind effects, which can decrease the combustion efficiency by required flare turndown ratio (more than 15,000 to 1) to accom-
reducing the residence time of the combustible materials. The modate various operating modes makes it difficult to maintain high
TCEQ’s ongoing evaluation of flare operations, therefore, is an efficiency at all conditions (Baukal and Schwartz, 2001; Allen and
important element of a future State Implementation Plan (SIP) Torres, 2011; Torres et al, 2012).

1328
Journal of the Air & Waste Management Association, 64(11):1328–1340, 2014. Copyright © 2014 A&WMA. ISSN: 1096-2247 print
DOI: 10.1080/10962247.2014.948229 Submitted January 26, 2014; final version submitted July 3, 2014; accepted July 9, 2014.
Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/uawm.
Singh et al. / Journal of the Air & Waste Management Association 64 (2014) 1328–1340 1329

Computational fluid dynamics (CFD) simulations of the com- three experiments are discussed in the next section. One of the
bustion process have received considerable attention due to their three setups was the Sandia flame in a vitiated coflow. Similar
potential to closely emulate real-world combustion scenarios. validation tests of Sandia Flames using various CFD models are
Such studies include, but are not limited to, the simulation of reviewed here. The PDF flamelet model using mixture fraction
internal combustion engines, industrial flares and burners, fire theory and (CH4–Air) GRI reaction mechanism 2.11 were used
and explosion, industrial furnaces, and others. The focus of this to simulate the Sandia flame (Nik et al., 2010). Tyliszczak (2013)
work is to model industrial flares and laboratory flames of used an unsteady-state LES model along with the PDF
similar fuel species and to compare the combustion efficiency turbulence-chemistry model coupled with GRI 2.11 reaction
and speciated emissions with experimental data. In recent years, mechanism to simulate the Sandia flame. The diameter of the
work on the simulation of open air combustion practices has lab-scale burner used for Sandia F flame was 18.9 mm. In
been done using different models. For example, Smith (2011) Vujanovic (2009), a reduced reaction mechanism was coupled
and Jatale et al. (2012) recently validated their large eddy simu- with the steady-state laminar flamelet combustion model. In all
lation (LES) model by comparing the CFD simulation results of these studies, the accuracy of the modeling was limited by the
with experimentally measured flare efficiencies of a 4-inch dia- PDF/flamelet model, which cannot predict slower reactions. In
meter laboratory flare at the CANMET wind tunnel flare facility. the present work, a rigorous combustion model, the eddy dis-
Castiñeira and Edgar (2008) used the probability density func- sipation concept (EDC) model, was coupled with the 50-species
tion (PDF) model to simulate small scale flares in a wind tunnel LU 1.1 reaction mechanism.
setup as well. The simulation results were validated with tem- To evaluate the reaction mechanism’s capability of modeling
perature and concentration profiles of major species. However, industrial scale flares, test cases from the TCEQ Flare Study project
due to a limited number of species (16) considered in the reaction were modeled. Industrial flares with diameters ranging from 24
mechanism, some intermediate species such as OH, HCHO, or inches to 36 inches were used during these tests. To our knowledge,
NO were not reported in Castiñeira’s work. On the other extreme, the previous results reported by Singh et al. in 2012 are the only
rigorous reaction mechanisms used for combustion modeling validation of CFD simulations with controlled industrial-scale
include as many as 700 reactions and more than 100 species flares. Prior simulation studies (e.g., Castiñeira 2008) mainly mod-
(Wang et al., 2007). Even though the computational power to eled laboratory scale flares with flare diameters of up to 6 inches.
simulate such reactive flows has increased drastically in recent Further, the TCEQ Flare Study tests modeled in this work and
years, such models are still very computationally expensive. earlier by Singh et al. (2012) were conducted at very low exit
Hence, reduced mechanisms are necessary to make numerical velocities and low heating values. These conditions represent flare
modeling practical for turbulent combustion applications. operations in a stand-by mode and have importance in representing
Lamar University’s research team has combined the GRI-3.0 regular flare activities in handling fugitive emissions, venting, and
mechanism (Smith et al., 2000) (optimized for methane) and the pressure relief operations.
USC-I mechanism (Davis et al., 1999) (optimized for ethylene In addition to the setups mentioned already, the McKenna
but without the NOx species) to obtain a mechanism containing burner, a laminar flat flame burner, was also modeled. Such
93 species and 600 reactions. This mechanism was reduced to laminar flame burners are often used for accurately measuring
create a series of 50-species mechanisms for the combustion of the concentration of various VOC species. Using the measured
C1–C3 light hydrocarbons for use in Fluent-Chemkin. concentration profiles, the reaction mechanism’s ability to pre-
Mechanisms LU 1.0 and LU 1.1 were developed based on dict the emission of HCHO during combustion process was
analysis of rate constant, maximum mass fraction, and number validated. Both formaldehyde and acetaldehyde, important radi-
of reactions involved in the reaction pathway. In LU 1.1, CN was cal producing photochemical species (Seinfeld and Pandis,
replaced by NO2 to model the 2010 John Zink flare test (Allen 2006), were also measured during the TCEQ’s 2010 Tulsa
and Torres, 2011). LU1.1 (Lou et al., 2011; Singh et al., 2012, Flare tests (e.g., Test Number S4.1, Run1).
2014) used in this study was validated with key laboratory test
data, that is, laminar flame speed (Davis and Law, 1998), adia-
batic flame temperature (Law et al., 2005), and ignition delay
Numerical Simulation
(Mayers and Bartle, 1969). The objective of this work was to All the already-mentioned case studies were simulated using
further validate LU 1.1 and the CFD models against three sets of the commercial CFD package ANSYS FLUENT 13.0. To reduce
experimental data involving the combustion of methane, ethy- the computational time, Fluent was run using parallel computing
lene, propylene, and a combination of the three species. LU1.1 settings; that is, each case was run on 8 or 12 local parallel
was validated by comparing the predicted concentration profiles processors.
of some important species with lab-scale flame experimental
data. For any detailed combustion reaction mechanism, the abil- Governing equations
ity to correctly predict concentrations of intermediate species is
important. To evaluate the accuracy of LU1.1, mass fractions of Computational fluid dynamics mainly involves solving sets of
species like NO, HCHO, and OH were measured along the transport equations using numerical methods like the Green–Gauss
height of the flame. These predicted concentration profiles or the least square method. The governing transport equations are
were then plotted and compared against the experimental data. solved for mass, momentum (turbulence), energy and chemical
Due to the different scale and setup of the three experiments, the species. Direct numerical simulation (DNS) for industrial flares is
model was validated for both laminar and turbulent flames. The computationally not feasible because of the time dependent
1330 Singh et al. / Journal of the Air & Waste Management Association 64 (2014) 1328–1340

governing equations. Large eddy simulation (LES) modeling in @


ðE Þ þ :ð~
vðE þ pÞÞ
which large eddies are explicitly computed in a time dependent @t !
simulation using the “filtered” Navier–Stokes equations can be X   (7)
applied for industrial flares, but it is computationally expensive. ¼ : keff T  hi~
J i þ teff :v þ Sh
Therefore, in this study, the more popular Reynolds-averaged j
Navier–Stokes (RANS) equations, which govern the transport of
the averaged flow quantities with the whole range of the scales of where keff is effective conductivity, ~ Ji is the diffusion flux of
turbulence being modeled, are used to model turbulence. The species i, hi is the enthalpy of species i, and teff is the effective
RANS model is widely used for its reduced computational time stress tensor.
and wide range of practical applications. The last and the most important governing transport equation
The basic governing equations to be solved for RANS com- to be solved for industrial flare modeling is the species transport
bustion modeling are the mass, momentum (turbulence), energy, equation,
and chemical species.
The continuity equation for the RANS model is given in eq 1: @
ðYi Þ þ :ðvYi Þ ¼ : Ji þ Ri þ Si (8)
@t
:ð:vÞ ¼ 0 (1)
where Yi is the local mass fraction of each species and J i is the
where r is the density of the fluid and v is the ensemble-averaged turbulent mass diffusion flux. Ri is the net rate of production of

velocity vector and is the sum of the mean (v Þ and fluctuating species i by chemical reaction and Si is the rate of creation by
velocities (v 0 ), shown in eq 2, addition from the dispersed phase and any user defined sources.
This equation is computed when the user-defined function of the
v ¼ v þ v0 (2) reduced mechanism is introduced in the simulation in this study.
The finite-volume method is used to discretize and solve the
The ensemble-averaged momentum equation is written as governing equations.
   
  2 Computational domain
: vi vj ¼ p þ :  vi þ vj  dij vk  v v0 0 (3)
3
All the simulations were performed on three-dimensional (3-
where p is the local pressure term, m is the viscosity of the fluid, D) computational domains. These 3-D domains were created and
dij is the Kronecker delta function, and v0 v 0 is the Reynolds meshed in GAMBIT 2.4.6. Both structured and unstructured
stress term. cells were used to discretize the domain. The mesh near the
A common approach, the Boussinesq hypothesis (eq 4), is fuel jet/burner/flare tip was kept very fine as compared to the
used to relate the Reynolds stresses to mean velocity gradients. rest of the domain. This helped in reducing the number of cells
This hypothesis is used for low computational cost associated and hence the computational time. To select the optimum num-
with the solving of the turbulent viscosity, mt. It is used in k-є ber of cells, grid independence studies were performed for all
turbulent modeling, which is followed in this work. In this case, three models. Grid independence studies help to avoid the use of
two additional transport equations, the turbulence kinetic energy an excessive number of cells while preserving the accuracy of the
k and the turbulence dissipation rate є, are solved and mt is final solution.
computed as a function of k and є.
In many cases, the models based on the Boussinesq hypoth- Boundary conditions
esis perform well:
The boundary conditions specified in the CFD model are
    discussed here. Fuel inlet, pilot inlet, and crosswind were speci-
@vi @vj 2 @vk
 vi vj ¼ t þ  k þ t dij (4) fied as “velocity inlets.” Each velocity-inlet surface was speci-
@xj @xi 3 @xk
fied by mass fractions, temperature, and a velocity magnitude.
The turbulent kinetic energy k is defined as The flow direction was kept normal to the surface. Turbulence of
the velocity inlet surfaces were specified by turbulence intensity
1 and hydraulic diameter. The bottom surface of all the domains,
k ¼ v0 v0 (5) flare stack, and burner surfaces were set as a non-slip wall. For
2
accurate prediction of turbulent flow near the flare stack bound-
For the k-є model, the eddy viscosity is calculated by the ary surface, the enhanced wall treatment function was used. All
Prandtl–Kolamagorov relationship as follows: the surfaces from which the flows exit the domain were set as the
pressure outlet. These pressure outlet surfaces were specified by
C k 2 a gauge pressure value of zero.
t ¼ (6)
"
CFD models
The other governing equation to be solved for is the energy
equation (eq 7). The general form of the energy equation is A pressure-based solver with double precision was used for
presented as modeling. The turbulence was modeled using the realizable k-e
Singh et al. / Journal of the Air & Waste Management Association 64 (2014) 1328–1340 1331

Table 1. Variables used in the model two is then used as the reaction rate. For this reason, the eddy
dissipation concept is an ideal model for simulating turbulence
Elemental mass balance error combustion. It should be noted that all the cases run using the
(C, H, O and N) <5% EDC approach were completed in two stages. Initially “cold
flow” was simulated, meaning that combustion chemistry was
High-temperature patch (initial ignition) 1800 K
disabled during this period. Once a converged cold flow was
Turbulence intensity (fuel) 5%
obtained, the region near the flare stack was patched with a
Turbulence intensity (crosswind) 10%
temperature of 2000 K. The EDC chemistry modeling was
Turbulence model k-epsilon realizable
then enabled and the combustion of the fuel started, which
further raised the plume temperature.

model. Radiation effects were neglected to reduce the computa- Radiation model. Due to the high computational cost, radia-
tional costs. For the pressure–velocity coupling, the widely used tion was ignored in the simulation. Radiation is generally used
SIMPLE (Patankar, 1980) algorithm was enabled. Discretization for modeling industrial furnaces, gas fired heaters, and other
of gradients for constructing values of scalars at the cell faces equipment where combustion is used for heat transfer and acts as
were computed using the Green–Gauss cell-based method, and source of heat energy. Most applications that utilize radiation
the pressure staggering option (PRESTO) was used for pressure models involve combustion in an enclosed domain. In open-air
discretization. For all the other equations, the first-order upwind combustion systems such as industrial flares and lab-scale bur-
scheme was used initially. However, for better accuracy, this was ners, the heat generated by the flare is of no use and is dissipated
later changed to the second-order upwind scheme. Similarly, the into the atmosphere. Using a radiation model can definitely
underrelaxation factors were initially set at 0.5 (except for pres- improve the simulation results but requires a considerable
sure, turbulent viscosity, and body forces, which were kept the amount of additional computational time.
same as default), and were then gradually brought to their default
values with convergence. Other variables defined in the model
Fluent postprocessing
are given in Table 1. Many specific results were obtained during the postproces-
sing step of the simulation to facilitate the comparison between
Turbulence model. For modeling turbulence in the domain, the simulations and experimental data. For the TCEQ flare setup
the realizable k-e model was used. The realizable k-e model involving a full-scale flare, the predicted and experimental flare
remains the most widely used model to simulate practical appli- efficiencies were compared. The modeled efficiencies were cal-
cations. It is derived from the standard k-e turbulence model. culated using the integral flow rates of various species over the
Modified equations for calculating the turbulent viscosity and inlet and outlet surfaces of the domain. For the lab-scale flames,
dissipation rate are introduced and added to the standard model. the predicted axial profiles of concentration and temperature
The realizable k-e model is suitable for a wider range of applica- were compared to the experimental ones. A “profile line” at
tions, gives a more stable solution, and is computationally inex- the center of the geometry was used in both cases. The profile
pensive at the same time. However, to closely emulate the open- line provided the measurements of temperature and mole/mass
air flaring process, the LES model should not be overlooked. It is fractions of various species along the height of the flame.
a transient model that can simulate large eddies in a turbulent
flow. However, due to its very high computational cost, it is not
commonly used for modeling complex combustion processes
Data Sources
involving hundreds of reactions. It normally takes hundreds of TCEQ flare: Industrial-scale flare
parallel processors and weeks of computational time to simulate
few seconds of a combustion process. For this reason, the realiz- The model to be validated was developed to simulate turbu-
able k-e model was preferred over the LES model in this work. lent industrial flares under various operating conditions. Hence,
the model was first checked against industrial-scale flare data
Combustion model. The turbulence–chemistry interaction obtained from the TCEQ 2010 Flare Study (Allen and Torres,
model used in this work was the eddy dissipation concept 2011) at the John Zink facility in Tulsa, OK. A similar study
(EDC) model. Due to their very low computational cost, non- (McDaniel, 1983; Pohl, 1984) was done under the auspices of
premixed models like probability density function (PDF) flame- the EPA in 1983–1984 and measured the effect of lower heating
let models are the most commonly used combustion models. value (LHV), exit velocity, and other parameters on the flare
However, their fundamental assumption of infinitely fast chem- performance. The 2010 TCEQ study was aimed to study the
istry renders them inapplicable for modeling low-Btu, low-exit- flare’s performance under low jet velocity, low Btu conditions
velocity flaring. On the other hand, the EDC model is one of the (stand-by mode) using the state-of-the-art measurement techni-
most rigorous chemistry–turbulence interaction models. The que. A 1.05-m-diameter flare was used to combust propylene
EDC model assumes that the reactions in the flame occur in gas, along with variable flows of Tulsa Natural Gas (TNG;
small turbulent structures called fine scales. To generate these Baukal and Schwartz, 2001) and nitrogen (N2). Operating para-
fine scales in a turbulent reacting flow, the model uses detailed meters like air-to-fuel ratio, combustion-zone heating value, and
reaction mechanisms. Unlike other combustion models, both the vent gas flow rate were varied during the flare tests. The mea-
kinetic rates and mixing rates are calculated. The slower of the surement of species concentration was done using redundant
1332 Singh et al. / Journal of the Air & Waste Management Association 64 (2014) 1328–1340

measurement systems. For example, an extractive sampling The flare was located at 5 m from the left side and at the center of
method with a moveable sample collector was used to collect the width. The CFD domain for the simulation of the TCEQ flare
plume samples during the tests to determine the flare efficien- is shown in Figure 1. This configuration provided enough time
cies. Remote-sensing technologies like passive and active for the combustion process to be completed inside the domain.
Fourier-transform infrared (PFTIR/AFTIR) spectroscopy and The height of the flare stack was kept as 10 m and the diameter as
GasFindIR passive infrared cameras were also used to analyze 1.05 m. To avoid complex geometry, and hence an unnecessarily
the flare plume. Therefore, the test data are believed to be more large number of cells, the fuel jet opening was modified. The jet
accurate than previous studies. The data points to compare used for previous study by Singh et al. (2011) had separate fuel
simulation results were taken from Appendix E of the TCEQ and pilot gas openings. In this work, the fuel and the pilot gas
2010 Flare Study Final Report. were provided from a common outlet at the center. The air assist
was injected from the outer ring of the flare stack opening and
Flare efficiencies. Two types of flare efficiencies were mon- the crosswind direction was from left to right.
itored and reported during the flare study: DRE and CE. The largest velocity and species concentration gradients were
located in regions near the stack. To increase the accuracy of the
(1) DRE (destruction and removal efficiency). DRE represents
reacting flow profile, the mesh density near the stack was kept
the percent of the fuel destroyed relative to the amount of
higher relative to other zones in the domain. The final grid of
fuel actually sent to the flare. Using C3H6 as an example, it
840,000 cells was successfully checked for skewness. Skewness
can be written as
is a commonly used parameter to check the quality of the grid.
C3 H6fed  C3 H6plume The skewness of the grid used for this work was kept under 0.4.
DREðC3 H6 Þ ¼ (9) Initially, several grids were prepared for the model. The grid with
C3 H6fed minimum computational time and with minimal loss of accuracy
(2) CE (combustion efficiency). CE, on the other hand, takes was selected.
into consideration the percentage of fuel successfully con-
verted into carbon dioxide, the final oxidation product. McKenna flame: Laminar premixed combustion
It is defined as
Laminar flat flames are one of the flames commonly used to
CO2plume study combustion chemistry. Flat flame burners like the
CE ¼ (10) McKenna burner (Holthuis & Associates, 2013) are often used
CO2plume þ COplume þ HCsplume

HCs plume in the preceding equation defines the amount of


hydrocarbons present in the plume. Using the data provided in
the TCEQ flare project final report, the flares were modeled in
the CFD Software, ANSYS Fluent (Fluent Inc., 2011). A total of
10 cases from Appendix E of the TCEQ 2010 Flare Study Final
Report were simulated. The details of cases modeled are pro-
vided in Table 2 (Allen and Torres, 2010).

Flare geometry. To keep the geometry simple, a rectangular


domain was built in GAMBIT. The length, width, and height of
the geometry were kept as 30 m, 10 m, and 30 m, respectively. Figure 1. Geometry used for the full-scale flare validation.

Table 2. Conditions used for TCEQ flare test cases

TCEQ case Air-assist flow rate, Wind speed,


number Propylene, lb/hr TNG, lb/hr Nitrogen, lb/hr LHV, Btu/scf lb/hr mph
A1.1 919 0 0 2,108 149,173 12.7
A2.1 355 0 0 2,125 83,818 12.8
A2.3 352 0 0 2,108 88,791 10.1
A2.4 353 0 0 2,113 148,799 10.0
A2.5 355 0 0 2,124 119,580 13.3
A3.3 181 18.4 701 334 60,121 11.1
A3.6 181 18.8 704 338 47,494 11.9
A4.3 299 30.3 591 563 66,472 10.7
A6.4 130 12.1 221 585 40,584 14.1
A6.5 130 12.1 221 584 56,594 15.5
Singh et al. / Journal of the Air & Waste Management Association 64 (2014) 1328–1340 1333

for the sake of simplifying the model, the porous surface of the
burner was replaced with small jet flames spread evenly across
the burner. These small flames coerced together to form a single
laminar and flat flame. The idea was to generate a flame having a
uniform axial profile of temperature and the species’ concentra-
tions. The temperature and the species concentration profiles
were measured along the length of the flame. As there were no
radial variations in the species concentration or temperature, the
shape of the burner (square or round) does not affect these
profiles. The geometry used for simulating the flat flame burner
is shown in Figure 3. A square geometry was used for both the
burner and the full domain. Due to the laminar characteristic of
the flame, the grid was finely meshed along the height of the
flame (normal to the burner surface). The rest of the domain had
a relatively coarser mesh. The grid had 611,008 cells and was
successfully checked for skewness.
Figure 2. Flat flame burner (McKenna burner). Sandia flame: Jet flame in a vitiated coflow
Unlike the previous experimental setup, this study simulated
a high jet velocity and a more turbulent flow. The experiment
to produce such flames. Stable flames from such burners provide
was performed by Cabra et al. (2005) at the Combustion
an accurate measurement of temperature and concentration of
Research Facility, Sandia National Laboratory, USA. In addition
species formed during the combustion process. A number of
to the flow conditions, the fuel composition used was also
researchers (Zhang et al., 2006; Bhargava and Westmoreland.,
different. In their work, Cabra et al. used CH4 and air as the
1998) have used these burners to study lab-scale flames. Zhang
reactants for combustion. The setup included a central burner
et al. at the National Synchrotron Radiation Laboratory, China,
surrounded by a perforated plate. The actual experimental setup
have studied low-pressure C2H4/O2/Ar laminar flames.
is shown in Figure 4. The jet fuel composition was kept at 33%
Synchrotron photoionization and molecular beam mass spectro-
CH4 and 66% air at a jet velocity of 100 m/sec. The premixed
metry (MBMS) techniques were used to measure species con-
fuel was maintained at a temperature of 320 K. The perforated
centrations. In this lab scale experimental setup, a 6 cm diameter
plate acted as a source of “coaxial flow of hot combustion
McKenna burner was used during the experiment. The burner
products from a lean premixed flame.” The combustion products
used by Zhang et al. is shown in Figure 2. The conditions used
were obtained from a lean H2/Air flame. The coaxial flow was
for this McKenna flame are shown in Table 3. The pore Reynolds
sent at a velocity of 5.4 m/sec and a temperature of 1350 K. The
number (d*v*r/m) for the modeled flame is 0.1065, which is
idea of providing a coaxial flow was to avoid any interference
well within the laminar region. The Richardson number (g*d/v2)
from the surrounding air. By keeping the composition and tem-
for the same flame is 2.96E-03. Note that the diameter of the
perature of the coaxial flow similar to the hot combustion pro-
micropores in a McKenna burner is given as 1.0 E–4 m (Holthuis
ducts, any interaction in the combustion chemistry could be
& Associates, 2013). Also, the Richardson number is the ratio of
avoided. Also, the coaxial flow reduced the mixing of surround-
buoyancy forces to inertia forces, where a number less than 1
ing air, which could have led to a leaner fuel and brought down
indicates a predominance of inertial forces (Pohl et al, 1984).
the flame temperature. Using this setup, Cabra et al. were able to
The two numbers were calculated using the data given in Table 4.
maintain a very stable and uniform flame. Raman/Rayleigh/
laser-induced fluorescence (LIF) measurement methods were
Geometry. Modeling lab-scale or industrial-scale flames using used to analyze the flame temperature and major species
a rigorous EDC (eddy dissipation concept) model is computa- concentrations.
tionally very expensive. The modeling becomes more complex Instead of pure O2 (as used in the McKenna flame setup),
when a comprehensive reaction mechanism like LU1.1, consist- air was used for the combustion process. Due to the addition
ing of 50 species and about 400 reactions, is employed. Hence, of N2 in the fuel, formation of NOx was also involved in the
combustion chemistry. Since almost all flares use air as their
oxidation medium, the Sandia flame is a closer representa-
Table 3. Conditions used for McKenna flame tion of the combustion chemistry during flaring. Also, the
N2 in the fuel provided an opportunity to measure the con-
Fuel composition (mole fractions) centration of NO, which is an important intermediate in the
C2H4 O2 Ar combustion process.

0.175 0.525 0.300 Geometry. A cylindrical domain was used for the simulation
Exit velocity (m/sec) 0.5757 of this flame. All of the geometry configurations and boundary
Temperature (K) 300 conditions were kept the same as the experimental setup. The
1334 Singh et al. / Journal of the Air & Waste Management Association 64 (2014) 1328–1340

Table 4. Conditions used to calculate the Reynolds/ Richardson numbers

d (Pore diameter) 1.00E–04 m McKenna Technical Support


r (Mixture density) 0.036 kg/m3 Aspen Plus Property Set
m (Mixture viscosity) 1.97E–05 kg/m-sec Aspen Plus Property Set
v (Gas velocity) 0.5757 m/sec
g 9.81 m/sec2 Perry’s Handbook

Results
The simulation results obtained in the form of flare efficien-
cies, temperature profiles, and concentrations of species (CO2,
CO, OH, H2O, O2, CH2O, and NO) were compared with the
experimental data. The details are given next.

TCEQ flare tests


The experimental data collected during the 2010 TCEQ flare
tests included mainly flare efficiencies. The predicted flare
Figure 3. Geometry used for the simulation of McKenna flame. destruction efficiencies, DRE and CE, are compared with field
measurements in Tables 6 and 7, respectively. A good agreement
was found for both flare efficiencies: The average absolute error
details are given in Table 5 and the geometry used is shown in for DRE was 4.50% with maximum error being around 10% (see
Figure 5. The grid had 86,360 cells and was successfully Table 6), while the absolute average error for CE was 3.04% with
checked for skewness. maximum error reaching around 6.7% (see Table 7). In the past,

Figure 4. Experimental setup used at Sandia National Laboratory.

Table 5. Conditions used for Sandia flame

Co-flow conditions
H2 O2 H2O CH4 N2
0.0001 0.1193 0.1516 0.0003 0.7287
Co-flow velocity Co-flow temperature
5.4 m/sec 1350 K
Jet conditions
CH4 O2 N2
0.3300 0.1518 0.5182
Jet velocity Jet temperature
100 m/sec 320 K
Singh et al. / Journal of the Air & Waste Management Association 64 (2014) 1328–1340 1335

absolute error in the DRE prediction was 13.3%. The EDC


model used in this work employed a new and simplified geome-
try. The new model improves DRE of air-assisted low LHV/low
jet velocity propylene/TNG flares significantly compared to the
DRE values reported in the AQRP report (Chen et al., 2011). It
also drastically improves CE prediction compared to the average
absolute error of 29.8% in the 2011 AQRP 10-022 report (for A/
F mass ratio < 28). The results clearly validate the CFD model
used, given the uncertainties in both the field measurements and
the numerical simulations.
Figure 6 presents the same comparisons in the form of
calculated versus experimental values for both DRE and
Figure 5. Three-dimensional geometry used to simulate Sandia flame. CE. As observed from the plot, the model closely predicts
the flare efficiencies. CE tends to have an evenly distributed
error, while DRE tends to have a higher predicted value. In
Table 6. Comparison of TCEQ measured and simulated DRE (%)
Figures 7 and 8, the DRE and CE versus combustion zone
2013 Simulation heating value (CZHV) are presented. CZHV is the resultant
heating value of the fuel gas when any assist medium like
TCEQ DRE Error air/steam is also taken into account. CZHV in this work was
TCEQ case number measurement (%) (%) calculated using eq (11). For higher CZHVs, the model gives
more accurate results compared with lower CZHVs. For this
A1.1 98.0% 99.8% 1.8% test case, the error is the largest (about 10%) for DREs
A2.1 97.2% 99.4% 2.2% predicted at the lowest heating values:
A2.3 96.1% 99.5% 3.6%
P 
A2.4 93.0% 98.0% 5.3% fi Hi þ m Hm
A2.5 95.1% 97.6% 2.6% CZHV ¼ P (11)
fi þ m þ s
A3.3 88.1% 91.7% 4.1%
A3.6 90.8% 99.3% 9.4% where fi is the volume flow rate of the ith component in vent gas,
A4.3 95.2% 94.2% 1.1% m the volume flow rate of makeup fuel, a the volume flow rate of
A6.4 92.9% 96.7% 4.0% assisted air, s the volume flow rate of assisted steam, Hi the
A6.5 87.9% 97.4% 10.8% heating value of the ith component in fuel gas (MJ/m3), Hm the
Average error (absolute) 4.50% heating value of the makeup fuel (MJ/m3), and CZHV the com-
bustion zone heating value (MJ/m3).

Table 7. Comparison of TCEQ measured and simulated CE (%) McKenna flame burner
CFD simulation Simulation of the flat flame burner using the LU 1.1 mechan-
ism was performed and the validation of the numerical model
TCEQ CE Error
TCEQ case number measurement (%) (%)
A1.1 96.9% 96.4% 0.5%
A2.1 95.9% 96.9% 1.1%
A2.3 94.4% 94.1% 0.4%
A2.4 89.3% 92.4% 3.4%
A2.5 92.6% 93.0% 0.5%
A3.3 85.2% 81.5% 4.4%
A3.6 88.2% 92.6% 5.0%
A4.3 93.6% 87.3% 6.7%
A6.4 89.2% 84.2% 5.7%
A6.5 81.5% 83.9% 2.9%
Average error (absolute) 3.04%

the same test cases were modeled and reported in the 2011 Texas
AQRP (Air Quality Research Program) 10-022 report (for A/F Figure 6. CFD Simulated flare efficiencies vs. TCEQ measured flare
mass ratio < 28) (Chen et al., 2011). In that report, the average efficiencies.
1336 Singh et al. / Journal of the Air & Waste Management Association 64 (2014) 1328–1340

Figure 7. Comparison of TCEQ measurements and experimental results: DRE


(%) vs CZHV.
Figure 9. Comparison of experimental and simulated temperature profiles of the
flat flame burner.

Figure 8. Comparison of TCEQ measurements and experimental results: CE (%) Figure 10. Comparison of experimental and simulated O2 mole fractions.
vs CZHV.

using this flame was done by comparing the axial profiles of


temperature and species concentration. The most important para-
meter to be observed is temperature as the combustion chemistry
is highly dependent on it. It can be seen in Figure 9 that the
general trend of the simulated temperature over the length of the
flame was correct. Figures 10–13 compare the mole fractions of
O2, C2H4, and CO2. A good agreement was found between the
predicted and experimental profiles of these major species con-
centrations. Although the exit mole fractions (10 mm above
burner) are the same, there is some discrepancy in the CO2 and
CH2O concentrations at the middle of the flame. For the radical-
producing CH2O, the EDC model underpredicts the maximum
mole fraction by a factor of 10 and the exiting mole fraction by a
factor of 1.6 (see Figure 14). These discrepancies can be seen as
a cumulative result of uncertainties in the reaction mechanism Figure 11. Comparison of experimental and simulated C2H4 mole fractions.
and the CFD model.
It should be noted that even though the maximum mole
fractions predicted for some of the species were not equal to Since only the exit mole fractions are used to calculate the flare
the experimental values, the exit mole fractions were about the efficiencies, the model can be used for that purpose. Further,
same. The model to be validated is aimed for flare combustion. even though the CFD model underpredicts the formaldehyde
Singh et al. / Journal of the Air & Waste Management Association 64 (2014) 1328–1340 1337

Figure 12. Comparison of experimental and simulated CO2 mole fractions. Figure 15. Comparison of experimental and simulated temperature profiles of
the jet flame.

Figure 13. Comparison of experimental and simulated CO mole fractions. Figure 16. Comparison of experimental and simulated CH4 mass fractions.

trend of observed data very well compared to the experimental


flame data. Figure 16 compares the CH4 mass fractions along the
height of the flame. It can be seen that the combustion of CH4 is
slow for the first few millimeters above the flame. However, at an
x/D value of 20 mm, the simulated and experimental results are
in total agreement. As shown by Figures 17–19, the mass frac-
tions of CO2, CO, and H2O are in very good agreement with the
experimental data. The maximum mass fractions of almost all
these products occur at the same height above the burner, that is,

Figure 14. Comparison of experimental and simulated CH2O mole fractions.

concentrations, the mere existence of formaldehyde (and maybe


at a higher concentration) as an incomplete combustion product
supports the view that flares can be a primary source of formal-
dehyde emission.

Sandia flame: High-velocity jet flow


Figure 15 compares the temperature profiles of the simulated
and experimental flames. The simulation predicted the general Figure 17. Comparison of experimental and simulated CO2 mass fractions.
1338 Singh et al. / Journal of the Air & Waste Management Association 64 (2014) 1328–1340

Figure 18. Comparison of experimental and simulated CO mass fractions.


Figure 21. Comparison of experimental and simulated NO mass fractions.

Similar to CO2, the exit mole fractions of both the radicals are
equal to the measured values. The discrepancy can be seen
around the middle of the flame, which again is due to the
uncertainty in the reaction mechanism.

Discussion
Though the simulation results were reasonably close to the
experimental data, there is some room for improvement. The
modeling of lab-scale and industrial-scale flares can be improved
by using more rigorous models. The most important factor
affecting the simulation results is the kinetic modeling. CFD
models and sufficient computational resources allow the use of
Figure 19. Comparison of experimental and simulated H2O mass fractions. a more detailed kinetic mechanism, which can improve the
accuracy of predicted concentration profiles. Improvements
can be made in the geometry of the flare tip. The ring-shaped
geometry used as the fuel and pilot gas source can be optimized
x/D ¼ 60 mm. In summary, the EDC model accurately predicts by increasing the flow area to improve the prediction of DREs. In
the profiles of temperature and concentrations of major species this work, the open air flaring process was modeled using steady-
(CH4, CO2, CO, and H2O). state models. For further improvements, unsteady-state models
Two important radicals, OH and NO, were also compared like LES models can be used.
with the experimental data. OH and NO are by-products of the In general, the CFD model used to simulate the three flames
combustion process and an important source for the formation of provided good results. In this work, a comprehensive reaction
atmospheric O3 (Seinfeld and Pandis, 2006). The model predicts mechanism LU1.1 was used in conjunction with the most rigor-
the mass fraction of trace species NO and OH (at the core of the ous turbulence model and turbulence–chemistry interaction
flame) within a factor of 3.5, as shown in Figures 20 and 21. models available to simulate industrial-scale flares at the stand-
by mode for the first time. The methodology proves to be
applicable to industrial-scale flaring in that the errors of the
predicted DREs and CEs are within 5% of the measured effi-
ciencies. For lab-scale flames, concentrations of major species,
for example, C2H4, CH4, CO2, CO, and H2O, were accurately
predicted. On the other hand, the study also observed some
discrepancies in the prediction of certain radicals and trace
species such as HCHO, OH, and NO.
The two possible reasons for these discrepancies are the
uncertainties in the original reaction mechanisms and error in
predicting the correct rate of reactions at high temperatures
(>2000 K). The latter may be a result of reduction of the com-
bined mechanism. Although the reaction mechanism was suc-
cessful in predicting the temperature profile and most of the exit
Figure 20. Comparison of experimental and simulated OH mass fractions. mole fractions, there is still some scope for improvement. A
Singh et al. / Journal of the Air & Waste Management Association 64 (2014) 1328–1340 1339

more comprehensive reaction mechanism that may include all Bhargava, A., and P.R. Westmoreland. 1998. Measured flame structure and
the species and reactions needed to accurately predict intermedi- kinetics in a fuel-rich ethylene flame. Combust. Flame 113:333–347.
doi:10.1016/S0010-2180(97)00208-3
ates can be employed. In addition to the reaction mechanism,
Bhargava, A., and P.R. Westmoreland. 1998. MBMS analysis of a fuel-lean
inclusion of radiation in the model can also help in improving the ethylene flame. Combust. Flame115(4): 456–467. doi:10.1016/S0010-2180
accuracy of the results. These changes can definitely help in (98)00018-2
better prediction of the combustion process, but will demand Cabra, R., J.Y. Chen, R.W. Dibble, A.N. Karpetis, and R.S. Barlow. 2005. Lifted
significantly higher computational resources. methane–air jet flames in a vitiated co-flow. Combust. Flame143(4):
Even though the CFD model underpredicts the formaldehyde 491–506. doi:10.1016/j.combustflame.2005.08.019
concentrations compared to the McKenna flat flame, the mere Castiñeira, D., and T.F. Edgar. 2008. Computational fluid dynamics simulation of
wind-tunnel experiments on flare combustion systems. Energy Fuels
existence of formaldehyde as an incomplete combustion product
22:1698–1706. doi:10.1021/ef700545j
supports the view that flares can be a primary source of formal- Chen, D., H. H. Lou, X. Li, K. Li, and C.B. Martin. 2011. Development of
dehyde emission. In fact, formaldehyde concentrations were speciated industrial flare emission inventories for air quality modeling in
measured in the range of 200–1200 ppbv in the 2010 flare Texas. Air Quality Research Program (AQRP) Project # 10-022, Final Report.
study (Test S4.1R1). Acetaldehyde was also detected in the Lamar University, Beaumont, TX, December 14.
similar concentration levels (Allen and Torres, 2011). These Chen, D., K. Singh, P. Gangadharan, H. Lou, X. Li, and P. Richmond. 2013. CFD
data and their ratio to CO and propylene can be utilized in future study of flare operating parameters. Paper 271b, AIChE Annual Meeting,
November 3–8, San Francisco, CA.
flare modeling work. Davis, S.G., C.K. Law, and H. Wang. 1999. Reaction mechanism of C3 fuel
combustion. http://ignis.usc.edu/Mechanisms/C3/c3.html
Conclusion Davis, S.G., and C.K. Law, 1998. Determination of and fuel structure effects on
laminar flame speeds of C1 to C8 hydrocarbons. Combust. Sci. Technol. 140
The EDC model with a new and simplified geometry (1): 427–449. doi:10.1080/00102209808915781
improves the accuracy of DRE calculations for air-assisted low Environ International Corporation. 2006. Comprehensive air quality model with
LHV/low jet velocity propylene/TNG flare tests with an average extension (CAMx) preprocessors. http://www.camx.com/down/support.php
Fluent, Inc. 2011. FLUENT 13.0 User’s Guide. http://aerojet.engr.ucdavis.edu/
error of 4.6%, compared to 13.3% published in the 2011 AQRP
fluenthelp/ (accessed September 22, 2014).
10-022 final report (for A/F mass ratio < 28; Singh et al., 2013). Holthuis & Associates. 2013. McKenna flat flame burner. http://www.flatflame.
The EDC model also improves the CE prediction accuracy with com
an average error of 2.6%, compared to 29.8% in the 2011 AQRP Jatale, A., P. Smith, J. Thornock, and S. Smith. 2012. A validation of flare
10-022 report (for A/F mass ratio < 28). combustion efficiency simulations. American Flame Research Committee,
For the lab-scale CH4/Air mixture (Sandia) flame, the EDC Salt Lake City, UT.
model accurately predicts the profiles of temperature and con- Law, C.K., A. Makino, and T.F. Lu. 2005. On the off-stoichiometric peaking of
adiabatic flame temperature with equivalence ratio. Presented at the 4th Joint
centrations of major species (CH4, CO2, CO). The model pre-
Meeting of the U.S. Sections of the Combustion Institute, Philadelphia,
dicts the mass fraction of trace species NO and OH (at the core of Pennsylvania, March 21.
the flame) within a factor of 3.5. In the case of C2H4/O2/Ar Lou, H.H., C.B. Martin, D. Chen, X. Li, K.Y. Li, H. Vaid, A.T. Kumar, K.D.
flame, a good agreement was found between the predicted and Singh, and D.P. Bean, Jr. 2011. A reduced reaction mechanism for the
experimental profiles of temperature, C2H4, CO2, and other simulation in ethylene flare combustion. Clean Technol. Environ. Policy 14
major species concentrations. The EDC model underpredicted (2): 229–239. doi:10.1007/s10098-011-0394-9
the maximum and exiting model fractions of CH2O by a factor of Mayers, B.F,. and E.R. Bartle, 1969. Reaction and ignition delay times in the
oxidation of propane. AIAA J. 7(10): 1862–1869. doi:10.2514/3.5473
10 and 1.6, respectively. OH, another intermediate species, was McDaniel, M. 1983. Flare efficiency study. U.S. Environmental Protection
found to be overpredicted in the model by a factor of 5. Agency, Report No. 600/2-83-052. July. http://www.epa.gov/ttn/chief/ap42/
ch13/related/ref_01c13s05_jan1995.pdf (accessed September 22, 2014).
Funding Nik, M.B., S.L. Yilmaz, P. Givi, M.R. Sheikhi, and S.B.Pope. 2010. Simulation of
Sandia Flame D using Velocity Scalar filtered density function, AIAA J. 48
This work is supported by the State of Texas and the authors (7): 1513–1522. doi:10.2514/1.J050154
gratefully acknowledge financial support from TCEQ Supplemental Patankar, S.V. 1980. Numerical Heat Transfer and Fluid Flow. New York, NY:
Environmental Program (SEP agreement 2009-009) and the Texas McGraw-Hill.
Air Research Center (TARC grant 079LUB0096A). Pohl, J., R. Payne, and J. Lee. 1984. Evaluation of the efficiency of industrial flares:
Test results. EPA-600/2-84-095. Prepared for U.S. EPA Office of Research and
References Development by Energy and Environmental Research Corporation. May. http://
www.tceq.state.tx.us/assets/public/implementation/air/rules/Flare/Resource_2.pdf
Al-Fadhli, F.M., Y. Kimura, E.C. McDonal Buller, and D.T. Allen. 2012. Impact (accessed September 22, 2014).
of flare destruction efficiency and products of incomplete combustion on Pohl, J., and N. Soelberg. 1985. Evaluation of the efficiency of industrial flares:
ozone formation in Houston, Texas. Ind. Eng. Chem. Res. 51:12663–12673. Flare head design and gas composition. EPA-600/2-85-106. Prepared for EPA
doi:10.1021/ie201400z Office of Air Quality Planning and Standards. September. http://nepis.epa.
Allen, D.T., and V.M. Torres. 2011. TCEQ flare study final report. TCEQ PGA gov/Exe/ZyPDF.cgi/P1003QL1.PDF?Dockey=P1003QL1.PDF (accessed
No. 582-8-86245-FY09-04 and Task Order No. UTA10-000924-LOAT-RP9. September 22, 2014).
Austin, TX: The University of Texas at Austin, The Center for Energy and Seinfeld, J. H., and S. Pandis. 2006. Atmospheric Chemistry and Physics—From
Environmental Resources. Air Pollution to Climate Change, 2nd ed. New York, NY: John Wiley and Sons.
Baukal, C.E., and R.E. Schwartz. 2001. The John Zink Combustion Handbook. Singh, K.D., T. Dabade, H. Vaid, P. Gangadharan, D. Chen, H. H. Lou, K.Y. Li, X.
New York, NY: CRC Press. Li, and C.B. Martin. 2012. Computational fluid dynamics modeling of
1340 Singh et al. / Journal of the Air & Waste Management Association 64 (2014) 1328–1340

industrial flares operated in a stand-by mode. Ind. Eng. Chem. Res. 51(39): Planning and Standards (OAQPS), April. http://www.epa.gov/ttn/atw/flare/
12611–12620. doi:10.1021/ie300639f 2012flaretechreport.pdf (accessed September 22, 2014).
Singh, K.D., P. Gangadharan, D. Chen, H. Lou, X. Li, P. Richmond. 2014. U.S. Government. 2009. Code of Federal Regulations—Standards of perfor-
Parametric study of ethylene flare operations and validation of a reduced mance for new stationary sources, general control device and work practice
combustion mechanism. Eng. Appl. Comput. Fluid Mech. 8(2): 2011–228. requirements. 40CFR § 60.18.http://edocket.access.gpo.gov/cfr_2009/julqtr/
Smith, G.P., G.M. Golden, M. Frenklach, N.W. Moriarty, B. Eiteneer, M. Goldenberg, pdf/40cfr60.18.pdf (accessed September 22, 2014).
T. Bowman, R.K. Hanson, S. Song, W.C. Gardiner, V.V. Lissianski, and Z. Qin, Vujanovic, M., N. Duic, and R. Tatschl. 2009. Validation of reduced mechanisms
2000. GRI-Mech. http://www.me.berkeley.edu/gri_mech for nitrogen chemistry in numerical simulation of a turbulent non-premixed
Smith, P. J., J. Thornock, D. Hinckley, and M. Hradisky. 2011. Large eddy flame. React. Kinetics Catal. Lett. 96(1): 125–138. doi:10.1007/s11144-009-
simulation of industrial flares, SC’11 Companion, November 12–18, 5463-2
Seattle, WA. ACM 978-1-4503-1030-7/11/11. Wang, H., X. You, A.V. Joshi, S.G. Davis, A. Laskin, F. Egolfopoulos, and
Texas Commission on Environmental Quality. 2000. Air permit technical guidance C.K. Law. 2007. USC Mech Version II. High-temperature combustion reac-
for chemical sources: Flares and vapors oxidizers. Air Permits Division, Texas tion model of H2/CO/C1-C4 compounds. http://ignis.usc.edu/USC_Mech_II.
Commission on Environmental Quality. October. http://www.tceq.state.tx.us/ html (accessed September 22, 2014).
assets/public/comm_exec/pubs/rg/rg360/rg36007/techsupp_4.pdf (accessed Zhang, Q., Y. Li, Z. Tian, T. Zhang, J. Wang, and F. Qi. 2006. Experimental study
September 22, 2014). of premixed stoichiometric ethylene/oxygen/argon flame. Chin. J. Chem.
Texas Commission on Environmental Quality. 2007. Technical supplement 4: Phys. 19(5): 379–385. doi:10.1360/cjcp2006.19(5).379.7
Flares. TCEQ Publication RG-360A. January. http://www.tceq.texas.gov/ ZEECO. 2007. Flare system emissions. Control. Texas Technology. Conference.
assets/public/comm_exec/pubs/rg/rg360/rg36010/rg-360a.pdf http://texasiof.ces.utexas.edu/texasshowcase/pdfs/presentations/b2/ssmith.
Texas Commission on Environmental Quality. 2009. Comprehensive flare study pdf (accessed September 22, 2014).
project. PGA No. 582- 8-862-45-FY09-04, Tracking no. 2008-81. http://
www.tceq.state.tx.us/assets/public/implementation/air/rules/Flare/Flare_
Research_Final_QAPP.pdf (accessed September 22, 2014).
University of Texas at Austin. 2010. Quality Assurance Project Plan, Texas About the Authors
Commission on Environmental Quality, Comprehensive Flare Study, PGA No.
582-8-862-45-FY09-04. http://www.tceq.state.tx.us/assets/public/implementation/ Kanwar Devesh Singh and Preeti Gangadharan each hold a Ph.D. in chemical
air/rules/Flare/Flare_Research_Final_QAPP.pdf (accessed September 22, 2014). engineering from Lamar University, Beaumont, TX.
Torres, V., S. Herndon, Z. Kodesh, and D.T. Allen. 2012. Industrial flare perfor-
mance at low flow conditions. 1. Study overview. Ind. Eng. Chem. Res. Daniel H. Chen and Helen H. Lou are professors and university scholars, and
51:12559–12568. doi:10.1021/ie202674t Peyton Richmond is an associate professor at Dan F. Smith Department of
Tyliszczak, A. 2013. LES-CMC & LES-flamelet simulation of non-premixed Chemical Engineering, Lamar University, Beaumont, TX.
methane flame (Sandia F). J. Theor. Appl. Mech. 51(4): 859–871.
U.S. Environmental Protection Agency. 2012. Parameters for properly designed Xianchang Li is an associate professor in the Department of Mechanical
and operated flares. Report for Flare Review Panel, EPA Office of Air Quality Engineering, Lamar University, Beaumont, TX.

View publication stats

You might also like