Sei sulla pagina 1di 35

This article was downloaded by: [Mount Allison University 0Libraries] On: 21 May 2013, At: 13:29 Publisher:

Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Combustion Science and Technology


Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/gcst20

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS WITH HIGH-TEMPERATURE AIR


STEFANO ORSINO , ROMAN WEBER* & UGO BOLLETTINI
a b a a b

International Flame Research Foundation, Research Station B.V., IN, The Netherlands

Energy Department- Engineering Division, ENEA Italian National Agency for NewTechnology, CR Casaccia, Rome, Italy Published online: 03 Apr 2007.

To cite this article: STEFANO ORSINO , ROMAN WEBER* & UGO BOLLETTINI (2001): NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS WITH HIGH-TEMPERATURE AIR, Combustion Science and Technology, 170:1, 1-34 To link to this article: http://dx.doi.org/10.1080/00102200108907848

PLEASE SCROLL DOWN FOR ARTICLE Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

Combust. Sci, andTech.. 170:134.2001 Copyright 02001 Taylor & Francis 0010-2202/01 512.00 + .OO

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS WITH HIGH-TEMPERATURE AIR


Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

STEFAN0 O R S I N O AND R O M A N WEBER*

International Flame Research F o u n d a t i o n ,


Research Station BY, IN,The
Netherlands

UGO B O L L E T T l N l

ENEA- Italian National Agency for N e w T e c h n o l o g y , Energy Department- Engineering Division, CR C a s a c c i a ,


Rome,

Italy

This article deals with the technology of burning natural gas with high-temperature air and large quantities of flue gas. The objective of this work is to assess the potential of several combustion models for their abilities to predict characteristics of the new combustion technology. Three numerical models have been used: (a) an Eddy-Break-Up model with a two-step reaction scheme, (b) an Eddy-Dissipation Concept model with chemical equilibrium and (c) pdf/mixture fraction model with equilibrium, non-adiabatic look up tables for chemistry. The nitric oxide post processors have incorporated thermal and prompt mechanisms as well as the NOx reburning mechanism. The computational results have been compared with both in-furnace measurements (temperature, 0 2 , COZ, CO. CHI, NO, velocities, radiative heat Hux) and with the measured furnace exit parameters. All three models have correctly reproduced the characteristics of the high temperature air combustion, namely the uniformity of the temperature field, high radiative Huxes and low NOx and C O emission. With the exception of a small region located within the natural gas jets, the chemical equilibrium assumption with respect to both natural gas combustion and equilibrium 0 , O H and N radicals have resulted in predictions of very good quality. The Eddy Dissipation Concept model and the pdf/mixture fraction model have provided almost identical results. The tested combustion models cannot describe the chemistry and Received 18 April 2001; accepted 26 June 2001 *E-mail: roman.weber@ifrf.net

S. ORSINO ET AL.
temperature field in the fuel jet region. This is partially caused by imperfections in predicting the entrainment of the weak methane jet interacting with the strong combustion air stream. A more comprehensive understanding of the chemistry of natural gas combustion under fuel-rich c o n d ~ t ~ o n with s comburent containing >4% oxygen is required to develop appropriate nonequilibrium sub-models for this fuel jet region.
Keywords: high temperature air combustion, flameless oxidation, mild combustion

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

INTRODUCTION Regulations of the Kyoto conference impose on industrialized nations reductions of carbon dioxide emissions. The most effective way of achieving the goal is to increase the thermal efficiency of industrial furnaces and, by doing so, to decrease the C 0 2 emissions per ton of product. The increase in the thermal efficiency requires both high process temperatures and efficient heat recovery systems. Technology Development Throughout the seventies and eighties, remarkable progress in heat recovery at the burner was achieved due to British Gas and Hotwork International (Masters et al., 1979). Further progress in the regenerator design was made in Japan, a t Nippon Furnace (NFK), at the beginning of this decade (Tanaka et al., 1979). New honeycomb type regenerators happen to be much more compact and possessed smaller thermal inertia (Tanaka et al., 1979, Katsuki and Hasegawa, 1998). The honeycomb regenerators operated a t a very small temperature difference (typically of 50- 100C)between the furnace exit temperature and the combustion air temperature. They offered possibilities of achieving air preheat up to 1200Cusing combustion products at temperature of 125G 1300C.This opened up new possibilities for further improvement of furnace efficiency and for minimizing fuel consumption. A high combustion air temperature usually results in high NOx emissions, thus measures are needed to overcome this disadvantage whilst simultaneously perfecting the heat regenerator design. In the last decade substantial developments in the low NOx combustion with highly preheated air have been achieved (Nakamachi et al., 1990, Wiinning and Wiinning, 1992). A fuel jet injected into hot combustion products is the

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

essential element of the technology. In some applications, burners are used in pairs; when the first burner is on fire the hot flue gases heat up the regenerator of the second burner. The firing cycle changes every 2CL50s. There are different ways of realization of this combustion technology and thus the process itself has been given several names. In Japan, the process was originally called Excess Enthalpy Combustion following the suggestion of Hardesty and Weinberg, 1974 and later on it was renamed into Colorless Combustion o r even HTAC (High Temperature Air Combustion). In Germany, it has been named Flameless Oxidation while in the USA, a term Low-NOx Injection is used. Many furnaces operating with these combustion techniques have been working with success in the last 5-10 years, especially in the steel industry (Suzukawa et al., 1998; Milani et al., 1997; Milani, 2000).
Fundamental and Applied Research

Although the technology of fuel combustion with highly preheated air has substantially advanced over the last decade o r so, there has been relatively little work undertaken by the fundamental combustion community to support the development. The work on fundamentals of the process was initiated probably by Mr. R. Tanaka of NKK sometime a t the beginning of the nineties. Soon afterwards (Gupta et al., 1997 and 1999) undertook research on the effect of combustion air temperature and oxygen concentration on flame color, visibility and thermal emission spectra. Propane was used as the fuel. The researchers observed a substantial increase in radiation intensity a t wavelengths corresponding to C2 radicals emission with the preheated air temperature under low oxygen concentration conditions. Mochida and Hasegawa (2000) have been developing a flame visualization technique based on the luminescence intensity ratio of C2 and C H radicals. Wunning and Wunning (1997) comprehensively described the main characteristic of flameless oxidation as promising high efficiency and low-NOx combustion technique. Blasiak et al. (2000) built an experimental facility for studying fuel jets immersing into a cross-flowing high temperature air stream and their preliminary findings have been reported. In 1997, the I F R F carried out semi-industrial scale experiments (Verlaan et al., 1998; and Weber et al., 1999) that identified principal characteristics of the combustion process. The furnace was operated almost like a well-stirred reactor. The measured radiative heat fluxes at the furnace walls were very high and uniform.

S. ORSINO ET AL.

Uniformity and homogenization of combustion of methane and propane with highly preheated air (1000C)have recently been studied by Ishiguro et al. (1998). They-obtained images of OH, C H and C2 emissions for a number of experimental conditions that differed in the air preheat level and oxygen content of the air. The investigators concluded that the increase in air temperature results in a decrease of flame temperature gradients (homogenization of flames). Furthermore, the higher the combustion air temperature the lower the flame fluctuations are. Plessing et al. (1998) used laser-induced predissociative fluorescence and Rayleigh thermometry to examine flameless oxidation a t laboratory scale. They observed that the flameless oxidation "takes place in the well-stirred reactor regime. The O H concentration in the combustion zones of flameless oxidation is lower than in non-preheated undiluted turbulent premixed flames." De Joannon et al. (2000) have examined the applicability of the existing chemical reaction schemes for combustion of hydrocarbons to high temperature air combustion conditions. In the most recent publication, Cavaliere and De Joannon (2000) have argued that flameless oxidation can be described as a two-staged combustion in which the first part is in rich conditions with plenty of inert gases. Mathematical modeling of the high temperature air combustion has also received attention. Ishii et al., 1997 and Hino et al., 1998 carried out simulations of the N K K experimental continuous slab reheat furnace with emphasis on the NOx formation. They show that numerical codes based on the mixture/fraction pdf approach are able to describe the main characteristic in terms of flow and temperature field of the high air temperature combustion and can be used to identify the best low-emissions furnace configuration. However, they found out that present NOx models might require an improvement to describe properly the NOx formation under "low-temperature" conditions. Yuang and Naruse (1998) reported the results of the simulation of a test furnace operated with a regenerative burner firing Liquefied Petroleum Gas with diluted (4% vol. 0 2 ) high temperature air. The numerical simulation was able to describe qualitatively the observed trends of NOx and soot emissions with the comburent dilution. Dong and Blasiak (2000) reported the modeling of the I F R F tests (Verlaan et al., I998 and Weber et al., 1999) on Natural Gas combustion with preheated air. Weber et al. (2000) reported preliminary simulations of the I F R F test and this paper may be seen as a continuation of this work. More recently, Coelho and Peters (2001) carried out numerical simulations of a mild combustion (FLOX) burner that operated at 10 kW

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

thermal input with a relatively low level of air preheat (500C). The experimental furnace was operated under the rather low temperature of 1000Cresulting in NOx emissions around 10 ppm. Coelho and Peters (2001) argued that the steady flamelet library was unable to "correctly describe the formation of N O since this is a chemically slow process, which is sensitive to transient effects" while the unsteady flamelet model was able to predict "the correct order of magnitude of N O emissions." Later on we will demonstrate that our steady-state calculations result in excellent N O predictions although for conditions of high air preheat (1300C), high furnace exit temperature (1220C) and for a steady-state operation of a 0.58 MW burner.
Objectives

Despite the valuable studies on the numerical simulation of the high temperature air combustion, some fundamental issues are still to be resolved. Contradictory conclusions have been drawn with respect to whether this new combustion technology operates as mixing or chemically controlled. The limitations of the available combustion models for predicting the novel combustion technology are not clear at all. Conditions of uniform and low temperature and low O2 concentration are substantially different from those encountered in conventional burner technology. Questions have been raised about whether the well-known thermal kinetics of methane oxidation that was mainly developed and tested under high temperature conditions, could perform well at temperatures below 1600C. Different opinions have been expressed on the importance of the NOx reburning mechanism and NO-prompt formation mechanism in the novel combustion technology. In this paper, we intend to clarify some of these issues by comparing the results of the numerical simulations with the in-furnace measurements.
EXPERIMENTAL

The semi-industrial scale experiments were carried out in a refractory lined furnace with a 2 m x 2 m cross-section and a length of 6.25 m, Figure 1. The furnace was equipped with one burner only which was operated under steady state conditions. The high temperature regenerator was replaced by a precombustor where natural gas was burned with air

S. ORSINO ET AL.

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

H
Figure 1. Experimental furnace.

under lean conditions. Oxygen was added to the hot gases leaving the precombustor to maintain its concentration at 21%. Such a comburent simulated the air preheat of the excess enthalpy combustion process. The experimental burner in Figure 2 was a I :I copy of the NFK burner that operated a t 0.58 MW fuel input and the combustion air was preheated to 1300C.

Figure 2. Experimental burner

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

Table 1. Experimental conditions


Flow rate kg/h TempT Natural Gas Comburent 47 830 kg/h 877 kg/h 25 1300 1220 Enthalpy MW 0.58 0.35 0.380

Composition %, vol CH4 87.8%, C2Ho 4.676, C,H8 1.6%, C.Hlo 5.5%, N2 LCV=44.76 MJ/kg 19.5%wel 02, 59.l%wel N2, i5%H20, 6.4%wet, C 0 2 , NO I IOppm vd 1.6% wet 02,140 ppm vd NOx

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

Furnace Exit Gases

The comburent was supplied through the central channel with an injection velocity of -85 m/s. The natural gas injectors were located 280 mni away from the burner centerline. The fuel was supplied through two injectors (although four injectors are shown in Figure 2 only two 100 m/s. The horizontal ones were used) with a n injection velocity of input-output conditions of the experiment are listed in Table 1 . During the test, the flue gas temperature and flue gas compositions were continuously monitored using a suction pyrometer and a gas sampling probe. Under steady operating conditions, a complete set of in flame measurements were acquired, including the flow and mixing patterns, the total radiative heat flux, the combustion gas temperatures and gas composition. Detailed mapping of the turbulent velocities was carried out using Laser Doppler Velocimetry (LDV). The in-furnace temperatures were measured using a suction pyrometer while the furnace wall temperatures were obtained using type B thermocouples. A cranked probe, operated under quenching rate of 1 0 ~ - 1 0 ~ K/s, was used to sample the gas, which was analyzed for oxygen, carbon monoxide, carbon dioxide, methane, hydrogen and nitric oxides. An ellipsoidal radiometer was used to measure hemispherical fluxes at the furnace wall. The in-furnace measurements, reported in this paper, were obtained by traversing the furnace in the horizontal plane of the two fuel injectors. A detailed description of the furnace and measurements technique can be found in Verlann et al.. 1998.

TURBULENT COMBUSTION MODEL

In the present study, three different combustion models were applied to simulate the natural gas combustion with high temperature air:

8
0
0 0

S. ORSlNO ET AL.

An eddy-break-up - mixed-is burnt model, An eddy dissipation - chemical equilibrium-model, A mixture fraction/pdf model with equilibrium look-up tables.

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

Since the NOx species have little impact on the combustion process, the calculation of the nitric oxides were post-processed. Two NOx models were used; one for the first two combustion models (the eddy-break-up and the eddy dissipation models) and a different one for the mixture fraction/pdf model.
Eddy-Break-Up Model

The I F R F EBU (eddy-break-up) combustion model (Peters and Weber, 1995) considers six major chemical species: hydrocarbons (C,H,), oxygen, carbon monoxide, carbon dioxide, water vapor and nitrogen. The combustion is modeled with a two-steps reaction scheme. The hydrocarbon C,Hy is first oxidized to C O and HzO while intermediate CO oxidizes to C 0 2 following the reaction scheme:

where:

In this reaction scheme, C,H,. is the Numerical Fuel (hypothetical single hydrocarbon) calculated from the natural gas composition, while fE, sl, H I and LCV are the C,H, carbon fraction, the stoichiometric oxygen requirement, the heat of reaction and the lower calorific value, respectively. The various source and sink terms entering the chemical species mass balance equation are described using the eddy-break-up model of

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

Magnussen and Hjertager (1976). The rates of formation and reduction of chemical species are related to the dissipation rate of eddies and expressed by means of the gaseous density p, the chemical species mass fraction mi, the turbulent kinetic energy k and the turbulent viscous dissipation rate E . Furthermore, the eddy-break-up concept makes use of a proportionality constant, the mixing rate coefficient Ami,. The maximum possible combustion rate of C,H,, and CO reads, in kg/m3 s:

where

PI Re, = -;

!-'

A = 12.93;

= 0.25;

C p = 0.09

In the case that the two combustibles burn according to the defined reaction scheme and at the above maximum possible rates, the required oxygen consumption rate reads:

The maximum possible oxygen consumption rate equals:

Thus, the following inequality must be obeyed:

If this inequality is obeyed, then the actual oxygen consumption rate equals the required one. If the inequality is not obeyed, then the combustion rates of the two combustibles (hydrocarbons and carbon monoxide) have to be limited according to:

S.ORSINO ET AL.

and the two combustion rates equal:

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

where the limiting is solely based on the oxygen requirements of the individual combustibles. Eddy Dissipation C o n c e p t w i t h Full Equilibrium Calculations The EDC model of Magnussen (1981) is derived using a detailed description of the dissipative process in the flow. The model derivation starts with the expressions describing the dissipation of turbulent kinetic energy. Further on, characteristic velocity and length scales for turbulent structures where the mixing is completed are derived. In the EDC, these turbulent structures are called "fine structures." Using the velocity scale and length scale associated with the fine structures, an expression for calculating the mass fraction they occupy is derived, i.e.,

where, v is the viscosity (m2/s), k the kinetic energy of the turbulence (m2/s2) and E is the dissipation rate of the kinetic energy (m2/s3). The characteristic residence time in the fine structures is defined by:

T o obtain a formula for the mean average reaction rate, an expression for the mass transfer between the fine structures and the surrounding is needed:

where Y! denotes the mass fraction of the species in the surrounding and Y ; is the mass fraction in the fine structure. In the above expressions, the parameter expresses the probability that conditions are suitable for reactions to occur within the fine structures. The relation between the

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

11

local mean value of mass fraction of species k , its mass fraction in the surrounding and in the fine structures is:

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

Species are assumed to be mixed on a molecular level within the fine structures. Thus Y;is obtained from solving a Perfectly Stirred Reactor (PSR) problem, where the mass fraction of component k in the inlet is Y! and the residence time of the PSR is T*. If the reactions in the reacting part are assumed to be infinitely fast and reversible, the conditions in the PSR can be obtained through the thermodynamic equilibrium approach. The equilibrium composition was calculated using the CEA code of NASA (Gordon and McBride, 1994). The program is based on a n algorithm for minimization of Gibbs free energy. In the calculations, thirteen species were considered: CzHs, CH4, H, OH, 0 , NO, NO2 and N2. The model used here is C02, 0 2 , H20, Hz, identical to the one used by Brink et al. (2000) and Breussin er al. (2000) for computing oxy-natural gas flames.
Mixture Fraction/Pdf Approach

For turbulent combustion flames, if the chemistry is sufficiently fast, the instantaneous thermochemical state of the fluid can be described as a function of a conserved scalar quantity known as mixture fraction (Pope, 1990; Peters,1986). For a fuel/oxidizer binary system, the mixture fraction can be defined as follows:

where ZK is the element mass fraction of the element K. Subscripts 0 and F denote the value a t the oxidizer and fuel stream inlets, respectively. Species concentrations are derived from the predicted mixture fraction distribution. Reaction mechanisms are not explicitly defined; instead, the reacting system is treated using chemical equilibrium calculations. The instantaneous thermochemical state of the fluid is a function of the mixture fraction ( f ) and instantaneous enthalpy (h):

12

S. ORSlNO ET AL.

where @ may represent the instantaneous species concentration, density or temperature. Interactions between chemistry and turbulence are taken into account using a joint probability density function (PDF), P (f,h ) . It can be assumed that heat loss does not significantly alter the turbulent enthalpy fluctuations so P = P (f). Then, the time-averaged quantities of any fluctuating variable can be calculated as follows:
Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

A p-function is used to describe the shape of the PDF. The /?-function P D F shape is calculated from the mean mixture fraction f and its v a r i a n c e p as follows:

where

and

The mean mixture fraction f and its variance solving their corresponding transport equations:

f'2 can be obtained

where the constant o,,C, and Cd take the values 0.7, 2.86 and 2.0.

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

13

The computational time is optimized using a pre-processor to obtain the equilibrium composition of all the chemical species considered in the calculation. During the pre-processing, three-dimensional look-up tables are generated for the mean flame temperature, density and species mole fraction as a function of the mean mixture fraction, its variance and the enthalpy. In the present work. eleven chemical species (CH4, C2H& C3H8, C4HI0,CO, C 0 2 , H 2 0 , OH, N2, 02, and HZ) have been considered.
Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

NOx M o d e l s

Chemical reaction rates of N O formation and N O concentrations are post-calculated using previously computed velocities, turbulence, temperature and chemistry fields. In the combustion of natural gases, the two principal formation mechanisms of NO are oxidation of molecular nitrogen (thermal-NO formation) and the mechanism in which hydrocarbon radicals attack molecular nitrogen to form nitrogen-radicals, the latter subsequently reacting with the molecular oxygen to form N O (prompt-NO formation). In a fuel rich region, N O can be reduced by CHi radicals to form species like HCN, which subsequently produces N O o r N2. This reduction mechanism is known as NO-reburning mechanism or fuel-staging mechanism.
IFRF NOx model (Peters and Weber.1995)

Thermul-NO formarion mechur~ism:It is generally accepted that, in combustion of fuel-lean and near stoichiometric fuel-air mixtures, the principal reactions governing the formation of N O from the oxidation of molecular nitrogen are those originally proposed by Zeldovich:

These two reactions are usually referred to as the thermal-NO formation mechanism o r the Zeldovich mechanism. The reaction

14

S. ORSINO ET AL.

may also contribute to the formation of NO, especially in fuel-rich and near stoichiometric fuel-air mixtures. The reactions (Rl), (R2) and (R3) are usually referred to as the extended Zeldovich mechanism. Then, under the assumption that the N-radicals concentration may be calculated from a steady state approximation, the chemical reaction rate of thermalN O formation reads, in g-mole/cm3s:
Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

Here

denotes the equilibrium constant of the overall reaction:

and may be expressed as

In the above equations, klr, k l bk2/, , k2hr k3/ and kjb stand for the forward and backward rate constants corresponding with the reactions (RI), (R2) and (R3). The constants of these rates are based on the work of Bowman (1975). The 0 - and OH-radical concentrations are directly obtained from the equilibrium calculations when the EDC model is applied. When the main calculations are performed using the mixed-is-burned model, those radical concentrations have to be estimated. The 0-radicals concentration is assumed to be equal to the equilibrium 0-radicals concentration, as in the case of molecular oxygen dissociation, k . :

where

is the equilibrium constant for the reaction equilibrium

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

15

For

Westenberg (1975) proposes the following expression, in atm:

The OH-radicals concentration is assumed to be equal to:

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

where kSJ and ksh are the forward and backward rate constants for the reaction:

For ksf,Westbrook and Dryer (1984) propose the following expression, in atm0.5:

and finally, for kSh, Baulch (1992) proposes, in atmO.':

Prompt-NO formmion mechat~isn~. In the combustion of hydrocarbon fuels, the actual chemical reaction rates of N O formation can exceed those attributable to the direct oxidation of molecular nitrogen by the (extended) Zeldovich mechanism. This is especially true under fuel-rich conditions. This rapidly formed N O was termed prompt-NO by Fenimore since the rapid N O formation he observed, occurred very early in the flame front. Reaction mechanisms involving hydrocarbon-radicals, like the reactions:

and

16

S. ORSINO ET AL.

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

together with the subsequent reaction of N-radicals via the reactions (R2) and (R3) to form NO, play an important role in the formation of promptNO. For gaseous fuels, De Soete (1975) proposes a roughly estimated chemical reaction rate appropriate for this prompt-NO formation mechanism. In this rate, Missaghi et al. (1990) have included a prompt factor to extend the expression to natural gas combustion. In terms of concentrations, it reads, in gmole/cm3s:

where M stands for the mixture molecular weight and p the mixture density. The oxygen power bl may vary between 0 and 1. For a temperature range of 1200 K to 2500 K, a value of 0.5 is appropriate. For the temperature power p,, a value of 1 is advised. For methane, CH4, which is the main component of natural gas, the constants C and E, take the following vlaues:

The prompt-factor f , , is calculated as follows (De Soete, 1975):

the where Nc is the number of carbon atoms in the hydrocarbon and lo, local stoichiometry.

NO reburning niechariism. Under fuel rich conditions, N O can be reduced by CHi radicals. The reactionsinvolved in thisglobal mechanismare:

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

For reaction R9, Chen (1994) proposes the rate


r9-i.m = 2.68

* ~ O ~ [ C , H ~ ~ ] [exp N O ] - Ego)

For reactions 10 and 11, De Soete (1975) proposes


Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

=0

( ) [ H C N ~ [ O ~ 67000 ] ~ ~ X ~ ( - ~ ) ;
Mmix
h

Pmix

where Mmi,and pmi, are the mixture molecular weight and the mixture density, respectively. The power b is locally calculated according to De Soete (1975):

Turbulence/Chernistry-Interocrion. The interaction between turbulent flow and chemistry is taken into account through the transport equation of N O chemical species. The source term of this equation, which is the time-mean chemical reaction rates of N O formation/reduction, is computed on the basis of the single-variable probability density function (pdf) model. Since the formation and reduction mechanisms of NO are strongly temperature-dependent, a "presumed, single-variable pdr'-approach has been chosen. For the instantaneous gaseous phase temperature, T, the Beta-pdf Bpdf(T;a,b ) has been chosen. This probability density function is defined as:

where the Beta-function B(a, b) is given by:

18

S. ORSINO ET AL.

Here, T, and Th, o r the unburned and the burned temperatures, are respectively the lowest-possible value and the highest-possible value of the instantaneous (read: fluctuating) gaseous phase temperature T. These temperature limits will be discussed further below. The expectation Exp(T) of the fluctuating temperature T is set to the time-mean temperature F, i.e.:
Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

while the variance Var(T) of the fluctuating temperature T is assumed to be approximately equal to a fraction of the maximum possible variance, i.e.

Here, the fractions s is called the variance coefficient. By means of the expectation and variance expressions, the Beta-pdf parameters a and b may be found:

Below, expressions will be given for the unburned and burnt temperatures T,, and T b . The assumption is made that, in oxygen rich regions, all the gaseous combustibles might burn instantaneously while in oxygen lean regions, all the oxygen might be used to burn a portion of the gaseous combustibles instantaneously. Under these conditions, the corresponding imaginary enthalpy increase Ahi,, may be calculated. Consequently, the imaginary temperature increase ATi,, may be algebraically obtained:

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

The unburned and the burnt temperature,. T, and Tb,are set to:

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

denotes the combustion air inlet temperature. The value of Here, TCai 0.5 is somewhat arbitrary. This "small" value is needed to make sure that T, is always less than T and that T h is always larger than F. The time-mean rates of thermal-NO, prompt-NO and NO-reburning are approximated by the expectations E ~ p ( r ~of - the ~ ~original ) chemical reaction rates rk-NO.

Nitric Oxide models used with the pdf combustion model. The NOx post-

processing used with the pdf/combustion models considers the thermal and prompt mechanisms. The NOx reburning mechanism is excluded. The extended Zeldovich mechanism is used and the rate constants are taken from Hanson and Salimian (1984). The prompt NOx mechanism is modeled following the overall simplified NO-formation rate proposed by the D e Soete (1975). The interactions with the turbulence are predicted using a pdf approach using a 8-function. The limits of the pdf integration are determined from the maximum and minimum values of the predicted temperature in the computational domain. The Fluent C F D code (version 5.4) is used in this paper to run the NOx post processor described in this paragraph. Flow Field Calculation The fluid flow is modeled solving the mass and momentum transport equations with the standard k-E turbulence model. The eddy-break-up and EDC combustion models have been implemented by the IFRF into the C F D code Fluent 4 as user-defined subroutines. F o r the pdf/mixture fraction approach, the original Fluent 5 formulation has been used. The radiation heat transfer has been simulated by the Discrete Ordinates radiation model (mixture fraction/pdf model) and by the Discrete Transfer radiation model (EBU and EDC models).

20

S.ORSlNO ET AL.

RESULTS

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

The experimental furnace had two symmetry planes, therefore, a grid reproducing only a quarter of the furnace was built. The grid was a structured hexahedral mesh of the geometrical domain with 92 x 42 x 129 (498,456) cells. Conjugated heat transfer boundary conditions were used so both the heat flux and the wall temperatures were calculated. Thus, radiation, convection inside the furnace, and conduction through the refractory were solved for. An overall heat transfer coefficient (at the outer furnace wall) of I I w / m 2 ~ was used. The value was calculated using the known (measured) amount of the heat extracted from the furnace. For the linear absorption coefficient, a constant value of 0.4 I/m was used. This value was derived from the relation:

where c: is the emission of a gas mixture of thickness L. The mean beam length L was evaluated according to Hottel and Sarofim (1967) as:

where V is the furnace volume and S is the total furnace surface. The emissivity was calculated using the exponential-wide band model (Edwards, 1976; Lallemant and Weber, 1996) for a gas mixture at the furnace exit temperature and composition. The predictions of the three combustion models were compared with the measurements performed a t five traverses on the plane passing through the natural gas injectors. In the following figures, radial distance 0 represents the furnace centerline (comburent injections) and the natural gas injection is located a t 0.28 m radial distance.
Flow Field

Figure 3 shows the measured and predicted axial velocity. The velocity measurements were taken using a Laser Doppler Anemometry and the measured time-mean values are accurate within 5% (Dugui: and Weber, 1992). The uncertainty originates from positioning of the LDA probe

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

21

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

inside such a large furnace. In this respect, positioning of the LDA probe inside the central (air) jet is less critical than within the small natural gas jets. The comburent was injected into the furnace a t a high velocity, as can be seen a t the first measured traverse where velocity as high as 90 m/s were measured and predicted. The strong central jet built a large zone of stagnation o r reverse flow where the axial velocity was close to zero. The fuel jets were injected into this stagnation zone. Figure 3 reports the results of the simulation performed with the three different combustion models. In all the three cases, the main features of the central jet flow field were correctly predicted. A deeper analysis of predictions in the fuel jet region revealed that the small fuel jet did not penetrate deeply enough into the furnace if compared to the experimental observations. The predicted natural gas jet was not strong enough and its entrainment of the surrounding flue gas was underpredicted (in Figure 4 the M, and Me represent the initial and entrained mass flow rate of the fuel jet, respectively). The predictions of the three combustion models showed similar trends, the predicted entrainment was 40-70% lower than the measured one. The eddy-break-up model predicted a slightly higher entrainment

00

AXIAL V E L O C I N ( d s )

IFRF-EBU IFRF-EDC

. . . flurntSpdf
A

00

murmnd

Figure 3. Measured and predicted axial \'elocity (mfs)

22

S. ORSINO ET AL.

NORMALISED POSITION X/D

$, 3000
v

10
I

20
I

30
1

40
I

50

IFRF-EBU

-6

A
200

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

v,

-4
A
0

LU

= n
z H

<

v) 1 5 0 -

1000

50-

-2

-0
AXIAL POSITION (m)
Figure 4. Natural gas jet entreinmenl

than the other two models. We attribute this effect to higher temperatures predicted within the (burning) fuel jet using the EBM model (see below). The imperfections in predicting the fuel jet are typical for numerical simulations of a small jet immersing into a hot stagnant flow and interacting with a much stronger - higher momentum jet. In the predictions, the momentum of the small jet was too quickly dissipated and the jet itself was too strongly influenced by the central comburent jet. This could be either due to a non-adequate modeling of the turbulence o r to an excessive numerical diffusion. A further investigation of the jet behavior was carried out using the realizable k-Eand RSM turbulence models, and both runs showed very similar results. The numerical diffusion was minimized using the second order interpolation scheme (QUICK) and several grid refinements were examined. All these extra runs resulted in insignificant changes to the predicted entrainment of the jet. There is, however, another factor that should be considered. We imposed the symmetry boundary conditions since only a quarter of the furnace was simulated. In reality, the interaction of the central strong jet and two natural gas weak jets leads to time-dependent departures from the symmetry. Forcing the asymmetry may result in a faster entrainment of the

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

23

natural gas jets into the central air jet. A simulation of the furnace without a symmetry boundary condition means maintaining similar grid fines would result in around two million cells. This is, however, beyond our computational capabilities.
Gaseous Species Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

Figures 5-7 show the measured and predicted 0 2 , C O and CH4 volumetric concentrations dry and the measured values are within a 5% accuracy range. The measurements showed uniformity in the chemistry field as it has already been discussed in previous works (Verlann et al., 1998; Weber et al., 1999). Almost the entire furnace volume was filled up with combusting products containing 2-3% 0 2 VOI.dry. Only in the central and fuel jet were strong gradients in the chemistry field observed. The predictions of the O2 field were of good quality and they were very similar for the three combustion models whereas larger differences could be observed in the C O predictions, see Figure 6. During the experiments, CO concentrations in excess of 2000 ppmvd were measured at the natural gas injection plane until the last measured traverse a t

OXYGEN (Xvol, dry)

IFRF-EBU IFRF-EDC
+

. . . .flucnt9pdf
A

meawed

Figure 5. Predicted and measured 02%VOI dry.

S. ORSINO ET AL.

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

CARBON MONOXIDE (%vo~, dry)

Figure 6. Predicted and measured COX vol dry

4.98 m distance downstream from the fuel injection (not shown in the figure). The eddy-break-up model predicted well the C O a t the first two traverses and than it showed a too fast CO-consumption. Carbon monoxide concentrations of the order of a few ppm were predicted a t the third traverse while the measured values were around 1.5% vol. dry. The two equilibrium-based approaches overpredicted the C O levels by far, showing a t the first two traverses C O concentrations above 10% vol. dry. Further downstream, C O was rapidly consumed and no C O was predicted at the third traverse. As Figure 7 shows, the methane concentration a t the first traverse was over-predicted by all three combustion models. Considering that the oxygen field was predicted well a t the same location, the methane predictions reinforced the observation that the deficiency in predicting the entrainment was the key Factor. Although all the oxygen entrained into the fuel jet reacted with the methane (see Figure 9,the fuel concentration remained too high if compared to the measurements. Thus, not enough oxygen was entrained into the fuel jet and consequently not enough methane reacted with the O2 (as it will be further discussed while analyzing the temperature field). Had the entrainment been calculated correctly, the E B U model predicted methane concentration would agree well with the measured values.

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

METHANE (%, dry)

IFRF-EBU IFRF-EDC

...A

f luent5-pdf measured

Figure 7. Predicted and measured CHI% vol dry

Temperature Field

The measured temperature field was almost uniform all over the furnace. The peak in-furnace temperature was as low as 1535 "C, only approximately 150 "C higher than the preheated comburent temperature. Since a suction pyrometer was used, we prescribe a 30 "C measurement accuracy for temperatures up to 1700 "C. The predictions showed both the uniformity in the temperature field in a large part of the furnace and a rather low peak in-furnace temperature, bearing in mind the high level of preheat. All three tested combustion models showed inaccuracies in the predictions of the fuel jet region, see Figure 8. The pdf and EDC models largely underpredicted the temperature at the first three measured traverses in the fuel jet region. The chemical equilibrium assumption is likely to be incorrect in this flame region. The endothermic reactions prevail producing a very low equilibrium temperature that is even lower than a non-reacting mixing temperature (De Joannon et ul., 2000). The equilibrium models were not able to reproduce the initial stages of the flame. The eddy-break-up model showed better predictions at

S.ORSINO ET AL.

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

TEMPERATURE (K)

Figure 8. Predicted and measured temperature

the third traverse, but at the first two traverses also, too low temperatures were predicted in the natural gas jet region. The above-discussed underprediction of the entrainment is likely to be the cause for the low temperatures at the first traverses in the eddy-break-up model predictions. As we have already discussed, a higher entrainment of the hot flue gas into the fuel jet could make more oxygen available for the combustion and would also increase the temperature by in-mixing a fluid of 1200 "C. It is believed that a higher entrainment into the fuel jet would not improve significantly the predictions using an equilibrium approach. Additional calculations were performed using the CEA code (Gordon and McBride, 1994) assuming an equilibrium in a well stirred reactor under two different conditions:

1. The initial not-reacting composition was based on the entrainment predicted by the equilibrium-based models, 2. The initial not-reacting composition was based on the measured entrainment.
In both cases, it was assumed that the starting temperature was 1473K (the temperature of the jet's surroundings). Table 2 reports the

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS Table 2. Temperature and CO at equilibrium assuming two dilTerent entrainments Composition based on measured entrainment Mixing temperature K Equilibrium temperature K Equilibrium CO mass% 1473 1063 17.83

27

Composition based on the predicted entrainment

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

equilibrium temperature and CO concentration for a perfectly stirred reactor with the measured and predicted entrainment (at the second measurement traverse). In both cases, the equilibrium temperature was lower than the starting temperature. Furthermore, increasing the entrainment resulted in only a moderate increase of the final equilibrium temperature. Even if more O2 had been available, the conditions would have remained understoichiometric with an equilibrium temperature lower than the mixing temperature. In the mathematical model, conjugate boundary conditions on the walls were used. The calculated wall temperatures were in a very good agreement with the measurements, see Figure 9. A radiative heat flux as high as 300-350 k w / m 2 was measured along the whole furnace. This value was substantially higher if compared with conventional natural gas combustion. The flatness of the heat flux is considered as one of the most

LEFT AXIS R A O I A T I M H E A T FLUX

RI6HT A X I S WALL TEMPERATURE

Figure 9. Predicted and measured radiative heat Rux at the furnace wall

28

S. ORSINO ET AL.

important characteristics of the HTAC combustion technique. This behavior was well represented in the predictions.
Emissions

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

The same NOx model was used with the EDC and EBU combustion model. Thus, the differences in the NOx predictions between these two cases are solely due to differences in the temperature and chemistry fields. Table 3 reports the comparison of the measured and predicted flue gas quantities for the three combustion models. The three cases showed predictions of a good quality in the flue gas properties. The outlet NOx emissions were predicted very well by both NOx models. Figure 10 shows the measured and predicted in-furnace NOx concentrations. In the experiments, no region of high N O formation was detected. The peak in the NOx concentration was a t the boundary of the air and natural gas jets where the highest temperatures was both measured and predicted. The predictions were able to capture well this behavior. The NOx concentration showed a peak between the air jet and the fuel jet at the two first traverses, then the profiles were flat and the NOx concentrations remained unaltered. Using the EBU and E D C combustion models, the influence of the NOx reburning mechanism was investigated. The NOx-reduction rates were 4-5 orders of magnitude lower than the NOx formation rates. The prompt and thermal NOx formations were investigated separately, and 95% of the NOx were formed through the thermal NOx path. Whilst comparing the measurements and the predictions, it should be considered that a large part of the outlet NOx were coming from the comburent that contained I I0 ppmvd NOx. Table 4 summarized the net NOx formation in the furnace. The net NOx formations were calculated

Table 3. Predicted and measured flue gas quanti~ies Measured Temperature ( K ) Oxygcn (%vol-dry) Carbon Dioxide (%vol-dry) CO ppmvd NOx (ppm-dry) 1493 2.2 20.5 0 140 IFRF-EBU 1522 2.5 20.5
0

FluentS-pdfl 1 1531 2.5 20.2 0 133

IFRF-EDC 1521 2.5 20.5 0 132

137

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

NIlRO6EN OXTDES (ppm, dry)

-1WF-EBU IWF-EDC

....-... flunrbpdf
m d

Figure 10. predicted and measured NOx concentration.

Table 4. Net NOx formation

Measurements NOx net for~nationkg/h


0.01048

EBU-combuslion model
0.00946

EDC combustion model


0.00616

PdffMixture fraction combustion model


-

0.00685

as the difference between the nitric oxides at the inlet and outlet. In all three simulations, the net-NOx formation in the furnace was underpredicted. With the EBU combustion model, the difference was only around lo%, in the other two cases (EDC and pdf/mixture fraction) the difference was higher, above 30%. These results may suggest that the NOx models tends to underestimate the NOx formation under the conditions (temperature and 0 2 ) typical of the high temperature air combustion technique. The Zeldovich mechanism can provide accurate evaluation of the NOx in a temperature region above 1800 K, but its accuracy a t lower temperatures is lower. In the case studied in the present paper. the temperature was below 1800 K in a large part of the furnace.

30

S. ORSINO ET AL.

CONCLUSION

A numerical simulation of a natural gas combustion with air preheated to 1300Chas been carried out. Three combustion models have been tested:
eddy-break-up model with a two-step reaction scheme, eddy dissipation concept model with chemical equilibrium, pdflmixture fraction model with equilibrium, non-adiabatic look up tables. These models are relatively simple turbulent combustion models and they have been already extensively validated for conventional combustion systems. The nitric oxide formation has been investigated with two different NOx models, both considering thermal and prompt NOx formation and one of them has included the NOx reburning mechanism. The model results have been compared with both the in-furnace measurements (including temperature, gaseous species, velocity and radiative heat flux) and with the furnace exit parameters. The following has been concluded: (a) the numerical models have correctly reproduced the characteristics of the high temperature air combustion, namely the uniformity of the temperature field, high radiative fluxes and low NOx and C O emissions. Since these characteristics are of paramount importance in engineering applications, this is indeed a highly encouraging observation. All three mathematical models have provided the predictions of similar quality, (b) with the exception of the small region located within the natural gas jet (see conclusion (c)), the chemical equilibrium assumption with respect to both the natural gas combustion and NO formation (equilibrium 0 , O H a n N radicals concentrations) have resulted in predictions of very good quality. The eddy dissipation concept model and the pdflmixture fraction model when coupled with chemical equilibrium have provided almost identical results, (c) the tested combustion models cannot describe the chemistry and temperature field in the fuel jet region. This is partially caused by imperfections in predicting the entrainment of the weak methane jet interacting with the strong combustion air stream. The two-step (eddy-break-up) model is too simple to describe the jet chemistry. However, more sophisticated chemical equilibrium models (pdf,

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

31

EDC) d o n o t offer any improvements since chemical equilibrium does


not prevail in this flame region. A more comprehensive understanding o f the chemistry o f natural gas combustion under fuel-rich conditions with comburent containing >4% oxygen is required in order t o develop appropriate non-equilibrium sub-models, (d) the predictions o f the flue gas N O x emissions a r e in good agreement with the measurements. Since the natural gas jet constitutes only a small fraction o f the furnace volume a n d relatively little NOx is formed within the jet, the imperfections in the predicted properties of the natural gas jet d o not affect the N O x predictions within the furnace a n d a t its exit. T h e study o n the N O x emission has shown that the thermal N O x formation is largely predominant a n d the NOx reburning mechanisms is o f little importance under the investigated conditions.

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

REFERENCES
Baulch, D. L. et al. 1992. Evaluated kinetic data for combustion modeling. Journal of Physical and Chemical Reference Data 2 l(3). Blasiak, W., and B. Lindblad. 2000. Highly preheated air combustion research in Sweden. In Proceedings of the 2nd International Seminar on High Temperature Combustion in Industrial Furnaces. Stockholm, Sweden: Jernkontoret-KTH. Bowman. C. T. 1975. Kinetics of pollutant formation and destruction in combustion. Prog. Energy Comhust. Sci. 1:3%47. Breussin, F . , N. Lallemant, and R. Weber. 2000. Computing of oxy-natural gas flames using both a global combustion scheme and a chemical equilibrium procedure. Combust. Sci, and Tech. 160:36%398. Brink, A,, M . Hupa, F. Breussin, N. Lallemant, and R. Weber. 2000. Modelling of oxy-natural gas combustion chemistry. AIAA Journal of Propulsion and Power 16(4):60%6 14. Cavaliere, A., and M. De Joannon. 2000. Detailed chemical kinetics in the reactor design for diluted high temperature combustion of air/paraffin mixtures. The 3rd CREST International Symposium, Yokohama, Japan. Chen, W. 1994. A global reaction rate for nitric oxide reburning. PI1.D. thesis, Bringham Young University. Coelho, P. J., and N. Peters. 2001. Numerical simulation of a mild combustion burner. Combustion and Flame 124:50&518. Dugue J., and R. Weber. 1992. Laser velocimetry in semi-industrial natural gas, oil and coal flames by means of a water-cooled LDV probe. The 6th International Symposium on Application of Laser in Anemometry in Fluid Mechanics, Lisbon, Portugal.

32

S. ORSINO ET AL.

De Joannon, M., A. Saponaro, and A. Cavaliere. 2000.Zero-dimensional analysis of methane diluted oxidation in rich conditions. Proceedings o f the Comhusrion Institute, 28, pp. 163%1646. De Soete, G . G. 1975. Overall reaction rates of N O and N2 formation from fuel nitrogen, Proceedings o f the Comhustion Insritute 15:1 0 9 s 1102. Dong, W., and W. Blasiak. 2000. Numerical modeling of highly preheated air combustion in a 580 kW testing furnace at IFRF. Tile 3rd C R E S T International Symposium, Yokohama, Japan. Edwards, D. K. 1976. Molecular gas band radiation. In Advances in Heat Trunsfer (Edited by T . F . Irvine, Jr. and J.P. Harnett). 12, pp. 115-193. New York: Academic Press. Gordon, S., and S. McBride. 1994. Computer program for calculations of complex equilibrium compositions and applications, analysis. N A S A Reference Publication 131 1 . Gupta, A. K.. and 2. Li. 1997. Effect of fuel property on the structure of highly preheated air flames. A S M E International Joint Power Generation Conference, Denver CO, ASME EC-Vol. 5, pp 247-257. Gupta, A. K., S. Bolz., and T. Hasegawa. 1999. Effect of air preheat temperature and oxygen concentration on flame structure and emissions. A S M E Journal of Energy Resources Technology 12 1 :20%216. Hanson, R. K., and S. Salimian. 1984. Survey of rate constants in H / N / O Systems. Comhustion Chemistry 36 1. Hardesty, D. R., and F. J. Weinberg. 1974. Burners producing large excess enthalpies. Combust. Sci. Technol. 8:201-214. Hino Y., C. Zhang, T. Ishii, and S. Sugiyama. 1998. Comparison of measurements and predictions of flame structure and NOx emissions in a gas-fired furnace. A I A A I A S M E Joint Thermophysics and Heut Transfer Conference, 1. Hottel. H. C., and A. F. Sarofim. 1967. Radiative Transfer. New York: McGrawHill Company. Ishi, T., C. Zhang, and S. Sugiyama. 1997. Numerical analysis of NOx formation rate in a regenerative furnace. Joint Power Generation Conference, EC-Vol. 5, pp. 267-278. Ishiguro, T., S. Tsuge, T. Furuhata, K. Kitigawa, N. Aral, T. Hasegawa, R. Tanaka, and A. K. Gupta. 1998. Homogenization and stabilization during combustion of hydrocarbons w ~ t hpreheated air. Proceedings o f the Com27, pp. 320S32 13. hustion Instir~rre, Katsuki, I . M., and T. Hasegawa. 1998. The science and technology of combustion in highly preheated air. Proceedings o f the Combustion Institute, 27, pp. 3135-3146. Lallemant, N.. and R. Weber. 1996. A computationally efficient procedure for calculating gas radiative properties using the exponential wide band model. 1nt. J . Hear Mass TransJer 39(15):327>3286.

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

NUMERICAL SIMULATION OF COMBUSTION OF NATURAL GAS

33

Magnussen, B. F. 1981. On the structure of turbulence and a generalized eddy dissipation concept for chemical reaction in turbulent flow. Nineteenth AIAA Aerospace Meeting, AIAA, New York. Magnussen, B. F., and B.H. Hjertager. 1976. On mathematical modeling of turbulent combustion with emphasis on soot formation and combustion. Proceedings of the Combustion Institute 17:719-729. Masters, J., R. J. Webb, and R. M. Davies. 1979. The use of modelling techniques in the design and application of recuperative burners. J. Inst. Energy 52: 196204. Milani, A. 2000. Mild combustion technique applied to regenerative firing in industrial furnaces. Proceedings of the 2nd International Seminar on High Temperature Combustion in Industrial Furnaces, Stockholm, Sweden: Jernkontoret-KTH. Milani, A., G . V. Salamone, and J. G. Wiinning, 1997. Low NOx- Regenerativbrenner in einer Anlage zum kontinuierlichen Gluhen von Edelstahlband. Gaswurme International, 46(12):60&612. Missaghi, M., M. Pourkashanian, A. Williams, and L. Yap. 1990. The prediction of NO emissions from an industrial burner. Proceeding of American Flame Days Cotfererrce, San Francisco. Mochida, S., and T. Hasegawa. 2000. Development of a combustion diagnostics method on advanced industrial furnaces utilizing high temperature air combustion. Proceedings of the T" International Seminar on High Temperature Combustion in Industrial Furnaces, Stockholm, Sweden: Jernkontoret-KTH. Nakamachi, I., K. Yasuzawa, T. Miyahata, and T. Nagata. 1990. Apparatus or method for carrying out combustion in a furnace. US Patent No. 4,945, 841. Peters, A,, and R. Weber. 1995. Mathematical modeling of a 2.25 MW, swirling natural gas flame. Part 1: Eddy break-up concept for turbulent combustion; Probability density function approach for nitric oxide formation. Combust. Sci. and Tech. 11&111:67-101. Peters, N. 1986. Laminar flamelet concepts in turbulent combustion. Proceedings of the Contbustion Institute 21: 123 1- 1250. Plessing. T., N. Peters, and J. G . Wunning. 1998. Laseroptical investigation of highly preheated combustion with strong exhaust gas recirculation. Proceedings of the Combustion Institute 27:3197-3204. Pope, S. B. 1990. Computations of turbulent combustion: progress and challenges. Proceedings of the Combustion Institute 23:591-612. Suzukawa, Y., S. Sugiyama, and T. Hasegawa. 1998. Development and npplication of direct fired regenerative burner heating system. The IFRF American-Japa~reseFlume Days, Maui, Hawaii. Tanaka, R., K. Kishirnoto, and T. Hasegawa. 1979. High efficiency heat transfer method with use of high temperature pre-heated air and gas recirculation. Combust. Sci. Technology 1(4):264-273, in Japanese.

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

34

S . ORSINO ET AL.

Downloaded by [Mount Allison University 0Libraries] at 13:29 21 May 2013

Verlaan, A. L., S. Orsino, N. Lallemant, and R. Weber. 1998. Fluidflow and mixing in a furnace equipped with the low-NOx regenerative burner of Nippon Furnace Kogyo. The results of the HTAC97 trials, I F R F Doc. No. F 46/y/l. Weber, R., S. Orsino, N. Lallemant, and A. Verlaan. 2000. Combustion of natural gas with high temperature air and large quantities of flue gas. Proceedings of the Combustion Institute 28, pp. 131 s 1321. Weber, R., A. L. Verlaan, S. Orsino, and N. Lallemant. 1999. On emerging furnace design methodology that provides substantial energy savings and drastic reductions in COz, C O and NOx emissions. J. Inst. Energy 72:77-83. Wesbrook, C. K., and F. L. Dryer. 1984. Chemical kinetic modeling of bydrocarbon combustion. Prog. Energy Comb. Sci. 1-57. Westenberg, A. A. 1975. Kinetics of NO and C O in Lean, Premixed Hydrocarbon-Air Flames. Comb. Sci. and Tech. 4:5%64. Wunning, J. A,, and J. G. Wunning. 1992. Brenner fur flammenlose Oxidation mit geringer NO-Bildung auch bei hochster Luftvorwarming. Gaswarme International 10:438. Wunning, J . A., and J. G. Wunning. 1997. Flameless oxidation to reduce thermal NO formation. Progr. Energy Combust. Science 23(12): 81-94. Yuang, J., and I. Naruse. 1998. Numerical study on combustion characteristics and NOx emissions in highly preheated and diluted air combustion. The IFRF American-Japanese Flame Days, Maui, Hawaii.

Potrebbero piacerti anche