the time dependent investigation of methane-air

10
Case Studies in Thermal Engineering 18 (2020) 100603 Available online 8 February 2020 2214-157X/© 2020 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/). The time dependent investigation of methane-air counterflow diffusion flames with detailed kinetic and pollutant effects into a micro/macro open channel Ali Edalati-nejad a , Sayyed Aboozar Fanaee a , Maryam Ghodrat b, * , Fatemeh Salehi c , Javad Khadem a a Department of Mechanical Engineering, University of Birjand, Birjand, Iran b School of Engineering and Information Technology, University of New South Wales Canberra, Canberra 2610 ACT, Australia c School of Engineering, Macquarie University, Australia A R T I C L E INFO Keywords: Counterflow diffusion flame Detailed mechanisms Time-dependent model Micro/macro open channel ABSTRACT In the present study, time-dependent numerical analysis of methane-air counterflow diffusion flame into a selected macro/micro open channel is investigated. The flame is simulated by multi and single-step reaction approaches into an open channel with a constant distance of 15 mm between air and the fuel inlet, and a hydraulic distance at the order of 0.1 mm. To solve the unsteady problem, a coupled pressure-velocity implicit division method is considered. The results show an acceptable agreement between numerical and experimental data that confirm the ac- curacy of the model. The results also revealed that the variation of the residence time to the inlet velocity is more sensitive than the inlet temperature. It is also found that at the larger inlet ve- locities, the flame is stabilized at a smaller value of hydraulic distance. This is a result of increasing the possibility of reactions between species. The generation rates of CO 2 , CO and H 2 O species are found to be nearly constant at t > 0.009s while for NO and NO 2 species the rates remain unchanged at t > 0.013s and t > 0.016s, respectively. 1. Introduction Recently the interest in the combustion of low-carbon hydrocarbons has grown rapidly, especially in power generation using methane due to its advantages over other common fuels [1]. Combustion systems are categorized as premixed and diffusion flame burners. Although premixed flames come with advantages such as better heat transfer and cleaner flames, there are some important concerns about their ignition stability and safety [2]. These issues can be solved by using diffusion flames in power and propulsion systems [3]. Diffusion flames are divided into two main categories of co-flow and counterflow structures. There are various experi- mental and simulation studies on diffusion flames. Yamaoka et al. [4] investigated the extinction limits of lean methane-air coun- terflow diffusion flames and showed an increasing trend for the flame strength with the increase of equivalence ratio in the high temperature zone. Yoshida et al. [5] presented the methane/air counterflow diffusion flame by fine water droplets and demonstrated that by increasing the density of water droplets, the maximum values of flame temperature decreased, confirming the maximum flame temperature was related to the scale of water droplets. Padilla et al. [6] developed a model for extinction limits and structure of * Corresponding author. E-mail address: [email protected] (M. Ghodrat). Contents lists available at ScienceDirect Case Studies in Thermal Engineering journal homepage: http://www.elsevier.com/locate/csite https://doi.org/10.1016/j.csite.2020.100603 Received 27 January 2020; Received in revised form 5 February 2020; Accepted 7 February 2020

Upload: others

Post on 28-May-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The time dependent investigation of methane-air

Case Studies in Thermal Engineering 18 (2020) 100603

Available online 8 February 20202214-157X/© 2020 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license(http://creativecommons.org/licenses/by-nc-nd/4.0/).

The time dependent investigation of methane-air counterflow diffusion flames with detailed kinetic and pollutant effects into a micro/macro open channel

Ali Edalati-nejad a, Sayyed Aboozar Fanaee a, Maryam Ghodrat b,*, Fatemeh Salehi c, Javad Khadem a

a Department of Mechanical Engineering, University of Birjand, Birjand, Iran b School of Engineering and Information Technology, University of New South Wales Canberra, Canberra 2610 ACT, Australia c School of Engineering, Macquarie University, Australia

A R T I C L E I N F O

Keywords: Counterflow diffusion flame Detailed mechanisms Time-dependent model Micro/macro open channel

A B S T R A C T

In the present study, time-dependent numerical analysis of methane-air counterflow diffusion flame into a selected macro/micro open channel is investigated. The flame is simulated by multi and single-step reaction approaches into an open channel with a constant distance of 15 mm between air and the fuel inlet, and a hydraulic distance at the order of 0.1 mm. To solve the unsteady problem, a coupled pressure-velocity implicit division method is considered. The results show an acceptable agreement between numerical and experimental data that confirm the ac-curacy of the model. The results also revealed that the variation of the residence time to the inlet velocity is more sensitive than the inlet temperature. It is also found that at the larger inlet ve-locities, the flame is stabilized at a smaller value of hydraulic distance. This is a result of increasing the possibility of reactions between species. The generation rates of CO2, CO and H2O species are found to be nearly constant at t > 0.009s while for NO and NO2 species the rates remain unchanged at t > 0.013s and t > 0.016s, respectively.

1. Introduction

Recently the interest in the combustion of low-carbon hydrocarbons has grown rapidly, especially in power generation using methane due to its advantages over other common fuels [1]. Combustion systems are categorized as premixed and diffusion flame burners. Although premixed flames come with advantages such as better heat transfer and cleaner flames, there are some important concerns about their ignition stability and safety [2]. These issues can be solved by using diffusion flames in power and propulsion systems [3]. Diffusion flames are divided into two main categories of co-flow and counterflow structures. There are various experi-mental and simulation studies on diffusion flames. Yamaoka et al. [4] investigated the extinction limits of lean methane-air coun-terflow diffusion flames and showed an increasing trend for the flame strength with the increase of equivalence ratio in the high temperature zone. Yoshida et al. [5] presented the methane/air counterflow diffusion flame by fine water droplets and demonstrated that by increasing the density of water droplets, the maximum values of flame temperature decreased, confirming the maximum flame temperature was related to the scale of water droplets. Padilla et al. [6] developed a model for extinction limits and structure of

* Corresponding author. E-mail address: [email protected] (M. Ghodrat).

Contents lists available at ScienceDirect

Case Studies in Thermal Engineering

journal homepage: http://www.elsevier.com/locate/csite

https://doi.org/10.1016/j.csite.2020.100603 Received 27 January 2020; Received in revised form 5 February 2020; Accepted 7 February 2020

Page 2: The time dependent investigation of methane-air

Case Studies in Thermal Engineering 18 (2020) 100603

2

counterflow non-premixed H2O-laden CH4/air flames and demonstrated that increasing strain rate led to an increase in the maximum flame temperature at the extinction limit. Chen et al. [7] investigated entropy generation in hydrogen-added ultra-lean methane/air counterflow diffusion flame and showed that with increasing fuel injection into combustion chamber heat transfer decreased irre-versibility in contrary with premixed combustion. Fuse et al. [8] analysed the generation of NO and NO2 pollutants from high-temperature methane/air counterflow diffusion flame and showed that NO and NO2 concentration cannot be measured sepa-rately with selective gas sampling systems, but the measurements of the total amounts of NOx (¼NO þ NO2) were sufficiently consistent with the predicted values. Sohn et al. [9] carried out numerical modeling and investigated the effects of air-dilution on NOx formation for methane/air counterflow diffusion flames with a high temperature preheated air. Their results demonstrated that by decreasing NOx emission the average values of pressure increased. This showed that an acceptable combination of air-dilution and pressure can control NOx emission. Sung et al. [10] presented the structure and pollutant emission of methane/air and propane/air counterflow diffusion flames with detailed kinetic reactions and concluded that the mole fraction of fuels is reduced with increasing the pressure of the chamber. Niemann et al. [11] developed a model to investigate the effect of pressure on methane/air counterflow diffusion flame and showed that pressure has a significant effect on critical quenching conditions for gaseous fuels. Glassman and Yaccarino [12] analysed the variation of flame temperature into diffusion flames and found that increasing flame temperature has an effect on soot generation into diffusion flames. Gülder and Snelling [13] studied the influences of preheating on the soot generation of co-flow diffusion flames and concluded that volume fractions of soot increase with increasing temperature. In Ref. [14] Axelbaum and Law investigated soot formation and inert addition in diffusion flames and showed that fuel concentration and flame temperature are related to pollutant emission of these flames and effects of the fuel concentration are more significant than flame temperature. Lim et al. [15] investigated the effect of air preheat on the structure of methane - air counterflow diffusion flames and revealed that at the temperature range of 300K–560K, CH4, CO2, O2, and N2 mole fractions decrease as a function of equivalence ratio. Puri et al. [16] studied the numerical and experimental modeling of counterflow methane/air diffusion flame and concluded that with increasing strain rate, the temperature is massively reduced. Smooke et al. [17] performed a comparison between experimental data and nu-merical calculations of diluted methane/air counterflow diffusion flame and showed a satisfactory agreement between all species data except H2 and CO. As claimed by the authors, these observed disagreements were ascribed to neglecting some species at the numerical calculations. In Ref. [18], the accurate parametric studies of micro-combustor with an analytical solution were carried out. A simple micro-sized channel was developed consisting of two parallel plates where premixed fuel and air entered the channel. It was concluded that with a non-reactive flow (similar to the preheat zone of a micro-combustor), non-dimensional temperature increased while the Peclet number reduced. In Refs. [19,20], the modeling of two-dimensional catalytic micro-combustors was investigated, considering a constant wall temperature. The results presented a calculation tool for exploring variations of mass and temperature distribution, flame characteristics and quenching distance into a micro-combustor. Hosseini and Wahid [21] studied the characteristics of lean premixed and non-premixed micro-flameless combustion of methane/air. In their work, the effect of bluff-body on temperature dis-tribution was simulated, revealing that the peak temperature of premixed conventional micro combustion was much higher (about 700K) than micro-flameless combustion value. However, the temperature of exhausting gases in conventional mode was lower (about 200K) than the micro-flameless regime. In Ref. [22], CFD modeling was adopted to analyze the flame stability of heat recirculation micro combustors, where inner and outer reactor recirculation were compared. The results showed that the inner reactor had a wider range of flame stability than the outer. Bagheri et al. [23] conducted a study on the combustion characteristics and flame stability in micro-scale combustion under various physical and chemical circumstances. They demonstrated that by increasing inlet velocity from 10 m/s to 20 m/s, the flame temperature raised. In the work of Fanaee [24], the modeling of a micro combustor with the stationary flame was studied. In his paper, Fanaee compared multi- and single-reaction combustion and calculations of noncatalytic and catalytic conditions. In a more recent study, Fanaee [25] presented wall temperature effects on the performance of finite-length homogenous micro-combustor using a two-dimensional model. The results demonstrated that the pathway of reacting flow has a considerable effect on the flame structure. Padilla et al. [26], carried out a detailed study on the structure of methane/air counterflow diffusion flames using a one-dimensional counterflow model combined with the GRI 3.0 mechanism and found that by increasing the amount of water, the OH intensity and reaction zone thickness decrease. In Ref. [27], Lee and his co-authors conducted a study on NOx emissions of counterflow non-premixed water-laden methane/air flames and showed that with increasing strain rates, NOx emissions in H2O-laden flames reduced. Most recently, Edalati-nejad et al. [28] conducted an unsteady investigation on counterflow flames into the proposed plus-shaped channel over a palladium catalyst. Their results presented at both non-catalytic and catalytic conditions considering the effect of reaction time and different equivalence ratios on flame characteristics and pollutant emissions. The results of these researchers showed that at lean flames the increment of equivalence ratio increased the CO2 mass fraction. Bidabadi and Esmaeilnejad [29] developed an analytical model for predicting counterflow flame propagation and analysed the effects of radiative heat loss and flow strain. They showed that the burning velocity of the counterflow flame increased as a function of non-dimensional vaporization temperature. Some researchers such as Yu et al. [30] also investigated the effects of hydrogen addition on the flame structure of premixed methane/air combustion and found that the flame front speed increased significantly with the hydrogen fraction. Delic [31] carried out a numerical simulation of combustion in a jet engine and used ReactingFoam solver of OpenFOAM software for steady-state condition simulation. Salehi et al. [32,33] used OpenFoam software in order to simulate the reacting flows. Busseti [34] conducted an experimental and numerical investigation on the propagation of compositional waves where the numerical simulation was performed using OpenFOAM.

In the aforementioned studies, the time-dependent flame characteristics and its structure have not been investigated. Furthermore, the effect of variation in hydraulic distance from micro to macro scale has not been studied. Hence, the main aim of this paper is conducting a time-dependent study of methane-air counterflow diffusion flame where the flame conditions, pollutants emission, and stability parameters are calculated within a rectangular micro/macro open channel. Furthermore, the effects of single and detailed

A. Edalati-nejad et al.

Page 3: The time dependent investigation of methane-air

Case Studies in Thermal Engineering 18 (2020) 100603

3

kinetic reactions are studied.

2. Methodology

In this section, the geometric properties and physical models are described and then, governing equations and mathematical models are presented.

2.1. Case set-up

In the presented study, the counterflow diffusion flame is passed through a rectangular open channel with different values of hydraulic distance at the macro and micro orders of 20.1, 10, 5 and 1 mm. The distance between air and fuel inlets is fixed at 15 mm. The origin of coordinate is considered at the fuel inlet and the depth of the rectangular open channel is equal to 1 mm for the two- dimensional simulation. The fuel and air velocity profiles at the inlet of the open channels are uniform. The initial internal field temperature is considered to be 1800K.

For model validation, the hydraulic distance of the open channel is set to be 20.1 mm and the distance between fuel and the air inlet is equal to 15 mm. The inlet temperature of fuel and air are respectively equal to 300K and 560K while their inlet velocities are respectively 0.7 m/s and 0.52 m/s. To further analyze the inlet velocity of air and fuel are considered from 0.001 m/s to 0.2 m/s and inlet temperature 300K and 800K. In order to study the micro and macro scaled geometry and find quenching distance parameter, hydraulic distance is considered from less than 0.4 mm–10 mm.

2.2. Numerical method

Due to the low range of Reynolds number, the Reynolds-Averaged Simulation (RAS) turbulent model and thin flame are selected here. The CFD solver is ReactingFOAM which is part of the OpenFOAM 4.1 platform. OpenFOAM is an open-source object-oriented cþþ software platform which has been widely used for reacting flows [32,33]. It has custom-written solvers that provide easy access to model extension/modification. Different solvers of OpenFOAM including ReactingFoam have been benchmarked with various experimental fire-related data [34]. ReactingFoam has also benchmarked with an analytical solution as well as a real Laser-Induced Thermal Grating Spectroscopy (LITGS) technique [34].

ReactingFoam solves the Favre filtered fully compressible Navier–Stokes equations [31] to capture turbulent structures of the flow. The k-ε turbulent model is adopted here. The governing equations consist of continuity, momentum, energy, state, and species transport expressions that are stated as below:

∂ρ∂tþ

∂∂xj

�ρeuj�¼ 0 (1)

∂∂tðρeuiÞþ

∂∂xj

�ρeuieuj

�¼ �

∂p∂xiþ

∂∂xj

2μeff

eSij �13

δijfSkk

��

(2)

∂∂tðρfYkkÞþ

∂∂xj

�ρeuj eYk

�¼

∂∂xj

μeff∂ eYk

∂xj

þ _ωk ; k¼ 1;… ;N (3)

∂∂tðρehsÞþ

∂∂xj

�ρeuj ehs

�¼

DpDtþ

∂∂xj

ραeff∂ehs

∂xj

þ _ωT (4)

where ui is the i-direction component of velocity, p is the pressure, ρ is the density, Yk shows the mass fraction of k-th species, N is the number of species and hs is the sensible enthalpy. μeff and αeff are the effective viscosity (summation of average dynamic and oscillation viscosity) and the effective thermal diffusivity (summation of molecular thermal diffusivity and oscillation), respectively, and δij is Kronecker delta function. The rate of the strain tensor, Sij, is defined as follows:

eSij¼12

�∂euj

∂xiþ

∂eui

∂xj

(5)

_ωk and _ωT are the average rate of species and combustion reactions. The heterogeneous and homogeneous reaction rates are calculated using:

kT ¼A Tβ exp�

�Ea

RT

(6)

where A, β, Ea and R, are respectively as the pre-exponential factor, the temperature exponent, the activation energy and the ideal gas constant. The residence time is defined to determine the transferring rate of mass species relative to flame front speed. This parameter is obtained using:

A. Edalati-nejad et al.

Page 4: The time dependent investigation of methane-air

Case Studies in Thermal Engineering 18 (2020) 100603

4

tres¼PV_m RT

(7)

where P, V, _m, and R, are the channel pressure, the channel volume, the mass flow rate and the gas constant, respectively. Besides, T is calculated for the average temperature between the inlet and maximum flame temperature using air properties.

A coupled pressure-velocity implicit division of operator, Piso, and semi-implicit method for pressure-related equations are adopted.

The solution is unsteady and, the convergence criteria for physical parameters are considered to be 10� 6 approximately. The GRI0.3 mechanism is employed which includes 324 reactions and 53 species for the CH4-air mixture (see Fig. 1).

3. Result and discussion

3.1. Validation and grid sensitivity study

A grid sensitivity analysis is first conducted using three grid sizes and structure distribution for the open channel case with inlet temperatures 300K and inlet velocities 0.05 m/s. The three meshes considered in this study include 1500 (60ˣ25), 4000(100ˣ40) and 6500 (130ˣ50) cells where the girds are refined near the flame zone. Fig. 2-A compares the temperature variation of the open channel for the considered grid sizes. The temperature variation is plotted along the length of the open channel (0.008 < x < 0.011). As can be seen, by increasing the number of grids, the maximum temperature initially decreases while with a further increase, the temperature marginally changes, confirming the independence of the grid to the solution and hence the second mesh with 4000 cells is adopted for all simulations.

Fig. 2-B compares the variation of calculated mole fraction of species N2, O2, and CH4 with the experimental data [15]. The nu-merical results show an acceptable agreement with the experimental measurements where the trends of the numerical calculation are aligned with the experimental results. Although the predicted N2, O2 and CH4 are slightly ahead of the measurement, the average difference between the simulation and the measurements in open channel for CH4, N2 and O2 is 8.2%, 6.4% and 4.1%, respectively.

3.2. Time scaled analyses

The mass fraction generation for H2O, NO2, CO and CO2 species at the inlet temperature of 800K, inlet velocity of 0.05 m/s and hydraulic distance of 1 mm are presented in Fig. 3-A The generation rate of CO2, CO and H2O species become nearly constant at t >0.009s while for NO and NO2 species, it takes longer times until they reach to their steady-state values - 0.013s and 0.016s, respec-tively. This is mainly due to the relatively lower mass diffusivity of NO species comparing to CO2, CO and H2O species.

Fig. 3-B shows the flame temperature against the reaction time for cases with Vi ¼ 0.05 m/s, dh ¼ 1 mm and the inlet temperatures of 300K and 800K. At both inlet temperatures, the flame temperature initially rises and then reaches almost a constant value. For the inlet temperature of 300K, there is a sharp rise at 0.015s, and then the temperature reaches a constant value at t ¼ 0.017s. The main reason for this trend is due to the change in equilibrium rates of reactive species before the finalized steady-state condition. As the inlet temperature increases to 800K, the finalized steady-state equilibrium occurs at t ¼ 0.01s. This is due to an increase in the energy content of the flame, and hence, increasing the equilibrium rates of the species into the open channel.

3.3. Steady-state analyses

The steady-state analyses are adapted from continuing the time-dependent solution until a stabilized form of the flame is achieved. Fig. 4-A shows the temperature contour of the counterflow diffusion flame with inlet temperature of fuel and air 300K and 560K into the conventional-sized channel. The stabilized flame is situated in a closer distance to the oxidizer side, because of larger values of mass

Fig. 1. The schematic modeling of counterflow diffusion flame pass through the open channel.

A. Edalati-nejad et al.

Page 5: The time dependent investigation of methane-air

Case Studies in Thermal Engineering 18 (2020) 100603

5

diffusivity of the fuel compared to that of the oxidizer species. The flame thickness is nearly 0.75 mm and the ratio of air/fuel mass flow rates is 2.73. The observed trend shows that, for the flame position, the effect of the combustion reaction rate is mainly governed by air/fuel mass flow rate.

Fig. 4-B presents the velocity contour of the diffusion flame in the case with inlet temperature of fuel and air 300K and 560K, inlet velocities 0.7 m/s and 0.52 m/s, respectively. Considering the couture plot of the velocity in the axial direction, it can be seen that the

Fig. 2. A) Grid independence for presented solution considering 1500, 4000 and 6500 grid points- B) Comparison between mole fractions of species for the presented numerical results and experimental data [15].

Fig. 3. A) Mass fraction generation for H2O, CO, NO, NO2 and CO2 species versus the period of reaction time, for Ti ¼ 800K Vi ¼ 0.05 m/s and dh ¼

1 mm - B) Flame temperature versus reaction time, for Vi ¼ 0.05 m/s, dh ¼ 1 mm and two inlet temperatures of 300K and 800K.

A. Edalati-nejad et al.

Page 6: The time dependent investigation of methane-air

Case Studies in Thermal Engineering 18 (2020) 100603

6

fuel penetrates faster than the oxidizer because of the larger values of fuel molecular mass relative to the oxidizer. The velocity magnitude of the flow is maximized at the outlet of the open channel where the flow progress is affected by the combustion reaction. This is mainly due to the acceleration of the fuel and air reacting flow which occurs because of the energy release of these reactions.

Fig. 5-A illustrates the temperature variations of counterflow diffusion flame along the axial direction of the open channel, from fuel inlet, for three different fuel and air inlet velocities at inlet temperature 800K and hydraulic distance of 1 mm. By reducing the inlet

Fig. 4. A) The temperature contour of counterflow diffusion flame into a conventional-sized channel - B)The velocity contour of counterflow diffusion flame into a conventional-sized channel.

Fig. 5. A) The flame temperature variations of counterflow diffusion flame versus distance from the fuel inlet – B) the flame temperature versus distance from fuel inlet, at the inlet temperature of 800K, inlet velocities of 0.05 m/s and hydraulic distance of 1 mm – C) Mass fraction of species versus distance from the fuel inlet, for multi and single-step reactions, where inlet temperatures of fuel and air equal to 800K, inlet velocity is 0.05 m/s and dh ¼ 5 mm – D) Maximum temperature versus hydraulic distance for fuel and air inlet temperatures of 300K and 800K and inlet velocities of 0.05 m/s.

A. Edalati-nejad et al.

Page 7: The time dependent investigation of methane-air

Case Studies in Thermal Engineering 18 (2020) 100603

7

velocity, the flame is shifted to a closer distance to the fuel inlet due to a decrease in the mixing rate. This phenomenon is compensated by forming a more combustible reacting flow in the vicinity of the fuel inlet.

Increasing the inlet velocity, on the other hand, leads to a decrease in the flame temperature and the trend is reversed at the range of 0.001 < X < 0.011. This is mainly because of locating the curve at the oxidizer side of the flame position.

The importance of a detailed multi-reaction mechanism relative to a single reaction is shown in Fig. 5-B The difference between a multi-step and single-step reaction is presented for the diffusion counterflow flame with a velocity of 0.05 m/s for both fuel and air. The position and value of maximum temperature, as a key parameter, is significantly different between the single and multi-step reactions that are the result of considering the accurate rate of detailed reaction into a multi-step solution. Therefore, the single reaction solution cannot be considered as an acceptable response to simplify the solution process of the counterflow diffusion flame. The maximum flame temperature for the multi-step analys is near to the air side because of fast peneteration of the fuel rather than the oxidizer due to the larger values of fuel molecular mass relative. This issue does not appear in single-step analyses because of lacking the effective species in the reactions, then it leads to inaccurate results.

To shed more lights on the importance of the chemistry, the mass fraction of species along the x-direction length of the open channel is shown in Fig. 5-C It is evident that the maximum difference between multi and single-step reaction for N2, CH4 and O2 species are 0.029, 0.08 and 0.192, respectively. The reported differences are acquired as results of accurate modeling of the reaction rates into the multi-step solution. As can be seen, the species of the multi-step analyze decay faster compared to the single-step species, because of considering the independent reactions and their effect on the enthalpy of reaction summation.

3.3.1. Micro-scaled analyses The micro-sized channel is defined as a channel height of the order of 10� 3 m or lower. To make a comparison between flame

structure within the macro and micro-sized channels, the variation of the hydraulic distance is at the range of 1–10 mm. Fig. 5-D presents the change in the maximum temperature of the flame versus hydraulic distance for two inlet temperatures of 300K

and 800K. As the hydraulic distance increases, the maximum temperature of the flame decreases which is because of the decrease of energy content with increasing thermal diffusivity of the reactive flow. Additionally, increasing fuel and air inlet temperatures results in a decrease in the maximum temperature and the pattern experiences a reverse trend at dh ¼ 9.8 mm.

The maximum temperature is influenced by the fuel and air inlet velocity in the case with dh ¼ 1 mm which is demonstrated in Fig. 6-A increasing fuel and air inlet velocity leads to a rise in the maximum temperature of the flame which is caused by the increase of the mixing rate of the fuel and the air and possibility of combustion occurrence. In other words, increasing the inlet temperature, at a constant velocity, increases the maximum temperature where the trend is conversely occurred, at a determined value of hydraulic diameter illustrated in Fig. 5-D.

The flame distance is defined as the longitudinal distance of the location where the maximum temperature occurs from fuel inlet and it is a key parameter to outline the structure of counterflow diffusion flame. The variation of this parameter against hydraulic distance is presented in Fig. 6-B for two inlet temperatures. With approaching to the micro-sized diameter, the flame is situated at a closer distance from the fuel inlet, moreover, the differences between the flame distance at two inlet temperatures decrease. Once the hydraulic diameter reaches to 1 mm, the values of the flame distance at two inlet temperatures become exclusive.

The residence time is employed to determine the transferring rate of mass species relative to the flame front speed as defined in Eq. (6). It is computed using the air properties and presented in Fig. 6-C and 6-D for different values of inlet velocities and temperatures. At higher velocity and inlet temperature, the residence time is shorter which is correspond to the higher maximum temperature shown in Fig. 6-A the observed trend is occurred due to the increased energy content of the open channel with enhancing the maximum temperature. That again leads to decreasing residence time of the reactive flow. It can be seen that the variation of residence time to the inlet velocity is more sensitive than that of the inlet temperature.

The minimum value of hydraulic distance at the flame stable limit is calculated to determine the quenching distance. The values of quenching distance and location of maximum temperature are presented in Table 1. For V ¼ 0.001 and 0.05 m/s quenching distance is equal to 0.5 mm and for V ¼ 0.2 m/s, the flame is stabilized at a smaller value of hydraulic distance due to the increase of the possibility of reactions between species as result of increase in inlet velocity. and The position of maximum temperature for the case with V ¼ 0.2 m/s is located at the longitudinal distance of 2.1 mm from the fuel inlet.

The flame thickness is defined as the longitudinal distance between two points with temperature differences of 200K, representing two points which are located at the fuel side and at the air side, respectively. Tables 2 and 3 present the flame thickness at different inlet temperatures and hydraulic distances for two inlet velocities. Decreasing the hydraulic distance leads to an increase in the flame thickness while the opposite is true for the inlet velocity and temperature. The computed results also show that the flame is over- expanded since the reaction process occurs rapidly in the open channel. The variation of the flame thickness is intensified with increasing the maximum inlet temperature approximately about 35%.

4. Conclusion

In this paper, a numerical time dependent micro/macro modeling of methane-air counterflow diffusion flame considering detailed kinetics is presented. The counterflow diffusion flame is passed through a rectangular open channel with different values of hydraulic distance at the macro and micro orders.

The outcomes of this study can be summarised as follows:

A. Edalati-nejad et al.

Page 8: The time dependent investigation of methane-air

Case Studies in Thermal Engineering 18 (2020) 100603

8

� The presented numerical results have an acceptable agreement with the experimental measurements. The calculated errors be-tween numerical results and experimental data are up to 14%, 7% and 21% respectively for nitrogen, oxygen and methane species. � The effect of combustion reaction rate is governed by (air/fuel) mass flow rate concerning the flame position

Fig. 6. A) Maximum temperature, versus fuel and air inlet velocity, for inlet temperatures of 300K and 800K and hydraulic distance of 1 mm - B) The position of maximum temperature versus hydraulic distance, for two inlet temperatures of 300K and 800K and inlet velocities of 0.05 m/s - Residence time versus C) different inlet velocities of fuel and air with the inlet temperature of 800K - D) inlet temperature with inlet velocities of 0.05 m/s where the hydraulic distance is equal to 1 mm.

Table 1 The variation of quenching distance and location of maximum temperature against different inlet velocities at the inlet temperature of 800K.

Inlet velocity, m/s 0.001 0.05 0.2

Quenching distance, mm 0.5 0.5 0.4 Location of maximum temperature, mm 1.95 2.1 2.1

Table 2 The variation of flame thickness versus hydraulic distance for two inlet temperatures equal to 300K and 800K where Vi ¼ 0.05 m/s.

Hydraulic distance (dh), mm 1 5 10

Flame thickness (δ) (Inlet temperature 800K), mm 3.75 3.45 3 Flame thickness (δ) (Inlet temperature 300K), mm 3.6 3.6 2.4

Table 3 The variation of flame thickness versus inlet velocities for two inlet temperatures equal to 300K and 800K where dh ¼ 1 mm.

Inlet velocity, m/s 0.001 0.05 0.2

Flame thickness (δ) (Inlet temperature 300K), mm 3.52 3.6 3.3 Flame thickness (δ) (Inlet temperature 800K), mm 3.15 3.75 4.42

A. Edalati-nejad et al.

Page 9: The time dependent investigation of methane-air

Case Studies in Thermal Engineering 18 (2020) 100603

9

� The velocity magnitude of the flow is maximized at the outlet of the open channel where flow progress is completely affected by combustion reaction rate. � The position and values of maximum temperature, is completely different between single and multi-step reactions. � The single reaction solution cannot be considered as an acceptable response to simplify solution process of the counterflow

diffusion flame. � Once hydraulic distance enlarged the maximum temperature of the flame is decreased accordingly, this occurs mainly due to the

decrease in energy content of the open channel. � The variation of residence time relative to inlet velocity is more sensitive than inlet temperature. � At larger inlet velocities, the flame is stabilized at a smaller value of hydraulic distance which is the result of increasing the

possibility of reactions between species. � The generation rate of CO2, CO and H2O species is nearly constant at t > 0.009s whereas for NO specie the rate is unchanged at t >

0.013s and for NO2 species the rate is fixed at t > 0.016s.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

[1] S. Roy, S. Zare, O. Askari, Understanding the Effect of Oxygenated Biofuels Addition on Combustion Characteristics of Gasoline, 2018, p. V06AT8A005. [2] S. Zare, S. Roy, A. El Maadi, O. Askari, An investigation on laminar burning speed and flame structure of anisole-air mixture, Fuel 244 (2019) 120–131. [3] B.A. Rabee, The effect of inverse diffusion flame burner-diameter on flame characteristics and emissions, Energy 160 (2018) 1201–1207. [4] I. Yamaoka, H. Tsuji, Y. Harigaya, Extinction and structure of methane/very lean methane-air counterflow diffusion flames, in: Symposium (International) on

Combustion, Elsevier, 1988, pp. 1837–1843. [5] A. Yoshida, R. Takasaki, K. Kashiwa, H. Naito, Y. Saso, Extinguishment of counterflow methane/air diffusion flame by polydisperse fine water droplets,

Combust. Flame 160 (2013) 1357–1363. [6] R. Padilla, O. Kwon, S. Lee, D. Dunn-Rankin, T. Pham, Extinction Limits and Structure of Counterflow Nonpremixed Water-Laden Methane/air Flames, ICDERS,

Irvine, CA, 2010. [7] S. Chen, Z. Liu, J. Liu, J. Li, L. Wang, C. Zheng, Analysis of entropy generation in hydrogen-enriched ultra-lean counter-flow methane–air non-premixed

combustion, Int. J. Hydrogen Energy 35 (2010) 12491–12501. [8] R. Fuse, H. Kobayashi, Y. Ju, K. Maruta, T. Niioka, NOx emission from high-temperature air/methane counterflow diffusion flame, Int. J. Therm. Sci. 41 (2002)

693–698. [9] C. Sohn, I. Jeong, S. Chung, Numerical study of the effects of pressure and air-dilution on no formation in laminar counterflow diffusion flames of methane in

high temperature air, Combust. Flame 130 (2002) 83–93. [10] C. Sung, B. Li, H. Wang, C. Law, Structure and sooting limits in counterflow methane/air and propane/air diffusion flames from 1 to 5 atmospheres, in:

Symposium (International) on Combustion, Elsevier, 1998, pp. 1523–1529. [11] U. Niemann, K. Seshadri, F.A. Williams, Methane, ethane, and ethylene laminar counterflow diffusion flames at elevated pressures: experimental and

computational investigations up to 2.0 MPa, Combust. Flame 161 (2014) 138–146. [12] I. Glassman, P. Yaccarino, The temperature effect in sooting diffusion flames, in: Symposium (International) on Combustion, Elsevier, 1981, pp. 1175–1183. [13] €O.L. Gülder, D. Snelling, Formation and temperature of soot particles in laminar diffusion flames with elevated temperatures, in: Symposium (International) on

Combustion, Elsevier, 1991, pp. 1509–1515. [14] R. Axelbaum, C. Law, Soot formation and inert addition in diffusion flames, in: Symposium (International) on Combustion, Elsevier, 1991, pp. 1517–1523. [15] J. Lim, J.P. Gore, R. Viskanta, Study of the effects of air-preheat on the structure of methane-air counterflow diffusion flames, ASME Heat Tran. Div. Publ. HTD

361 (1998) 39–47. [16] I. Puri, K. Seshadri, M. Smooke, D. Keyes, A comparison between numerical calculations and experimental measurements of the structure of a counterflow

methane-air diffusion flame, Combust. Sci. Technol. 56 (1987) 1–22. [17] M. Smooke, I. Puri, K. Seshadri, A comparison between numerical calculations and experimental measurements of the structure of a counterflow diffusion flame

burning diluted methane in diluted air, in: Symposium (International) on Combustion, Elsevier, 1988, pp. 1783–1792. [18] A. Fanaee, J. Esfahani, The investigation of semi-three-dimensional heat transfer modeling in microcombustors, J. Therm. Sci. Eng. Appl. 3 (2011), 011007. [19] M. Pourali, J. Esfahani, S. Fanaee, Two-dimensional analytical investigation of conjugate heat transfer in a finite-length planar micro-combustor for a hydrogen-

air mixture, Int. J. Hydrogen Energy 44 (2019) 12176–12187. [20] S. Fanaee, J. Esfahani, Analytical two-dimensional modeling of hydrogen–air mixture in catalytic micro-combustor, Meccanica 50 (2015) 1717–1732. [21] S.E. Hosseini, M.A. Wahid, Investigation of bluff-body micro-flameless combustion, Energy Convers. Manag. 88 (2014) 120–128. [22] G. Bagheri, S.E. Hosseini, Impacts of inner/outer reactor heat recirculation on the characteristic of micro-scale combustion system, Energy Convers. Manag. 105

(2015) 45–53. [23] G. Bagheri, S.E. Hosseini, M.A. Wahid, Effects of bluff body shape on the flame stability in premixed micro-combustion of hydrogen–air mixture, Appl. Therm.

Eng. 67 (2014) 266–272. [24] S. Fanaee, Self-similar nonasymptotic solution of multireaction stationary flow in catalytic microcombustor, J. Thermophys. Heat Tran. (2017) 1–10. [25] S. Fanaee, The analytical modeling of finite-length homogonous micro-combustor for a hydrogen-oxygen mixture with wall temperature effects, J. Mechan. 32

(2016) 631–642. [26] R.E. Padilla, D. Escofet-Martin, T. Pham, W.J. Pitz, D. Dunn-Rankin, Structure and behavior of water-laden CH4/air counterflow diffusion flames, Combust.

Flame 196 (2018) 439–451. [27] S. Lee, C.H. Shin, S. Choi, O.C. Kwon, Characteristics of NOx emissions of counterflow nonpremixed water-laden methane/air flames, Energy 164 (2018)

523–535. [28] A. Edalati-nejad, S.A. Fanaee, J. Khadem, The unsteady investigation of methane-air premixed counterflow flame into newly proposed plus-shaped channel over

palladium catalyst, Energy 186 (2019) 115833. [29] M. Bidabadi, A. Esmaeilnejad, An analytical model for predicting counterflow flame propagation through premixed dust micro particles with radiative heat loss,

J. Loss Prev. Process. Ind. 35 (2015) 182–199. [30] M. Yu, K. Zheng, L. Zheng, T. Chu, P. Guo, Effects of hydrogen addition on propagation characteristics of premixed methane/air flames, J. Loss Prev. Process.

Ind. 34 (2015) 1–9. [31] Delic F. Three-Dimensional Numerical Simulation of Combustion in a Jet Engine Combustion Chamber.

A. Edalati-nejad et al.

Page 10: The time dependent investigation of methane-air

Case Studies in Thermal Engineering 18 (2020) 100603

10

[32] F. Salehi, M. Cleary, A. Masri, Y. Ge, A. Klimenko, Sparse-Lagrangian MMC simulations of an n-dodecane jet at engine-relevant conditions, Proc. Combust. Inst. 36 (2017) 3577–3585.

[33] F. Salehi, M. Talei, E.R. Hawkes, C.S. Yoo, T. Lucchini, G. D’Errico, et al., Conditional moment closure modelling for HCCI with temperature inhomogeneities, Proc. Combust. Inst. 35 (2015) 3087–3095.

[34] A. Busseti, An Experimental and Numerical Investigation on the Propagation of Compositional Waves, Master’s Thesis, University of Genova, 2018.

A. Edalati-nejad et al.