Next Article in Journal
Effects of Numerical Schemes of Contact Angle on Simulating Condensation Heat Transfer in a Subcooled Microcavity by Pseudopotential Lattice Boltzmann Model
Next Article in Special Issue
An Ab Initio RRKM-Based Master Equation Study for Kinetics of OH-Initiated Oxidation of 2-Methyltetrahydrofuran and Its Implications in Kinetic Modeling
Previous Article in Journal
Decoupling of Electricity Consumption Efficiency, Environmental Degradation and Economic Growth: An Empirical Analysis
Previous Article in Special Issue
Transfer Functions of Ammonia and Partly Cracked Ammonia Swirl Flames
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical and Experimental Investigations of CH4/H2 Mixtures: Ignition Delay Times, Laminar Burning Velocity and Extinction Limits

1
Institute of Technical Thermodynamics, Karlsruhe Institute of Technology (KIT), 76131 Karlsruhe, Germany
2
Institute of Thermal Engineering, Technische Universität Bergakademie Freiberg (TUBAF), 09599 Freiberg, Germany
*
Authors to whom correspondence should be addressed.
Energies 2023, 16(6), 2621; https://doi.org/10.3390/en16062621
Submission received: 15 December 2022 / Revised: 25 February 2023 / Accepted: 28 February 2023 / Published: 10 March 2023
(This article belongs to the Special Issue Low-Carbon Fuel Combustion from Fundamentals to Applications)

Abstract

:
In this work, the influence of H2 addition on the auto-ignition and combustion properties of CH4 is investigated experimentally and numerically. Experimental ignition delay times (IDT) are compared with simulations and laminar burning velocities (LBVs), and extinction limits/extinction strain rates (ESRs) are compared with data from the literature. A wide variety of literature data are collected and reviewed, and experimental data points are extracted for IDT, LBV and ESR. The results are used for the validation of existing reaction mechanisms. The reaction mechanisms and models used are able to reproduce the influence of H2 addition to CH4 (e.g., shortening IDTs, increasing ESRs and increasing LBVs). IDTs are investigated in a range from 6 to 15 bar and temperatures from 929 to 1165 K with H2 addition from 10 to 100 mol%. We show that LBV and ESR are predicted in a wide range by the numerical simulations. Moreover, the numerical simulations using detailed Aramco Mech 3.0 (581 species) are compared with the derived reduced reaction mechanism UCB Chen (49 species). The results show that the reduced chemistry obtained by considering only the IDT is also valid for LBV and ESR.

1. Introduction

Auto-ignition processes of methane (CH4) and hydrogen (H2) have been investigated for the pure substances. However, there is an increasing interest in mixtures of CH4 and H2. For example, there are investigations by several groups [1,2,3] on combustion in SI engines, using CH4/H2 mixtures investigating the peak temperature and NOx production. Petersen et al. [4] investigated lean CH4/H2 mixtures for gas turbine combustion, and Adolf et al. [5] and Melaina et al. [6], for example, investigated the addition of H2 to the natural gas grid. In addition to these technical uses of CH4/H2 mixtures, there is also a safety aspect [7], for example, in nuclear power plants, as Rudy et al. [8] shows, or in pipelines/gas-processing sites as studied by Lowesmith et al. [9]. In addition, CH4 (a main component of natural gas) and H2 are very common fuels and are handled as fuels for greenhouse gas reduction [10,11].
A brief overview of studies investigating the ignition delay times (IDT) of CH4/H2 is summarized in Table 1 without claiming completeness. The studies are categorized by their setup (ST: shock tube and RCM: rapid compression machine) and sorted by year of publication. Regarding the fuel composition, the column named H2 represents the mole fraction of H2 in the CH4/H2 mixture (0 corresponds to pure CH4, and 100 corresponds to pure H2).
Hu et al. [24] measured IDT in an ST and compared the results to twelve reaction mechanisms. Based on [24], the H2 IDT data is simulated in this work by the reaction mechanism from [25]. Focusing on the CH4/H2 mixtures from Table 1, Gersen et al. [14] investigated IDT in a pressure range between 15 and 70 bar and showed that the CH4/C3H8 reaction mechanism of Petersen et al. [26] predicts the IDT for pure CH4 and also for pure H2 better than for their CH4/H2 mixtures. Donohoe et al. [15] measured IDT in an ST and a RCM, as well as flame speeds, of CH4/H2 mixtures and validated a former version of the AramcoMech (version 1.3) with these data.
As mentioned above, Petersen et al. [4] investigated the IDT of two CH4/H2 mixtures in gas turbines in comparison to other CH4/alkanes mixtures. They found a significantly reduced IDT by adding H2. This effect is further investigated in this current work. Zhang et al. [23] studied lean ( ϕ = 0.5 ) CH4/H2 mixtures in a shock tube. Under their conditions, they found a mixture of 40/60 CH4/H2 (molar) was almost independent of pressure. Under the aspect of new technical applications, it makes sense to improve the data on the ignition and combustion properties of CH4/H2 mixtures through dedicated experiments.
For the laminar burning velocity (LBV), numerous experiments have been conducted under different conditions over the last fifteen years. The addition of H2 to CH4 regarding the LBV has been investigated both numerically, e.g., [27,28], and experimentally. An overview of these experiments is given in [29] and in Table 2. The experimental data (see Table 2) were generated with different setups that are described in more detail in [30]. It can be seen that all methods except SEF were applied only under atmospheric conditions. The measurements that are known to us terminate at a pressure of 7.5 bar [31] and a temperature of 650 K [32]. In the high-temperature region, however, only hydrogen admixtures up to 50 mol−% are considered.
For higher pressures, values up to 100 mol−% of H2 in fuel are available. As expected, the trend indicates that the LBV increases over the entire mixing range when temperatures rise. A clear increase in LBV can also be observed for increasing hydrogen fractions with an increase in admixture up to about 50% in a linear fashion and exponentially beyond this point. For the pressure dependence, inconsistencies are found in the literature [30]; this is particularly applicable for higher (80% H2) hydrogen admixtures. For low admixtures, the LBV decreases with increasing pressure.
In comparison to LBV, the extinction strain rate has only been considered in very few experimental studies. An overview of these experiments is given in in Table 3. In the premixed flames [7,65], this may primarily relate to the very high outflow velocities that occur. In the non-premixed case [66,67], there are also only two known studies, for which the increased complexity in the precise dosing of very small amounts of hydrogen (mass-related) could be the determining factor. To the best of our knowledge, neither premixed nor non-premixed studies are known that have investigated a content of above 40 mol−% H2 admixture in CH4. This represents a clear gap in the literature.
As can be seen from Table 1, Table 2 and Table 3, there are several experimental data available for CH4 and H2 mixtures. Recently, Eckart et al. [67] covered the range 0–100% of H2 mixed to CH4 for ESRs; however, the setup was limited to low temperatures. In Eckart et al. [63], they used two setups for the LBV measurements with setup limits ( 5 bar and 373 K). In a small range of conditions, Troshin et al. [51] measured LBVs up to 10 bar (293– 573 K), and Berwal et al. measured up to 650 K ( 1 bar) [64]. Considering the IDT, data for CH4/H2 mixtures are still missing in wide areas. Gersen et al. [14] already provided preliminary work but only higher than 15 bar and only for a limited temperature range. This range will be therefore widened by the recent work.
In this work, we mainly investigated four stoichiometric CH4/H2 mixtures (0/100, 50/50, 80/20 and 90/10 CH4/H2, molar) in regard to their auto-ignition properties. For this purpose, IDTs were measured experimentally in an RCM and compared with simulations. All experimental results were compared with a detailed (AramcoMech 3.0 [68]) and an associated reduced (UCB Chen [69]) reaction mechanism.
As stated earlier, IDT is only one of the relevant validation targets for detailed and reduced reaction mechanisms. Therefore, after a detailed sensitivity analysis for the IDTs, the influence of the reduction of the mechanism on the other validation parameters of the laminar burning velocity and the extinction strain rate were again examined.

2. Methodology

2.1. Rapid Compression Machine: Experiment

The experiments were performed in a rapid compression machine (RCM), which is explained in detail in [70,71,72]; a brief overview of that RCM is given here. Figure 1 depicts the experimental setup of the RCM and the experiment.
The RCM is a piston-cylinder device with a compression time between 20 and 40 ms. The compression takes place under well-defined, reproducible initial and boundary conditions with regard to the pressure, temperature and gas composition. At top dead center (TDC), the piston is fixed with a mechanical lock, consisting of a knee-lever construction, whereby an isochoric state of the combustion chamber is forced. The initial pressure p 0 in the combustion chamber is measured with an absolute pressure sensor (MKS Baratron 121A, [73]) with a relative accuracy of 0.5%. The initial temperature T 0 is measured with a type-K thermocouple (accuracy of 2.2 K) [74] and can be adjusted by heating between 320 and 470 K. The gas mixtures to be investigated are prepared in a mixing vessel (about 10 L).
The volume and the pressure in the mixing vessel are higher than the initial pressure and the initial volume in the combustion chamber. This minimizes the measurement inaccuracy in the production of the initial gas. The combustion chamber of the RCM is filled with gas from the mixing vessel. After filling, an RCM experiment is performed. The temporal pressure trace in the combustion chamber is measured with a quartz pressure sensor (Kistler 6061 B [75]) with a linearity of ≤0.5%. In the investigated temperature range, the mentioned uncertainties result in temperature uncertainties of approximately ± 7 K. The time-resolved piston position is detected with a linear position sensor/potentiometer (Burster type 8712 [76]). The linear position sensor has a linearity of 0.1%.
Figure 2 (left) shows the pressure trace as measured in an experiment.
In this study, an RCM experiment is characterized by the pressure and temperature after the end of compression (index EOC, Figure 2 (left)). At TDC and isochoric conditions, however, heat losses occur, and this results in a temperature and pressure drop after compression. Ignition is detected by a rapid increase in pressure. The time measured between the end of compression and ignition is the ignition delay time (IDT, τ I G ).
In measuring the IDT in an RCM, effects (such as heat losses, dilution and pressure) can vary from facility/experiment to facility/experiment and, thus, influence the IDT. Therefore, for validation purposes, the H2 IDT data from this study is compared to the results of other publications (results for stoichiometric mixtures at pressure p C > 10   bar ). Experiments listed in Table 1 matching these criteria are listed with additional information in Table 4.
To minimize the effects of different dilutions (Inert/O2 ratio) and different compression pressures ( p E O C ) of the compared studies, we scaled the IDT by the O2 concentration ( [ O 2 ] ) as proposed by [12]:
τ I G [ O 2 ] = τ I G p E O C X O 2 R T E O C
with τ I G as the IDT of a dedicated experiment, X O 2 as the corresponding mole fraction, R as the gas constant and T E O C as the temperature at the end of compression. For these calculations, the ideal gas law was assumed. The comparison of the studies by the scaled IDT is shown in Figure 2 (right), whereby different colors denote different RCM facilities with different heat losses (except [24] (blue bullets) with a shock tube). H2 is one of the most frequently measured fuels in RCMs and STs apart from CH4; however, our CH4 data were already published in [69], and we, therefore, chose H2 here. The IDT results of the mentioned studies and our work show good agreement. Based on the depicted IDTs from RCM experiments, an Arrhenius-like fitting curve was calculated,
τ I G [ O 2 ] = A * exp ( E * / T ) ,
with A * = 3.3 × 10 17 smol / m 3 , E * = 3.6 × 10 4   K and T as the temperature. The data of the different RCM studies are in a band of ± 60 % for this fit. The shock tube data from Hu et al. [24] vary from this band—caused by the different setup and the very high dilution [77].

2.2. Rapid Compression Machine: Model

Simulations were performed using the in-house code HOMREA [78]. IDTs were simulated using the adiabatic core assumption [79,80]. We assumed that, despite the heat losses of the hot gas to the combustion chamber walls, an adiabatic gas (adiabatic core) remains in the middle of the combustion chamber. This adiabatic core delivers work to the surrounding gas layer, which causes a change of the internal energy and, thus, in the temperature of the core gas. The effective volume v ( τ ) of the adiabatic core is calculated from a measured pressure curve as follows,
v 0 v ( t ) c p ( T ) c v ( T ) 1 v d v = ln p 0 p ( t ) .
Here, c p ( T ) and c v ( T ) are the temperature-dependent specific heat capacities of the gas mixture, v 0 is the initial volume, p 0 is the initial pressure, and p ( t ) is the experimentally measured pressure of a non-reacting gas mixture. In the experiment with the non-reacting gas mixture, O2 was substituted with N2, which resulted in a similar heat capacity and thermal conductivity to the corresponding oxygen-containing mixture; however, there was no chemical reaction at the conditions of interest [81,82].
Due to the non-reacting gas mixture, only the influence of heat losses, which are not influenced by the chemistry, was measured. The adiabatic core assumption is sufficient for ignition delay times less than 100 ms [83]. Figure 2 (left) shows the pressure trace measured in the experiment and the pressure trace simulated with the adiabatic core model under nominally the same initial conditions. The pressure trace—in particular, the pressure drop after compression by heat loss and the pressure increase by the combustion after ignition—is reproduced by the simulation.

2.3. Heat-Flux Burner: Model

For comparison with the experimental LBV data, freely propagating one-dimensional flames were investigated numerically with the laminar premixed flamespeed code of Chemkin-Pro 2020 R2 [84]. The laminar burning velocity is a fundamental quantity present in a flat unstretched flame; it can be adequately calculated with a 1D prediction [30]. The LBV was calculated using multicomponent transport coefficients with a thermal diffusion option (Soret effect). The number of grid points was set to 1300 with respect to the gradient (GRAD = 0.04) and curvature (CURV = 0.08) adaptive mesh parameters. Further details are explained in the work of Eckart et al. [85].

2.4. Extinction Limits: Model

The in-house code INSFLA [78,86] was used to calculate the extinction strain rate (ESR) for the counterflow flame. In our model, the detailed transport models, including differential diffusion of different species and the thermal diffusion (the Soret effect) were included [87]. The optical thin model (OTM) for thermal radiation was also considered as documented in [88]. The detailed description of the ESR determination is consistent with that in our previous paper [89]. Note that the extinction limits were determined by using one-dimensional simulations. It was shown in our previous paper [89] that this is sufficient for the accuracy of the extinction limits. Nevertheless, two-dimensional simulations would further improve the accuracy.

2.5. Reaction Mechanisms

In this work, two different reaction mechanisms are used. First, the detailed reaction mechanism ArmacoMech 3.0 [68] was applied. This reaction mechanism consists of 581 species and 3037 reactions. The AramcoMech 3.0 was built by a H2/CO sub-reaction mechanism [17] and a C1−C2 reaction mechanism [90]. The (sub)-reaction mechanisms have already been validated for a large database, such as IDTs, LBVs and species formation in flow reactors [17,68,90]. Due to the size of the AramcoMech 3.0, it is less suitable for very complex models, such as DNS simulations, which are usually performed with smaller reaction mechanisms [91,92]. Therefore, the AramcoMech 3.0 was reduced in a previous work to better reflect the C1−C3 chemistry (UCB Chen [69]). The UCB Chen consists of 49 species and 332 reactions [72] and can also be used as a base reaction mechanism (C1−C3) for other reduced reaction mechanisms, e.g., it was extended for ethanol combustion [72].

3. Results

3.1. Influence of CH4/H2 Ratio

Figure 3 (left) depicts the IDT results from experiments and 0D simulations with an effective volume profile. Four different fuel blends at a pressure of 10 bar are shown. The AramcoMech 3.0 reaction mechanism shows good agreement for the 90/10 CH4/H2 mixture and for pure H2. The IDT results of the experiment and simulation of the other two mixtures diverge with decreasing temperature. The IDTs of mixtures with an intermediate ratio of CH4/H2 were predicted as too short by the AramcoMech 3.0 [68]. A similar effect was observed by Gersen et al. [14] for the reaction mechanism of Petersen et al. [26].
Panigrahy et al. [93] found several reactions that are important for CH4/H2 mixtures but less important for pure CH4 or pure H2. One reaction is CH4 + H· ⇌ CH3· + H2 [93]. The parameters of this reaction are different in the Aramco Mech 3.0 [68] in comparison to the NUI Galway Mech studied in the work of Panigrahy et al. [93]. However, the trend was still captured well for all mixtures. The trend of the experimental data suggests that the IDTs merge at low temperatures. This trend was also predicted by the numerical simulations.
At lower temperatures, the difference in IDT for CH4 and H2 decreased. When the temperature decreased further, the H2 ignition became slower than for CH4 as shown by Panigrahy et al. [93]. This is explained by the higher reactivity of CH3· in comparison to H atoms at low temperatures [93]. Figure 3 (left) demonstrates, on the one hand, the ignition-enhancing effect of H2 in the fuel and, on the other hand, the increase of the apparent activation energy with increasing H2 fraction. To analyze the ignition-enhancing effect of H2 in a CH4/H2 mixture in more detail, Figure 3 (right) compares the IDTs of different simulations with varying H2 mole fractions in fuel (black line). The simulations were each performed in an adiabatic, isochoric and homogeneous reactor with the same initial pressure and the same initial temperature. The H2 content in fuel α mentioned in Figure 3 (right) is
α = X H 2 / ( X H 2 + X CH 4 )
with X i as mole fractions. A stoichiometric mixture with a H2 content of α leads to an oxygen ratio β , defined as
β = ν H 2 α + ν CH 4 ( 1 α ) = 0.5 α + 2 ( 1 α ) = 2 1.5 α
with the stoichiometric coefficients ν i and the assumption of a complete reaction to H2O and CO2. The Ar dilution is derived from air with a molar Ar/O2 ratio of 3.76, resulting in a gas mixture with the following molar shares:
( 1 α ) / α / β / 3.76 β CH 4 / H 2 / O 2 / Ar ( molar ) .
For pure CH4 ( α = 0 ), the gas composition CH4/H2/O2/Ar is 1 / 0 / 2 / 7.52 (molar), and for a pure H2 mixture ( α = 1 ), the gas composition is 0 / 1 / 0.5 / 1.88 (molar).
The simulations were performed with the AramcoMech 3.0 reaction mechanism [68], and they highlight the ignition-accelerating effect of H2. The ignition-accelerating effect can be quantified by the sensitivity S of the IDT with respect to the change in the H2 content α
S = τ α τ α + 0.01 τ α 0.01 ,
τ ( α ) is the calculated IDT shown in Figure 3 (right) as a black line. The corresponding sensitivity according to Equation (7) is shown in Figure 3 (right) as a black dashed line. Low values of the H2 content α (low H2 mole fractions) were the most sensitive to increasing α , meaning that small additions of H2 had the highest impact on reducing IDT at constant pressure/temperature. The more H2 mixed into the fuel, the less sensitive the system becomes for H2 addition according to the IDT. Note that, if CH4 is treated as an inert species and everything else is kept constant, the reactants (H2/O2) no longer form a stoichiometric mixture but lean mixtures instead and, thus, ignite later than their stoichiometric counterparts [24].
Furthermore, the dilution of inert/O2 increases, which generally leads to longer IDT—all other conditions being unchanged. CH4 as an intert species adds a high heat capacity to the system without any ignition-supporting effect and, therefore, slows down the temperature rise during self-ignition.
Figure 4 shows the relative sensitivities of the OH radical (concentration). Full lines refer to positive sensitivities, and dashed lines refer to negative sensitivities. All sensitivities were calculated for an adiabatic and isochoric system for the time right before the gas mixture ignites ( 0.9 τ I G ). The initial conditions were p = 10   bar and T = 1000   K with a variable H2 content α (CH4/H2 ratio, Equation (4) and a constant Ar/O2 = 3.76 (molar) ratio.
For pure CH4 mixtures ( α = 0 ), the most sensitive reaction was
CH 3 + O 2 + M CH 3 O 2 + M .
For a high CH4 mole fraction in the fuel, the following reaction producing methyl radicals (CH3·) was sensitive
CH 4 + HO 2 CH 3 + H 2 O 2 .
On the other hand, with an increasing H2 fraction, reaction
O 2 + H O + OH
became the most sensitive, followed by two other hydrogen reactions. With respect to negative sensitivities, the formation of hydroperoxyl (HO2) was the most sensitive reaction for compositions α > 21 %
H + O 2 + M HO 2 + M
However, this reaction loses sensitivity with an increasing CH4 mole fraction (left side). Then, the consumption of CH3· and CH3O2· radicals become important. A notable fact is that even an α as small as 5 % led to high sensitivity with respect to the reaction O2 + H → OH + O. This sensitive reaction also explains the high sensitivity of the IDT according the H2 mole fraction in Figure 3.

3.2. Pressure Influence

Figure 5 (left) depicts the IDT of pure H2 (full symbols) in comparison to the 80/20 CH4/H2 fuel mixture (open symbols) at pressures of 10 and 15 bar. All data points from the simulations show shorter or the same IDT as the experiments. In the H2 case, the experiments are compared to two reaction mechanisms, first the AramcoMech 3.0 reaction mechanism and a H2 reaction mechanism [25]. Both reaction mechanisms reproduced the trend well and also calculated comparable values for the ignition delay time.
When comparing the 80/20 CH4/H2 mixture versus pure H2, both fuels show a linear decrease in the ignition delay time as a function of the temperature increase. Increasing the pressure results in a decrease in the ignition delay time in the investigated pressure/temperature regimes. In comparison, the 80/20 CH4/H2 mixture ignites slower than pure H2 at equivalent pressure and temperature. The temperature sensitivity decreases for the CH4/H2 mixture compared to pure H2 (the slope in the Arrhenius diagram).
Figure 5 (right) compares the ignition delay times of a fuel mixture of 90/10 CH4/H2 and pure H2 as a function of pressure. The compression temperatures for both gas mixtures are in a narrow temperature range since the initial temperature is held constant. The compression temperature of the CH4/H2 mixture is T C = 1040–1070 K, and the compression temperatures for the pure H2 experiments are between T C = 970 and 1000 K. Increasing the compression pressure from p C = 6 to 15 bar, the IDT drops for both fuels. Here, the CH4/H2 mixture shows a stronger pressure dependence for the IDT when compared with H2.

3.3. Laminar Burning Velocity

In addition to the new IDT measurements, the numerical models were also compared against other experiments from the literature. Therefore, LBV and ESR experiments, as listed in Table 2, were selected. In Figure 6, the laminar burning velocities of CH4/H2 mixtures are shown.
The mixtures are 100/0, 80/20, 60/40, 40/60 and 20/80 CH4/H2 (molar). In addition, the LBVs were calculated for temperatures of the gases at 298, 373, 473 and 573 K. Here, the two numerical models, AramcoMech 3.0 [68] (dashed line) and UCB Chen [69] (solid line) are compared. Furthermore, the experimental data from Table 2 are presented. The data are not assigned to the individual measurement series but only serve to determine the range in which a scattering of experimental data is present in the literature.
When evaluating the data, it is clear that the LBV increases with higher temperatures in all equivalence ratio ranges as expected. For temperatures above 373 K, only very few data sets are available [29,63], which clearly complicates the validation of the detailed reaction mechanisms. For CH4, experimental data are available for all four temperature ranges—at least in the lean range. However, above 373 K, the few data for CH4 are clearly overestimated by the models. For the range below 373 K, a good agreement between the reaction mechanisms and the experimental data can be found for all stages of CH4/H2 mixtures.
However, the agreement decreases for 60–80 mol−% H2, and a slight underestimation of certain data sets is observed. Comparing AramcoMech 3.0 with the UCB Chen reaction mechanism, no significant differences were found. This can be assumed to be because the UCB Chen reaction mechanism was developed by reducing AramcoMech 3.0. During development, laminar burning velocity values were also a target to be reproduced by the reduced mechanism.
For higher temperatures (above 373 K) there are currently not enough data available to make a comprehensive comparison. However, it is already evident from the CH4 data that the two mechanisms significantly overestimate the values in the lean range. Simulations of the stoichiometric 80 mol−% H2 mixtures and 473 K show an underestimation of the measurement results. As already mentioned, there are currently only a few studies available that have measured both high temperatures and high H2 contents (see Table 2).
From these comparisons follow some suggestions for future work. Since a validation of the reaction mechanisms for LBV above 373 K is hardly possible, experiments should be performed in these temperature ranges for which some existing spherical chambers as well as diverging channel setups are suitable. Furthermore, this is particularly the case for H2 admixtures above 60%, where significantly less data are available. Only then could the two reaction mechanisms be validated reliably in these ranges; currently, they show a trend towards a slight underestimation of the LBV for higher H2 admixtures. The reason behind the found discrepancies and their improvement should be discussed extensively by means of the rate of production and sensitivity analysis.

3.4. Extinction Strain Rate

Having shown the comparison for the LBV, the extinction of premixed flames is also a key parameter for the validation of reaction mechanisms. Therefore, the experiments of Jackson et al. [65] were used for validation. The experimental setup was a laminar, premixed flame in counterflow flame configuration. The gases were premixed CH4/air and CH4/H2/air mixtures. The authors stated that the double flame created by the opposed jet was a benchmark for assessing the effects of stretch on premixed flames for a range of fuel/air mixtures. In their setup, they investigated the lean extinction for equivalence ratios from 0.42 to 0.9.
The results and comparison to the numerical investigations performed in this work are shown in Figure 7. Both reaction mechanisms reproduce the trend of the pure CH4 case very well (initial temperature of the gas mixture: T g a s = 573   K ). There are almost no visible differences for the numerical results obtained by the two reaction mechanisms. Neither reaction mechanism can accurately reproduce the values at equivalence ratios of 0.9, and both overestimated these values. Considering the case of the H2 admixture (90/10 CH4/H2, molar), it can be seen that both reaction mechanisms were again able to reproduce the trend well; however, it can be observed that the reduced mechanism shows slightly higher ESRs at equivalence ratios above 0.6 compared with the detailed reaction mechanism, which has an over-estimation of less than 5%. Still, both numerical models accurately describe the effects of the addition of H2.

4. Conclusions

In this work, the auto-ignition of different CH4/H2 mixtures under stoichiometric conditions was investigated experimentally using rapid compression machine measurements. Four different fuel blends were studied in a pressure range from 6 to 15 bar and in a temperature range from 929 to 1165 K. Further, the ignition delay time, the extinction strain rate and the laminar burning velocity data in the literature were reviewed, extracted and compared to new numerical simulations for a wide range of experimental conditions with a detailed and reduced mechanism. The following observations were made:
1.
In the investigated temperature range, all fuel compositions showed a linear correlation of log ( τ ) to 1 / T . The apparent activation energy increased with increasing H2 content.
2.
Even small additions of H2 to CH4 enhanced the ignition process.
3.
With further increase of the H2 mole fraction, chain termination reactions, including C-atoms, became slightly less important.
4.
In the investigated pressure/temperature range, an increase in pressure enhanced the ignition of all fuel mixtures.
5.
The ignition delay times (IDT) of mixtures with a high CH4/low H2 mole fraction, as well as the opposite constellation, were predicted well by the reaction mechanism.
6.
The laminar burning velocities (LBV) of mixtures with a high CH4/low H2 mole fraction were predicted well by the reaction mechanism; for a higher temperature and higher H2 mole fraction, the mechanism showed discrepancies. A similar effect was observed by Zettervall et al. [94].
7.
The extinction strain rates (ESR) for pure CH4 and the 90/10 CH4/H2 were well reproduced by both reaction mechanisms.
The measurement of new IDTs over a wide temperature range showed that the reduced reaction mechanism UCB Chen [69] was able to reproduce the entire range of H2 enrichment in CH4. Furthermore, it was shown that satisfactory results were also obtained for the laminar burning velocity and the extinction strain rates. For some conditions, there is a lack of experimental measurements for higher hydrogen admixtures and higher temperatures, which should be the primary focus of research for future investigations. The accurate prediction of IDT, LBV and ESR encourages the use of the reaction mechanisms for simulating turbulent flames where ignition and local extinction take place.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en16062621/s1.

Author Contributions

Conceptualization, S.D., S.E., R.S. and U.M.; Methodology, S.D., S.E. and R.S.; Software, U.M.; Validation, S.D.; Formal analysis, U.M.; Investigation, S.D., S.E. and C.Y.; Resources, R.S. and U.M.; Writing–original draft, S.D., S.E. and C.Y.; Writing–review & editing, R.S., H.K. and U.M.; Visualization, S.D. and S.E.; Supervision, R.S. and U.M.; Project administration, R.S., H. K. and U.M.; Funding acquisition, R.S., H.K. and U.M. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support by the Deutsche Forschungsgemeinschaft within the framework of the DFG research group FOR 1993 “Multi-functional conversion of chemical species and energy” Project Number 229243862 and SCHI 647/3-3 is gratefully acknowledged. The authors from Freiberg gratefully acknowledge the financial support from the European Union and the state of Saxony in the ESF project (FKZ: 100284311).

Data Availability Statement

The data presented in this study are available in Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Abbreviations

BB: Bunsen burner; CEV: Cardiff explosion vessel; CF: counterflow; CV: combustion vessel; EHC: external heated converging channel; ESR: extinction strain rate; HF: heat flux burner; IDT: ignition delay time; LBV: laminar burning velocity; RCM: rapid compression machine; SB: slotburner; SEF: spherical expanding flame; and ST: shock tube.

References

  1. Kamil, M.; Rahman, M.M. Performance prediction of spark-ignition engine running on gasoline-hydrogen and methane-hydrogen blends. Appl. Energy 2015, 158, 556–567. [Google Scholar] [CrossRef]
  2. Diéguez, P.M.; Urroz, J.C.; Sáinz, D.; Gandía, L.M. Experimental study of the performance and emission characteristics of an adapted commercial four-cylinder spark ignition engine running on hydrogen—Methane mixtures. Appl. Energy 2014, 113, 1068–1076. [Google Scholar] [CrossRef]
  3. Akansu, S.O.; Bayrak, M. Experimental study on a spark ignition engine fueled by CH4/H2 ( 70/30 ) and LPG. Int. J. Hydrogen Energy 2011, 36, 9260–9266. [Google Scholar] [CrossRef]
  4. Petersen, E.L.; Hall, J.M.; Smith, S.D.; de Vries, J.; Amadio, A.R.; Crofton, M.W. Ignition of Lean Methane-Based Fuel Blends at Gas Turbine Pressures. J. Eng. Gas Turbines Power 2007, 129, 937–944. [Google Scholar] [CrossRef] [Green Version]
  5. Adolf, J.; Balzer, C.H.; Louis, J.; Fischedick, M.; Arnold, K.; Patowski, A.; Schüwer, D. Energy of the Future? Sustainable Mobility through Fuel Cells and H2; Shell Deutschland Oil GmbH: Hamburg, Germany, 2017. [Google Scholar]
  6. Melaina, M.W.; Antonia, O.; Penev, M. Blending Hydrogen into Natural Gas Pipeline Networks: A Review of Key Issues; Technical Report March; National Renewable Energy Laboratory (NREL): Golden, CO, USA, 2013. [Google Scholar] [CrossRef] [Green Version]
  7. Yang, X.; Wang, T.; Zhang, Y.; Zhang, H.; Wu, Y.; Zhang, J. Hydrogen effect on flame extinction of hydrogen-enriched methane/air premixed flames: An assessment from the combustion safety point of view. Energy 2022, 239, 122248. [Google Scholar] [CrossRef]
  8. Rudy, W.; Zbikowski, M.; Teodorczyk, A. Detonations in hydrogen-methane-air mixtures in semi confined flat channels. Energy 2016, 116, 1479–1483. [Google Scholar] [CrossRef]
  9. Lowesmith, B.J.; Hankinson, G.; Johnson, D.M. Vapour cloud explosions in a long congested region involving methane/hydrogen mixtures. Process Saf. Environ. Prot. 2011, 89, 234–247. [Google Scholar] [CrossRef] [Green Version]
  10. Bains, P.; Psarras, P.; Wilcox, J. CO2 capture from the industry sector. Prog. Energy Combust. Sci. 2017, 63, 146–172. [Google Scholar] [CrossRef]
  11. Schlögl, R. The role of chemistry in the energy challenge. ChemSusChem 2010, 3, 209–222. [Google Scholar] [CrossRef]
  12. Lee, D.; Hochgreb, S. Hydrogen autoignition at pressures above the second explosion limit (0.6-4.0 MPa). Int. J. Chem. Kinet. 1998, 30, 385–406. [Google Scholar] [CrossRef]
  13. Mittal, G.; Sung, C.J.; Yetter, R.A. Autoignition of H2/CO at elevated pressures in a rapid compression machine. Int. J. Chem. Kinet. 2006, 38, 516–529. [Google Scholar] [CrossRef]
  14. Gersen, S.; Anikin, N.B.; Mokhov, A.V.; Levinsky, H.B. Ignition properties of methane/hydrogen mixtures in a rapid compression machine. Int. J. Hydrogen Energy 2008, 33, 1957–1964. [Google Scholar] [CrossRef]
  15. Donohoe, N.; Heufer, A.; Metcalfe, W.K.; Curran, H.J.; Davis, M.L.; Mathieu, O.; Plichta, D.; Morones, A.; Petersen, E.L.; Güthe, F. Ignition delay times, laminar flame speeds, and mechanism validation for natural gas/hydrogen blends at elevated pressures. Combust. Flame 2014, 161, 1432–1443. [Google Scholar] [CrossRef] [Green Version]
  16. Hashemi, H.; Christensen, J.M.; Gersen, S.; Levinsky, H.; Klippenstein, S.J.; Glarborg, P. High-pressure oxidation of methane. Combust. Flame 2016, 172, 349–364. [Google Scholar] [CrossRef] [Green Version]
  17. Kéromnès, A.; Metcalfe, W.K.; Heufer, K.A.; Donohoe, N.; Das, A.K.; Sung, C.J.; Herzler, J.; Naumann, C.; Griebel, P.; Mathieu, O.; et al. An experimental and detailed chemical kinetic modeling study of hydrogen and syngas mixture oxidation at elevated pressures. Combust. Flame 2013, 160, 995–1011. [Google Scholar] [CrossRef] [Green Version]
  18. Bhaskaran, K.A.; Gupta, M.C.; Just, T. Shock tube study of the effect of unsymmetric dimethyl hydrazine on the ignition characteristics of hydrogen-air mixtures. Combust. Flame 1973, 21, 45–48. [Google Scholar] [CrossRef]
  19. Slack, M.W. Rate coefficient for H + O2 + M = HO2 + M evaluated from shock tube measurements of induction times. Combust. Flame 1977, 28, 241–249. [Google Scholar] [CrossRef]
  20. Wang, B.L.; Olivier, H.; Grönig, H. Ignition of shock-heated H 2 -air-steam mixtures. Combust. Flame 2003, 133, 93–106. [Google Scholar] [CrossRef]
  21. Huang, J.; Hill, P.G.; Bushe, W.K.; Munshi, S.R. Shock-tube study of methane ignition under engine-relevant conditions: Experiments and modeling. Combust. Flame 2004, 136, 25–42. [Google Scholar] [CrossRef]
  22. Pang, G.A.; Davidson, D.F.; Hanson, R.K. Experimental study and modeling of shock tube ignition delay times for hydrogen-oxygen-argon mixtures at low temperatures. Proc. Combust. Inst. 2009, 32, 107–118. [Google Scholar] [CrossRef]
  23. Zhang, Y.; Huang, Z.; Wei, L.; Zhang, J.; Law, C.K. Experimental and modeling study on ignition delays of lean mixtures of methane, hydrogen, oxygen, and argon at elevated pressures. Combust. Flame 2012, 159, 918–931. [Google Scholar] [CrossRef]
  24. Hu, E.; Pan, L.; Gao, Z.; Lu, X.; Meng, X.; Huang, Z. Shock tube study on ignition delay of hydrogen and evaluation of various kinetic models. Int. J. Hydrogen Energy 2016, 41, 13261–13280. [Google Scholar] [CrossRef]
  25. Ó Conaire, M.; Curran, H.J.; Simmie, J.M.; Pitz, W.J.; Westbrook, C.K. A comprehensive modeling study of hydrogen oxidation. Int. J. Chem. Kinet. 2004, 36, 603–622. [Google Scholar] [CrossRef]
  26. Petersen, E.L.; Kalitan, D.M.; Simmons, S.; Bourque, G.; Curran, H.J.; Simmie, J.M. Methane/propane oxidation at high pressures: Experimental and detailed chemical kinetic modeling. Proc. Combust. Inst. 2007, 31, 447–454. [Google Scholar] [CrossRef]
  27. Lamioni, R.; Bronzoni, C.; Folli, M.; Tognotti, L.; Galletti, C. Feeding H2-admixtures to domestic condensing boilers: Numerical simulations of combustion and pollutant formation in multi-hole burners. Appl. Energy 2022, 309, 118379. [Google Scholar] [CrossRef]
  28. Zhang, Y.; Wu, J.; Ishizuka, S. Hydrogen addition effect on laminar burning velocity, flame temperature and flame stability of a planar and a curved CH4-H 2—Air premixed flame. Int. J. Hydrogen Energy 2009, 34, 519–527. [Google Scholar] [CrossRef]
  29. Eckart, S.; Prieler, R.; Hochenauer, C.; Krause, H. Application and comparison of multiple machine learning techniques for the calculation of laminar burning velocity for hydrogen-methane mixtures. Therm. Sci. Eng. Prog. 2022, 32, 101306. [Google Scholar] [CrossRef]
  30. Konnov, A.A.; Mohammad, A.; Kishore, V.R.; Kim, N.I.; Prathap, C.; Kumar, S. A comprehensive review of measurements and data analysis of laminar burning velocities for various fuel+air mixtures. Prog. Energy Combust. Sci. 2018, 68, 197–267. [Google Scholar] [CrossRef]
  31. Hu, E.; Huang, Z.; He, J.; Jin, C.; Zheng, J. Experimental and numerical study on laminar burning characteristics of premixed methane–hydrogen–air flames. Int. J. Hydrogen Energy 2009, 34, 4876–4888. [Google Scholar] [CrossRef]
  32. Berwal, P.; Shawnam; Kumar, S. Laminar burning velocity measurement of CH4/H2/NH3-air premixed flames at high mixture temperatures. Fuel 2023, 331, 125809. [Google Scholar] [CrossRef]
  33. Braun-Unkhoff, M.; Kick, T.; Frank, P.; Aigner, M. An Investigation on Laminar Flame Speed as Part of Needed Combustion Characteristics of Biomass-Based Syngas Fuels. Proc. ASME Turbo Expo 2009, 2, 343–352. [Google Scholar] [CrossRef]
  34. Boushaki, T.; Dhué, Y.; Selle, L.; Ferret, B.; Poinsot, T. Effects of hydrogen and steam addition on laminar burning velocity of methane–air premixed flame: Experimental and numerical analysis. Int. J. Hydrogen Energy 2012, 37, 9412–9422. [Google Scholar] [CrossRef] [Green Version]
  35. Göckeler, K.; Albin, E.; Krüger, O.; Paschereit, C.O. Burning velocities of hydrogen-methane-air mixtures at highly steam-diluted conditions. In Proceedings of the fourth International Conference on Jets, Wakes and Separated Flows, ICJWSF-2013, Nagoya, Japan, 17–21 September 2013. [Google Scholar]
  36. Lhuillier, C.; Oddos, R.; Zander, L.; Lückoff, F.; Göckeler, K.; Paschereit, C.O.; Djordjevic, N. Hydrogen-Enriched Methane Combustion Diluted with Exhaust Gas and Steam: Fundamental Investigation on Laminar Flames and NOx Emissions. In Proceedings of the ASME Turbo Expo 2017: Turbomachinery Technical Conference and Exposition, Charlotte, NC, USA, 26–30 June 2017. Part F1300. [Google Scholar] [CrossRef]
  37. Ilbas, M.; Crayford, A.P.; Yilmaz, I.; Bowen, P.J.; Syred, N. Laminar-burning velocities of hydrogen–air and hydrogen–methane–air mixtures: An experimental study. Int. J. Hydrogen Energy 2006, 31, 1768–1779. [Google Scholar] [CrossRef]
  38. Park, O.; Veloo, P.S.; Liu, N.; Egolfopoulos, F.N. Combustion characteristics of alternative gaseous fuels. Proc. Combust. Inst. 2011, 33, 887–894. [Google Scholar] [CrossRef]
  39. Li, Z.; Cheng, X.; Wei, W.; Qiu, L.; Wu, H. Effects of hydrogen addition on laminar flame speeds of methane, ethane and propane: Experimental and numerical analysis. Int. J. Hydrogen Energy 2017, 42, 24055–24066. [Google Scholar] [CrossRef]
  40. Cammarota, F.; Di Benedetto, A.; Di Sarli, V.; Salzano, E.; Russo, G. Combined effects of initial pressure and turbulence on explosions of hydrogen-enriched methane/air mixtures. J. Loss Prev. Process Ind. 2009, 22, 607–613. [Google Scholar] [CrossRef]
  41. Halter, F.; Chauveau, C.; Djebaïli-Chaumeix, N.; Gökalp, I. Characterization of the effects of pressure and hydrogen concentration on laminar burning velocities of methane–hydrogen–air mixtures. Proc. Combust. Inst. 2005, 30, 201–208. [Google Scholar] [CrossRef]
  42. Halter, F.; Chauveau, C.; Gökalp, I. Characterization of the effects of hydrogen addition in premixed methane/air flames. Int. J. Hydrogen Energy 2007, 32, 2585–2592. [Google Scholar] [CrossRef]
  43. Shy, S.S.; Chen, Y.C.; Yang, C.H.; Liu, C.C.; Huang, C.M. Effects of H2 or CO2 addition, equivalence ratio, and turbulent straining on turbulent burning velocities for lean premixed methane combustion. Combust. Flame 2008, 153, 510–524. [Google Scholar] [CrossRef]
  44. Miao, H.; Jiao, Q.; Huang, Z.; Jiang, D. Measurement of laminar burning velocities and Markstein lengths of diluted hydrogen-enriched natural gas. Int. J. Hydrogen Energy 2009, 34, 507–518. [Google Scholar] [CrossRef]
  45. Tahtouh, T.; Halter, F.; Samson, E.; Mounaïm-Rousselle, C. Effects of hydrogen addition and nitrogen dilution on the laminar flame characteristics of premixed methane–air flames. Int. J. Hydrogen Energy 2009, 34, 8329–8338. [Google Scholar] [CrossRef]
  46. Hu, E.; Huang, Z.; He, J.; Zheng, J.; Miao, H. Measurements of laminar burning velocities and onset of cellular instabilities of methane–hydrogen–air flames at elevated pressures and temperatures. Int. J. Hydrogen Energy 2009, 34, 5574–5584. [Google Scholar] [CrossRef]
  47. Hu, E.; Huang, Z.; He, J.; Miao, H. Experimental and numerical study on lean premixed methane–hydrogen–air flames at elevated pressures and temperatures. Int. J. Hydrogen Energy 2009, 34, 6951–6960. [Google Scholar] [CrossRef]
  48. Fairweather, M.; Ormsby, M.P.; Sheppard, C.G.; Woolley, R. Turbulent burning rates of methane and methane–hydrogen mixtures. Combust. Flame 2009, 156, 780–790. [Google Scholar] [CrossRef] [Green Version]
  49. Salzano, E.; Cammarota, F.; Di Benedetto, A.; Di Sarli, V. Explosion behavior of hydrogen–methane/air mixtures. J. Loss Prev. Process Ind. 2012, 25, 443–447. [Google Scholar] [CrossRef]
  50. Okafor, E.C.; Hayakawa, A.; Nagano, Y.; Kitagawa, T. Effects of hydrogen concentration on premixed laminar flames of hydrogen–methane–air. Int. J. Hydrogen Energy 2014, 39, 2409–2417. [Google Scholar] [CrossRef]
  51. Troshin, K.Y.; Borisov, A.A.; Rakhmetov, A.N.; Arutyunov, V.S.; Politenkova, G.G. Burning velocity of methane-hydrogen mixtures at elevated pressures and temperatures. Russ. J. Phys. Chem. B 2013, 7, 290–301. [Google Scholar] [CrossRef]
  52. Reyes, M.; Tinaut, F.V.; Horrillo, A.; Lafuente, A. Experimental characterization of burning velocities of premixed methane-air and hydrogen-air mixtures in a constant volume combustion bomb at moderate pressure and temperature. Appl. Therm. Eng. 2018, 130, 684–697. [Google Scholar] [CrossRef]
  53. Khan, A.R.; Ravi, M.R.; Ray, A. Experimental and chemical kinetic studies of the effect of H2 enrichment on the laminar burning velocity and flame stability of various multicomponent natural gas blends. Int. J. Hydrogen Energy 2019, 44, 1192–1212. [Google Scholar] [CrossRef]
  54. Morovatiyan, M.; Shahsavan, M.; Baghirzade, M.; Hunter Mack, J. Effect of Hydrogen and Carbon Monoxide Addition to Methane on Laminar Burning Velocity. In Proceedings of the ASME 2019 Internal Combustion Engine Division Fall Technical, Conference, ICEF 2019, Chicago, IL, USA, 20–23 October 2019. [Google Scholar] [CrossRef] [Green Version]
  55. Cai, X.; Wang, J.; Bian, Z.; Zhao, H.; Zhang, M.; Huang, Z. Self-similar propagation and turbulent burning velocity of CH4/H2/air expanding flames: Effect of Lewis number. Combust. Flame 2020, 212, 1–12. [Google Scholar] [CrossRef]
  56. Nguyen, M.T.; Luong, V.V.; Pham, Q.T.; Phung, M.T.; Do, P.N. Effect of Ignition Energy and Hydrogen Addition on Laminar Flame Speed, Ignition Delay Time, and Flame Rising Time of Lean Methane/Air Mixtures. Energies 2022, 15, 1940. [Google Scholar] [CrossRef]
  57. Coppens, F.H.; De Ruyck, J.; Konnov, A.A. Effects of hydrogen enrichment on adiabatic burning velocity and NO formation in methane + air flames. Exp. Therm. Fluid Sci. 2007, 31, 437–444. [Google Scholar] [CrossRef]
  58. Hermanns, R. Laminar Burning Velocities of Methane-Hydrogen-Air Mixtures. Ph.D. Thesis, Technische Universiteit Eindhoven, Eindhoven, The Netherlands, 2007. [Google Scholar] [CrossRef]
  59. Dirrenberger, P.; Le Gall, H.; Bounaceur, R.; Herbinet, O.; Glaude, P.A.; Konnov, A.; Battin-Leclerc, F. Measurements of laminar flame velocity for components of natural gas. Energy Fuels 2011, 25, 3875–3884. [Google Scholar] [CrossRef] [Green Version]
  60. Nilsson, E.J.; van Sprang, A.; Larfeldt, J.; Konnov, A.A. The comparative and combined effects of hydrogen addition on the laminar burning velocities of methane and its blends with ethane and propane. Fuel 2017, 189, 369–376. [Google Scholar] [CrossRef]
  61. Jithin, E.V.; Varghese, R.J.; Velamati, R.K. Experimental and numerical investigation on the effect of hydrogen addition and N2/CO2 dilution on laminar burning velocity of methane/oxygen mixtures. Int. J. Hydrogen Energy 2020, 45, 16838–16850. [Google Scholar] [CrossRef]
  62. Wang, Z.; Wang, S.; Whiddon, R.; Han, X.; He, Y.; Cen, K. Effect of hydrogen addition on laminar burning velocity of CH4/DME mixtures by heat flux method and kinetic modeling. Fuel 2018, 232, 729–742. [Google Scholar] [CrossRef]
  63. Eckart, S.; Pizzuti, L.; Fritsche, C.; Krause, H. Experimental study and proposed power correlation for laminar burning velocity of hydrogen-diluted methane with respect to pressure and temperature variation. Int. J. Hydrogen Energy 2022, 47, 6334–6348. [Google Scholar] [CrossRef]
  64. Berwal, P.; Solagar, S.; Kumar, S. Experimental investigations on laminar burning velocity variation of CH4+H2+air mixtures at elevated temperatures. Int. J. Hydrogen Energy 2022, 47, 16686–16697. [Google Scholar] [CrossRef]
  65. Jackson, G.S.; Sai, R.; Plaia, J.M.; Boggs, C.M.; Kiger, K.T. Influence of H2 on the response of lean premixed CH4 flames to high strained flows. Combust. Flame 2003, 132, 503–511. [Google Scholar] [CrossRef]
  66. Niemann, U.; Seshadri, K.; Forman, A. Effect of addition of a nonequidiffusional Effect of addition of a nonequidiffusional reactant to an equidiffusional diffusion flame. Combust. Theory Model. 2013, 17, 563–576. [Google Scholar] [CrossRef]
  67. Eckart, S.; Rong, F.Z.; Hasse, C.; Krause, H.; Scholtissek, A. Combined experimental and numerical study on the extinction limits of non-premixed H2/CH4 counterflow flames with varying oxidizer composition. Int. J. Hydrogen Energy 2022, in press. [Google Scholar] [CrossRef]
  68. Zhou, C.W.; Li, Y.; Burke, U.; Banyon, C.; Somers, K.P.; Ding, S.; Khan, S.; Hargis, J.W.; Sikes, T.; Mathieu, O.; et al. An experimental and chemical kinetic modeling study of 1,3-butadiene combustion: Ignition delay time and laminar flame speed measurements. Combust. Flame 2018, 197, 423–438. [Google Scholar] [CrossRef]
  69. Drost, S.; Aznar, M.S.; Schießl, R.; Ebert, M.; Chen, J.Y.; Maas, U. Reduced reaction mechanism for natural gas combustion in novel power cycles. Combust. Flame 2021, 223, 486–494. [Google Scholar] [CrossRef]
  70. Drost, S.; Schießl, R.; Werler, M.; Sommerer, J.; Maas, U. Ignition delay times of polyoxymethylene dimethyl ether fuels (OME2 and OME3 and air: Measurements in a rapid compression machine. Fuel 2019, 258, 116070. [Google Scholar] [CrossRef]
  71. Drost, S.; Werler, M.; Schießl, R.; Maas, U. Ignition delay times of methane/diethyl ether (DEE) blends measured in a rapid compression machine (RCM). J. Loss Prev. Process Ind. 2021, 71, 104430. [Google Scholar] [CrossRef]
  72. Drost, S.; Kaczmarek, D.; Eckart, S.; Herzler, J.; Schießl, R.; Fritsche, C.; Fikri, M.; Atakan, B.; Kasper, T.; Krause, H.; et al. Experimental Investigation of Ethanol Oxidation and Development of a Reduced Reaction Mechanism for a Wide Temperature Range. Energy Fuels 2021, 35, 14780–14792. [Google Scholar] [CrossRef]
  73. Pressure Measurement & Control Baratron; Technical report; ® Manometer: Andover, MA, USA, 2008.
  74. National Institute of Standards and Technology. NIST Standard Reference Database 60; National Institute of Standards and Technology: Gaithersburg, MD, USA, 1993. [Google Scholar] [CrossRef]
  75. Kistler. Kistler Gruppe: ThermoCOMP®-Quartz Pressure Sensor, Type 6061B; Kistler Holding AG: Winterthur, Switzerland, 2013; pp. 1–3. [Google Scholar]
  76. Burster. Potentiometric Displacement Sensors. Models 8712, 8713; Technical Report; Gernsbach, Germany, 2020; Available online: https://www.burster.com/fileadmin/user_upload/redaktion/Documents/Products/Data-Sheets/Section_8/8712_EN.pdf (accessed on 3 April 2022).
  77. Würmel, J.; Silke, E.J.; Curran, H.J.; Ó Conaire, M.S.; Simmie, J.M. The effect of diluent gases on ignition delay times in the shock tube and in the rapid compression machine. Combust. Flame 2007, 151, 289–302. [Google Scholar] [CrossRef]
  78. Maas, U.; Warnatz, J. Ignition processes in hydrogen-oxygen mixtures. Combust. Flame 1988, 74, 53–69. [Google Scholar] [CrossRef]
  79. Goldsborough, S.S.; Hochgreb, S.; Vanhove, G.; Wooldridge, M.S.; Curran, H.J.; Sung, C.J. Advances in rapid compression machine studies of low-and intermediate-temperature autoignition phenomena. Prog. Energy Combust. Sci. 2017, 63, 1–78. [Google Scholar] [CrossRef] [Green Version]
  80. Sung, C.J.; Curran, H.J. Using rapid compression machines for chemical kinetics studies. Prog. Energy Combust. Sci. 2014, 44, 1–18. [Google Scholar] [CrossRef]
  81. Chen, C.J.; Back, M.H.; Back, R.A. The Thermal Decomposition of Methane. I. Kinetics of the Primary Decompositionto C2H6 + H2; Rate Constant for the Homogeneous Unimolecular Dissociation of Methane and its Pressure Dependence. Can. J. Chem. 1975, 53, 3580–3590. [Google Scholar] [CrossRef] [Green Version]
  82. Holmen, A.; Rokstad, O.A.; Solbakken, A. High-Temperature Pyrolysis of Hydrocarbons. 1. Methane to Acetylene. Ind. Eng. Chem. Process Des. Dev. 1976, 15, 439–444. [Google Scholar] [CrossRef]
  83. Desgroux, P.; Gasnot, L.; Sochet, L.R. Instantaneous temperature measurement in a rapid-compression machine using laser Rayleigh scattering. Appl. Phys. B 1995, 61, 69–72. [Google Scholar] [CrossRef]
  84. ANSYS Inc.; ANSYS Europe Ltd. ANSYS Chemkin-Pro Getting Started Guide, Release 2020 R2b; Number January; Ansys®: Canonsburg, PA, USA, 2020. [Google Scholar]
  85. Eckart, S.; Cai, L.; Fritsche, C.; vom Lehn, F.; Pitsch, H.; Krause, H. Laminar burning velocities, CO and NOx emissions of premixed polyoxymethylene dimethyl ether flames. Fuel 2021, 293, 120321. [Google Scholar] [CrossRef]
  86. Maas, U. Coupling of chemical reaction with flow and molecular transport. Appl. Math. 1995, 40, 249–266. [Google Scholar] [CrossRef]
  87. Hirschfelder, J.O.; Curtiss, C.F.; Bird, R.B.; Mayer, M.G. Molecular Theory of Gases and Liquids; Wiley: New York, NY, USA, 1964; Volume 165. [Google Scholar]
  88. Barlow, R.; Karpetis, A.; Frank, J.; Chen, J.Y. Scalar profiles and NO formation in laminar opposed-flow partially premixed methane/air flames. Combust. Flame 2001, 127, 2102–2118. [Google Scholar] [CrossRef]
  89. Eckart, S.; Yu, C.; Maas, U.; Krause, H. Experimental and numerical investigations on extinction strain rates in non-premixed counterflow methane and propane flames in an oxygen reduced environment. Fuel 2021, 298, 120781. [Google Scholar] [CrossRef]
  90. Metcalfe, W.K.; Burke, S.M.; Ahmed, S.S.; Curran, H.J. A Hierarchical and Comparative Kinetic Modeling Study of C1–C2 Hydrocarbon and Oxygenated Fuels. Int. J. Chem. Kinet. 2013, 45, 638–675. [Google Scholar] [CrossRef]
  91. Van Oijen, J.A. Direct numerical simulation of autoigniting mixing layers in MILD combustion. Proc. Combust. Inst. 2013, 34, 1163–1171. [Google Scholar] [CrossRef] [Green Version]
  92. Miyata, E.; Fukushima, N.; Naka, Y.; Shimura, M.; Tanahashi, M.; Miyauchi, T. Direct numerical simulation of micro combustion in a narrow circular channel with a detailed kinetic mechanism. Proc. Combust. Inst. 2015, 35, 3421–3427. [Google Scholar] [CrossRef]
  93. Panigrahy, S.; Abd El-Sabor Mohamed, A.; Wang, P.; Bourque, G.; Curran, H.J. When hydrogen is slower than methane to ignite. Proc. Combust. Inst. 2022, 1–11, in press. [Google Scholar] [CrossRef]
  94. Zettervall, N.; Fureby, C.; Nilsson, E.J.K. Evaluation of Chemical Kinetic Mechanisms for Methane Combustion: A Review from a CFD Perspective. Fuels 2021, 2, 210–240. [Google Scholar] [CrossRef]
Figure 1. Schematic of the main components of the RCM experiment. Adapted from [70].
Figure 1. Schematic of the main components of the RCM experiment. Adapted from [70].
Energies 16 02621 g001
Figure 2. (Left): Pressure trace of an RCM experiment with its corresponding simulation (same initial conditions, p 0 = 0.6 bar, T 0 = 411 K). (Right): Ignition delay times of H2 at stoichiometric conditions scaled by the O2 concentration compared with data from the literature. Data from: Gersen, 2008 (1&2) [14], Hu, 2016 [24], Lee, 1998 [12] and Mittal, 2006 [13].
Figure 2. (Left): Pressure trace of an RCM experiment with its corresponding simulation (same initial conditions, p 0 = 0.6 bar, T 0 = 411 K). (Right): Ignition delay times of H2 at stoichiometric conditions scaled by the O2 concentration compared with data from the literature. Data from: Gersen, 2008 (1&2) [14], Hu, 2016 [24], Lee, 1998 [12] and Mittal, 2006 [13].
Energies 16 02621 g002
Figure 3. (Left): IDT experiments and simulations of different fuels at a constant pressure. Simulations were performed with the AramcoMech 3.0 [68] and UCB Chen [69] reaction mechanism. (Right): Sensitivity analysis of stoichiometric mixtures with varying CH4/H2 ratios in fuel. Simulations were performed with AramcoMech 3.0 [68].
Figure 3. (Left): IDT experiments and simulations of different fuels at a constant pressure. Simulations were performed with the AramcoMech 3.0 [68] and UCB Chen [69] reaction mechanism. (Right): Sensitivity analysis of stoichiometric mixtures with varying CH4/H2 ratios in fuel. Simulations were performed with AramcoMech 3.0 [68].
Energies 16 02621 g003
Figure 4. Sensitivity analysis of stoichiometric mixtures with varying CH4/H2 ratio α . Simulations were performed with AramcoMech 3.0 [68]. Full lines: positive sensitivities and dashed lines: negative sensitivities.
Figure 4. Sensitivity analysis of stoichiometric mixtures with varying CH4/H2 ratio α . Simulations were performed with AramcoMech 3.0 [68]. Full lines: positive sensitivities and dashed lines: negative sensitivities.
Energies 16 02621 g004
Figure 5. (Left): Comparison of a stoichiometric H2 mixture (bullets) vs. stoichiometric 80/20 CH4/H2 mixture (open circles) at pressures of 10 ± 0.5 bar and 15 ± 0.5 bar. Simulations: AramcoMech 3.0 [68] (full lines) and UCB Chen [69] (dashed lines). (Right): IDT experiments of two fuels with a constant initial temperature and varying pressure. Lines are drawn to guide the eye.
Figure 5. (Left): Comparison of a stoichiometric H2 mixture (bullets) vs. stoichiometric 80/20 CH4/H2 mixture (open circles) at pressures of 10 ± 0.5 bar and 15 ± 0.5 bar. Simulations: AramcoMech 3.0 [68] (full lines) and UCB Chen [69] (dashed lines). (Right): IDT experiments of two fuels with a constant initial temperature and varying pressure. Lines are drawn to guide the eye.
Energies 16 02621 g005
Figure 6. Comparison of the laminar burning velocity of a methane–hydrogen mixture ( p = 1   bar ) for different temperatures ( 298, 373, 473 and 573 K) to the experimental literature data from Table 2.
Figure 6. Comparison of the laminar burning velocity of a methane–hydrogen mixture ( p = 1   bar ) for different temperatures ( 298, 373, 473 and 573 K) to the experimental literature data from Table 2.
Energies 16 02621 g006
Figure 7. Experimental [65] and modeling results for the extinction strain rate of 0 and 10 mol−% H2 in a premixed CH4 flame ( T g a s = 573   K ) as a function of the equivalence ratio.
Figure 7. Experimental [65] and modeling results for the extinction strain rate of 0 and 10 mol−% H2 in a premixed CH4 flame ( T g a s = 573   K ) as a function of the equivalence ratio.
Energies 16 02621 g007
Table 1. Overview of the studies that investigated ignition delay times for CH4/H2 mixtures. Column H2 represents the H2 mole fraction in the fuel mixture of CH4/H2. RCM: rapid compression machine and ST: shock tube.
Table 1. Overview of the studies that investigated ignition delay times for CH4/H2 mixtures. Column H2 represents the H2 mole fraction in the fuel mixture of CH4/H2. RCM: rapid compression machine and ST: shock tube.
SetupAuthor, YearT / Kp / barH2 in fuel/
mol−%
Source
RCMLee, 1998950–10506–40100[12]
RCMMittal, 2006950–110015–50100[13]
RCMGersen, 2008950–106015–700–100[14]
RCMDonohoe, 2014930–105010, 3060, 80[15]
RCMHashemi, 2016800–125015–800[16]
RCMKéromnès, 2013914–10148–70100[17]
STBhaskaran, 1973800–14002.5100[18]
STSlack, 1977980–11762100[19]
STWang, 2003900–13503–17100[20]
STHuang, 20041000–135016–400[21]
STPetersen, 20071132–155318–3020, 40[4]
STPang, 2009908–11183–4100[22]
STZhang, 20121000–20005–200–100[23]
STKéromnès, 2013925–21001–33100[17]
STDonohoe, 20141100–17001–3030, 80[15]
STHu, 2016850–15001–16100[24]
Table 2. Overview of the studies that investigated the laminar burning velocity for CH4/H2 mixtures. Column H2 represents the H2 mole fraction in the fuel mixture of CH4/H2. BB: Bunsen burner; SB: slotburner; CEV: Cardiff explosion vessel; CF: counterflow; CV: combustion vessel; SEF: spherical expanding flame; HF: heat flux burner; and EHC: external heated converging channel.
Table 2. Overview of the studies that investigated the laminar burning velocity for CH4/H2 mixtures. Column H2 represents the H2 mole fraction in the fuel mixture of CH4/H2. BB: Bunsen burner; SB: slotburner; CEV: Cardiff explosion vessel; CF: counterflow; CV: combustion vessel; SEF: spherical expanding flame; HF: heat flux burner; and EHC: external heated converging channel.
SetupAuthor, YearT / Kp / barH2 / mol−%Source
BBBraun-Unkhoff, 2009373150[33]
SBBoushaki, 201230010–30[34]
SBGöckeler, 201344010–50[35]
SBLhuillier, 2017393–47310–80[36]
CEVIlbas, 200629810–100[37]
CFPark, 2011298195[38]
CFLi, 201728310–50[39]
CVCammarota, 2009298110[40]
SEFHalter, 20052983,50–20[41]
SEFHalter, 200729810–20[42]
SEFShy, 200829810–30[43]
SEFMiao, 2009298120[44]
SEFTahtouh, 200930010–30[45]
SEFHu, 2009a30310–100[31]
SEFHu, 2009b303–4431–7.50–80[46]
SEFHu, 2009c303–4431–7.50–80[47]
SEFFairweather, 200936010–50[48]
SEFSalzano, 20122931–60–100[49]
SEFOkafor, 201435010–100[50]
SEFDonohoe, 2014300150[15]
SEFTroshin, 2014293–5731–100–20[51]
SEFReyes, 201730010–100[52]
SEFKhan, 2019300120[53]
SEFMorovatiyan, 2019298–44010–20[54]
SEFCai, 2020298120–80[55]
SEFNguyen, 202229810–50[56]
HFCoppens, 200729810–35[57]
HFHermanns, 2007298–43310–40[58]
HFDirrenberger, 201129810–67[59]
HFNilsson, 201729810–50[60]
HFJithin, 202030010–40[61]
HFWang, 201829810–40[62]
HF/SEFEckart, 2022298–3731–50–50[63]
EHCBerwal, 2022300–65010–50[64]
Table 3. Overview of the studies that investigated extinction strain rates for CH4/H2 mixtures. Column H2 represents the H2 mole fraction in the fuel mixture of CH4/H2. CF: counterflow.
Table 3. Overview of the studies that investigated extinction strain rates for CH4/H2 mixtures. Column H2 represents the H2 mole fraction in the fuel mixture of CH4/H2. CF: counterflow.
SetupAuthor, YearT / Kp / barH2 / mol−%ModeSource
CFJackson, 200357310–20premixed[65]
CFYang, 202229810–40premixed[7]
CFNiemann, 201329810–18non-premixed[66]
CFEckart, 202229810–100non-premixed[67]
Table 4. Overview of IDT as shown in Figure 2 for pure H2 and stoichiometric conditions.
Table 4. Overview of IDT as shown in Figure 2 for pure H2 and stoichiometric conditions.
SetupInert/O2p / barSource
RCM3.76 10 This work (1)
RCM9 10 This work (2)
RCM516–21Gersen (1) [14]
RCM520–50Gersen (2) [14]
ST30.810Hu [24]
RCM5 10 Lee [12]
RCM1315Mittal [13]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Drost, S.; Eckart, S.; Yu, C.; Schießl, R.; Krause, H.; Maas, U. Numerical and Experimental Investigations of CH4/H2 Mixtures: Ignition Delay Times, Laminar Burning Velocity and Extinction Limits. Energies 2023, 16, 2621. https://doi.org/10.3390/en16062621

AMA Style

Drost S, Eckart S, Yu C, Schießl R, Krause H, Maas U. Numerical and Experimental Investigations of CH4/H2 Mixtures: Ignition Delay Times, Laminar Burning Velocity and Extinction Limits. Energies. 2023; 16(6):2621. https://doi.org/10.3390/en16062621

Chicago/Turabian Style

Drost, Simon, Sven Eckart, Chunkan Yu, Robert Schießl, Hartmut Krause, and Ulrich Maas. 2023. "Numerical and Experimental Investigations of CH4/H2 Mixtures: Ignition Delay Times, Laminar Burning Velocity and Extinction Limits" Energies 16, no. 6: 2621. https://doi.org/10.3390/en16062621

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop