Next Article in Journal
Modelling Aero-Structural Deformation of Flexible Membrane Kites
Previous Article in Journal
Battery Storage Systems Control Strategies with Intelligent Algorithms in Microgrids with Dynamic Pricing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental and Kinetic Study on Laminar Burning Velocities of High Ratio Hydrogen Addition to CH4+O2+N2 and NG+O2+N2 Flames

State Key Laboratory of Clean Energy Utilization, Zhejiang University, Hangzhou 310027, China
*
Author to whom correspondence should be addressed.
Energies 2023, 16(14), 5265; https://doi.org/10.3390/en16145265
Submission received: 18 June 2023 / Revised: 29 June 2023 / Accepted: 1 July 2023 / Published: 9 July 2023
(This article belongs to the Section A5: Hydrogen Energy)

Abstract

:
In 2020, energy-related CO2 emissions reached 31.5 Gt, leading to an unprecedented atmospheric CO2 level of 412.5 ppm. Hydrogen blending in natural gas (NG) is a solution for maximizing clean energy utilization and enabling long-distance H2 transport through pipelines. However, insufficient comprehension concerning the combustion characteristics of NG, specifically when blended with a high proportion of hydrogen up to 80%, particularly with minority species, persists. Utilizing the heat flux method at room temperature and 1 atm, this experiment investigated the laminar burning velocities of CH4/NG/H2/air/He flames incorporating minority species, specifically C2H6 and C3H8, within NG. The results point out the regularity of SL enhancement, reaching its maximum at an equivalence ratio of 1.4. Furthermore, the propensity for the enhancement of laminar burning velocity aligned with the observed thermoacoustic oscillation instability during fuel-rich regimes. The experimental findings were contrasted with kinetic simulations, utilizing the GRI 3.0 and San Diego mechanisms to facilitate analysis. The inclusion of H2 augments the chemical reactions within the preheating zone, while the thermal effect from temperature is negligible. Both experimental and simulated results revealed that CH4 and NG with a large proportion of H2 had no difference, no matter whether from a laminar burning velocity or a kinetic analysis aspect.

1. Introduction

A significant energy demand has prompted profound apprehension due to the substantial surge in carbon dioxide (CO2) emissions. In the year 2020, global CO2 emissions from energy sources reached 31.5 Gt, resulting in an unprecedented level of atmospheric CO2 of 412.5 ppm—exhibiting approximately a 50% increase compared to the levels observed at the onset of the Industrial Revolution [1]. Hydrogen energy, described as a clean, efficient, and multi-application energy carrier, could be a connection between traditional fossil and renewable energy. However, the challenge of hydrogen storage and transportation remains a big obstacle. The mixing of hydrogen with natural gas (NG) will be an important solution to transport H2 in existing long-distance NG pipelines. Moreover, as an engine fuel, the synergy between natural gas and hydrogen is complementary. The incorporation of hydrogen amplifies the efficiency of natural gas combustion, diminishes carbon output, and expands its ignition threshold. Natural gas mitigates the challenges posed by hydrogen, including a low ignition energy, rapid combustion rate, and susceptibility to flash ignition and deflagration in realistic applications [2,3]. Therefore, it is crucial to explore the fundamental combustion properties of hydrogen/natural gas blends under more realistic conditions.
Laminar burning velocity, SL, is a fundamental parameter of laminar premixed combustion and serves as the basis for studying turbulent combustion [4]. The investigation into laminar burning velocity aids in enhancing the comprehension of diverse combustion processes, including quenching, blow-off, and flashback [5]. Moreover, laminar burning velocity serves as a valuable tool for validating chemical kinetic mechanisms.
Methane is the dominant component of NG. The SL of CH4 with added H2 has already been thoroughly investigated in diverse publications [6,7,8,9,10,11]. De Goey et al. [7] expanded upon the classical theoretical framework proposed by Peters and Williams for CH4/air flames to examine the structural characteristics of CH4/H2/air flames at ϕ = 1. According to the theoretical analysis, with hydrogen addition a decreasing inner-layer temperature was observed, and it dominated the increasing burning velocity. Halter et al. [8] investigated the impact of H2 addition (up to a 20% mixing ratio) and pressure on the properties of methane/air laminar flames across diverse equivalence ratios. The findings revealed that the increase in hydrogen resulted in an elevation of the laminar burning velocity and a decline in the flame’s sensitivity to stretch. Hermanns et al. [9] studied the impact of adding H2 to CH4 flames, with H2 volumetric proportions reaching up to 40%, employing the heat flux method. Several studies on CH4+H2+air SL under increasing conditions (300–650 K) have been complemented by Berwal et al. [6]. A power-law correlation was established to create a description of the temperature dependency, while at ϕ = 1.1 the temperature exponent exhibited a parabolic trend with a minimum value. Hu et al. [10,11] investigated the SL of methane/hydrogen/air mixtures, utilizing a constant volume combustion chamber method. These experimental outcomes indicated two typical characteristics of CH4+H2+air combustion, where in χ H 2 = 0.1–0.5 and χ H 2 = 0.9–1.0 conditions, in H2 combustion zones, there exists a linear increment, and in transition regimes with H2 mixed ( χ H 2 = 0.6–0.8) there exists a logarithmic amplification in SL.
Typically, besides methane, there always exist minority species in natural gas, like ethane and propane. The mixture of these components with hydrogen has not been investigated well. Nilsson et al. [4] examined the influence of H2 incorporation on the SL of methane/ethane/propane blends. It was found that, compared to pure methane, the heavier hydrocarbon exhibited a reduced augmentation in SL with hydrogen addition. Huang et al. [12] analyzed the SL of natural gas with hydrogen volumetric fractions between 0 and 100% and ϕ ranging from 0.6 to 1.4. With increasing hydrogen content, SL increased exponentially, while the flame instability intensified.
However, research about the effects of H2 on real composition NG, especially with large proportions of H2 (up to 0.8), has not been investigated and understood well. Furthermore, the responses of different parts of the components of natural gas to H2 addition have rarely been investigated, as well. Therefore, it is necessary to investigate the effect of real composition NG mixed with high proportions of H2.
Nevertheless, as mentioned above, high SL arising from a large proportion of H2 addition to NG will generate combustion instability. It is acknowledged that the high reactivity of hydrogen will raise multiple safety concerns when blended with ethane, methane, and propane. These effects also lead to significant changes in the ignition and propagation characteristic parameters of NG/H2/air flames: a reduction in minimum ignition energies and self-ignition delay times, alongside the growth in adiabatic flame temperatures and laminar burning velocity [13,14,15,16]. Several studies have revealed the effects of H2 on natural gas components, including the broadening of the flammability limits of CH4/air flames [13,17,18] and increases in SL [19,20,21,22]. Meanwhile, research has extensively focused on hydrogen-enriched combustion instability, not only from acoustically coupled pulsating motion aspects [23], but also from intense interactions between hydrodynamic and acoustic modes [24]. Following these theories, combustion instability has also been observed and taken into consideration in this study.
Based on the above-mentioned motivations, this research aims to fill in gaps regarding a large proportion of hydrogen, which can reach up to 0.8, mixed with a complex composition of real natural gas. The aims of this study are as follows: (1) to measure the precise laminar burning velocities for H2 mixed with Chinese typical Natural Gas, as Table 1 lists (#97 volume fraction: 96.68% CH4 + 1.53% C2H6 + 0.47% C3H8) [25], with varying volumetric proportions at 1 atm and room temperature, and to analyze thermoacoustic oscillation between different experimental working conditions; and (2) to validate and understand the profoundly kinetic mechanism influencing the SL of blended fuels with H2 added.

2. Materials and Methods

2.1. Experimental Setup

The SL of H2/CH4/NG/He/air flames were experimentally measured through the heat flux method under 1 atm and room temperature conditions. Figure 1 displays the current laboratory configuration schematic representation in this study. The heat flux method, including its principles and associated uncertainties, has been comprehensively described in our prior publications [26,27,28,29]. The setup was mainly composed of a heat flux burner and a gas supply system. Gas flow rates were regulated using 5 mass flow controllers (MFC, Alicat) to generate unburnt gas mixtures based on experimental parameters. The gas mixture was directed through the burner chamber and ignited above the burner plate, resulting in the formation of a planar laminar flame. The burner plate featured a 30 mm diameter perforated hexagonal pattern with 0.3 mm diameter apertures and a 0.4 mm spacing [30].
Table 1 lists the three types of natural gas adopted in the present investigation, named as NGOne, NGTwo, and NGThree. NGOne is a typical Natural Gas in China, and the others are artificial ones, for exploring the effects of different hydrocarbon fractions on SL and the combustion process. Table 2 outlines the experimental conditions employed. Notably, in conditions involving high hydrogen mixing ratios (up to 70% and 80%), the unburned gas mixture was diluted with helium to maintain laminar burning velocities within the heat flux SL tolerance range of approximately 70 cm/s. The volumetric fraction of helium, χ H e , in the total gas mixture was 30%, as described in Equation (1). The mole fractions of hydrogen, methane, or natural gas within the fuel composition, χ M , are described in Equation (2) with the subscript [M] denoting H2, CH4, NGOne, NGTwo, or NGThree.
χ H e = V H e / V f u e l + V a i r
χ M = V M / V H 2 + V C x H y
The equivalence ratio ϕ is defined as follows, in Equation (3):
ϕ = V H 2 + V C x H y / V a i r V H 2 + V C x H y / V a i r S t o i c h i o m e t r i c

2.2. Modeling Details

For modeling, the PREMIX module of Chemkin Pro was utilized with curvature and gradient parameters set at 0.02 for grid independence. Several kinetic mechanisms, the GRI 3.0 mechanism [31], San Diego mechanism [32], the USC mech Version II [33], Aramco 2.0 [34], and the CRECK mechanism [35], have been developed to describe hydrogen and hydrocarbon chemical reaction processes. The most mature and complete mechanism is GRI 3.0, where the validation of the model for C1–C4 hydrocarbons has demonstrated its accurate prediction of laminar burning velocities for methane, ethane, propane, and methane/hydrogen blends. In the present work, GRI 3.0 was therefore selected for the following kinetic analysis, and the San Diego mechanism was adopted to investigate the cases with helium addition.

3. Results and Discussion

3.1. Experimental Results of Methane and Hydrocarbon Mixtures at 298 K

The laminar burning velocity of methane/air at 298 K and 1 atm has been studied completely. Through extensive experimental procedures and statistical analysis, the scatter in SL values for methane/air across different setups and environments was within ±2 cm/s. To assess the reliability of the present laboratory setup, Figure 2a presents the measured SL of methane/air flames, considering varying equivalence ratios at room temperature and 1 atm from the current work and the ones by Han et al. [28], Wang et al. [29], and Nilsson et al. [4], who also used the heat flux method. The measured results of this study align with published data, affirming the experimental setup’s validity. Figure 2c,d show the measured SL of natural gas/air flames, which were simulated using the five selected mechanisms. Generally, there was good agreement between the simulations and experimental findings, except for a slight overprediction at ϕ < 0.8 and ϕ > 1.3. At NGOne, for the fuel-lean and fuel-rich side, simulations seemed to overpredict a little compared with other fuels. Moreover, the GRI 3.0 matched well with NGOne experimental results around stoichiometric conditions, where the laminar burning velocities were 37.99 cm/s, 39.45 cm/s, and 40.7 cm/s, respectively, for NGOne, NGTwo, and NGThree. All of the experimental data are provided in Table S1 of the Supplementary Material.

3.1.1. Effect of H2 Mole Fraction on Laminar Burning Velocity

Figure 3a illustrates the experimental and modeled SL of CH4/H2/air/He mixtures at room temperature and 1 atm, graphed as a function of the equivalence ratio, with χ H 2 spanning 0 to 0.6 without helium and 0.7 to 0.8 with helium dilution. The laminar burning velocities of NGOne, NGTwo, and NGThree mixed with H2 are displayed in Figure 3b–d. Detailed experimental data can be found in Table S1 of the Supplementary Material.
In Figure 3, the literature data were obtained from previous works of Han et al. [28], Wang et al. [29], Halter et al. [8], Hu et al. [10] and Huang et al. [12]. When comparing experimental and simulated results, the GRI 3.0 mechanism demonstrated better accuracy in predicting the SL of methane/ethane/propane/hydrogen blends compared to methane/hydrogen blends. Furthermore, it is crucial to emphasize that as χ H 2 increased, the discrepancy between experimental results and simulations decreased, no matter whether it was with helium dilution or not. As for methane, NGOne, NGTwo, and NGThree blended with hydrogen, the equivalence ratios of peak values gradually rose from ϕ = 1.05 to around ϕ = 1.1 under the 60% H2 mole fraction, even up to about ϕ = 1.15 under the 80% H2 mole fraction. The corresponding SL for methane/hydrogen blends were 36.56 cm/s, 74.28 cm/s, and 79.53 cm/s, respectively. Even based on helium addition, the tendency of increasing ϕ with H2 blends remained, and this phenomenon may be derived from the peak SL of pure hydrogen occurring at around ϕ = 1.7, so that its addition to CxHy affects the equivalence ratio, shifting the peak velocity towards the fuel-rich side. Helium, in this case, acts as a diluent in the fuel and oxidizer mixture.
The hydrogen addition to hydrocarbon blends led to non-monotonic impacts on the SL at diverse equivalence ratios. To assess the magnitude of the SL enhancement resulting from hydrogen addition to CH4- or CH4/C2H6/C3H8-blended flames, a parameter ξ was introduced to quantify the normalized effect of enhancement; in Equation (4), taking S L 0.4 , ϕ as an example, ξ 0.4 0.2 , ϕ can be denoted as [27]:
ξ 0.4 0.2 , ϕ = S L 0.4 , ϕ S L 0.2 , ϕ S L 0.2 , ϕ × 100 %
where S L 0.4 , ϕ indicates the laminar burning velocity of H2/CH4/NGOne/NGTwo/NGThree mixed with a H2 ratio of 0.4, and ξ 0.4 0.2 , ϕ represents the enhancement between S L 0.4 , ϕ and S L 0.2 , ϕ . The potential systematic errors associated with the heat flux method can be considered negligible, and the consistent experimental setup across different conditions further reduces potential uncertainties. According to Equation (5), the quantification of ξ uncertainty can be determined by [30]:
ξ = d ξ d S L χ H 2 , ϕ · S L χ H 2 , ϕ
Figure 4 displays the enhancement parameter ξ for CH4/air, NGOne/air, NGTwo/air and NGThree/air flames with 20–80% H2 added at room temperature and 1 atm with respect to the equivalence ratio. Apart from extremely fragmented data points, it can be pointed out that the results from both experimental and simulation analyses revealed that the most prominent enhancements of SL occurred at fuel-rich sides for all hydrocarbon blends at different χ H 2 . Hydrogen addition significantly amplified the sensitivity of off-stoichiometric flames, owing to the reduced reactivity and chemical heat release observed in lean and rich premixed flames, as compared to stoichiometric conditions. In addition, the fuel-rich side contained a large amount of hydrogen, so peak values occurred at these conditions. The equivalence ratio of peak SL enhancement achieved around ϕ = 1.4, and its peak values continually increased with hydrogen addition, where the ξ value was attained from 37.99% to 114.71%. With hydrogen added, the maximum values of SL enhancement tended to the fuel-rich side. Helium addition did not affect this trend, indicating that it mainly contributes to the physical characteristics of SL rather than chemical effects. Notably, the largest SL enhancement was observed for the case with 80% hydrogen addition.

3.1.2. Effect of CH4/C2H6/C3H8 Mole Fraction on Laminar Burning Velocity

Figure 3 shows different levels of fitness between experimental SL and predicted ones. From these figures, these five mechanisms are both consistent with CH4, NGOne, NGTwo, and NGThree blended with H2. The trends of equivalence ratios corresponding to peak SL increase with H2 addition did not change between the four studied fuels. This means that the influence of H2 is barely irreversible under different saturated hydrocarbon conditions.
Figure 4 shows the enhancement parameter ξ of four experimental gasses. The simulation and measured data revealed the non-monotonic rule of SL enhancement with H2 increasing. The best agreement between the experiments and modeling of the NGTwo condition is notable compared with the other three types of fuel. Meanwhile, the laminar flames generated constant and regular beeps under fuel-rich conditions. At the same time, the combustion stability was weakened, manifested as jumping flames as well as irregular flame shapes, which indicates that the current SL is undesirable. The corresponding beeps phenomenon is detailed in Table 3.
The constant and regular beeps during combustion on heat flux could be explained by the thermoacoustic oscillation instability introduced before by Matsuyama et al. [23] and Hemchandra et al. [24] According to the discussion stated above, the condition of 70% hydrogen addition was more unstable, and under the condition of NGThree mixed with H2, which mainly consisted of CH4 and C3H8, beeps occurred ahead at χ H 2 = 0.6. This phenomenon provides valuable guidance for determining the appropriate volumetric fraction of H2 in hydrocarbon blends to avoid unstable and hazardous conditions in future industrial hydrogen-doped combustion applications.

3.2. Kinetic Analysis

To obtain the underlying patterns of key intermediate species during combustion, the behaviors of species, H, O, OH, NO, and CH3 were presented. The profiles obtained from H2/CH4/NGOne/air flames at ϕ = 1.4 (where SL enhancement attained its maximum value) are displayed in Figure 5, where the solid line represents the χ H 2 = 0 condition, the dot-dash line represents the χ H 2 = 0.4 condition, and the dotted line represents the χ H 2 = 0.6 condition. Figure 5 illustrates the significantly elevated mole fraction of the H radical, and the OH radical, CH3 radical, and O radical appeared subsequently. With the H2 addition increasing, the analysis revealed that the maximum mole fractions of these species predominantly concentrated in the upstream region of the flame, displaying an increasing trend. This suggests that these radicals substantially impacted combustion within the preheating zone (0.05–0.1 cm). Such regularity can interpret the thermoacoustic oscillations mentioned above and the areas where thermoacoustic oscillation instability occurs in the preheating zone (0.05–0.1 cm), according to previous analysis.
Meanwhile, the NO component had a trend of rising first and then falling from χ H 2 = 0 to χ H 2 = 0.6, thus meaning that although H2 addition may result in higher NO emission initially, due to high temperature, NO emission will decline under the condition of a large proportion of hydrogen doping due to relatively complete combustion, which can even be lower than the 0% H2 condition in the downstream region of the flame. Additionally, when comparing H2/CH4/air flames, the radicals in H2/NGOne/air flames reached their peak values further upstream in the flame, and the reaction areas of CH3 radicals became narrower. This indicates that the reaction of H2/NGOne/air occurs earlier and intensifies, explaining the higher instability observed in NGOne blended with H2, as discussed in Table 3. H2/NGTwo/air and H2/NGThree/air flames had the same regularity compared to H2/CH4/air flame.
Figure 6 gives a clear display of H2/CH4/air/He and H2/NGOne/air/He flames, which revealed the same trend consistent with these radicals in H2/CH4/NGOne/air flames without He. In comparison to the H2/CH4/air/He flame, the H radical in the H2/NGOne/air/He flame was slightly higher at the 70% H2 condition and lower at the 80% H2 condition.
Figure 7 depicts the peak mole fractions of multiple crucial radicals at various positions within the flame, which represent the stoichiometric and H2/CH4/air conditions. In Figure 7, Figure 8 and Figure 9, the flame temperature profile is also depicted to illustrate the thermal effect arising from H2 addition. Corresponding to the discussion mentioned above, the growth of the H2 ratio in H2/CH4/air would prompt the maximum mole fraction of H, O, and OH radicals but reduce peak mole fraction of the CH3 radical, which could potentially arise from the limited availability of CxHy radicals in the hydrogen-rich flame, for stoichiometric H2/CH4/air flames. The slopes of O, OH radicals’ growth and CH3 radical decline were much gentler and flatter than the slope of H radical growth. The violent slope of the H radical growth matched well with the increment of SL as the H2 proportions grew, which indicates that the SL enhancement exhibited a robust positive correlation with it. The slopes of H, O, OH, and CH3 radicals in H2/NGOne/NGTwo/NGThree/air/He exhibited the same trends as H2/CH4/air/He.
Figure 8 represents the maximum mole fraction, adiabatic temperature, and laminar burning velocity of studied radicals at χ H 2 = 0.2, 0.8 conditions of H2/CH4/air/He flames at diverse equivalence ratios. To aid in visualization, the mole fraction of CH3 radical was multiplied by 10, and denoted as “CH3×10”. H2/NGOne/air/He, H2/NGTwo/air/He, and H2/NGThree/air/He had the same regularity as H2/CH4/air/He. According to Figure 8, which illustrates the equivalence-ratio-dependent variations in radical concentration, adiabatic temperature and SL, the chemical effect, represented by H radical concentration, exerted a more pronounced influence on laminar burning velocity than adiabatic temperature as H2 proportions ranged from 0.2 to 0.8. With the H2 proportion growing, the equivalence ratios of the peak temperature barely changed, while the equivalence ratios corresponding to the maximum H radical concentrations tended to be higher. Note that the mole fractions of the CH3 radical declined violently because of the decrease of CxHy in mixtures. Meanwhile, the mole fraction of the species under investigation initially rose and subsequently decreased as the equivalence ratio varied, and equivalence ratios of their peak values had a deviation to the fuel-rich sides, consistent with the regularity of SL presented above.
Figure 9 portrays the concentration of NOx with H2 ratio from 0 to 0.8 in H2/CH4/air/He flames. EINOx, as a NOx emission index, is reported as grams of NO2-equivalent formed per kilogram of fuel consumed. In Figure 9a, with the H2 volumetric ratio rising from 0% to 60%, EINOx tended to generate downstream while pure CH4 EINOx reached its local peak value upstream. All generation origin zones were far from the preheating zone (0.05–0.1 cm). In Figure 9b, the EINOx of the stoichiometric condition increased with the H2 ratio, consistent with Tad growth. Furthermore, with the H2 fraction ranging from 0% to 30%, EINOx grew and then, with the H2 fraction ranging from 30% to 60%, EINOx declined at ϕ = 1.4. Under helium dilution, the extreme point of EINOx occurred at χ H 2 = 0.72. This phenomenon could be attributed to incomplete combustion on the fuel-rich side. It is interesting to point out that in Figure 9c, although the H2 ratio attains 0.7 and 0.8, EINOx is even smaller than other experimental conditions because of the helium dilution added, resulting in a low temperature and mole fraction of total fuel in the reactant.
Figure 10 illustrates the fuel-rich side’s rate of production (ROP) of H radicals at ϕ = 1.4, providing insights into the mechanism underlying the observed behavior of various radicals in the upstream flame zone with H2 and He added. The top 10 vital elementary reactions involving active radical production or consumption are represented by different colors. Generally, reactions R1 and R2 have a prevailing influence on H radical formation at different conditions, which were associated with CO/H2 chemistry. With hydrogen addition increasing without helium, the H radical main reaction areas kept moving upstream of the flame and the reaction zones corresponding to peak values narrowed, with peak values growing accompanied. Under 30% He blended conditions, the H radical main reaction areas moved downstream and widened slightly, but were still angled towards upstream with the H2 volumetric fraction rising. Associated with discussions of mole fraction mentioned above, H2/NGOne/air/He, H2/NGTwo/air/He, and H2/NGThree/air/He had the same regularity of H radical reactions, which indicates that the H radicals in H2/NGOne/NGTwo/NGThree/air/He flames attained the peak further upstream. To sum up, the trends of H radicals, particularly under 30% He blending conditions, provide further insights into the thermoacoustic oscillation observed in large-scale hydrogen doping experiments.
O H + H 2 = H + H 2 O ( R 1 )
O + H 2 = H + O H ( R 2 )
To further elucidate the mechanisms underlying the augmentation of laminar burning velocity with H2 blends with CH4 and natural gases, detailed kinetic analyses have been carried out through GRI 3.0 without He blends and the San Diego mechanism with He blends. Figure 11 presents the A-factor sensitivities of SL for the 10 most influential reactions in H2/CH4/air flames under the specified conditions ( ϕ = 0.6, 1.0, 1.3, and 1.4) for χ H 2 = 0 and 0.4 individually. Figure 11 reveals that the reaction R3, which has been considered the pivotal chain-branching reaction in hydrocarbon flame [36], plays a dominant role in positive sensitivity under various circumstances. This phenomenon may arise from the significance of the reaction R3 in H2/CO chemistries, which could analogize kinetic processes between H2/CO and H2/CxHy. The chain termination reaction R4 exhibits the most significant negative sensitivity, indicating the notable impact of C-containing species like CH3 radical on laminar burning velocity.
H + O 2 = O + O H ( R 3 )
C H 3 + H + M = C H 4 + M ( R 4 )
H O 2 + C H 3 = O H + C H 3 O ( R 5 )
O + C H 3 = > H + H 2 + C O ( R 6 )
With the hydrogen addition increasing from 0 to 0.4, the positive sensitivity factors of the reaction R5 decreased significantly, while the positive sensitivity factors of the reaction R6 increased substantially. This indicates that the reactions R5 and R6 have a competing relationship to consume CH3 radical, and under large proportions of hydrogen doping circumstances CH3 radical has been transferred to generate the H radical instead of OH radical. Of particular note is the growth rate of the positive sensitivity factors of the reaction R6, which is particularly much larger at ϕ = 1.4, with the H2 mole fraction increasing. This tendency may be ascribed to the high temperature and reactivity of highly active free radicals under large proportions of hydrogen doping reactions at fuel-rich conditions.
As to the instability of thermoacoustic oscillation mentioned above at a 70% and 80% H2 volumetric fraction with helium blended, Figure 12 points out that the reaction R7 makes a replacement for the reaction R6, which means that the chain-branching reaction R6 was no more vital. The disappearance of such a chain-branching reaction could track back the temperature drop due to He addition, corresponding to the analysis of H ROP. According to Figure 12, it can also be found that the reaction R8 has the second most negative sensitivity factors accelerating kinetic processes through this H-atom abstraction reaction. Other sensitivity analyses on three types of H2/NG/air/He flames have similar trends compared with Figure 11 and Figure 12.
O + C H 3 = H + C H 2 O ( R 7 )
H + O H + M = H 2 O + M ( R 8 )

4. Conclusions

This study investigated the augmentation of H2 addition into CH4/NGOne/NGTwo/NGThree/air with and without He at room temperature and 1 atm. The experimental measurements adopted the heat flux method and kinetic modeling using GRI 3.0 and the San Diego mechanism to compare the consistency between laboratory results and simulations. In summary, the primary findings are summarized as follows.
  • The SL measured by the heat flux method fits well with GRI 3.0 and San Diego mechanism simulations at various conditions. Furthermore, SL has much more consistency with simulated lines in the case of large proportions of hydrogen doping. At the same time, combustion stability weakens the manifestation in jumping flames as well as irregular flame shapes for the conditions of large H2 volumetric fraction. The constant and regular beeps accompanying combustion on heat flux could be explained by thermoacoustic oscillation instability.
  • The significant factor, SL enhancement, has been observed to reach peak values when fuel-rich in contrast to lean and stoichiometric fuel blends. The equivalence ratio of peak SL enhancement achieved around ϕ = 1.4 and continued to increase with hydrogen addition, where the ξ could reach from 37.99% to 114.71%, respectively.
  • With H2 addition growing, the analysis revealed that the maximum mole fractions of these species predominantly concentrate in the upstream region of the flame, presenting an increasing trend. This suggests that these radicals substantially impact combustion within the preheating zone. (0.05–0.1 cm). On the other hand, in hydrogen blended mixtures, the chemical effect outweighs the thermal effect arising from increased temperature.
  • When χ H 2 varied from 0.3 to 0.6, the peak SL enhancement achieved around ϕ = 1.4 and, while at this condition, EINOx descended. Blending with a H2 ratio beyond 0.7 is not recommended because of the intense thermoacoustic oscillation instability, especially for natural gas mainly composed of CH4 and C3H8. These phenomena could inform the amount of H2 volumetric fraction (recommended range is 0.3–0.6) used in hydrocarbon blends to achieve high efficiency combustion and low emission and avoid unstable and dangerous conditions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en16145265/s1, Table S1: Laminar burning velocities of H2/CH4/NGOne/NGTwo/NGThree/air/He flames.

Author Contributions

Conceptualization, Z.Z. and R.Z.; methodology, Z.Z.; validation, Z.Z. and R.Z.; formal analysis, Z.Z.; investigation, Y.Z.; resources, Z.W.; data curation, Z.Z.; writing—original draft preparation, Z.Z.; writing—review and editing, W.W., Y.H. and Z.W.; visualization, Y.Z.; supervision, W.W., Y.H. and Z.W.; project administration, Z.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key R&D Program of China (2022YFB4003900), the National Natural Science Foundation of China (52125605), and the Zhejiang Provincial Natural Science Foundation of China (LZ21E060003, LR23E060001).

Data Availability Statement

Data are contained within supplementary material. The data presented in this study are available in the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. IEA. Global Energy Review 2021; IEA: Paris, France, 2021; Available online: https://www.iea.org/reports/global-energy-review-2021 (accessed on 1 May 2021).
  2. Xiang, L.; Jiang, H.; Ren, F.; Chu, H.; Wang, P. Numerical study of the physical and chemical effects of hydrogen addition on laminar premixed combustion characteristics of methane and ethane. Int. J. Hydrogen Energy 2020, 45, 20501–20514. [Google Scholar] [CrossRef]
  3. Lamioni, R.; Bronzoni, C.; Folli, M.; Tognotti, L.; Galletti, C. Impact of H2-enriched natural gas on pollutant emissions from domestic condensing boilers: Numerical simulations of the combustion chamber. Int. J. Hydrogen Energy 2023, 48, 19686–19699. [Google Scholar] [CrossRef]
  4. Nilsson, E.J.K.; 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]
  5. Fruzza, F.; Lamioni, R.; Tognotti, L.; Galletti, C. Flashback of H2-enriched premixed flames in perforated burners: Numerical prediction of critical velocity. Int. J. Hydrogen Energy, 2023; in press. [Google Scholar] [CrossRef]
  6. 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]
  7. De Goey, L.P.H.; Hermanns, R.T.E.; Bastiaans, R.J.M. Analysis of the asymptotic structure of stoichiometric premixed CH4–H2–air flames. Proc. Combust. Inst. 2007, 31, 1031–1038. [Google Scholar] [CrossRef]
  8. 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]
  9. Hermanns, R.T.E.; Konnov, A.A.; Bastiaans, R.J.M.; de Goey, L.P.H.; Lucka, K.; Köhne, H. Effects of temperature and composition on the laminar burning velocity of CH4+H2+O2+N2 flames. Fuel 2010, 89, 114–121. [Google Scholar] [CrossRef]
  10. 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]
  11. 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]
  12. Huang, Z.; Zhang, Y.; Zeng, K.; Liu, B.; Wang, Q.; Jiang, D. Measurements of laminar burning velocities for natural gas–hydrogen–air mixtures. Combust. Flame 2006, 146, 302–311. [Google Scholar] [CrossRef]
  13. Askar, E.; Schröder, V.; Seemann, A.; Schuetz, S. Power-to-Gas: Safety Characteristics of Hydrogen/NaturalGas-Mixtures. Chem. Eng. Trans. 2016, 48, 397–402. [Google Scholar]
  14. Brower, M.; Petersen, E.L.; Metcalfe, W.; Curran, H.J.; Füri, M.; Bourque, G.; Aluri, N.; Güthe, F. Ignition Delay Time and Laminar Flame Speed Calculations for Natural Gas/Hydrogen Blends at Elevated Pressures. J. Eng. Gas Turbines Power 2013, 135, 4007763. [Google Scholar] [CrossRef] [Green Version]
  15. Kuppa, K.; Goldmann, A.; Schöffler, T.; Dinkelacker, F. Laminar flame properties of C1-C3 alkanes/hydrogen blends at gas engine conditions. Fuel 2018, 224, 32–46. [Google Scholar] [CrossRef]
  16. Mosisa Wako, F.; Pio, G.; Salzano, E. The Effect of Hydrogen Addition on Low-Temperature Combustion of Light Hydrocarbons and Alcohols. Energies 2020, 13, 3808. [Google Scholar] [CrossRef]
  17. Molnarne, M.; Schroeder, V. Hazardous properties of hydrogen and hydrogen containing fuel gases. Process Saf. Environ. Prot. 2019, 130, 1–5. [Google Scholar] [CrossRef]
  18. Wang, T.; Zhou, Y.; Luo, Z.; Wen, H.; Zhao, J.; Su, B.; Cheng, F.; Deng, J. Flammability limit behavior of methane with the addition of gaseous fuel at various relative humidities. Process Saf. Environ. Prot. 2020, 140, 178–189. [Google Scholar] [CrossRef]
  19. 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]
  20. Hu, G.; Zhang, S.; Li, Q.F.; Pan, X.B.; Liao, S.Y.; Wang, H.Q.; Yang, C.; Wei, S. Experimental investigation on the effects of hydrogen addition on thermal characteristics of methane/air premixed flames. Fuel 2014, 115, 232–240. [Google Scholar] [CrossRef]
  21. 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]
  22. Tang, C.L.; Huang, Z.H.; Law, C.K. Determination, correlation, and mechanistic interpretation of effects of hydrogen addition on laminar flame speeds of hydrocarbon–air mixtures. Proc. Combust. Inst. 2011, 33, 921–928. [Google Scholar] [CrossRef]
  23. Matsuyama, S.; Hori, D.; Shimizu, T.; Tachibana, S.; Yoshida, S.; Mizobuchi, Y. Large-Eddy Simulation of High-Frequency Combustion Instability in a Single-Element Atmospheric Combustor. J. Propuls. Power 2016, 32, 628–645. [Google Scholar] [CrossRef]
  24. Hemchandra, S.; Shanbhogue, S.; Hong, S.; Ghoniem, A.F. Role of hydrodynamic shear layer stability in driving combustion instability in a premixed propane-air backward-facing step combustor. Phys. Rev. Fluids 2018, 3, 063201. [Google Scholar] [CrossRef] [Green Version]
  25. Huang, B.; Peng, T.; Xia, C.; Zhai, Y.; Shi, J.; Sun, Z.; Zheng, F.; Wu, Y. Identification of Natural Gas Components Using the Support Vector Machine Model. Chem. Technol. Fuels Oils 2021, 57, 713–723. [Google Scholar] [CrossRef]
  26. Alekseev, V.A.; Naucler, J.D.; Christensen, M.; Nilsson, E.J.K.; Volkov, E.N.; de Goey, L.P.H.; Konnov, A.A. Experimental Uncertainties of the Heat Flux Method for Measuring Burning Velocities. Combust. Sci. Technol. 2016, 188, 853–894. [Google Scholar] [CrossRef]
  27. Chen, C.; Wang, Z.; Yu, Z.; Han, X.; He, Y.; Zhu, Y.; Konnov, A.A. Experimental and kinetic modeling study of laminar burning velocity enhancement by ozone additive in NH3+O2+N2 and NH3+CH4/C2H6/C3H8+air flames. Proc. Combust. Inst. 2023, 39, 4237–4246. [Google Scholar] [CrossRef]
  28. Han, X.; Wang, Z.; Wang, S.; Whiddon, R.; He, Y.; Lv, Y.; Konnov, A.A. Parametrization of the temperature dependence of laminar burning velocity for methane and ethane flames. Fuel 2019, 239, 1028–1037. [Google Scholar] [CrossRef]
  29. Wang, S.; Wang, Z.; Elbaz, A.M.; Han, X.; He, Y.; Costa, M.; Konnov, A.A.; Roberts, W.L. Experimental study and kinetic analysis of the laminar burning velocity of NH3/syngas/air, NH3/CO/air and NH3/H2/air premixed flames at elevated pressures. Combust. Flame 2020, 221, 270–287. [Google Scholar] [CrossRef]
  30. Zhu, R.; Han, X.; Zhang, Z.; He, Y.; Wang, Z. Experimental and kinetic study on laminar burning velocities of NH3/CH4/H2S/air flames. Fuel 2023, 332, 126174. [Google Scholar] [CrossRef]
  31. Smith, G.P.; Frenklach, M.; Moriarty, N.W.; Eiteneer, B.; Goldenberg, M.; Bowman, C.T.; Hanson, R.K.; Song, S.; Gardiner, W.C., Jr.; Lissianski, V.V.; et al. Available online: http://combustion.berkeley.edu/gri-mech/version30/text30.html (accessed on 1 December 1998).
  32. Chemical-Kinetic Mechanisms for Combustion Applications, Mechanical and Aerospace Engineering (Combustion Research), University of California at San Diego. Available online: http://combustion.ucsd.edu (accessed on 1 May 2001).
  33. Wang, H.; You, X.; Joshi, A.V.; Davis, S.G.; Laskin, A.; Egolfopoulos, F.; Law, C.K.; USC Mech Version II. USC Mech Version II, High-Temperature Combustion Reaction Model of H2/CO/C1-C4 Compounds. Available online: https://ignis.usc.edu:80/Mechanisms/USC-Mech%20II/USC_Mech%20II.htm (accessed on 1 May 2007).
  34. Li, Y.; Zhou, C.-W.; Somers, K.P.; Zhang, K.; Curran, H.J. The oxidation of 2-butene: A high pressure ignition delay, kinetic modeling study and reactivity comparison with isobutene and 1-butene. Proc. Combust. Inst. 2017, 36, 403–411. [Google Scholar] [CrossRef]
  35. Ranzi, E.; Frassoldati, A.; Stagni, A.; Pelucchi, M.; Cuoci, A.; Faravelli, T. Reduced Kinetic Schemes of Complex Reaction Systems: Fossil and Biomass-Derived Transportation Fuels. Int. J. Chem. Kinet. 2014, 46, 512–542. [Google Scholar] [CrossRef]
  36. Ranzi, E.; Frassoldati, A.; Grana, R.; Cuoci, A.; Faravelli, T.; Kelley, A.P.; Law, C.K. Hierarchical and comparative kinetic modeling of laminar flame speeds of hydrocarbon and oxygenated fuels. Prog. Energy Combust. Sci. 2012, 38, 468–501. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the experimental setup of the heat flux method.
Figure 1. Schematic diagram of the experimental setup of the heat flux method.
Energies 16 05265 g001
Figure 2. Laminar burning velocities of CH4/natural gas (NG)/air flames as a function of ϕ at 298 K and 1 atm [4,28,29].
Figure 2. Laminar burning velocities of CH4/natural gas (NG)/air flames as a function of ϕ at 298 K and 1 atm [4,28,29].
Energies 16 05265 g002
Figure 3. Laminar burning velocities of CH4/NGOne/NGTwo/NGThree/H2/air flames as a function of ϕ for different χ H 2 at 298 K and 1 atm.
Figure 3. Laminar burning velocities of CH4/NGOne/NGTwo/NGThree/H2/air flames as a function of ϕ for different χ H 2 at 298 K and 1 atm.
Energies 16 05265 g003
Figure 4. SL enhancement for CH4/NGOne/NGTwo/NGThree/H2/air flames with 20–80% H2 added, as a function of ϕ. Symbols: experimental data.
Figure 4. SL enhancement for CH4/NGOne/NGTwo/NGThree/H2/air flames with 20–80% H2 added, as a function of ϕ. Symbols: experimental data.
Energies 16 05265 g004
Figure 5. Simulated results of the mole fractions of key radicals for H2/CH4/NGOne/air flames with different H2 contents at ϕ = 1.4.
Figure 5. Simulated results of the mole fractions of key radicals for H2/CH4/NGOne/air flames with different H2 contents at ϕ = 1.4.
Energies 16 05265 g005
Figure 6. Simulated results of the mole fractions of key radicals for H2/CH4/NGOne/air/He flames with different H2 contents at ϕ = 1.4.
Figure 6. Simulated results of the mole fractions of key radicals for H2/CH4/NGOne/air/He flames with different H2 contents at ϕ = 1.4.
Energies 16 05265 g006
Figure 7. Peak mole fraction of intermediate species as a function of H2 ratio at 1 atm and 298 K.
Figure 7. Peak mole fraction of intermediate species as a function of H2 ratio at 1 atm and 298 K.
Energies 16 05265 g007
Figure 8. Peak mole fraction, adiabatic temperature, and laminar burning velocities of intermediate species as a function of equivalence ratio in H2/CH4/air/He flames: (a) χ H 2 = 0.2, χ C H 4 = 0.8, (b) χ H 2 = 0.8, χ C H 4 = 0.2 at 1 atm and 298 K.
Figure 8. Peak mole fraction, adiabatic temperature, and laminar burning velocities of intermediate species as a function of equivalence ratio in H2/CH4/air/He flames: (a) χ H 2 = 0.2, χ C H 4 = 0.8, (b) χ H 2 = 0.8, χ C H 4 = 0.2 at 1 atm and 298 K.
Energies 16 05265 g008
Figure 9. EINOx as a function of distance, H2 ratio, and equivalence ratio in H2/CH4/air/He flames: (a) ϕ = 1.4, (b) solid line: ϕ = 1, dash line: ϕ = 1.4, (c) ϕ = 0.6–1.4, χ H 2 = 0–0.8.
Figure 9. EINOx as a function of distance, H2 ratio, and equivalence ratio in H2/CH4/air/He flames: (a) ϕ = 1.4, (b) solid line: ϕ = 1, dash line: ϕ = 1.4, (c) ϕ = 0.6–1.4, χ H 2 = 0–0.8.
Energies 16 05265 g009
Figure 10. Simulated results for the rate of production of H radical for H2/CH4/air/He flames with different H2 contents at ϕ = 1.4.
Figure 10. Simulated results for the rate of production of H radical for H2/CH4/air/He flames with different H2 contents at ϕ = 1.4.
Energies 16 05265 g010
Figure 11. Normalized sensitivity coefficients in predicting laminar burning velocities of H2/CH4/air flames under lean, stoichiometric, and fuel conditions.
Figure 11. Normalized sensitivity coefficients in predicting laminar burning velocities of H2/CH4/air flames under lean, stoichiometric, and fuel conditions.
Energies 16 05265 g011aEnergies 16 05265 g011b
Figure 12. Normalized sensitivity coefficients in predicting laminar burning velocities of H2/CH4/air/He flames under fuel-rich conditions.
Figure 12. Normalized sensitivity coefficients in predicting laminar burning velocities of H2/CH4/air/He flames under fuel-rich conditions.
Energies 16 05265 g012
Table 1. Composition of studied natural gas (NG).
Table 1. Composition of studied natural gas (NG).
Types of Natural GasMethane (%)Ethane (%)Propane (%)
NGOne981.530.47
NGTwo98.471.530
NGThree99.5300.47
Table 2. Experimental conditions adopted in this work.
Table 2. Experimental conditions adopted in this work.
CaseMole Fraction of Fuel ComponentsHeliumEquivalence
Ratio   ϕ
Unburnt Gas Temperature
Tu (K)
H2Natural Gas *
a1/b1/c1/d101No0.6–1.4298
a2/b2/c2/d20.20.8No0.6–1.4298
a3/b3/c3/d30.40.6No0.6–1.4298
a4/b4/c4/d40.60.4No0.6–1.4298
a5/b5/c5/d50.70.3Yes0.6–1.25298
a6/b6/c6/d60.80.2Yes0.6–1.3298
* Note: Natural gas indicates pure methane (a1–6), NGOne (b1–6), NGTwo (c1–6), or NGThree (d1–6).
Table 3. The beeps phenomenon occurs in this work.
Table 3. The beeps phenomenon occurs in this work.
Mole Fraction of H2 (%)Natural Gas TypeBeeps Occurrence Conditions
0.6NGThree ϕ = 1.4
0.7CH4, NGThree ϕ = 1.3
0.7NGOne, NGTwo ϕ = 1.25
0.8CH4, NGOne, NGTwo, NGThree ϕ = 1.35
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

Zhang, Z.; Zhu, R.; Zhu, Y.; Weng, W.; He, Y.; Wang, Z. Experimental and Kinetic Study on Laminar Burning Velocities of High Ratio Hydrogen Addition to CH4+O2+N2 and NG+O2+N2 Flames. Energies 2023, 16, 5265. https://doi.org/10.3390/en16145265

AMA Style

Zhang Z, Zhu R, Zhu Y, Weng W, He Y, Wang Z. Experimental and Kinetic Study on Laminar Burning Velocities of High Ratio Hydrogen Addition to CH4+O2+N2 and NG+O2+N2 Flames. Energies. 2023; 16(14):5265. https://doi.org/10.3390/en16145265

Chicago/Turabian Style

Zhang, Ziyue, Runfan Zhu, Yanqun Zhu, Wubin Weng, Yong He, and Zhihua Wang. 2023. "Experimental and Kinetic Study on Laminar Burning Velocities of High Ratio Hydrogen Addition to CH4+O2+N2 and NG+O2+N2 Flames" Energies 16, no. 14: 5265. https://doi.org/10.3390/en16145265

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