Next Article in Journal
A Two-Terminal Directional Protection Method for HVDC Transmission Lines of Current Fault Component Based on Improved VMD-Hilbert Transform
Previous Article in Journal
Sliding Mode Control of Buck DC–DC Converter with LC Input Filter
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effects of Cracking Ratio on Ammonia/Air Non-Premixed Flames under High-Pressure Conditions Using Large Eddy Simulations

1
State Key Laboratory of Clean Energy Utilization, Zhejiang University, Hangzhou 310027, China
2
Shanghai Institute for Advanced Study, Zhejiang University, Shanghai 200120, China
*
Author to whom correspondence should be addressed.
Energies 2023, 16(19), 6985; https://doi.org/10.3390/en16196985
Submission received: 2 September 2023 / Revised: 30 September 2023 / Accepted: 4 October 2023 / Published: 7 October 2023

Abstract

:
Ammonia is a promising carbon-free fuel. However, one of the main challenges for ammonia combustion is the high level of NO emissions. In this study, simulations were conducted for ammonia/air laminar counterflow flames and turbulent non-premixed jet flames in the KAUST high-pressure combustion duct (HPCD) at a pressure of 5 bar, with two ammonia cracking ratios of 14% and 28%. The influence of ammonia cracking ratio on the flame structure and NO formation mechanism were examined. The laminar counterflow flame results showed that HNO is one of the most critical species related to NO formation and NO is mainly generated through the path of NH 2 NH HNO NO . For the turbulent flames, the flamelet/progress variable (FPV) approach was employed in the context of large eddy simulations (LES) for high-fidelity simulations. The simulation results were compared with the measured data with promising agreements, which proves the accuracy of the FPV method for the present flames. It was shown that with increasing cracking ratio, not only the flame reactivity is enhanced, but also the generation of NO is increased. The correlation between NO and HNO is weaker when compared to that between NO and radicals such as O , H and OH in the entire flame. Through the distribution of NO source terms, it was found that the NO source term has a higher absolute value in the upstream region and the absolute value rapidly decreases with increasing streamwise distance. The total NO source term is positive in the fuel-lean zone and shows negative values in the fuel-rich zone.

1. Introduction

Ammonia has been widely applied in the global industrial chain. Its production and transportation infrastructures are relatively mature [1,2]. Most importantly, as a carbon-free fuel, ammonia does not produce CO 2 in its chemical conversion process, greatly reducing greenhouse gas emissions [3]. However, ammonia combustion features a flame speed much lower than traditional hydrocarbon fuels or hydrogen, low calorific value [4] and large amounts of harmful nitrogen monoxide produced during combustion, which presents challenges for its industrial application [5].
Previous studies showed that using blend fuels such as NH 3 / CH 4 or NH 3 / H 2 can increase flame speed and improve heat release rate [6,7] compared with pure ammonia, thereby extending the flame stability limits [8,9,10]. Khateeb et al. [11] showed that in a laboratory-scale swirl burner NH 3 / CH 4 mixtures could generate a larger stability range than pure methane and also reduce the tendency of flame flashback. Otomo et al. [12] found that adding H 2 to NH 3 can increase flame temperature and improve flame speed. Frigo et al. [13] demonstrated that ammonia could directly burn in internal combustion engines, but required a combustion promoter; the best carbon-free promoter is hydrogen, with a very high combustion speed and a wide combustion range.
Obtaining a mixture of ammonia and hydrogen via on-board (in situ) cracking of ammonia is an ideal choice to avoid the risks and costs of storing and distributing hydrogen [14]. Mei et al. [15] conducted experimental and numerical simulations on partially cracked NH 3 /air mixtures under high-pressure conditions. They found that as the ammonia cracking ratio increases, chemical and transport effects dominate the flame speed, whereas the thermal effect plays a minor role. The formation of NO with an increase in the cracking ratio exhibits non-monotonic behavior. Shohdy et al. [16] investigated how the cracking ratio of ammonia affects the overall flame speed, flammable limits and NO x concentration in a swirling premixed combustor. Mercier et al. [17] explored the impact of ammonia cracking levels on the performance and emissions of a single-cylinder spark ignition engine, finding that a very low cracking ratio of ammonia (10%) can significantly expand the engine’s operating range. Kim et al. [18] employed a chemical reactor network to numerically study the NO x emission characteristics of air-staged and fuel-staged combustion of partially cracked NH 3 /air mixtures, revealing the impact of ammonia cracking ratio, residence time and staging strategy on NO x emissions. It is clear that with partial ammonia cracking, there is a clear trade-off: high ammonia cracking ratios lead to high nitrogen oxide levels, while low ratios result in high unburned ammonia levels.
Recently, Tang et al. [19] at KAUST reported experimental results of ammonia/air non-premixed flames with different ammonia cracking ratios of 14% and 28%. Both laminar counterflow flames and turbulent jet flames were considered. The experiments provided detailed quantitative information about the temperature, mass fraction and mixture fraction of the flames. Later, Wang et al. [20] used the one-dimensional NO laser-induced fluorescence ( NO -LIF) method combined with one-dimensional Raman spectroscopy to quantitatively obtain the NO mole fraction for these flames. The experimental configurations are featured by well-defined boundary conditions and the measurements of the species mass fractions, temperature and main pollutant NO mole fractions are comprehensive. Therefore, they are chosen as the target flames of the present work.
Abdelwahid et al. [21] proposed a PC-transport model based on principal component analysis (PCA) and deep neural networks (DNN) for simulating the above mentioned flames. The authors showed the potential of the proposed model for turbulent ammonia/air non-premixed combustion. On the other hand, the flamelet model represented by the flamelet/progress variable (FPV) method is based on the fast chemistry assumption. It is suggested that the thermochemical state in turbulent flames can be described by a few control variables [22]. The flamelet model has been widely used in large eddy simulations (LES) of turbulent premixed [23,24,25,26] and non-premixed [27,28,29] flames. However, it has been rarely used for the study of non-premixed NH 3 /air flames with partial ammonia cracking.
In this context, the focus of the present work is on the flame structure and NO formation mechanism of turbulent ammonia/air non-premixed combustion with various ammonia cracking ratios, which was not covered by the study of Abdelwahid et al. [21]. The objectives are as follows. First, the relative importance of different NO formation pathways with various ammonia cracking ratios is examined in laminar counterflow flames. Second, the LES results of turbulent non-premixed jet flames using the FPV model are compared with experimental results, validating the accuracy of the numerical methods. The effects of the ammonia cracking ratio on the flame structure are studied. Finally, the mechanisms of NO formation under different ammonia cracking ratio conditions are explored.

2. Experimental and Numerical Setup

2.1. Experimental Details

In this work, the experimental configurations of laminar counterflow flames and turbulent non-premixed jet flames are considered. The experiments were conducted in the High-Pressure Combustion Duct (HPCD) at the KAUST Clean Combustion Research Center [30], with an operating pressure of 5 bar and inflow temperature of 294 K. The counterflow burner is horizontally installed inside the HPCD, composed of two identical nozzles with a diameter of 5 mm and a distance of 5 mm. Two cases with different ammonia cracking ratios were investigated. Specifically, the ammonia cracking ratio is 14% for CACF14, while it is 28% for CACF28. The fuel composed of an NH 3 / H 2 / N 2 mixture is injected from one nozzle and the air from the other, with a strain rate of 130 1/s [31].
The turbulent jet burner consists of a 0.57-m-long stainless steel tube, supplying the fuel mixture NH 3 / H 2 / N 2 , surrounded by an air co-flow nozzle with a diameter of 0.25 m, as shown in Figure 1. The inner and outer diameters of the central jet tube are 4.58 mm and 6.35 mm, respectively. Again, two cases with different ammonia cracking ratios of 14% and 28% are considered in the experiments, which are denoted as CAJF14 and CAJF28, respectively. The Reynolds numbers of the turbulent flames are Re = 11,200 for both cases based on the jet velocity and nozzle diameter. Detailed experimental parameters for various cases are provided in Table 1.

2.2. Computational Setup

Figure 2 shows a schematic of the configurations for the laminar counterflow flames and turbulent jet flames. For the laminar counterflow flames, the length of the computational domain is the same as the distance between the two nozzles. The domain is discretized using 150 grids. The information of the composition and velocity in the two streams is provided in Table 1. The strain rate is 130 1/s, which is consistent with the experiment [19]. As for the turbulent jet flames, the length of the computational domain is 80D, which is sufficient to avoid the influence of the outlet boundary on the region of interest, and the diameter is 35D. The simulation uses a high-quality structured hexahedral grid with 258,000 cells. The grid is refined in the streamwise and radial directions to achieve a high resolution in the reaction zone. To make the turbulence at the inlet more realistic, a fully developed circular tube flow is used to generate the turbulence at the inlet. A fixed reference pressure with a zero-gradient boundary condition is set at the outlet of the burner.
The one-dimensional simulations of laminar counterflow flames are performed using ANSYS Chemkin-PRO [32], in which the OPPDIF model developed by Kee et al. [33] was used. This model is based on a similarity solution for the governing equations, simplifying two-dimensional or three-dimensional flows mathematically to one dimension. Unity Lewis number transport model is used for the one-dimensional calculations. In the laminar flames, a focus is on the distributions of main species mass fractions, temperature and NO mole fractions. Four different mechanisms of Mathieu [7], Otomo [12], Okafor [34] and Jiang [35] are employed for the simulations. The Mathieu mechanism consists of 55 species and 275 elementary reactions, and accurately predicts the ammonia ignition delay time and laminar flame propagating velocity, as validated in refs. [36,37]. The Otomo mechanism includes 32 species and 213 elementary reactions, which was developed based on the mechanism of Song et al. [38], and can also predict the laminar flame propagating velocity and ignition delay times of NH 3 combustion very well [39]. The Okafor mechanism, featured by 59 species and 356 elementary reactions, is a reduced reaction mechanism developed for the combustion of CH 4 / NH 3 flames optimized for laminar burning velocities [40]. It has been verified for both the pure NH 3 /air flame and NH 3 / CO / air flames [41]. The Jiang mechanism was derived from San Diego’s short nitrogen mechanism [42] and consists of 19 species and 60 elementary reactions, which effectively reduces time consumption in reactive flow simulations while still adhering to acceptable accuracy thresholds.
The three-dimensional large-eddy simulations are carried out using a solver called FPVFoam, which is developed based on the open-source code OpenFOAM [43]. The PIMPLE algorithm is adopted to solve the transport equations. The second-order backward scheme is used for the unsteady term and the second-order Gauss linear is used for diffusion and convective terms. The time step is set to 2 × 10 6 s to keep the maximum Courant–Friedrich–Lewy (CFL) number less than 0.5 under all conditions. Only the Okafor mechanism is selected for the LES, which provides promising predictions for the turbulent flames. For both cases, the simulations ran for at least 20 flow-through times to reach a statistically steady state. The physical time for collecting statistical data exceeds 100 ms, equivalent to three flow-through times.

3. Numerical Methodology

3.1. Conservation Equations for LES

In the framework of large eddy simulation, the filtered equations of mass, momentum, mixture fraction, mixture fraction variance [44] and progress variable [45] are written as follows:
ρ ¯ t + ρ ¯ u ˜ j x j = 0
ρ ¯ u ˜ i u ˜ i t + ρ u ˜ i u ˜ j x j = p ¯ x i + x j μ ¯ + μ t u ˜ i x j + u ˜ j x i 2 3 δ i j u ˜ k x k
( ρ ¯ Z ˜ ) t + ρ ¯ u ˜ i Z ˜ x i = x i ρ ¯ D ˜ + μ t S c t Z ˜ x i
ρ ¯ Z ˜ 2 t + ρ ¯ u ˜ i Z ˜ 2 x i = x i ρ ¯ D ˜ + μ t S c t Z ˜ 2 x i + 2 ρ ¯ D ˜ + μ t S c t Z ˜ x i 2 2 ρ ¯ χ ˜
ρ ¯ Y ˜ c t + ρ ¯ u ˜ j Y ˜ c x j = x j ρ ¯ D ˜ + μ t S c t Y c ˜ x j + ρ ¯ ω Y c , ˙ ˜
where ϕ ˜ and ϕ ¯ represent the density-weighted Favre filtering and spatial filtering for a variable ϕ , respectively. ρ denotes the density, u i the velocity component in the ith direction, p the pressure, and μ dynamic viscosity. μ t is the subgrid eddy viscosity, which is closed by the dynamic Smagorinsky model [46]. δ i j denotes the Kronecker delta function. The mixture fraction Z is calculated based on the Bilger’s method [47] as follows:
Z = 2 Y C Y C , o / M W C + 0.5 Y H Y H , o / M W H Y O Y O , o / M W O 2 Y C , f Y C , o / M W C + 0.5 Y H , f Y H , o / M W H Y O , f Y O , o / M W O ,
where Y and M W are the elemental mass fraction and atomic mass, respectively. The subscripts f and o represent fuel and oxidizer, respectively. S c t is the turbulent Schmidt number. ω Y c ˙ is the source term of the progress variable, which is extracted from the flamelet library. In the equations, the molecular diffusivity is calculated via the unity Lewis number assumption D = λ / ρ c p . χ represents the scalar dissipation rate. The filtered scalar dissipation rate is modeled according to the method in Ref. [48], which is composed of two parts, i.e., the resolved large-scale component and the unresolved sub-grid scale (SGS) component:
χ ˜ = D ˜ | Z ˜ | 2 + μ t S c t C χ 2 Δ 2 Z ˜ 2 ,
where Δ represents the filter width and C χ is a model constant.

3.2. Flamelet/Progress Variable (FPV) Model

In the FPV approach, a flamelet library is required [45]. In this work, the library is constructed using FlameMaster [49] by solving one-dimensional steady-state counterflow flames. In this approach, all the thermochemical variables (including temperature, species mass fractions, thermodynamic and transport properties) can be parameterized using two control variables, i.e., the mixture fraction Z and the progress variable Y c . The thermochemical scalars can be extracted from the flamelet library as follows:
ϕ = ϕ Z , Y c .
The progress variable Y c is typically defined as the sum of the main species mass fractions of the products [50]. In this work, the progress variable is defined based on the mass fraction of H 2 O . The normalized progress variable is calculated as follows:
C = Y c Y c m i n ( Z ) Y c m a x ( Z ) Y c m i n ( Z ) ,
where Y c m a x ( Z ) and Y c m i n ( Z ) are the local maximum and minimum of the progress variable, respectively, which can be extracted from the flamelet library. With the normalized progress variable C, the flamelet library solutions can be parameterized as follows:
ϕ = ϕ Z , C .
In the FPV model, the interactions of turbulence and combustion are considered by combining the thermochemical quantities with the joint probability density function (PDF) [44]:
ϕ ˜ = 0 1 0 1 ϕ Z , C P ˜ Z , C d Z d C .
Here, the δ -PDF is employed for the progress variable C, while the β -PDF is used for the mixture fraction Z. Therefore, all thermochemical scalars in the flamelet table, including fluid density, flame temperature and chemical species mass fractions, are represented in terms of Z ˜ , Z ˜ 2 and C ˜ :
ϕ ˜ = ϕ ˜ Z ˜ , Z ˜ 2 , C ˜ .
In the flamelet table, the coordinates for mixture fraction, mixture fraction variance and progress variable are discretized using 251, 7 and 251 uniformly spaced grid points, respectively. In total, over 350 flamelets were computed to construct the table.

3.3. The NO Prediction Model

The flamelet model is based on the fast chemistry assumption, where chemical reactions occur at a sufficiently fast rate compared to the turbulent Kolmogorov time scale [51]. However, the reaction rate of NO is relatively slow, and in actual turbulent flames, the concentration of NO often deviates far from its equilibrium state within a short residence time. Previous studies have shown that directly calculating NO using lookup tables in the flamelet model can lead to very serious errors [52,53].
In order to overcome the inherent defects of inaccurate prediction of NO in traditional flamelet models and predict the NO in the ammonia/air combustion accurately, the approach of Ihme et al. [53] is used, in which an extra transport equation is solved for NO . Specifically, the forward and backward NO reaction rates are modeled with the reaction rates taken from the flamelet library, respectively. The backward NO reaction rates directly related to NO concentration are corrected by NO obtained in the flow field.
ρ ¯ Y ˜ NO t + ρ ¯ u ˜ j Y ˜ NO x j = x j ρ ¯ D ˜ + μ t S c t Y ˜ NO x j + ρ ¯ ω NO ˙ ˜
ω NO ˙ ˜ = ω NO ˙ + ˜ + Y ˜ NO ω NO ˙ ˜ Y ˜ NO FPV
In the above equations, Y ˜ NO FPV is the mass fraction of species NO obtained directly from the flamelet table. ω NO ˙ + ˜ and ω NO ˙ ˜ are the filtered source terms of NO generation and consumption, respectively. To calculate the NO source term, all the reactions that involve NO are considered. The reactions involving NO in the Okafor mechanism are listed in Table 2.

4. Results and Discussions

4.1. 1D Laminar Counterflow Flame Simulation

4.1.1. Mechanism Comparison

Figure 3 shows the profiles of major species mass fractions and temperature in the mixture fraction space. It can be seen that the predictions of the four mechanisms almost completely overlap. The simulation results for the species mass fractions and temperature match the experimental results quite well. Some minor discrepancies are observed in the fuel-rich region for the H 2 mass fraction. This can be attributed to the high mass fraction of H 2 at the positions with large mixture fractions, leading to a pronounced preferential diffusion effect.
Figure 4 shows the profiles of NO mole fractions in the mixture fraction space. The predictions generally agree with the experimental results, especially in the fuel-lean side. In the fuel-rich side, the NO mole fraction is under-predicted for all mechanisms. The experimental results show that the peak NO mole fraction of CACF14 is larger than that of CACF28, so that the Mathieu and Jiang mechanisms, where the peak mole fraction of CACF14 is less than that of CACF28, are excluded for further study. The Okafor mechanism has a peak NO mole fraction closer to the experimental measurement value under the current unity Lewis number assumption. Therefore, the Okafor mechanism is chosen for subsequent study.

4.1.2. Reaction Path Analysis

In order to explore the impact of ammonia cracking ratio on the formation of nitrogen-containing species, especially NO formation, this section performs nitrogen flow analysis, which is implemented by integrating the reactions across the entire field to calculate the nitrogen flow fraction between relevant species [54]. The nitrogen flow between species A and B is calculated by summing the average integrated reaction rates of all elementary reactions that contribute to A → B. Under the current conditions, almost 100% of NH 3 is converted to NH 2 ; thus, the nitrogen flow is normalized using the sum of the average integrated reaction rates of all elementary reactions that contribute to NH 3 NH 2 .
Figure 5 presents the nitrogen flow diagrams of major paths for CACF14 and CACF28. It is worth noting that, for simplicity, the reaction paths in Figure 5 are not complete. As can be seen, the species directly related to NO formation are HNO , N and NH . Among them, the species that contributes the most to NO generation is HNO . The second most important species is N , while NH contributes the least to NO generation. As the cracking ratio increases, the amount of NO generated through N increases, but the amount of NO generated through species HNO and NH remains basically unchanged. It can also be seen that NO is mainly generated through the following four paths: NH 2 HNO NO ( path a ) , NH 2 NH HNO NO ( path b ) , NH 2 NH NO ( path c ) and NH 2 NH N NO ( path d ) . On the other hand, the products of NO consumption are N 2 O , NNH and N 2 . As the cracking ratio increases, the amount of NO directly converted to N 2 increases, but the conversion to NNH and N 2 O remains basically unchanged. NO is mainly consumed through NO N 2 O N 2 ( path e ) , NO N 2 O NNH N 2 ( path f ) , NO N 2 ( path g ) and NO NNH N 2 ( path h ) .
In order to study the influence of the ammonia cracking ratio on the conversion of nitrogen-containing species and the relative magnitudes of the NO pathways, a quantitative analysis is conducted in terms of the conversion rate, which is defined as the ratio of the net outflow of nitrogen from species A to species B to the total net outflow of nitrogen from species A. Figure 6a–c present the conversion rates of NH 2 , NH , HNO and N for various flames.
Figure 6a shows that most NH 2 is directly converted to NH , and the ratio of this conversion increases with the increasing cracking ratio. According to Figure 6b, NH is mostly converted to HNO and N . As the cracking ratio increases, the conversion ratio of NH to these two species increases, with the conversion to N being more significant. Figure 6c shows that HNO is entirely converted to NO , while N is mostly converted to NO and N 2 .
The relative importance of NO pathways is quantitatively assessed, as shown in Figure 6d. It is obvious that the most important pathway for NO generation is path b, followed by path d. As the ammonia cracking ratio increases, the generation of NO through paths b and d increases, while that through path a decreases. For path c, the change in the cracking ratio does not lead to significant change in NO formation.
Figure 7 shows the profiles of mole fractions for various radicals in the mixture fraction space. It is seen that the concentrations of HNO and NH in CACF28 are lower than those in CACF14, which is due to the higher hydrogen concentration and lower NH i ( i = 1 , 2 , 3 ) concentration related to the higher cracking ratio in CACF28, which further leads to a lower concentration of HNO [20]. According to Figure 5, HNO is an important radical for NO generation [55]. Therefore, the mole fraction of NO in CACF28 is lower than that in CACF14, as shown in Figure 4. In addition, it can be seen that the concentrations of radicals such as H , OH and O in CACF28 are higher than those in CACF14, which is attributed to the higher H 2 concentration associated with the larger cracking ratio, so the reactivity of CACF28 is stronger than that of CACF14.

4.2. 3D Turbulent Jet Flame Simulation

4.2.1. Simulation Verification

In this section, the LES results of the two turbulent flames, i.e., CAJF14 and CAJF28, are compared with the experimental results. Figure 8 shows mean and root-mean-square (RMS) values of the mixture fraction Z and progress variable Y c at various streamwise locations. For the mean values, it can be observed that the simulation results demonstrate a favorable agreement with the measured data across the entire domain. However, the mean of Y c is under-predicted near the peaks in the upstream region X / D = 5 , which is more pronounced in the CAJF28 flame with a higher cracking ratio. This is attributed to the more prominent differential diffusion effects near the nozzle region [19,21]. It is evident that at the streamwise locations X / D > 5 , turbulent diffusion overtakes molecular diffusion, leading to a good performance of simulations. For the RMS values of the mixture fraction, the peaks initially experience a slight increase along the streamwise direction and subsequently decrease as the flame progresses. It can be observed that for the progress variable, two RMS peaks are observed, corresponding to the lean and rich sides, as shown in the experiments. The RMS peak in the rich side becomes less pronounced with increasing streamwise distance, attributed to the more uniform distribution of jet temperature and composition due to combustion product entrainment. The RMS peak in the lean side gradually increases with the streamwise location.
Figure 9 presents the mean and RMS profiles of Y NH 3 , Y H 2 and the temperature T. In terms of the mean values, similar to the mixture fraction, the measured and predicted values of Y NH 3 and Y H 2 exhibit a good match throughout the entire field. This indicates the satisfactory performance of the unit Lewis number transport model under the current operating conditions. The profile of temperature T resembles that of the progress variable. Both flames exhibit varying degrees of under-prediction for temperature in the upstream region X / D = 5 , with this effect being more pronounced for CAJF28. Regarding the RMS values, the profiles of Y NH 3 and Y H 2 show a single peak, similar to the mixture fraction Z. The peak value increases first with the streamwise position and then decreases. The temperature T profile exhibits a double-peak distribution in the upstream, similar to that of the progress variable.
Figure 10 shows the mean and RMS profile of X NO . Notably, for the mole fraction of NO , the numerical method of the present wrok also produces satisfactory results, affirming the accuracy of the employed NO model. Nevertheless, in the downstream region at X / D = 40 , both cases display varying degrees of over-prediction in the NO mole fraction. The peak mole fraction of NO appears at X / D = 5 . With increasing streamwise distance due to the expansion of the jet, the radial position of the peak NO mole fraction gradually increases. Meanwhile, the peak mole fraction of NO decreases along the streamwise direction. The RMS values of X NO also agree well with the experiment results.
In summary, both the mean and RMS values of the simulations successfully capture the characteristics of the experimental flame [19]. This confirms the rationality of the numerical methods and boundary conditions, providing a solid foundation for subsequent analyses.

4.2.2. Flame structure analysis

By examining the distributions of various quantities in mixture fraction space, a deeper understanding of the flame structure can be achieved. Figure 11 and Figure 12 show the scatter plots of NO mole fraction and temperature, respectively, in mixture fraction space at two streamwise locations of X / D = 5 and X / D = 20 . The conditional means and flamelet solutions with different scalar dissipation rates are also presented. It is shown that as the scalar dissipation rate increases, the mole fraction of NO in the steady flamelet branch increases while the temperature gradually decreases. This indicates that in both flames, the pathway of thermal NO is not dominant. It is also evident that with an increase in streamwise distance, the conditional mean values of the NO mole fraction approach the flamelet solutions with lower scalar dissipation rates.
In Figure 12, dashed lines illustrate the unsteady flamelet solutions for different scalar dissipation rates. It can be seen that at the streamwise location X / D = 5 , a considerable number of data points are distributed over the unsteady flamelet solutions. However, as the streamwise distance increases, the level of local extinction decreases. By comparing CAJF14 and CAJF28, it is apparent that CAJF28 shows less local extinction compared to CAJF14. This is attributed to the elevated H 2 concentration resulting from a higher cracking ratio, which enhances the reaction activity and thus promotes the completeness of combustion.

4.2.3. NO Formation Mechanism

Figure 13 displays the time-averaged distributions of NO mole fraction for CAJF14 and CAJF28. It can be seen that NO is primarily observed in a narrow region upstream near the stoichiometric mixture fraction Z s t where fuel and oxidizer meet, and a broader region downstream as the jet flame develops. In the downstream region, the peak NO concentration for CAJF28 occurs at a more upstream location and the peak concentration is higher. This means that when the cracking ratio increases, not only is the flame reactivity enhanced but also the generation of NO is increased for the turbulent flames. In addition, compared with Figure 4, it can be found that the impact of changes in the ammonia cracking ratio on NO production seems to be more significant in turbulent flames. This is because in the turbulent jet flame, there are complex interactions between turbulence and the formation of NO , which might result in a more pronounced difference of NO formation for the two turbulent flames with different ammonia cracking ratios.
In the analysis of laminar flames, it has been shown that intermediates such as HNO , NH and N are directly associated with NO generation. Figure 14 shows time-averaged distributions of key species in the turbulent jet flames. As can be seen, in the upstream region of the flames, similar to the laminar flames, a higher cracking ratio results in a lower NH concentration, which in turn leads to a reduced HNO level. As the streamwise distance increases, the magnitude of nitrogen-containing species concentrations decreases rapidly, indicating that the downstream regions approach the equilibrium state more closely than the upstream regions. Radicals such as OH are highly indicative of reaction zones and are commonly used to characterize turbulent flames [24], which also play a role in NO generation in ammonia flames [56,57]. Time-averaged distributions of O , H and OH concentrations are presented in Figure 14d–f. It can be observed that in the upstream regions of the flames, similar to the laminar flame results, the higher cracking ratio in CAJF28 results in higher radical concentrations compared to CAJF14. The differences in O , H and OH concentrations between the two flames remain distinct in the downstream regions. Furthermore, based on the distributions of radicals, it is concluded that the flame length of CAJF28 is shorter than that of CAJF14. This observation is consistent with the long exposure duration pictures obtained from the experimental results [20], indicating that with increasing cracking ratio, more intense combustion is achieved and the fuel is consumed more rapidly.
To gain a deeper understanding of NO formation in turbulent jet flames, two representative locations are chosen for the following analysis, i.e., X / D = 5 and 60. From Figure 13, it is evident that these two locations correspond to the regions where the NO mole fraction peaks. Figure 15 illustrates the scatter plots between important radical concentrations with the NO concentration at various locations. It is apparent that the correlation between NO and HNO is weaker when compared to that between NO and radicals such as O , H and OH , both in the upstream and downstream regions. Figure 15b–d indicate that the radicals O , H and OH exhibit a positive correlation with NO , confirming the a role that these radicals play in NO generation in ammonia flames.
The following analysis focuses on the influence of varying ammonia cracking ratios on the distribution of NO source terms, the information of which is not available from the experiments. The combustion of NH 3 / H 2 mixture mainly involves two mechanisms for NO formation. The first is the thermal NO mechanism, also known as the Zeldovich mechanism, which includes three reactions, i.e., N + NO N 2 + O , N + O 2 NO + O and N + OH NO + H . The second mechanism is the fuel mechanism, i.e., NH i + OX NO + H i X , where OX is an oxygen-containing species [58]. Figure 16 shows the mean distributions of the fuel, thermal and total NO source terms of both flames. It can be seen that the NO source term has a higher absolute value in the upstream region, and the absolute value of the source term rapidly decreases with the increase in the streamwise distance. The total NO source term is positive in the fuel-lean zone and shows negative values in the fuel-rich zone. In addition, it can be observed that the fuel NO in most regions is negative, while the thermal NO is positive.

5. Conclusions

In this study, simulations of laminar counterflow and turbulent non-premixed ammonia/air flames in HPCD were conducted. Two cases with different ammonia cracking ratios of 14% and 28% were investigated. The influence of changes in the ammonia cracking ratio on flame structure and NO formation mechanism under high-pressure conditions were examined. The main findings are summarized as follows:
  • Through nitrogen flow analyses, the conversion of nitrogen-containing species and the NO pathways were examined. It was found that the species that contributes the most to NO generation is HNO , and the most important pathway for NO generation is NH 2 NH HNO NO .
  • The simulation and experimental results agree quite well, and the simulations provided fairly accurate predictions for NO , proving the applicability of the FPV method in hydrogen–ammonia flames. The flame structure and NO formation of turbulent jet flames were examined. It was found that with increasing cracking ratio, not only is the flame reactivity enhanced but the generation of NO is also increased for the turbulent flames.
  • The correlation between important radical concentrations and NO concentration was examined. We showed that the correlation between NO and HNO is weaker when compared to that of NO and radicals such as O , H and OH throughout the entire field. Moreover, the increased cracking ratio also results in a higher level of radical concentrations.
  • The distribution of NO source terms were also explored. It was found that the NO source term has a higher absolute value upstream, and the absolute value rapidly decreases with the increase in the streamwise distance. The total NO source term is positive in the fuel-lean zone and shows negative values in the fuel-rich zone. In addition, the fuel NO in most regions is negative, while the thermal NO is positive.

Author Contributions

Conceptualization, C.W. and H.W.; methodology, C.W.; software, C.W.; validation, C.W. and H.W.; formal analysis, C.W.; investigation, C.W.; resources, C.W.; data curation, C.W. and H.W.; writing—original draft preparation, C.W. and H.W.; writing—review and editing, C.W., H.W., K.L. and J.F.; visualization, C.W.; supervision, H.W.; project administration, H.W., K.L. and J.F.; funding acquisition, H.W., K.L. and J.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of China (Grant Nos.: 52022091, 51976185, 51925603). The work was also supported by the Fundamental Research Funds for the Central Universities, China (Grant No: 2021FZZX001-10).

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Smith, C.; Hill, A.K.; Torrente-Murciano, L. Current and Future Role of Haber–Bosch Ammonia in a Carbon-Free Energy Landscape. Energy Environ. Sci. 2020, 13, 331–344. [Google Scholar] [CrossRef]
  2. Yapicioglu, A.; Dincer, I. A Review on Clean Ammonia as a Potential Fuel for Power Generators. Renew. Sustain. Energy Rev. 2019, 103, 96–108. [Google Scholar] [CrossRef]
  3. Awad, O.I.; Zhou, B.; Harrath, K.; Kadirgama, K. Characteristics of NH3/H2 Blend as Carbon-Free Fuels: A Review. Int. J. Hydrogen Energy 2022. is press. [Google Scholar] [CrossRef]
  4. Hayakawa, A.; Goto, T.; Mimoto, R.; Arakawa, Y.; Kudo, T.; Kobayashi, H. Laminar Burning Velocity and Markstein Length of Ammonia/Air Premixed Flames at Various Pressures. Fuel 2015, 159, 98–106. [Google Scholar] [CrossRef]
  5. Kobayashi, H.; Hayakawa, A.; Somarathne, K.D.K.A.; Okafor, E.C. Science and Technology of Ammonia Combustion. Proc. Combust. Inst. 2019, 37, 109–133. [Google Scholar] [CrossRef]
  6. Li, J.; Huang, H.; Kobayashi, N.; He, Z.; Nagai, Y. Study on Using Hydrogen and Ammonia as Fuels: Combustion Characteristics and NOx Formation: Hydrogen and Ammonia as Fuels. Int. J. Energy Res. 2014, 38, 1214–1223. [Google Scholar] [CrossRef]
  7. Mathieu, O.; Petersen, E.L. Experimental and Modeling Study on the High-Temperature Oxidation of Ammonia and Related NOx Chemistry. Combust. Flame 2015, 162, 554–570. [Google Scholar] [CrossRef]
  8. Li, S.; Zhang, S.; Zhou, H.; Ren, Z. Analysis of Air-Staged Combustion of NH3/CH4 Mixture with Low NOx Emission at Gas Turbine Conditions in Model Combustors. Fuel 2019, 237, 50–59. [Google Scholar] [CrossRef]
  9. Xiao, H.; Lai, S.; Valera-Medina, A.; Li, J.; Liu, J.; Fu, H. Study on Counterflow Premixed Flames Using High Concentration Ammonia Mixed with Methane. Fuel 2020, 275, 117902. [Google Scholar] [CrossRef]
  10. Kurata, O.; Iki, N.; Matsunuma, T.; Inoue, T.; Tsujimura, T.; Furutani, H.; Kobayashi, H.; Hayakawa, A. Performances and Emission Characteristics of NH3–Air and NH3CH4–Air Combustion Gas-Turbine Power Generations. Proc. Combust. Inst. 2017, 36, 3351–3359. [Google Scholar] [CrossRef]
  11. Khateeb, A.A.; Guiberti, T.F.; Zhu, X.; Younes, M.; Jamal, A.; Roberts, W.L. Stability Limits and Exhaust NO Performances of Ammonia-Methane-Air Swirl Flames. Exp. Therm. Fluid Sci. 2020, 114, 110058. [Google Scholar] [CrossRef]
  12. Otomo, J.; Koshi, M.; Mitsumori, T.; Iwasaki, H.; Yamada, K. Chemical Kinetic Modeling of Ammonia Oxidation with Improved Reaction Mechanism for Ammonia/Air and Ammonia/Hydrogen/Air Combustion. Int. J. Hydrogen Energy 2018, 43, 3004–3014. [Google Scholar] [CrossRef]
  13. Frigo, S.; Gentili, R. Analysis of the Behaviour of a 4-Stroke Si Engine Fuelled with Ammonia and Hydrogen. Int. J. Hydrogen Energy 2013, 38, 1607–1615. [Google Scholar] [CrossRef]
  14. Comotti, M.; Frigo, S. Hydrogen Generation System for Ammonia–Hydrogen Fuelled Internal Combustion Engines. Int. J. Hydrogen Energy 2015, 40, 10673–10686. [Google Scholar] [CrossRef]
  15. Mei, B.; Zhang, J.; Shi, X.; Xi, Z.; Li, Y. Enhancement of Ammonia Combustion with Partial Fuel Cracking Strategy: Laminar Flame Propagation and Kinetic Modeling Investigation of NH3/H2/N2/Air Mixtures up to 10 Atm. Combust. Flame 2021, 231, 111472. [Google Scholar] [CrossRef]
  16. Shohdy, N.N.; Alicherif, M.; Lacoste, D.A. Transfer Functions of Ammonia and Partly Cracked Ammonia Swirl Flames. Energies 2023, 16, 1323. [Google Scholar] [CrossRef]
  17. Mercier, A.; Mounaïm-Rousselle, C.; Brequigny, P.; Bouriot, J.; Dumand, C. Improvement of SI Engine Combustion with Ammonia as Fuel: Effect of Ammonia Dissociation Prior to Combustion. Fuel Commun. 2022, 11, 100058. [Google Scholar] [CrossRef]
  18. Kim, N.; Lee, M.; Park, J.; Park, J.; Lee, T. A Comparative Study of NOx Emission Characteristics in a Fuel Staging and Air Staging Combustor Fueled with Partially Cracked Ammonia. Energies 2022, 15, 9617. [Google Scholar] [CrossRef]
  19. Tang, H.; Yang, C.; Wang, G.; Krishna, Y.; Guiberti, T.F.; Roberts, W.L.; Magnotti, G. Scalar Structure in Turbulent Non-Premixed NH3/H2/N2 Jet Flames at Elevated Pressure Using Raman Spectroscopy. Combust. Flame 2022, 244, 112292. [Google Scholar] [CrossRef]
  20. Wang, G.; Tang, H.; Yang, C.; Magnotti, G.; Roberts, W.L.; Guiberti, T.F. Quantitative Laser-Induced Fluorescence of NO in Ammonia-Hydrogen-Nitrogen Turbulent Jet Flames at Elevated Pressure. Proc. Combust. Inst. 2023, 39, 1465–1474. [Google Scholar] [CrossRef]
  21. Abdelwahid, S.; Malik, M.R.; Al Kader Hammoud, H.A.; Hernández-Pérez, F.E.; Ghanem, B.; Im, H.G. Large Eddy Simulations of Ammonia-Hydrogen Jet Flames at Elevated Pressure Using Principal Component Analysis and Deep Neural Networks. Combust. Flame 2023, 253, 112781. [Google Scholar] [CrossRef]
  22. Tang, J. Study and Application of NOx Prediction Model and Flamelet Tabulation Method for Aeroengine Combustion Chamber. Ph.D. Thesis, Northwestern Polytechnical University, Xi’an, China, 2018. [Google Scholar]
  23. Tang, Y.; Raman, V. Large Eddy Simulation of Premixed Turbulent Combustion Using a Non-Adiabatic, Strain-Sensitive Flamelet Approach. Combust. Flame 2021, 234, 111655. [Google Scholar] [CrossRef]
  24. An, Z.; Zhang, M.; Zhang, W.; Mao, R.; Wei, X.; Wang, J.; Huang, Z.; Tan, H. Emission Prediction and Analysis on CH4/NH3/Air Swirl Flames with LES-FGM Method. Fuel 2021, 304, 121370. [Google Scholar] [CrossRef]
  25. Honzawa, T.; Kai, R.; Okada, A.; Valera-Medina, A.; Bowen, P.J.; Kurose, R. Predictions of NO and CO Emissions in Ammonia/Methane/Air Combustion by LES Using a Non-Adiabatic Flamelet Generated Manifold. Energy 2019, 186, 115771. [Google Scholar] [CrossRef]
  26. Kai, R.; Tokuoka, T.; Nagao, J.; Pillai, A.L.; Kurose, R. LES Flamelet Modeling of Hydrogen Combustion Considering Preferential Diffusion Effect. Int. J. Hydrogen Energy 2022, 48, 11086–11101. [Google Scholar] [CrossRef]
  27. Jiang, X.; Guo, J.; Wei, Z.; Quadarella, E.; Im, H.G.; Liu, Z. A Species-Weighted Flamelet/Progress Variable Model with Differential Diffusion Effects for Oxy-Fuel Jet Flames. Combust. Flame 2023, 251, 112674. [Google Scholar]
  28. Gao, Z.; Wang, H.; Song, C.; Luo, K.; Fan, J. Large-Eddy Simulation of Hydrothermal Flames Using Extended Flamelet/Progress Variable Approach. J. Supercrit. Fluids 2020, 163, 104843. [Google Scholar] [CrossRef]
  29. Wang, S.; Wang, H.; Luo, K.; Fan, J. The Effects of Differential Diffusion on Turbulent Non-Premixed Flames LO2/CH4 under Transcritical Conditions Using Large-Eddy Simulation. Energies 2023, 16, 1065. [Google Scholar] [CrossRef]
  30. Boyette, W.R.; Guiberti, T.F.; Magnotti, G.; Roberts, W.L. Structure of Turbulent Nonpremixed Syngas Flames at High Pressure. Proc. Combust. Inst. 2019, 37, 2207–2214. [Google Scholar] [CrossRef]
  31. Tang, H.; Yang, C.; Wang, G.; Guiberti, T.F.; Magnotti, G. Raman Spectroscopy for Quantitative Measurements of Temperature and Major Species in High-Pressure Non-Premixed NH3/H2/N2 Counterflow Flames. Combust. Flame 2022, 237, 111840. [Google Scholar] [CrossRef]
  32. Design, R. CHEMKIN-PRO 18.2; ANSYS, Inc.: San Diego, CA, USA, 2017. [Google Scholar]
  33. Kee, R.J.; Miller, J.A.; Evans, G.H.; Dixon-Lewis, G. A Computational Model of the Structure and Extinction of Strained, Opposed Flow, Premixed Methane-Air Flames. Symp. (Int.) Combust. 1989, 22, 1479–1494. [Google Scholar] [CrossRef]
  34. Okafor, E.C.; Naito, Y.; Colson, S.; Ichikawa, A.; Kudo, T.; Hayakawa, A.; Kobayashi, H. Experimental and Numerical Study of the Laminar Burning Velocity of CH4–NH3–Air Premixed Flames. Combust. Flame 2018, 187, 185–198. [Google Scholar] [CrossRef]
  35. Jiang, Y.; Gruber, A.; Seshadri, K.; Williams, F. An Updated Short Chemical-kinetic Nitrogen Mechanism for Carbon-free Combustion Applications. Int. J. Energy Res. 2020, 44, 795–810. [Google Scholar] [CrossRef]
  36. Cheng, M.; Wang, H.; Xiao, H.; Luo, K.; Fan, J. Emission Characteristics and Heat Release Rate Surrogates for Ammonia Premixed Laminar Flames. Int. J. Hydrogen Energy 2021, 46, 13461–13470. [Google Scholar] [CrossRef]
  37. Chai, W.S.; Bao, Y.; Jin, P.; Tang, G.; Zhou, L. A Review on Ammonia, Ammonia-Hydrogen and Ammonia-Methane Fuels. Renew. Sustain. Energy Rev. 2021, 147, 111254. [Google Scholar] [CrossRef]
  38. Song, Y.; Hashemi, H.; Christensen, J.M.; Zou, C.; Marshall, P.; Glarborg, P. Ammonia Oxidation at High Pressure and Intermediate Temperatures. Fuel 2016, 181, 358–365. [Google Scholar] [CrossRef]
  39. Xu, L.; Chang, Y.; Treacy, M.; Zhou, Y.; Jia, M.; Bai, X.S. A Skeletal Chemical Kinetic Mechanism for Ammonia/n-Heptane Combustion. Fuel 2023, 331, 125830. [Google Scholar] [CrossRef]
  40. Alnasif, A.; Mashruk, S.; Shi, H.; Alnajideen, M.; Wang, P.; Pugh, D.; Valera-Medina, A. Evolution of Ammonia Reaction Mechanisms and Modeling Parameters: A Review. Appl. Energy Combust. Sci. 2023, 15, 100175. [Google Scholar]
  41. Han, X.; Wang, Z.; Costa, M.; Sun, Z.; He, Y.; Cen, K. Experimental and Kinetic Modeling Study of Laminar Burning Velocities of NH3/Air, NH3/H2/Air, NH3/CO/Air and NH3/CH4/Air Premixed Flames. Combust. Flame 2019, 206, 214–226. [Google Scholar] [CrossRef]
  42. Petrova, M.; Williams, F. A Small Detailed Chemical-Kinetic Mechanism for Hydrocarbon Combustion. Combust. Flame 2006, 144, 526–544. [Google Scholar] [CrossRef]
  43. Weller, H.G.; Tabor, G.; Jasak, H.; Fureby, C. A Tensorial Approach to Computational Continuum Mechanics Using Object-Oriented Techniques. Comput. Phys. 1998, 12, 620. [Google Scholar] [CrossRef]
  44. Peters, N. Turbulent Combustion. Meas. Sci. Technol. 2001, 12, 2022. [Google Scholar] [CrossRef]
  45. Pierce, C.D.; Moin, P. Progress-Variable Approach for Large-Eddy Simulation of Non-Premixed Turbulent Combustion. J. Fluid Mech. 2004, 504, 73–97. [Google Scholar] [CrossRef]
  46. Lilly, D.K. A Proposed Modification of the Germano Subgrid-scale Closure Method. Phys. Fluids A Fluid Dyn. 1992, 4, 633–635. [Google Scholar] [CrossRef]
  47. Bilger, R.; Stårner, S.; Kee, R. On reduced mechanisms for methane-air combustion in nonpremixed flames. Combust. Flame 1990, 80, 135–149. [Google Scholar] [CrossRef]
  48. Domingo, P.; Vervisch, L.; Veynante, D. Large-Eddy Simulation of a Lifted Methane Jet Flame in a Vitiated Coflow. Combust. Flame 2008, 152, 415–432. [Google Scholar] [CrossRef]
  49. Pitsch, H. A C++ Computer Program for 0-D and 1-D Laminar Flame Calculations; RWTH: Aachen, Germany, 1998. [Google Scholar]
  50. Ihme, M.; Shunn, L.; Zhang, J. Regularization of Reaction Progress Variable for Application to Flamelet-Based Combustion Models. J. Comput. Phys. 2012, 231, 7715–7721. [Google Scholar] [CrossRef]
  51. Peters, N. Laminar Diffusion Flamelet Models in Non-Premixed Turbulent Combustion. Prog. Energy Combust. Sci. 1984, 10, 319–339. [Google Scholar] [CrossRef]
  52. Coelho, P.J.; Peters, N. Unsteady Modelling of a Piloted Methane/Air Jet Flame Based on the Eulerian Particle Flamelet Model. Combust. Flame 2001, 124, 444–465. [Google Scholar] [CrossRef]
  53. Ihme, M.; Pitsch, H. Modeling of Radiation and Nitric Oxide Formation in Turbulent Nonpremixed Flames Using a Flamelet/Progress Variable Formulation. Phys. Fluids 2008, 20, 055110. [Google Scholar] [CrossRef]
  54. Wang, H.; Hawkes, E.R.; Savard, B.; Chen, J.H. Direct Numerical Simulation of a High Ka CH4/Air Stratified Premixed Jet Flame. Combust. Flame 2018, 193, 229–245. [Google Scholar] [CrossRef]
  55. Okafor, E.C.; Somarathne, K.D.K.A.; Ratthanan, R.; Hayakawa, A.; Kudo, T.; Kurata, O.; Iki, N.; Tsujimura, T.; Furutani, H.; Kobayashi, H. Control of NOx and Other Emissions in Micro Gas Turbine Combustors Fuelled with Mixtures of Methane and Ammonia. Combust. Flame 2020, 211, 406–416. [Google Scholar] [CrossRef]
  56. Somarathne, K.D.K.A.; Okafor, E.C.; Sugawara, D.; Hayakawa, A.; Kobayashi, H. Effects of OH Concentration and Temperature on NO Emission Characteristics of Turbulent Non-Premixed CH4/NH3/Air Flames in a Two-Stage Gas Turbine like Combustor at High Pressure. Proc. Combust. Inst. 2021, 38, 5163–5170. [Google Scholar] [CrossRef]
  57. Somarathne, K.D.K.A.; Okafor, E.C.; Hayakawa, A.; Kudo, T.; Kurata, O.; Iki, N.; Kobayashi, H. Emission Characteristics of Turbulent Non-Premixed Ammonia/Air and Methane/Air Swirl Flames through a Rich-Lean Combustor under Various Wall Thermal Boundary Conditions at High Pressure. Combust. Flame 2019, 210, 247–261. [Google Scholar] [CrossRef]
  58. Duynslaegher, C.; Jeanmart, H.; Vandooren, J. Flame Structure Studies of Premixed Ammonia/Hydrogen/Oxygen/Argon Flames: Experimental and Numerical Investigation. Proc. Combust. Inst. 2009, 32, 1277–1284. [Google Scholar] [CrossRef]
Figure 1. Instantaneous temperature cloud map of CAJF28 (left) and schematic diagram of KAUST jet burner in HPCD (right) [19].
Figure 1. Instantaneous temperature cloud map of CAJF28 (left) and schematic diagram of KAUST jet burner in HPCD (right) [19].
Energies 16 06985 g001
Figure 2. Computational domain and boundary conditions of (a) counterflow flame and (b) jet flame.
Figure 2. Computational domain and boundary conditions of (a) counterflow flame and (b) jet flame.
Energies 16 06985 g002
Figure 3. Measured mass fractions of major species and temperature [19] (symbols) as a function of mixture fraction in CACF14 and CACF28. The solid line is the simulation results using different chemical mechanisms.
Figure 3. Measured mass fractions of major species and temperature [19] (symbols) as a function of mixture fraction in CACF14 and CACF28. The solid line is the simulation results using different chemical mechanisms.
Energies 16 06985 g003
Figure 4. Measured mole fractions of NO [20] (symbols) in mixture fraction space of CACF14/28. The solid line is the simulation results using different chemical mechanisms.
Figure 4. Measured mole fractions of NO [20] (symbols) in mixture fraction space of CACF14/28. The solid line is the simulation results using different chemical mechanisms.
Energies 16 06985 g004
Figure 5. Nitrogen flow diagrams for CACF14/CACF28. The thickness of the arrows represents the relative size of the flow. The numbers on the arrows indicate the relative conversion rate, with CACF14 in red and CACF28 in blue.
Figure 5. Nitrogen flow diagrams for CACF14/CACF28. The thickness of the arrows represents the relative size of the flow. The numbers on the arrows indicate the relative conversion rate, with CACF14 in red and CACF28 in blue.
Energies 16 06985 g005
Figure 6. (a) The ratios of NH 2 converted to nitrogen-containing species in CACF14 (red) and CACF28 (blue); (b) ratios of NH converted to nitrogen-containing species; (c) ratios of HNO and N converted to nitrogen-containing species; (d) conversion rates of NO generated from NH 2 via four pathways: a–d.
Figure 6. (a) The ratios of NH 2 converted to nitrogen-containing species in CACF14 (red) and CACF28 (blue); (b) ratios of NH converted to nitrogen-containing species; (c) ratios of HNO and N converted to nitrogen-containing species; (d) conversion rates of NO generated from NH 2 via four pathways: a–d.
Energies 16 06985 g006
Figure 7. The mole fraction results of radicals (a) HNO , (b) NH , (c) N , (d) H , (e) O and (f) OH in the mixture fraction space in CACF14 (red) and CACF28 (blue).
Figure 7. The mole fraction results of radicals (a) HNO , (b) NH , (c) N , (d) H , (e) O and (f) OH in the mixture fraction space in CACF14 (red) and CACF28 (blue).
Energies 16 06985 g007
Figure 8. Radial distribution of time-averaged (circle points, solid lines) and RMS values (square points, dashed lines) of mixture fraction, progress variable at different streamwise locations ( X / D = 5, 10, 20, 40) for CAJF14 (red) and CAJF28 (blue), as measured in experiments [19] (points) and computed with FPV (lines).
Figure 8. Radial distribution of time-averaged (circle points, solid lines) and RMS values (square points, dashed lines) of mixture fraction, progress variable at different streamwise locations ( X / D = 5, 10, 20, 40) for CAJF14 (red) and CAJF28 (blue), as measured in experiments [19] (points) and computed with FPV (lines).
Energies 16 06985 g008
Figure 9. Radial distribution of time-averaged (circle points, solid lines) and RMS values (square points, dashed lines) of Y NH 3 , Y H 2 and temperature T at different streamwise locations ( X / D = 5 , 10 , 20 , 40 ) for CAJF14 (red) and CAJF28 (blue), as measured in experiments [19] (points) and computed with FPV (lines).
Figure 9. Radial distribution of time-averaged (circle points, solid lines) and RMS values (square points, dashed lines) of Y NH 3 , Y H 2 and temperature T at different streamwise locations ( X / D = 5 , 10 , 20 , 40 ) for CAJF14 (red) and CAJF28 (blue), as measured in experiments [19] (points) and computed with FPV (lines).
Energies 16 06985 g009
Figure 10. Radial distribution of time-averaged (circle points, solid lines) and RMS values (square points, dashed lines) of X NO at different streamwise locations ( X / D = 5 , 10 , 20 , 40 ) for CAJF14 (red) and CAJF28 (blue), as measured in experiments [20] (points) and computed with FPV (lines).
Figure 10. Radial distribution of time-averaged (circle points, solid lines) and RMS values (square points, dashed lines) of X NO at different streamwise locations ( X / D = 5 , 10 , 20 , 40 ) for CAJF14 (red) and CAJF28 (blue), as measured in experiments [20] (points) and computed with FPV (lines).
Energies 16 06985 g010
Figure 11. The figure of instantaneous NO mole fraction (black dots), NO conditional mean value (green dots) and steady flamelet solutions at different scalar dissipation rates (colored solid lines according to the value of scalar dissipation rate) versus mixture fraction at streamwise locations X / D = 5 , X / D = 20 for CAJF14 and CAJF28. The arrows point towards the flamelets where the scalar dissipation rate increases.
Figure 11. The figure of instantaneous NO mole fraction (black dots), NO conditional mean value (green dots) and steady flamelet solutions at different scalar dissipation rates (colored solid lines according to the value of scalar dissipation rate) versus mixture fraction at streamwise locations X / D = 5 , X / D = 20 for CAJF14 and CAJF28. The arrows point towards the flamelets where the scalar dissipation rate increases.
Energies 16 06985 g011
Figure 12. The figure of instantaneous temperature (black dots), temperature conditional mean value (red dots), steady flamelet solutions of temperature at different scalar dissipation rates (colored solid lines according to the size of scalar dissipation rate) and unsteady flamelet solutions (colored dashed lines according to the value of scalar dissipation rate) versus mixture fraction at streamwise locations X / D = 5 and X / D = 20 for CAJF14 and CAJF28. Arrows pointing towards the flamelets where the scalar dissipation rate increases; solid black arrow indicates steady flamelets and dashed white arrow indicates unsteady flamelets.
Figure 12. The figure of instantaneous temperature (black dots), temperature conditional mean value (red dots), steady flamelet solutions of temperature at different scalar dissipation rates (colored solid lines according to the size of scalar dissipation rate) and unsteady flamelet solutions (colored dashed lines according to the value of scalar dissipation rate) versus mixture fraction at streamwise locations X / D = 5 and X / D = 20 for CAJF14 and CAJF28. Arrows pointing towards the flamelets where the scalar dissipation rate increases; solid black arrow indicates steady flamelets and dashed white arrow indicates unsteady flamelets.
Energies 16 06985 g012
Figure 13. Time-averaged distributions of NO mole fraction for CAJF14 (top) and CAJF28 (bottom). The Z s t isoline is represented by a white dotted line.
Figure 13. Time-averaged distributions of NO mole fraction for CAJF14 (top) and CAJF28 (bottom). The Z s t isoline is represented by a white dotted line.
Energies 16 06985 g013
Figure 14. Time-averaged distributions of mole fractions for (a) HNO , (b) NH , (c) N , (d) H , (e) O and (f) OH in CAJF14 (top) and CAJF28 (bottom).
Figure 14. Time-averaged distributions of mole fractions for (a) HNO , (b) NH , (c) N , (d) H , (e) O and (f) OH in CAJF14 (top) and CAJF28 (bottom).
Energies 16 06985 g014
Figure 15. Scatter plots showing the distribution of mole fractions of NO (a), O (b), H (c) and OH (d) in the mixture fraction space, where CAJF14 is represented in red and CAJF28 in blue.
Figure 15. Scatter plots showing the distribution of mole fractions of NO (a), O (b), H (c) and OH (d) in the mixture fraction space, where CAJF14 is represented in red and CAJF28 in blue.
Energies 16 06985 g015
Figure 16. Time-averaged distributions of fuel, thermal and total NO source terms for CAJF14 (top) and CAJF28 (bottom).
Figure 16. Time-averaged distributions of fuel, thermal and total NO source terms for CAJF14 (top) and CAJF28 (bottom).
Energies 16 06985 g016
Table 1. Operating condition parameters.
Table 1. Operating condition parameters.
Flame TypeFlame Case NH 3 H 2 N 2 U fuel  (m/s) U air  (m/s)
LaminarCACF140.7540.1840.0620.360.29
CACF280.5630.3280.1090.360.29
TurbulentCAJF140.7540.1840.0628.60.24
CAJF280.5630.3280.10910.10.24
Table 2. The reactions involving NO in the Okafor mechanism.
Table 2. The reactions involving NO in the Okafor mechanism.
SequenceReactionSequenceReaction
R220:N + NO = N 2 + OR255: NH 2 + HNO = NH 3 + NO
R221:N + O 2 = NO + OR256: NH 2 + NO = N 2 + H 2 O
R222:N + OH = NO + HR257: NH 2 + NO = NNH + OH
R224: N 2 O + O = 2 NOR263:NNH + O = NH + NO
R228: HO 2 + NO = NO 2 + OHR269:NNH + NO = N 2 + HNO
R229:NO + O + M = NO 2 + MR271:H + NO + M = HNO + M
R230: NO 2 + O = NO + O 2 R272:HNO + O = NO + OH
R231: NO 2 + H = NO + OHR273:HNO + H = H 2 + NO
R233:NH + O = NO + HR274:HNO + OH = NO + H 2 O
R237:NH + O 2 = NO + OHR275:HNO + O 2 = HO 2 + NO
R240:NH + NO = N 2 O + HR276:HNO + NO 2 = HONO + NO
R241:NH + NO = N 2 O + HR290: N 2 H 2 + O = NH 2 + NO
R242:NH + NO = N 2 + OH
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

Wang, C.; Wang, H.; Luo, K.; Fan, J. The Effects of Cracking Ratio on Ammonia/Air Non-Premixed Flames under High-Pressure Conditions Using Large Eddy Simulations. Energies 2023, 16, 6985. https://doi.org/10.3390/en16196985

AMA Style

Wang C, Wang H, Luo K, Fan J. The Effects of Cracking Ratio on Ammonia/Air Non-Premixed Flames under High-Pressure Conditions Using Large Eddy Simulations. Energies. 2023; 16(19):6985. https://doi.org/10.3390/en16196985

Chicago/Turabian Style

Wang, Chengming, Haiou Wang, Kun Luo, and Jianren Fan. 2023. "The Effects of Cracking Ratio on Ammonia/Air Non-Premixed Flames under High-Pressure Conditions Using Large Eddy Simulations" Energies 16, no. 19: 6985. https://doi.org/10.3390/en16196985

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