Next Article in Journal
Mechanical Characteristics and Precision Analysis of Inflatable Deployable Parabolic Membrane Antenna Structures
Previous Article in Journal
Wind-Induced Responses of Nonlinear Angular Motion for a Dual-Spin Rocket
Previous Article in Special Issue
Experimental Investigation of Pure Hydrogen Flame in a Matrix Micro-Mixing Combustor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Low-Order Modelling of Extinction of Hydrogen Non-Premixed Swirl Flames

by
Hazem S. A. M. Awad
*,
Savvas Gkantonas
and
Epaminondas Mastorakos
Department of Engineering, University of Cambridge, Cambridge CB2 1PZ, UK
*
Author to whom correspondence should be addressed.
Aerospace 2025, 12(8), 676; https://doi.org/10.3390/aerospace12080676
Submission received: 15 June 2025 / Revised: 25 July 2025 / Accepted: 27 July 2025 / Published: 29 July 2025
(This article belongs to the Special Issue Scientific and Technological Advances in Hydrogen Combustion Aircraft)

Abstract

Predicting the blow-off (BO) is critical for characterising the operability limits of gas turbine engines. In this study, the applicability of a low-order extinction prediction modelling, which is based on a stochastic variant of the Imperfectly Stirred Reactor (ISR) approach, to predict the lean blow-off (LBO) curve and the extinction conditions in a hydrogen Rich-Quench-Lean (RQL)-like swirl combustor is investigated. The model predicts the blow-off scalar dissipation rate (SDR), which is then extrapolated using Reynolds-Averaged Navier–Stokes (RANS) cold-flow simulations and simple scaling laws, to determine the critical blow-off conditions. It has been found that the sISR modelling framework can predict the BO flow split ratio at different global equivalence ratios, showing a reasonable agreement with the experimental data. This further validates sISR as an efficient low-order modelling flame extinction tool, which can significantly contribute to the development of robust hydrogen RQL combustors by enabling the rapid exploration of combustor operability during the preliminary design phases.

1. Introduction

Hydrogen propulsion can be considered an efficient solution for the decarbonization of the aviation industry. Designing aircraft engines powered by hydrogen presents unique challenges due to hydrogen’s distinct combustion properties. One major issue is hydrogen’s high adiabatic flame temperature, which can lead to high nitrogen oxide (NOx) emissions [1]. Since strict emissions regulations govern aircraft engine certification, controlling NOx output becomes critical for the development of hydrogen-powered aircraft. To address that, advanced combustor configurations like the Rich-Quench-Lean (RQL) design have shown promise [2]. The RQL combustor minimizes NOx emissions by, first, burning fuel in a rich primary zone, then rapidly cooling the hot gases with dilution air jets [2]. This approach helps control flame temperatures and reduces harmful emissions, making it a promising solution for hydrogen combustion in aviation. One of the key parameters that governs the combustion in an RQL combustor is the dilution ratio. The dilution ratio (R) is defined as the fraction of the dilution air flow rate ( m d ˙ ) to the total air flow rate, which includes both primary and dilution air ( m p ˙ + m d ˙ ) [3,4]. The precise control of dilution is critical for managing flame stability, temperature distribution, and emissions, particularly NOx. Despite hydrogen’s high reactivity and wide flammability range, excessive dilution poses a critical stability risk. Injecting too much dilution air cools the reaction zone and increases turbulent mixing/scalar dissipation rates, which can locally quench the flame [3]. If the dilution ratio exceeds a critical threshold, localized extinctions can result in global flame blow-off (BO), even for highly reactive hydrogen flames. Therefore, accurately predicting the critical blow-off dilution ratio is essential for designing reliable hydrogen RQL combustors. Moreover, lean blow-off (LBO) remains a fundamental operability limit, irrespective of fuel type, and accurately predicting the LBO curve remains important during the preliminary combustor design phases.
Large Eddy Simulation (LES) has shown some success in the prediction of localized extinction in laboratory-scale flames, such as the Sandia jet, Sydney swirl flame, or Cambridge swirling flames. Zhang and Mastorakos [5] used LES with Conditional Moment Closure (LES-CMC) to predict the entire lean blow-off (LBO) curve of a swirling flame, achieving around 25% accuracy in blow-off limits and durations. However, such comprehensive studies are rare, mainly due to the expensive computational cost of LES. To make BO prediction more feasible during the design phase, there is a growing need for lower-cost modelling methods that can assess a wider range of geometries and conditions without relying entirely on high-fidelity CFD. Gkantonas and Mastorakos [6] developed a low-order modelling tool that can predict LBO behaviour. The developed model is based on a stochastic representation of the Imperfectly Stirred Reactor (ISR) model [7,8,9], termed the sISR method. The model provides a map that shows the extinction probability (i.e., the proportion of a flame’s stoichiometric iso-surface undergoing local extinctions) at different values of scalar dissipation rates (SDRs). It relies on a key assumption, supported by experimental and numerical studies [5,10,11,12,13,14], which suggests that global flame blow-off occurs when 30% of the stochiometric iso-surface experiences local extinction. Thus, from the extinction threshold, the model predicts the blow-off SDR, which is then extrapolated using reference CFD data and simple scaling laws, to determine critical blow-off conditions such as velocity or dilution ratio. Gkantonas and Mastorakos validated the sISR against atmospheric Cambridge swirl spray flames (ethanol/kerosene fuels), demonstrating its capability to predict blow-off velocities within 13% of the experimental values [6]. Although sISR has been previously validated for hydrocarbon fuels, the sISR model itself is fuel-agnostic and provides a general framework extinction prediction at a wide range of fuels and combustor geometries [6]. Several core insights—such as the stochasticity of extinction events and a critical extent of localized extinction before full blow-off—are conceptually valid regardless of the used fuel. However, hydrogen’s strong differential diffusion effects are known to influence flame extinction and can pose modeling challenges to the sISR approach. The inclusion of differential diffusion to the sISR approach is possible following previous studies using the Imperfectly Stirred Reactor Network (ISRN) framework [15], which show that differential diffusion effects can be approximately recovered through additional terms in the conditional transport equations and species-specific Lewis numbers (e.g., H2 and H). This study extends the sISR validation by investigating its applicability in predicting the LBO curve and blow-off dilution ratio in a swirl hydrogen Rich-Quench-Lean (RQL)-like combustor. The paper is structured as follows: Firstly, the investigated model and its governing equation are discussed. Then, the experimental step, numerical implementation, and methodology are introduced. The results are then discussed, and, finally, the conclusions are summarised.

2. Methods

2.1. Governing Equation

The Imperfectly Stirred Reactor (ISR) model is based on the spatially integrated form of the Conditional Moment Closure (CMC) equations [9,15,16]. The model is called ‘stochastic’ as it is based on a stochastic representation of the SDR. Unlike a perfectly stirred reactor, the ISR allows for mixture fraction inhomogeneities, which is essential when modelling local extinction and finite-rate chemical effects. The governing equations for the conditional enthalpy and mass fractions are as follows [15]:
Q h t = N η 2 Q h η 2
Q α t = N η 2 Q α η 2 + ω α ˙ | η
where η is the space sample in mixture fraction space, Q h indicates a conditional specific enthalpy (i.e., Q h = h | η ), and Q α indicates a conditional mass fraction (i.e., Q α = Y α | η ). ω α ˙ | η stands for the conditional chemical source term for a given species α and N η is the conditional profile of the SDR. In the present analysis, the effect of differential diffusion is neglected as the investigated combustor involves high Reynolds numbers. Therefore, species diffusion effects are less significant compared to turbulent mixing and the LBO curve and extinction dilution ratio predictions are not expected to significantly change when including the effect of preferential diffusion. To model this term, the Amplitude Mapping Closure (AMC) model [16,17] is used, which expresses it as follows:
N η = N 0 t G ( η )
with G ( η ) being a shaping factor defined as follows:
G η = e x p ( 2 [ erf 1 ( 2 η 1 ) ] 2 )
Here, N 0 t is a representative SDR that varies in time according to a stochastic differential equation (SDE), capturing temporal fluctuations. Following similar approaches to Pitsch and Fedorov [18] and Gkantonas and Mastorakos [19], the stochastic behaviour of N 0 is governed by the following [6]:
d N 0 = f N 0 d t + σ N ϕ N 0 d W t
where d W t is a Wiener process, and the functions f and ϕ define the deterministic and stochastic components, respectively. They are given by the following [6]:
f N 0 = ( log N 0 log N 0 exp 0.5 σ N 2 N 0 τ N
ϕ N 0 = N 0 2 τ N
In this formulation, σ N defines the lognormal distribution width and is computed from the fluctuation parameter F = ( N 0 ) 2 N 0 as follows [6]:
σ N = l o g ( F 2 + 1 )
The fluctuation parameter F can be approximated using the turbulent Reynolds number R e l as follows [6]:
F = 15 R e l 1 8
Finally, the characteristic timescale τ N introduces a memory effect, capturing the autocorrelation time of SDR fluctuations. Since such fluctuations are concentrated in the smallest turbulent scales, prior work [19,20] suggests that τ N 1.6 τ k , with τ k being the Kolmogorov timescale. More details regarding the model are found in Ref. [6].

2.2. Extraction of sISR Parameters from CFD

The sISR prediction approach relies on three global parameters extracted from the CFD simulation: (1) mean scalar dissipation rate N 0 , (2) fluctuation scaling parameter F, and (3) characteristic timescale τ N . These parameters are computed using volume-weighted averages of standard CFD quantities (density, turbulence properties, etc.) within a control region surrounding the stoichiometric mixture fraction iso-surface. The mean scalar dissipation rate (SDR) is estimated as follows [2]:
N 0 = N 0 1 P η G η d η
where G η is a prescribed profile in mixture fraction space, derived from the AMC model, and N and P η are found based on a mass-based integration of all CFD cells within a volume surrounding the iso-surface of the stoichiometric mean mixture fraction. This is carried out as follows [2]:
P η = ρ ¯ P η d V ρ ¯ d V
N = ρ ¯ N d V ρ ¯ d V
with N defined at the CFD level as follows [2]:
N = C d 2 ε k ξ 2 ; w i t h C d = 2.0
where ρ ¯ is the mean density, ε is the turbulence dissipation rate, k is the turbulent kinetic energy, and ξ 2 is the mixture fraction variance. The integration region is typically a band surrounding the stoichiometric iso-surface and, in this analysis, is taken as ξ s t + / 70%. The fluctuation parameter F and the characteristic time scale τ N are obtained based on mass-based integration over the same region as in Equations (11) and (12) and are approximated using the local Reynolds number and the turbulent timescale, respectively [2].
F = ρ ¯ R e l 1 8 d V ρ ¯ d V = ρ ¯ 15 k 2 ν ε 1 8 d V ρ ¯ d V
τ N = ρ ¯ ( 1.6 τ k ) d V ρ ¯ d V = ρ ¯ 1.6 ν ε 1 2 d V ρ ¯ d V

2.3. Investigated Combustor

The RQL swirl combustor developed at Cambridge University is here investigated [3,4]. A schematic representation of the burner is presented in Figure 1. The primary air enters the combustor through a swirler, which introduces the flow at an 80-degree angle relative to the main axis. Fuel injection occurs through a 4 mm diameter hole positioned at the centre of a bluff body. Additional dilution air is supplied via four side jets, each inclined at 45 degrees to the combustor wall and located 54 mm downstream from the inlet. Both the air and fuel are introduced at an initial temperature of 300 K, and the system operates under atmospheric pressure.
Two equivalence ratios are used to characterize the operating condition: the primary equivalence ratio ( ϕ p ), which is defined as the ratio of the fuel flow to the primary airflow; and the global equivalence ratio ( ϕ g ), which is defined as the ratio of the fuel flow to the total airflow. The aim of the present investigation is to predict the LBO curve using a limited number of CFD simulations far from BO. These CFD simulations are selected at conditions far from BO at a fixed dilution ratio and different global equivalence ratios, which span the LBO curve ϕ g range. The operating conditions for the simulated CFD cases are shown in Table 1.

2.4. Methodology and Numerical Implementation

The sISR model is solved using the CLIO code [16], which solves the governing equation of the CMC equations. The sISR model is initialized using a 0D-CMC solution, where the initial scalar dissipation rate N 0 is set to be equal to the time-averaged value N 0 = N 0 ( t ) . For the numerical implementation, an operator splitting approach is applied, using the same solvers and discretization methods as those described in Ref. [16]. The predicted extinction probability depends on the choice of the chemical mechanism, as it directly affects the scalar dissipation rate at which extinction occurs [6]. Thus, the sISR predictions can be highly sensitive to the chosen chemical kinetics. A chemical mechanism, which consists of 10 species and 31 reactions, has been employed for hydrogen chemistry [21]. This mechanism has been validated against experimental data in terms of the ignition delay time and the laminar flame speed, revealing good agreement with the experimental data [21]. Moreover, the used mechanism shows better performance compared to other hydrogen chemical mechanism [21]. The mixture fraction space is discretized into 76 grid points refined close to the stochiometric mixture fraction for hydrogen (i.e., ξ s t = 0.028). The stochastic differential equation (SDE) is solved using the Milstein method [19] and the total simulation time is taken as 0.15 s, which corresponds to approximately 800 τ N , to provide enough sampling time for the SDE and the sISR equations to develop. Given that N 0 evolves stochastically, multiple sISR simulations are necessary in order to estimate the probability of extinction. Here, an upper limit on N 0 is imposed by clipping its lognormal distribution. A statistical limit is used, which is defined relative to the distribution’s percentiles (e.g., the 99.999th percentile). This helps eliminate rare events and outliers. In this work, an extinction event is defined based on the temperature at the stoichiometric mixture fraction ( T s t o i = T | η s t ) falling below a predefined threshold—typically within the range of 1000 K to 1500 K. If n r denotes the total number of realisations and n e x t is the number of those in which extinction occurs, then the extinction probability p e x t is calculated (i.e., p e x t = n e x t / n r ). In the present work, 100 different realisations are used to estimate the extinction probability. In this work, the flame is considered extinguished if the extinction probability is higher than 30% following Ref. [6].
The cold reference CFD cases are solved using the in-house Rolls-Royce CFD code PRECISE-UNS [22,23]. The simulations are conducted using the RANS framework with the k ϵ as a turbulence model. The LES mesh consists of approximately 4M tetrahedral cells. An in-flow boundary condition is employed at the air and fuel inlet, whereas an outlet boundary condition is used at the combustor exit. And second-order accuracy scheme is used for the spatial discretization of all the variables and convergence is assessed by monitoring the relative change in key solution quantities over successive iterations. The procedure for predicting the extinction dilution ratio and lean blow-off (LBO) curve using the sISR framework involves the following steps:
(1)
Parameter Extraction from CFD: Key parameters—including SDR, fluctuation parameter F , and a characteristic timescale τ N —are extracted from a reference CFD simulation corresponding to a known operating condition.
(2)
The Construction of Extinction Probability Map: A series of sISR simulations is performed over a broad range of SDR values at the fixed values of F and τ N obtained in Step 1. These simulations yield extinction probabilities, from which an extinction probability map is constructed. The map is then refined by averaging the extinction probabilities over additional sets of F and τ N to ensure robustness. The extinction SDR is obtained at an extinction probability that corresponds to 30%.
(3)
The Identification of Extinction SDR: From the probability map, the critical SDR corresponding to an extinction probability of 30% is identified and defined as the extinction SDR.
(4)
Estimation of Extinction Dilution Ratio: Finally, the extinction dilution ratio R e x t is estimated by scaling the reference dilution ratio R r e f using the ratio of extinction SDR (step 3) to the reference SDR (step 1). A schematic of the sISR prediction methodology is shown in Figure 2 and more details can be found in Ref. [6].

3. Results

3.1. Cold-Flow Reference Simulations

Figure 3 and Figure 4 show the velocity magnitude, mixture fraction, and scalar dissipation rate (SDR) fields for three operating conditions with increasing global equivalence ratio—Case A ( ϕ g = 0.05), Case B ( ϕ g = 0.07), and Case C ( ϕ g = 0.09)—at a 45-degree oblique slice spanning the combustor corners. In all cases, the fuel mass flow rate remains constant, and ϕ g is varied by reducing the air mass flow. It is also worth noting that all cases have the same dilution ratio. Each column corresponds to a different case, while the rows display (from top to bottom) the velocity, mixture fraction, and SDR fields. At the lowest equivalence ratio (Case A), the high primary air flow results in a strong jet momentum, which induces a pronounced recirculation zone downstream of the jet. This intense recirculation limits the expansion of the fuel jet, confining the mixture and resulting in a higher local velocity and sharper scalar gradients. As ϕ g increases (Cases B and C), the reduced airflow leads to a weaker jet momentum and a less intense recirculation zone, allowing the fuel to spread more broadly into the combustion chamber. The SDR fields also reflect this behaviour: Case A shows the most intense dissipation rates near the injector, driven by a strong velocity and mixture gradients, while the SDR magnitude decreases with increasing ϕ g . This can be confirmed by estimating N 0 within a volume surrounding the stochiometric mixture fraction iso-surface according to Equation (9), which reveal N 0 = 21.1, 11.3, and 7.8 s 1 for case A, B, and C, respectively. The SDR behaviour is also consistent with the blow-off dilution ratio at each ϕ g , with case A being closer to the blow-off compared to cases B and C.

3.2. Extinction Probability

Figure 5 and Figure 6 show the change in temperature at the stoichiometric mixture fraction and N 0 with time for a fully burning realisation and a blow-off realisation. As shown in Figure 6, the stochastic representation of the SDR allows for temperature fluctuations for both realisations. However, in some instances, the SDR attains higher values for the blow-off case compared to the fully burning cases, resulting in a flame blow-off after about 40 ms. These are two realisations out of 100 realisations used to obtain the extinction probability at N 0 = 60   s 1 . In that case, the extinction probability is 30%, suggesting that 30 realisations out of 100 result in a blow-off behaviour. Following the sISR approach and as depicted in Figure 2, a “probability map” for extinction events can be constructed, by obtaining the extinction probability at different values of SDRs. Figure 7 illustrates the numerical estimate of the mean extinction probability, p e x t , as a function of the mean SDR, N 0 for different cases spanning various combinations of parameters F and τ N . The figure also shows the associated uncertainty in the estimated p e x t , represented by confidence intervals derived from binomial statistics across multiple parameter scenarios. It can be seen in Figure 7 that the relationship resembles a sigmoid, with p e x t increasing with a higher N 0 , indicating a greater likelihood of blow-off (BO) as one moves to the right along the curve. By adopting a critical probability threshold of approximately 30% (indicated by the red star symbol), the right region of the plot can be identified as an unstable regime where BO is likely to occur, whereas the left region indicates a stable flame regime. It is worth noting that the influence of hydrogen’s high laminar flame speed and adiabatic flame temperature is implicitly accounted for in the model through the extinction scalar dissipation rate threshold N 0 , e x t , which directly affects the extinction probability curve. For instance, in the present analysis, the SDR at which blow-off occurs is approximately 60   s 1 (i.e., S D R B O 60   s 1 ) for hydrogen flames. In contrast, for hydrocarbon fuels such as ethanol at atmospheric conditions, the corresponding threshold is significantly lower (e.g., S D R B O 8   s 1 ) [6].

3.3. LBO Curve and Blow-Off Dilution Ratio Predictions

As outlined by Klimenko and Bilger [9], the volume-weighted conditional scalar dissipation rate (SDR) within a control volume, denoted as N η , can be derived by performing a double integration of the mixture fraction probability density function (PDF) transport equation. This results in the following expression [6]:
N η b u r n e r = 1 τ r e s P η 0 η 0 η [ P η i n P η o u t ] d η d η
where τ r e s is the residence time within the burner and P η is the mixture fraction PDF at either the inlet or outlet. If we assume that changes in operating conditions do not significantly affect the shape or magnitude of N η b u r n e r through variations in the PDF, then the conditional SDR becomes primarily governed by the residence time, which is inversely proportional to the total mass flow rate. Assuming a constant fuel mass flow, the total mass flow rate—and, hence, N 0 —is directly proportional to the bulk velocity U b . In this work, the dilution ratio is changed by increasing the dilution jet velocity, while maintaining a constant total mass flow rate. Thus, in the present investigation, N 0 is directly proportional to the dilution ratio. This allows the extraction of the extinction dilution ratio from the reference SDR and the dilution ratio as follows:
R e x t = S D R B O × R r e f S D R r e f
The prediction of the LBO curve and extinction dilution ratio for the hydrogen swirl RQL-like combustor is shown in Figure 8. It can be observed that the extinction low-order modelling approach, sISR, is capturing the trending behaviour of the LBO curve compared to the experimental data. As the global equivalence ratio increases, the extinction dilution ratio also increases, a behaviour that is predicted by sISR. This further confirms the applicability of sISR in predicting the LBO curve using a limited number of CFD simulations, which can be far from BO and can also be based on a cheap CFD (i.e., cold flow using RANS).

4. Conclusions

The applicability of the low-order modelling extinction prediction tool, the stochastic Imperfectly Stirred Reactor (sISR), to predict the lean blow-off (LBO) curve and the extinction condition of a flow split between the primary and secondary zone is investigated for a hydrogen-fueled RQL-like swirl combustor. The model predicts the flame global extinction using a limited set of Computational Fluid Dynamics (CFD) simulations as reference conditions, which is, here, obtained using cold-flow RANS. It provides a map that shows the extinction probability (i.e., the proportion of a flame’s stoichiometric iso-surface undergoing local extinctions) at different values of scalar dissipation rates (SDRs). It has been found that the sISR modelling framework is capable of predicting blow-off (BO) dilution ratios across various global equivalence ratios, with the predictions showing a good agreement with the experimental data. The results support the applicability of the sISR model as a low-order tool for predicting flame extinction in RQL-like hydrogen combustors. Its application can play a valuable role in the early-stage design of hydrogen-rich RQL combustors by facilitating rapid assessments of combustor stability and operability.

Author Contributions

H.S.A.M.A.: writing—original draft, visualisation, software, methodology, investigation, formal analysis, data curation, and conceptualisation. S.G.: software, visualisation, methodology, investigation, formal analysis, and conceptualisation. E.M.: writing—review and editing, supervision, resources, funding acquisition, and conceptualisation. All authors have read and agreed to the published version of the manuscript.

Funding

The research is funded partially by the UKRI (Horizon Europe Guarantee), contract number 10062550, EU Clean AviationJoint Undertaking grant number 101102000, project CAVENDISH. The research is also partially funded by the Rolls-Royce Group.

Data Availability Statement

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

Acknowledgments

This work was performed using resources provided by the Cambridge Service for Data Driven Discovery (CSD3) operated by the University of Cambridge Research Computing Service.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Alhuyi Nazari, M.; Fahim Alavi, M.; Salem, M.; Assad, M.E.H. Utilization of hydrogen in gas turbines: A comprehensive review. Int. J. Low-Carbon Technol. 2022, 17, 513–519. [Google Scholar] [CrossRef]
  2. Straub, D.L.; Casleton, K.H.; Lewis, R.E.; Sidwell, T.G.; Maloney, D.J.; Richards, G.A. Assessment of rich-burn, quick-mix, lean-burn trapped vortex combustor for stationary gas turbines. J. Eng. Gas Turbines Power 2005, 127, 36–41. [Google Scholar] [CrossRef]
  3. Choe, J.; Mastorakos, E. Influence of side dilution jets on flame extinction and nitric oxide reduction in a hydrogen-fueled model gas turbine combustor. In Proceedings of the Mediterranean Combustion Symposium, Corfu, Greece, 1–5 June 2025. [Google Scholar]
  4. Awad, H.S.A.M.; Gkantonas, S.; Choe, J.; Mastorakos, E. The influence of side dilution jets on the flame stabilization mechanism and NOx emissions in a hydrogen-fuelled model gas turbine combustor. In Proceedings of the 12th European Combustion Meeting, Edinburgh, UK, 7–10 April 2025. [Google Scholar]
  5. Zhang, H.; Mastorakos, E. Prediction of global extinction conditions and dynamics in swirling non-premixed flames using LES/CMC modelling. Flow Turbul. Combust. 2016, 96, 863–889. [Google Scholar] [CrossRef]
  6. Gkantonas, S.; Mastorakos, E. Development of a low-order model of turbulent non-premixed flame global flow-off. Int. J. Spray Combust. Dyn. 2025. submitted. [Google Scholar]
  7. Mobini, K.; Bilger, R.W. Imperfectly Stirred Reactor Model Predictions of Reaction in a Burner with Strong Recirculation. Combust. Sci. Technol. 2004, 176, 45–70. [Google Scholar] [CrossRef]
  8. Gough, A.J.; Mobini, K.; Chen, Y.C.; Bilger, R.W. Measurements and Predictions in a Confined Bluff-Body Burner Modeled as an Imperfectly Stirred Reactor. Proc. Combust. Inst. 1998, 27, 3181–3188. [Google Scholar] [CrossRef]
  9. Klimenko, A.Y.; Bilger, R.W. Conditional moment closure for turbulent combustion. Prog. Energy Combust. Sci. 1999, 25, 595–687. [Google Scholar] [CrossRef]
  10. Sevault, A.; Dunn, M.; Barlow, R.S.; Ditaranto, M. On the structure of the near field of oxy-fuel jet flames using Raman/Rayleigh laser diagnostics. Combust. Flame 2012, 159, 3342–3352. [Google Scholar] [CrossRef]
  11. Barlow, R.S.; Frank, J.H. Effects of turbulence on species mass fractions in methane/air jet flames. Symp. Combust. 1998, 27, 1087–1095. [Google Scholar] [CrossRef]
  12. 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]
  13. Wandel, A.P.; Lindstedt, R.P. Hybrid multiple mapping conditioning modeling of local extinction. Proc. Combust. Inst. 2013, 34, 1365–1372. [Google Scholar] [CrossRef]
  14. Xu, J.; Pope, S.B. PDF Calculations of Turbulent Nonpremixed Flames with Local Extinction. Combust. Flame 2000, 123, 281–307. [Google Scholar] [CrossRef]
  15. Gkantonas, S. Predicting Soot Emissions with Advanced Turbulent Reacting Flow Modelling. Ph.D. Thesis, University of Cambridge, Cambridge, UK, 2021. [Google Scholar]
  16. Awad, H.S.A.M.; Gkantonas, S.; Mastorakos, E. A spark energy deposition model in mixture fraction space for simulations of turbulent non-premixed flame ignition. Appl. Energy Combust. Sci. 2025, 21, 100317. [Google Scholar] [CrossRef]
  17. O’Brien, E.E.; Jiang, T.L. The conditional dissipation rate of an initially binary scalar in homogeneous turbulence. Phys Fluids 1991, 12, 3121–3123. [Google Scholar] [CrossRef]
  18. Pitsch, H.; Fedotov, S. Investigation of scalar dissipation rate fluctuations in non-premixed turbulent combustion using a stochastic approach. Combust. Theory Model. 2001, 51, 41. [Google Scholar] [CrossRef]
  19. Gkantonas, S.; Mastorakos, E. Stochastic modelling for the effects of micromixing on soot in turbulent non-premixed flames. Combust. Theory Model. 2024, 28, 849–872. [Google Scholar] [CrossRef]
  20. Vedula, P.; Yeung, P.K.; Fox, R.O. Dynamics of scalar dissipation in isotropic turbulence: A numerical and modelling study. J. Fluid Mech. 2001, 433, 29–60. [Google Scholar] [CrossRef]
  21. Hong, Z.; Davidson, D.F.; Hanson, R.K. An improved H2/O2 mechanism based on recent shock tube/laser absorption measurements. Comb. Flame 2011, 158, 633–644. [Google Scholar] [CrossRef]
  22. Anand, M.S.; Eggels, R.L.G.M.; Staufer, M.; Zedda, M.; Zhu, J. An advanced unstructured-grid finite-volume design system for gas turbine combustion analysis. In Proceedings of the Gas Turbine India Conference, Bangalore, Karnataka, India, 5–6 December 2013; ASME: New York, NY, USA, 2013; Volume 35161, p. V001T03A003. [Google Scholar]
  23. Awad, H.S.A.M.; Gkantonas, S.; Mastorakos, E.; Eggels, R. Parametric analysis of hydrogen swirl combustors with incompletely stirred reactor network modelling. J. Propuls. Power 2025. submitted. [Google Scholar]
Figure 1. Schematic of the investigated combustor [3,4].
Figure 1. Schematic of the investigated combustor [3,4].
Aerospace 12 00676 g001
Figure 2. Schematic of the sISR extinction prediction methodology.
Figure 2. Schematic of the sISR extinction prediction methodology.
Aerospace 12 00676 g002
Figure 3. Velocity magnitude (1st row), and mixture fraction (2nd row) for case A (1st column), case B (2nd column), and case C (3rd column).
Figure 3. Velocity magnitude (1st row), and mixture fraction (2nd row) for case A (1st column), case B (2nd column), and case C (3rd column).
Aerospace 12 00676 g003
Figure 4. SDR for case A (1st column), case B (2nd column), and case C (3rd column).
Figure 4. SDR for case A (1st column), case B (2nd column), and case C (3rd column).
Aerospace 12 00676 g004
Figure 5. T s t o i vs. time for a fully burning and blow-off realisation at N 0 = 60   s 1 .
Figure 5. T s t o i vs. time for a fully burning and blow-off realisation at N 0 = 60   s 1 .
Aerospace 12 00676 g005
Figure 6. N 0 vs. time for a fully burning and blow-off realisation at N 0 = 60   s 1 .
Figure 6. N 0 vs. time for a fully burning and blow-off realisation at N 0 = 60   s 1 .
Aerospace 12 00676 g006
Figure 7. Mean probability of extinction with vertical bars representing binominal confidence intervals. The critical threshold for LBO is indicated by the red star symbol.
Figure 7. Mean probability of extinction with vertical bars representing binominal confidence intervals. The critical threshold for LBO is indicated by the red star symbol.
Aerospace 12 00676 g007
Figure 8. LBO curve and extinction dilution ratio prediction. Blue symbol: reference cold-flow RANS CFD simulation.
Figure 8. LBO curve and extinction dilution ratio prediction. Blue symbol: reference cold-flow RANS CFD simulation.
Aerospace 12 00676 g008
Table 1. Operating conditions for the reference CFD cases. R is the dilution ratio.
Table 1. Operating conditions for the reference CFD cases. R is the dilution ratio.
Case ϕ g R   =   m d ˙ / ( m d ˙ + m p ˙ )
A0.050.1
B0.070.1
C0.090.1
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

Awad, H.S.A.M.; Gkantonas, S.; Mastorakos, E. Low-Order Modelling of Extinction of Hydrogen Non-Premixed Swirl Flames. Aerospace 2025, 12, 676. https://doi.org/10.3390/aerospace12080676

AMA Style

Awad HSAM, Gkantonas S, Mastorakos E. Low-Order Modelling of Extinction of Hydrogen Non-Premixed Swirl Flames. Aerospace. 2025; 12(8):676. https://doi.org/10.3390/aerospace12080676

Chicago/Turabian Style

Awad, Hazem S. A. M., Savvas Gkantonas, and Epaminondas Mastorakos. 2025. "Low-Order Modelling of Extinction of Hydrogen Non-Premixed Swirl Flames" Aerospace 12, no. 8: 676. https://doi.org/10.3390/aerospace12080676

APA Style

Awad, H. S. A. M., Gkantonas, S., & Mastorakos, E. (2025). Low-Order Modelling of Extinction of Hydrogen Non-Premixed Swirl Flames. Aerospace, 12(8), 676. https://doi.org/10.3390/aerospace12080676

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