Fact-checked by Grok 2 weeks ago

Premixed flame

A premixed flame is a type of flame that occurs when and oxidizer are thoroughly mixed to form a homogeneous premixture prior to ignition, resulting in a propagating front—typically a —where the is governed primarily by and rates rather than fuel-oxidizer mixing at the flame. In premixed flames, the laminar burning velocity, a key metric representing the speed at which the flame propagates through the unburned mixture under quiescent conditions, typically peaks near stoichiometric equivalence ratios and can reach values such as 45 cm/s for methane-air mixtures at and temperature. This velocity increases with reactant temperature (e.g., up to ~2.5 m/s at 1000 K) and decreases with , while can enhance the effective by wrinkling the flame surface, leading to regimes like flamelet or distributed reaction zones depending on the turbulent Damköhler and Karlovitz numbers. The structure features distinct zones: a preheat layer where conducts ahead of the reaction, a thin reaction zone (often 0.1–1 mm thick) with rapid exothermic reactions and radical formation (e.g., H-atom peaks in hydrogen flames), and a post-flame recombination layer. Premixed flames are fundamental to numerous applications due to their and over emissions, including spark-ignition internal combustion engines, Bunsen and flat-flame burners for laboratory studies, combustors (especially lean-premixed systems to reduce ), and rocket engines. Theoretical models, such as the Zel'dovich-Frank-Kamenetskii theory, describe propagation as proportional to the square root of times , providing a basis for predicting and under stretch or . These flames can transition to detonations under high-speed conditions, posing safety considerations in industrial processes like prevention.

Fundamentals

Definition and Characteristics

A premixed flame occurs when and oxidizer are uniformly mixed prior to ignition, leading to a thin reaction zone where propagates through the mixture as a wave. This uniform mixing distinguishes premixed flames from diffusion flames, enabling a well-defined flame front driven primarily by rates rather than mixing processes. Key characteristics of premixed flames include a thin flame front, typically on the order of millimeters, comprising distinct preheat and zones; an representing the theoretical maximum temperature under ideal no-loss conditions; the equivalence ratio, which dictates mixture and thus flame behavior; and a release rate that measures the energy liberation from exothermic reactions. The equivalence ratio \phi, defined as the ratio of the actual fuel-to-oxidizer to the stoichiometric value, classifies mixtures as (\phi < 1), stoichiometric (\phi = 1), or rich (\phi > 1); it profoundly affects the , with maximum values occurring near \phi = 1 (slightly fuel-rich for hydrocarbons), and influences flame stability and completeness of . The release rate, a fundamental quantity in premixed , quantifies the volumetric or mass-specific rate of production in the zone, impacting dynamics and acoustic interactions. The basic structure of a premixed flame features a preheat zone ahead of the reaction front, where conductive from downstream raises the of the unburned from the initial value T_0 toward the ignition , with minimal chemical activity and gradual depletion of reactants. This is followed by the narrow zone, where intense chain-branching reactions rapidly consume and oxidizer, producing intermediates (e.g., radicals like H and OH) and releasing heat, causing a sharp rise to the T_{ad} and the formation of primary products. Behind the zone lies the post-combustion region, where slower recombination reactions equilibrate concentrations, with profiles showing a monotonic increase across the front and profiles exhibiting steep gradients—reactants dropping abruptly, intermediates peaking briefly, and products accumulating to levels. The adiabatic flame temperature T_{ad} arises from the first-law energy balance for an adiabatic, constant-pressure process, approximated under constant specific heat. For lean mixtures (\phi \leq 1), T_{ad} = T_0 + \frac{\Delta H_c}{c_p} Y_f \phi, where \Delta H_c is the magnitude of the heat of combustion per unit mass of fuel, c_p is the average specific heat capacity of the mixture, and Y_f is the stoichiometric fuel mass fraction; for rich mixtures (\phi > 1), \phi is effectively replaced by 1 due to oxidizer limitation. This approximation assumes complete reaction and neglects dissociation effects, which lower actual temperatures at high T_{ad}.

Comparison to Other Flame Types

In premixed flames, fuel and oxidizer are thoroughly mixed prior to ignition, resulting in a uniform reactive that propagates as a thin front driven by the intrinsic laminar burning velocity. In contrast, flames, also known as non-premixed flames, involve separate streams of and oxidizer that mix primarily through molecular and turbulent at the flame sheet, where the burning rate is controlled by the rate of reactant supply rather than an inherent propagation speed. This fundamental difference in mixture preparation leads to distinct flame structures: premixed flames exhibit a single, well-defined reaction zone with sharp gradients in and , while flames feature broader zones shaped by counter- gradients. Partially premixed flames represent an intermediate mode, where fuel and oxidizer are mixed to some extent upstream of the flame but continue to mix during combustion, combining elements of premixed propagation with diffusion-controlled regions. Unlike fully premixed flames, which rely on complete upstream homogeneity for consistent reactivity, partially premixed flames often develop stratified equivalence ratios, leading to dual flame branches—a premixed leading edge and a trailing diffusion flame. Lifted flames, typically a subset of non-premixed configurations, occur when the flame stabilizes away from the burner due to partial premixing in the shear layer, contrasting with the anchored, propagating nature of premixed flames. These distinctions have key implications for combustion behavior. Premixed flames produce minimal soot due to controlled stoichiometry that avoids localized fuel-rich pockets, whereas diffusion flames generate higher soot levels from such pockets formed during mixing. Flame speeds in premixed systems are intrinsically higher, typically on the order of 0.3–0.5 m/s for hydrocarbon-air mixtures, compared to the diffusion-limited rates in non-premixed flames that depend on flow and geometry. Additionally, premixed flames are more susceptible to intrinsic instabilities, such as Darrieus–Landau hydrodynamic and thermo-diffusive modes, arising from their uniform reactivity, while non-premixed flames exhibit greater stability due to the stabilizing effect of differential diffusion. The recognition of premixed flames as distinct from open diffusion flames dates to 19th-century studies, notably Humphry Davy's investigations into explosions for his 1815 safety lamp, which highlighted propagation limits in homogeneous gas-air mixtures versus uncontrolled open s.

Propagation Mechanisms

Laminar Propagation

In laminar premixed flames, propagation occurs steadily and without wrinkles through quiescent reactant mixtures, driven by the coupled processes of heat conduction from the reaction zone to the preheat region, mass diffusion of reactants toward the flame front, and exothermic chemical reactions confined to a thin zone of order 0.1–1 mm thickness. This mechanism ensures a planar flame front advances at a constant speed relative to the unburned gas, with the preheat zone facilitating thermal activation and the reaction zone consuming fuel and oxidizer efficiently. The , defined as Le = \alpha / D where \alpha is the and D is the of the deficient reactant, critically influences the behavior by governing the relative rates of and . For Le > 1, typically in hydrocarbon-air mixtures, diffusion outpaces , stabilizing the planar front and promoting uniform . Conversely, for Le < 1, as in rich mixtures or hydrogen-doped fuels, dominates, leading to diffusive- instabilities that perturb the front and can locally reverse or accelerate directions through cellular structures. The laminar burning velocity S_L, the normal component of the flame speed relative to the unburned gases ahead of the front, quantifies this propagation rate and serves as a fundamental property for characterizing . S_L depends strongly on unburned gas temperature (increasing roughly linearly due to enhanced reaction rates), pressure (decreasing as S_L \propto p^{-n} with n \approx 0.1–0.3 from third-body recombination effects), and mixture composition (peaking near stoichiometric equivalence ratios where adiabatic flame temperature is maximized). For instance, in methane-air mixtures at standard temperature and pressure (298 K, 1 atm), stoichiometric S_L values range from 0.35 to 0.42 m/s, reflecting variations in measurement precision and minor impurities. Experimental determination of S_L avoids turbulence influences by employing techniques like the constant volume bomb, where a spherical flame expands in a sealed vessel and S_L is derived from the early-stage pressure rise via thermodynamic relations assuming adiabaticity. Alternatively, the heat flux method stabilizes a quasi-one-dimensional flame on a heated porous burner, balancing conductive heat losses upstream to extrapolate the unstretched S_L from measured profiles. These methods yield uncertainties below 3% under controlled conditions, enabling reliable data for model validation. The Zeldovich-Frank-Kamenetskii theory provides a foundational one-dimensional model for S_L, treating the flame as an eigenvalue problem balancing diffusion and reaction timescales under large activation energy asymptotics. The resulting expression approximates the propagation speed as S_L = \sqrt{ \frac{\lambda}{\rho_u c_p} \left( \frac{d \ln \omega}{dT} \right)_{\max} }, where \lambda is the thermal conductivity, \rho_u the unburned gas density, c_p the specific heat at constant pressure, and \left( \frac{d \ln \omega}{dT} \right)_{\max} the maximum logarithmic derivative of the reaction rate \omega with respect to temperature, capturing the sensitivity of chemistry to thermal gradients in the reaction zone. This formulation highlights S_L's square-root dependence on transport properties and exponential sensitivity to kinetics, aligning with experimental trends while assuming unity Lewis number for simplicity.

Turbulent Propagation

In turbulent premixed flames, fluid turbulence interacts with the flame front by imposing unsteady straining and wrinkling, which dramatically increases the flame's surface area and accelerates propagation speeds beyond laminar levels. This interaction is delineated into distinct regimes using the Karlovitz number Ka = \frac{u'}{S_L} \cdot \frac{l_t}{\delta_L}, representing the ratio of large-eddy turnover time to chemical time, and the Damköhler number Da = \frac{l_t / S_L}{\tau_\mathrm{chem}}, capturing the balance between turbulent diffusion and reaction timescales, where u' denotes turbulence intensity, l_t the integral length scale, \delta_L the laminar flame thickness, S_L the laminar burning velocity, and \tau_\mathrm{chem} the chemical timescale. These parameters define key regimes of turbulence-flame coupling. The wrinkled flamelet regime occurs at low Ka (< 1) and high Da (> 1), where turbulence wrinkles the thin flame sheet without disrupting its internal structure, leading to enhanced burning primarily through geometric surface enlargement. As Ka rises to intermediate levels (around 1–10) with sustained high Da, the flamelet regime emerges, featuring more intense surface distortions but preserved local laminar-like . In the thin reaction zone regime (higher Ka, Da > 1), small-scale eddies penetrate the preheat layer (\eta < \delta_L < l_t), broadening it while the narrow reaction sheet (~0.1 \delta_L) remains intact, avoiding widespread extinction. At elevated Ka (>> 10) and low Da (< 1), the distributed (or broken reaction zone) prevails, where turbulence fully fragments the flame, thickening and homogenizing the reaction zone across the volume. Turbulence markedly boosts the turbulent burning velocity S_T relative to S_L, with empirical scalings in the flamelet regimes approximating \frac{S_T}{S_L} \approx 1 + c \left( \frac{u'}{S_L} \right)^n, where the constant c depends on flow conditions and the exponent n varies from ~0.5 for moderate wrinkling (emphasizing diffusive effects) to ~1 for strong turbulence (approaching volumetric burning). These power laws arise from the increased flame area due to multi-scale wrinkling, transitioning toward linear dependence on u' in distributed regimes per Damköhler's hypothesis. The flamelet regimes maintain a thin, propagating front akin to a collection of strained laminar elements, enabling closure models based on surface area augmentation, whereas distributed combustion shifts to volumetric reaction rates with potential local extinctions and reignitions. The integral length scale l_t primarily drives wrinkling amplitude, with larger l_t yielding greater surface folding and higher S_T / S_L for fixed u', as it allows broader convective distortions. In contrast, the Kolmogorov scale \eta governs fine-scale quenching risks; when \eta \ll \delta_L, intense small eddies impose excessive straining, causing local flame quenching and cusps, particularly in high-Ka thin reaction zones, though recovery often occurs via reignition in unburned mixture pockets. Recent experimental investigations in 2025 have refined scaling laws for S_T in expanding turbulent premixed flames, highlighting deviations from stationary assumptions and the role of radial growth in regime transitions, as detailed in Physical Review Fluids.

Experimental Configurations

Bunsen Flame

The Bunsen flame is established on a cylindrical tube burner where a premixed fuel-oxidizer mixture exits at a controlled velocity into quiescent ambient air, forming a stable, conical flame shape due to the balance between the laminar burning velocity of the mixture and the axial flow velocity at the burner exit. This configuration assumes laminar premixed flame propagation, as detailed in related sections on propagation mechanisms. The burner typically features a straight or contoured nozzle to ensure uniform exit velocity, often with water cooling to prevent preheating of the incoming mixture and optional turbulence-generating elements upstream for controlled studies. The geometry of the Bunsen flame manifests as an axisymmetric cone, with the half-cone angle \alpha determined by the relation S_L = u_0 \sin \alpha, where S_L is the laminar burning velocity and u_0 is the unburned mixture velocity at the nozzle exit. At the flame tip, curvature increases the local stretch rate, which can elevate the burning velocity compared to the planar flame value due to hydrodynamic effects. The cone base is anchored at the burner rim, while the surface remains relatively smooth under low stretch conditions, enabling clear visualization of the flame front via schlieren imaging or chemiluminescence. In combustion research, the Bunsen flame serves as a fundamental tool for visualizing the premixed flame front and quantifying laminar burning velocities, achieved by measuring the cone angle \alpha for a given u_0 or by adjusting u_0 to maintain a constant \alpha across conditions. This angle method, often combined with high-speed imaging, allows precise determination of S_L for various mixtures, pressures, and temperatures, providing data for validation of chemical kinetic models. Despite its simplicity, the Bunsen flame exhibits limitations from edge effects, where preferential diffusion of heat and species at the flame periphery into the surrounding air causes local thickening of the reaction zone and deviations in measured burning velocities. Quenching at the burner rim further perturbs the flow and angle measurements, introducing errors up to several percent, particularly at higher temperatures. Historically, the configuration was introduced by in 1855 to produce a clean, soot-free flame for laboratory gas analysis and spectroscopic studies of elements.

Stagnation Flame

The stagnation flame configuration, commonly referred to as the counterflow premixed flame, employs two opposed jets: one delivering the premixed fuel-oxidizer mixture and the other supplying an inert gas or additional oxidizer, resulting in a collision that establishes a stagnation plane for flame anchoring. This setup generates a controlled aerodynamic strain on the flame, with the strain rate defined as a = \frac{2u}{L}, where u represents the jet exit velocity and L the nozzle separation distance, approximating the potential flow field near the stagnation point. The flame stabilizes at or near this plane, enabling precise control over flow conditions without significant gravitational interference in compact apparatus. Key characteristics of the stagnation flame include a relatively flat front, which minimizes geometric distortions and allows direct observation of strain-induced responses. By incrementally raising the strain rate through increased jet velocities, researchers can measure extinction limits, where the flame fails to sustain beyond a critical strain value specific to the mixture. This configuration is advantageous for isolating aerodynamic influences, as buoyancy effects are negligible in small-scale, horizontally oriented burners or under microgravity conditions. In research applications, the stagnation flame serves to quantify the Markstein number, a parameter indicating the flame's sensitivity to stretch at extinction, by analyzing flame position shifts relative to the stagnation plane under varying strain. It also enables evaluation of the laminar burning velocity S_L under strained conditions, providing insights into stretch effects on propagation. Such measurements are typically conducted in burner-stabilized opposed-jet apparatuses, facilitating detailed diagnostics like laser-induced fluorescence for species profiling. Relative to the Bunsen flame, the stagnation configuration reduces curvature impacts on the flame front, promoting more accurate one-dimensional theoretical modeling and validation. This approach traces its development to mid-20th-century opposed-jet studies, which pioneered controlled strain investigations in combustion.

Spherical Flame

The spherical premixed flame configuration involves igniting a quiescent, homogeneous combustible mixture at the center of a closed spherical vessel, commonly referred to as a constant-volume bomb. Upon ignition, typically via a weak spark or laser-induced breakdown, the flame propagates outward as a thin, expanding front, capturing the evolution from inception through acceleration. In the early stages, before appreciable pressure buildup, the flame radius R(t) grows linearly with time as R(t) \approx S_L t, where S_L is the unstrained laminar burning velocity, approximating constant-pressure conditions. The dynamics of this expanding flame are governed by transient stretch effects primarily due to geometric curvature, with the mean curvature \kappa = 2/R inducing diffusive losses at the convex flame surface. This curvature modifies the local flame speed via the relation S_L(\kappa) = S_L^0 (1 - \mathrm{Ma} \, \kappa \, \delta_L), where S_L^0 is the unstrained speed, \mathrm{Ma} is the Markstein number quantifying sensitivity to curvature (typically positive for lean mixtures, stabilizing the front), and \delta_L is the flame thickness. As the flame expands and R increases, \kappa decreases, reducing stretch and allowing S_L to approach S_L^0; however, this evolution can accelerate the front nonlinearly if stabilizing effects weaken. This setup holds significant research value for accurately determining S_L by extrapolating stretched flame speeds to zero stretch in constant-pressure-like early phases, offering advantages over other geometries for high-pressure measurements. It also enables direct visualization of cellular structures emerging from the onset of , where density gradients across the flame drive perturbations into growing modes once the flame radius exceeds a critical value related to \delta_L. Recent 2025 experiments on turbulent spherical flames have derived scaling relations for speed enhancement, linking turbulence intensity to instability-amplified surface area in expanding fronts.

Theoretical Modeling

Burning Velocity

The laminar burning velocity, denoted S_L, represents the fundamental speed at which a planar premixed flame propagates into an unburned mixture under unstretched, adiabatic conditions. Its theoretical foundation stems from asymptotic analysis assuming large activation energy, which confines the reaction zone to a thin layer within the flame structure. This approach, pioneered by , simplifies the governing conservation equations by separating the preheat and reaction zones. In the asymptotic limit of large activation energy E / (R T_b) \gg 1, where E is the activation energy, R is the gas constant, and T_b is the burned gas temperature, the flame structure consists of a relatively thick preheat zone followed by an infinitesimally thin reaction sheet. The energy conservation equation in one dimension for steady propagation is \rho_u S_L c_p \frac{dT}{dx} = \frac{d}{dx} \left( \lambda \frac{dT}{dx} \right) + Q \omega, where \rho_u is the unburned density, c_p is the specific heat, \lambda is the thermal conductivity, Q is the heat release per unit mass, and \omega is the reaction rate. Neglecting the source term in the preheat zone (T < T_i, where T_i is the ignition temperature) and integrating yields the temperature gradient at the reaction sheet. In the reaction zone, convection is negligible, and using temperature as the independent variable, the equation becomes \lambda \left( \frac{dT}{dx} \right)^2 = - \rho_u S_L c_p (T - T_u) Q \omega / \left( \frac{dT}{dx} \right), but matching gradients at T = T_i leads to the burning velocity expression. For constant properties, the derivation results in S_L^2 = \frac{2 \lambda}{\rho_u c_p (T_b - T_u)^2} \int_{T_u}^{T_b} \omega \, dT, where the integral captures the integrated reaction rate over the temperature range from unburned T_u to burned T_b. This form arises from balancing conduction in the preheat zone with the exothermic reaction, assuming Lewis number unity and a one-step irreversible reaction. The laminar burning velocity depends strongly on mixture composition and thermodynamic conditions. Equivalence ratio \phi (fuel-to-oxidizer ratio relative to stoichiometric) influences S_L through the reaction rate and flame temperature, typically peaking near \phi = 1 and following an approximate form S_L \propto \phi^n (1 - \phi)^m for lean mixtures, with exponents n and m varying by fuel (e.g., n \approx 2, m \approx 1 for hydrocarbons). Initial temperature T_0 affects S_L via enhanced diffusivity and reaction rates, often scaling as S_L \propto T_0^\alpha with \alpha \approx 1.5 to 2 for many fuels. Pressure p impacts dissociation and collision rates, leading to S_L \propto p^\beta where \beta \approx -0.1 to -0.25, reflecting reduced velocity at higher pressures due to three-body recombination. These power-law dependencies are derived from asymptotic sensitivity and validated across fuels like methane and hydrogen. The Lewis number Le = \lambda / (\rho c_p D), the ratio of thermal to mass diffusivity D, modulates S_L and flame stability. For Le = 1, the flame is stable; Le < 1 (e.g., lean hydrogen mixtures) promotes Darrieus-Landau hydrodynamic instability enhanced by preferential diffusion, increasing local S_L at convex regions and leading to cellular structures; conversely, Le > 1 (e.g., lean hydrocarbons) stabilizes the flame by reducing S_L at protrusions. This thermodiffusive effect couples with hydrodynamics, influencing overall propagation without altering the planar S_L but affecting wrinkled flame dynamics. Extensions to non-planar flames account for stretch effects, where and alter local . The Markstein correction linearizes this as the stretched velocity S_T = S_L - [M](/page/M) a \cdot \kappa, with [M](/page/M) a the Markstein number (positive for stabilizing effects) and \kappa = (1/A) (dA/dt) + S_L / R the stretch rate combining tangential and $1/R. Derived from weakly curved flame asymptotics, [M](/page/M) a depends on Le and transport properties, quantifying sensitivity to flow-induced deformation. In flamelet models for mildly turbulent or stretched premixed combustion, the turbulent burning velocity S_T is modeled as S_T = S_L \cdot A_T^{1/2}, where A_T is the ratio of wrinkled flame area to planar area, assuming thin flames without extinction. This arises from integrating local S_L over the increased surface, valid in the flamelet regime where turbulence wrinkles but does not thicken the reaction zone. The model, rooted in asymptotic flamelet theory, bridges laminar and turbulent propagation. Modern theoretical advancements incorporate numerical (CFD) models to compute S_L beyond asymptotic approximations, using detailed chemistry mechanisms. Software like solves 1D steady flame equations with multi-step kinetics, , and , yielding S_L profiles sensitive to species and —omissions in classical . As of 2025, these simulations support high-fidelity predictions for complex fuels, integrating over thousands of reactions for accuracy unattainable analytically.

Flame Instabilities

Flame instabilities in premixed flames refer to perturbations that cause deviations from a planar propagating front, resulting in wrinkling, cellular patterns, or fingering that can accelerate burning rates or lead to . These instabilities stem from the inherent properties of the flame, such as density gradients, differential , and external accelerations, and are analyzed through theory to determine growth rates of perturbations with wavenumber k. The hydrodynamic instability arises from the abrupt jump across the interface, where the expansion of burned gases (\rho_b < \rho_u) induces divergences that amplify initial . Seminal analyses by Darrieus and Landau established that planar premixed are unconditionally unstable to long-wavelength , with the growth rate approximated as \sigma_{DL} \approx S_L k \sqrt{\frac{\rho_u - \rho_b}{\rho_u + \rho_b}} - D k^2, where S_L is the laminar burning velocity, k is the perturbation , and D represents diffusive stabilization for short wavelengths. This mechanism dominates in mixtures with significant ratios, such as hydrocarbon-air , leading to corrugated surfaces observed in various configurations. Diffusive-thermal instability complements hydrodynamic effects when the Lewis number Le = \alpha / D_f < 1, where \alpha is and D_f is the fuel diffusion coefficient, as in lean hydrogen-air mixtures. Here, diffuses faster than the deficient reactant, creating local hot spots that accelerate local burning and form cellular structures on the front. This is particularly pronounced in low-Le fuels, promoting self-sustained oscillations and enhanced propagation speeds without external forcing. The Rayleigh-Taylor (RT) instability occurs under of the flame front, such as during rapid expansion in explosions, where the lighter burned products are pushed into the denser unburned . The follows \sigma_{RT} \approx \sqrt{A g k} - D k^2, with Atwood number A = (\rho_u - \rho_b)/(\rho_u + \rho_b) and g the effective , leading to mixing and potential if the perturbation overwhelms the reaction zone thickness. Recent investigations into sustainable fuels have highlighted unique instability behaviors in ammonia-hydrogen (NH3/H2) premixed flames, where blended low-Le effects exacerbate cellular formation and thermo-acoustic coupling at elevated pressures, as explored in 2025 studies published in Combustion and Flame. These findings underscore gaps in understanding multi-fuel instabilities for decarbonization applications.

Applications and Developments

Industrial and Engineering Applications

Premixed flames have been integral to industrial processes since the , with early applications in town gas burners for lighting, heating, and cooking using coal-derived gases that were mixed with air prior to . These systems, developed in the early in and later adopted across and , enabled efficient fuel use in factories and households, marking a shift from solid fuels to gaseous ones for controlled burning. The invention of the in 1857 further standardized premixed for laboratory and small-scale industrial heating, where and air are precisely mixed to produce a stable, adjustable flame. In modern engineering, premixed flames are employed in gas turbines through lean premix combustors, such as dry low (DLN) systems, where and are thoroughly mixed upstream to achieve . Similarly, spark-ignition engines rely on homogeneous charge premixed , forming a - in the intake manifold before spark initiation for power generation in automotive and industrial applications. Bunsen-type burners continue to serve as foundational designs in laboratory and pilot-scale industrial setups for processes like flame testing and material heating. The primary benefits of premixed flames in these devices include reduced emissions from uniform mixing and lower peak temperatures; for instance, lean premix in gas turbines can limit to 15–25 ppm while also decreasing and volatile organic compounds compared to diffusion flames. Enhanced efficiency arises from complete , with spark-ignition engines achieving up to 10% lower fuel consumption through optimized charge homogeneity. However, challenges persist, including flashback, where the flame propagates upstream into the premix zone if the flow velocity falls below the laminar burning velocity, and autoignition in high-temperature regions of the , both of which pose risks to equipment integrity.

Modern Research and Sustainable Fuels

Recent research on premixed flames has increasingly focused on sustainable fuels to support decarbonization efforts, particularly () and ()/ blends as zero-carbon alternatives. premixed /air flames exhibit high laminar burning velocities exceeding 2 m/s, attributed to the fuel's high , alongside (Le) values below 1 that promote thermo-diffusive instabilities, leading to enhanced flame wrinkling and accelerated propagation under turbulent conditions. These characteristics make /air mixtures promising for low-emission combustion, though instabilities require careful control to prevent flashback. Similarly, / blends offer carbon-free combustion with stable flame holding comparable to when hydrogen content reaches 30 vol.%, but they face challenges from elevated emissions due to N2O intermediates formed during oxidation. A 2025 study in Combustion and Flame detailed parametric analyses of NO and N2O pathways in flames, highlighting the need for optimized blending ratios to mitigate while maintaining low-carbon benefits. Advances in premixed flame applications emphasize partially premixed configurations for gas turbines, where enrichment enables net-zero carbon operations by boosting reactivity and reducing unburned hydrocarbons. Hydrogen addition to methane-air mixtures in these setups increases laminar burning velocities and pools, improving across a wider equivalence ratio range, as demonstrated in high-fidelity simulations of swirling flames. Numerical modeling has progressed with unsteady Reynolds-averaged Navier-Stokes (URANS) and () approaches to capture turbulent premixed flames, incorporating thermo-diffusive effects for accurate prediction of interactions and formation. These models, validated against direct numerical simulations, reveal synergistic turbulence- effects that enhance burning rates in lean mixtures. Emerging trends in 2025 research include advanced diagnostics and theoretical analyses for alternative fuels near operational limits. has been applied to quantify species profiles and temperature in premixed NH3/O2/Ar flames, enabling non-intrusive measurements of intermediates and validation of kinetic mechanisms despite fluorescence interferences. Additionally, studies on chain-branching reactions near flammability limits, using asymptotic models, address nonadiabatic effects like radiative losses that influence in weakly burning premixed flames, providing insights into stability for dilute or low-reactivity mixtures. These efforts underscore the potential of NH3 and H2-based premixed flames for systems while highlighting ongoing challenges in emissions control and modeling fidelity.

References

  1. [1]
  2. [2]
  3. [3]
  4. [4]
  5. [5]
    Premixed Flame - an overview | ScienceDirect Topics
    A premixed flame is defined as a type of flame where the fuel has been mixed with the oxidizer prior to reaching the flame front, with its flame speed primarily ...
  6. [6]
    None
    ### Summary of Laminar Premixed Flames from Lecture Notes
  7. [7]
    Heat release rate markers for premixed combustion - ScienceDirect
    Heat release rate (HRR) is a very important quantity in the study of laminar and turbulent reacting flows. From a practical view point, the spatial distribution ...
  8. [8]
    15.5 Adiabatic Flame Temperature
    ### Summary of Adiabatic Flame Temperature Equation
  9. [9]
    11.6: Adiabatic Flame Temperature
    ### Summary of Adiabatic Flame Temperature
  10. [10]
    The premixed plane flame (Chapter 2) - Theory of Laminar Flames
    In a premixed flame the reactants are constituents of a homogeneous mixture that burns when raised to a sufficiently high temperature. In a diffusion flame the ...
  11. [11]
    Thermodynamics Glossary - Premixed and Diffusion Flames
    Thermodynamics Glossary - Premixed and Diffusion Flames. In a premixed flame, the fuel and oxidizer are mixed before reaching the flame.
  12. [12]
    Effects of confinement on partially premixed flames
    Partially premixed combustion is an intermediate regime between the limiting cases of premixed and nonpremixed combustion. Although combustion problems are ...
  13. [13]
    Flame Radiation and Soot Emission From Partially Premixed ...
    It was found that with progressive partial premixing, the peak soot loading and the thickness of the soot zone first decreased and then increased, and while the ...
  14. [14]
    Comparison of flame response characteristics between Non ...
    Nov 1, 2022 · It is clear that the stability of non-premixed flame is superior to that of premixed flame. Accordingly, this study verifies flame stability by ...
  15. [15]
    Humphrey Davy and the safety lamp: the use of metal gauze as a ...
    Oct 23, 2015 · The 'safety lamp' invented by Humphrey Davy in 1815 utilised the cooling effect of metal gauze to prevent the flame of a candle or oil lamp.<|control11|><|separator|>
  16. [16]
    None
    ### Summary of Zeldovich-Frank-Kamenetskii Theory Section
  17. [17]
    Lewis number effect on the propagation of premixed flames in ...
    For Le < 1, the burning rate always increases as the flame travels from one end of the tube to the other. For Le > 1, the evolution is extremely sensitive to ...
  18. [18]
    [PDF] Theory of premixed-flame propagation in large-scale turbulence
    Specifically, the Lewis number (the ratio of thermal to molecular diffusivity) is set equal to unity, and the density change associated with the heat release of ...
  19. [19]
    Effect of CH4, Pressure, and Initial Temperature on the Laminar ...
    Apr 26, 2021 · The CH4 in the fuel affects the flame speed by changing the main species of free radicals in the flame; the high pressure not only increases the ...
  20. [20]
    Laminar Burning Velocities of Hydrogen-Blended Methane–Air and ...
    Nov 12, 2021 · For this stoichiometric CH4–air mixture, the earlier reported LBVs vary within 35 and 41.5 cm s−1 [4,5,6,62] at ambient initial conditions, a ...
  21. [21]
    Laminar burning velocity measurements in constant volume vessels
    Jan 1, 2018 · Laminar burning velocity measurements have been made in a constant volume vessel using both flame front imaging and the pressure rise ...
  22. [22]
    Accurate measurements of laminar burning velocity using the Heat ...
    The Heat Flux method was further developed to significantly reduce its experimental uncertainty and used to determine burning velocities under conditions ...
  23. [23]
    Karlovitz Numbers and Premixed Turbulent Combustion Regimes ...
    The present paper aims to overview different definitions of K a , comparing them, and suggesting the most appropriate choice of K a for each combustion regime ...
  24. [24]
    Scaling of turbulent flame speed for expanding flames with ...
    Sep 9, 2013 · In this paper we clarify the role of Markstein diffusivity, which is the product of the planar laminar flame speed and the Markstein length, on the turbulent ...
  25. [25]
    Turbulence Intensity and Length Scale Effects on Premixed ...
    Dec 29, 2021 · The turbulent flame speed is found to increase with increasing turbulence intensity and also with increasing integral length scale.
  26. [26]
    Effect of turbulent motions at different length scales on turbulent ...
    However, whether local turbulent motions scaled with the Kolmogorov length scale can actually influence flame topology is not clear. A recent study [13] has ...Missing: quenching | Show results with:quenching
  27. [27]
    [PDF] Measurement of Flame Speeds by a Nozzle Burner Method 1
    The nozzle burner method uses stable Bunsen-type flames above nozzles, with controlled mixture composition, temperature, and pressure, and varying initial ...Missing: formula | Show results with:formula
  28. [28]
    Burning Velocity of Turbulent Methane/Air Premixed Flames in ...
    Sep 21, 2020 · The addition of syngas increases the turbulent burning velocity up to 10.3%. As previously discussed, studies at high pressure conditions have ...Introduction · Methodology · Conclusions · References
  29. [29]
    Aerodynamics of a slender axisymmetric Bunsen flame with large ...
    The shape of a Bunsen flame in a given flow has been computed by Lewis and von Elbe [1] for a constant burning velocity and by Sivashinsky [18] and Matalon et ...Missing: formula | Show results with:formula
  30. [30]
    [PDF] C. K. Law
    The influence of stretch and preferential diffusion on premixed flame extinction and stability have been investigated via two model flame configurations, namely ...
  31. [31]
    The Origin of the Bunsen Burner | Journal of Chemical Education
    Abstract: In response to a reader query, the column traces the origins of the famous Bunsen burner.
  32. [32]
    Determination of Markstein numbers in counterflow premixed flames
    We have proposed a method for measuring the Markstein number, relative to both the unburned and burned gases, in flames with chemical zones of finite width.
  33. [33]
    Experiments and modelling of premixed laminar stagnation flame ...
    Jun 23, 2011 · The jet-wall geometry yields a stable, steady flame and allows for precise measurement and specification of all boundary conditions on the flow.
  34. [34]
    The effect of strain rate on a premixed laminar flame - ScienceDirect
    Increasing strain rate causes the flame front to move away from the stagnation point, and then back towards it for very large strain rates. The partial ...
  35. [35]
    LES of premixed and non-premixed combustion in a stagnation ...
    In the premixed mode, flame is attached to the injector, while in the non-premixed mode, flame is lifted from the injector. Results are used to highlight some ...
  36. [36]
  37. [37]
    Determination of Markstein numbers in counterflow premixed flames
    Aug 9, 2025 · We show how to measure the Markstein number of the flame with respect to both the unburned and burned gases. The numerical values of these ...
  38. [38]
    Non-premixed flame characteristics of opposed methane jets in ...
    Before these methods, many previous studies were conducted with a counter flowing jet flame configured in an ambient or a wide air stream (Tsuji and Yamaoka, ...Missing: early | Show results with:early
  39. [39]
    The constant-volume propagating spherical flame method for ...
    This method can be used to measure laminar flame speed at high pressures and temperatures which are close to engine-relevant conditions.Missing: setup | Show results with:setup
  40. [40]
    Finding the markstein number using the measurements of ...
    The linear equation with respect to the flame stretch rate cannot approximate the experimental data corresponding to high flame curvature, while the linear ...
  41. [41]
    The curvature Markstein length and the definition of flame ...
    The flame speed-flame stretch relation is modulated by a coefficient on the order of flame thickness, which mimics the effects of diffusion and chemical ...
  42. [42]
    Modeling the formation and growth of instabilities during spherical ...
    A freely expanding spherical flame can experience significant acceleration due to the growth of the Darrieus–Landau instability, increasing the severity of the ...
  43. [43]
    Experimental study and proposed power correlation for laminar ...
    Jan 29, 2022 · A power-law correlation for temperature and pressure is proposed for methane/hydrogen-air mixtures, which has a practical application in estimating the LBV.
  44. [44]
    The temperature dependence of the laminar burning velocity of ...
    The adiabatic laminar burning velocity is a fundamental parameter of each combustible mixture, which depends on the stoichiometric ratio, pressure and ...
  45. [45]
    Lewis number effects on lean premixed combustion characteristics ...
    The Lewis number (Le), defined as the ratio of thermal to mass diffusivity of the deficient reactant, represents a key property in premixed combustion systems, ...
  46. [46]
    Turbulent burning velocity and thermodiffusive instability of ...
    These results imply that thermodiffusive instability of laminar premixed flames substantially affects burning velocity in weak turbulence only.
  47. [47]
    Curved and stretched flames: the two Markstein numbers
    Sep 28, 2011 · It is shown in this paper that two different Markstein numbers characterize usual wrinkled flames sustained by a multiple-step chemical network ...
  48. [48]
    Impacts of the Lewis and Markstein numbers on premixed flame ...
    Jan 5, 2022 · While the Markstein number may provide a substantial impact on the flame acceleration rate in narrow channels, this effect diminishes with ...
  49. [49]
    Rayleigh–Taylor instability: Modelling and effect on coherent ...
    Modelling of Rayleigh–Taylor instability during premixed combustion is presented. · Rayleigh–Taylor instability introduced as a combustion enhancing mechanism.Missing: quenching | Show results with:quenching
  50. [50]
    Bright Light - Science History Institute
    May 20, 2013 · Coal gas was used for lighting, heating, and cooking from the early 19th century, first in Britain, then in the rest of Europe and in North America.
  51. [51]
    Bunsen burner – Knowledge and References - Taylor & Francis
    A Bunsen burner is a laboratory equipment that is a gas burner designed to efficiently mix air with fuel gas, commonly used for heating in laboratories.
  52. [52]
    [PDF] Lean Pre-Mixed Combustion 3.2.1.2-1 Introduction 3.2.1.2-2 ...
    Lean premix combustion design not only produces lower NOx than diffusion flame technology, but also lowers. CO and volatile organic compounds (VOC), due to ...
  53. [53]
    Premixed Combustion in Spark Ignition Engines and the Influence of ...
    The premixed, homogeneous charge gasoline combustion process in SI engines is influenced by the thermo-chemical state of the cylinder charge.
  54. [54]
    Flames | Research Starters - EBSCO
    Common premixed flames include Bunsen burners, furnace pilot lights, and the internal combustion engines in personal and industrial motor vehicles and airplanes ...
  55. [55]
    Visualization of Different Flashback Mechanisms for H 2 /CH 4 ...
    Flame flashback from the combustion chamber to the premixing section is a major operability issue when using high H2 content fuels in lean premixed.
  56. [56]
    Three-dimensional dynamics of unstable lean premixed hydrogen ...
    The increased wrinkling and flame speed due to external turbulence can be attributed to the synergistic effects between thermo-diffusive instabilities and ...
  57. [57]
    LES combustion model for premixed turbulent hydrogen flames with ...
    Jan 24, 2025 · Premixed hydrogen flames are prone to thermodiffusive instabilities due to strong differential diffusion effects.
  58. [58]
  59. [59]
    A comprehensive parametric study on NO and N2O formation in ...
    Understanding the mechanism of NOx formation and destruction is a prerequisite for the development of effective NOx mitigation techniques in ammonia flames.
  60. [60]
    The impact of H2-enrichment on flame structure and combustion ...
    Nov 15, 2024 · Hydrogen addition to hydrocarbon-air gaseous mixtures results in an increase in the laminar burning velocity as well as in an increase in the radical pool.
  61. [61]
    Predictive URANS/PDF Modeling of Unsteady-State Phenomena in ...
    In this study, we present a numerical investigation of an unsteady hydrogen–air diffusion flame. The main objective is to explore flame propagation dynamics, ...
  62. [62]
    Predictive URANS/PDF Modeling of Unsteady-State Phenomena in ...
    Sep 30, 2025 · This study investigates the viability of hydrogen as a clean energy vector by simulating an unsteady, turbulent, non-premixed hydrogen jet flame ...
  63. [63]
    Flame structure study of premixed NH3/O2/Ar flames using Raman ...
    Quantitative laser-diagnostic studies in NH3 flames have previously been performed with laser-induced fluorescence (LIF) and Coherent Anti-Stokes Raman ...
  64. [64]
    Asymptotic analysis of nonadiabatic chain-branching premixed ...
    Oct 6, 2025 · The present study analyzes the behavior of chain-branching premixed flames near or below the flammability limit.