Fact-checked by Grok 2 weeks ago

Combustor

![Combustor diagram showing key components][float-right] A combustor is the component of a engine, such as a or , where from the upstream is mixed with injected and ignited to generate high-temperature, high-pressure gases that drive the downstream . The primary function of the combustor is to convert the stored in the into through continuous , while maintaining stable propagation, minimizing losses, and ensuring uniform distribution to protect components from thermal damage. Key design challenges include achieving complete , which typically exceeds 99% in modern systems, and controlling emissions of oxides of formed at high temperatures. Common configurations encompass can-type, where individual combustion cans are arranged circumferentially; annular-type, featuring a single ring-shaped chamber; and can-annular (or tubo-annular), combining multiple cans within an annular casing for improved compactness and . These designs enable the combustor to operate under extreme conditions, with inlet air temperatures up to 800°C and temperatures reaching 2000°C or more, facilitating the engine's overall .

Fundamentals

Definition and Operating Principles

A combustor, also referred to as a or burner, is the section of a engine where compressed air from the upstream is mixed with injected and continuously combusted to generate high-temperature, high-pressure gases that expand through the downstream turbine blades to produce mechanical power. This process distinguishes from reciprocating internal combustion engines by employing steady-state, continuous rather than intermittent cycles, enabling rotary motion and high . The combustor must maintain stable flame propagation under varying operating conditions while minimizing pressure losses, typically limited to 4-7% of inlet pressure to preserve overall . Operationally, the combustor receives at pressures up to 40 and temperatures around 400-600°C, depending on the ratio. Approximately 15-25% of this airflow enters the primary zone through liner perforations or swirl vanes to facilitate fuel-air mixing and initial , while the remainder serves as dilution air (30-50%) to moderate exit temperatures to turbine-compatible levels (typically 1200-1600°C) and as film-cooling air to protect the combustor liner from thermal damage. Fuel, usually liquid derivatives or gaseous , is atomized via pressure-atomizing nozzles or vaporized in premixers and injected axially or radially into the primary zone, where ignition—initially via plugs—establishes a self-sustaining stabilized by aerodynamic recirculation zones created by swirl or bluff-body flameholders. occurs predominantly as a in conventional designs, releasing through exothermic oxidation reactions (e.g., C_xH_y + (x + y/4)O_2 → xCO_2 + (y/2)H_2O), adding at near-constant pressure per the Brayton , which underpins efficiency. Key challenges in combustor operation include achieving uniform exit profiles to avoid hot streaks that could overstress components, suppressing combustion instabilities like acoustic oscillations that arise from unsteady heat release coupling with flow acoustics, and controlling emissions such as formed via thermal mechanisms at high s above 1800 K. Modern designs incorporate lean-premixed or staged fueling to lean out the mixture and reduce peak s, thereby lowering while maintaining stability across part-load conditions from idle to full power. The resulting hot gas stream, with velocities around 50-100 m/s, exits the combustor with minimal radial gradients (pattern factor <0.25) to optimize performance and longevity.

Thermodynamic and Chemical Fundamentals

In gas turbine combustors, combustion primarily involves the exothermic oxidation of hydrocarbon fuels, such as kerosene-based , with compressed air supplied from the upstream compressor. The stoichiometric air-fuel mass ratio for kerosene combustion is approximately 15:1, representing the precise amount of air required for complete conversion of fuel to carbon dioxide and water without excess oxygen or unburned hydrocarbons. In practice, combustors operate with significantly leaner mixtures, often exceeding 50:1 air-fuel ratio overall, to dilute the combustion products and limit temperatures that would exceed material tolerances in downstream turbine components. This lean operation promotes stability while minimizing formation through lower peak temperatures, though it risks incomplete combustion if mixing is inadequate. The chemical kinetics of these reactions are governed by chain-branching mechanisms involving radicals like hydroxyl (OH) and hydrogen (H), which propagate flame fronts at speeds on the order of 0.3-1 m/s for premixed hydrocarbon-air flames under combustor conditions. Primary combustion zones achieve near-stoichiometric local equivalence ratios (φ ≈ 1, where φ = stoichiometric air-fuel / actual air-fuel) for efficient energy release, while dilution zones introduce excess air to quench reactions and cool gases. The adiabatic flame temperature for a stoichiometric jet fuel-air mixture reaches approximately 2093°C at standard pressure, derived from equilibrium calculations accounting for dissociation of products like CO2 and H2O at high temperatures. Thermodynamically, the combustor embodies the constant-pressure heat addition process (3-4) of the Brayton cycle, where fuel combustion increases the enthalpy of the working fluid (air plus combustion products) while maintaining near-isobaric conditions due to the low Mach number flow (typically <0.3). The temperature rise ΔT_{3-4} is determined by the energy balance: Δh = q_in = f * LHV / (1 + f), where f is the fuel-air mass ratio, LHV is the lower heating value (≈43 MJ/kg for Jet-A), and dissociation effects reduce the effective heat release by 5-10% at combustor exit temperatures. Actual turbine inlet temperatures are constrained to 1400-1700°C by material limits, far below the stoichiometric adiabatic flame temperature, necessitating film cooling and excess air dilution to achieve thermal efficiency gains from higher overall pressure ratios (up to 40:1 in modern engines). Entropy generation arises from irreversible mixing and finite-rate chemistry, but the process maximizes work extraction in the subsequent expansion by elevating the temperature ratio T_4 / T_3, directly influencing cycle efficiency η = 1 - (T_1 / T_3) * (r_p^{(γ-1)/γ} - 1) / (T_4 / T_3 - 1), where r_p is the compressor pressure ratio and γ ≈1.33 for products.

Historical Development

Early Inventions and Patents

The concept of a continuous-flow combustor emerged in early gas turbine designs, with the first relevant patent issued to English engineer on June 14, 1791 (British Patent No. 1643). Barber's invention described a system compressing atmospheric air via a piston, injecting and igniting fuel in a combustion chamber to produce high-temperature gases, which then expanded through turbine blades connected to a crankshaft for mechanical power or direct propulsion. This layout anticipated modern combustor functions by mixing fuel with compressed air for steady combustion, though limitations in metallurgy and ignition reliability prevented practical operation. Subsequent 19th-century efforts, such as those by engineers like in steam turbines, did not directly advance combustor technology, as focus remained on external combustion or reciprocating engines. Interest in internal continuous combustion revived in the early 20th century amid aviation demands, but no major combustor-specific patents preceded the turbojet era. A pivotal advancement occurred with British Royal Air Force officer , who filed the foundational patent for a turbojet engine on January 16, 1930 (British Patent Application No. 25960/30, granted as GB 347206 in 1931). Whittle's design specified a reverse-flow combustor where compressed air from a centrifugal compressor entered multiple flame tubes, mixed with vaporized liquid fuel, and ignited to generate hot gases for turbine drive and exhaust thrust, achieving combustion efficiencies suitable for sustained flight. This patent emphasized annular or can-annular configurations to ensure uniform burning and minimize pressure losses, addressing instabilities in early flame propagation. Whittle's work, tested in prototypes by 1937, directly influenced subsequent jet engine combustors despite initial challenges with fuel atomization and liner cooling.

Mid-20th Century Advancements

Following World War II, combustor designs evolved to accommodate the transition from centrifugal to axial compressors in jet engines, enabling more uniform airflow distribution and higher combustion efficiency. The can-annular configuration emerged as a significant advancement, combining the modularity of individual cans with an encircling outer casing to reduce overall engine length and weight compared to traditional can-type setups, while facilitating easier maintenance and testing of discrete combustion zones. This design was first widely implemented in high-thrust turbojets, such as the , which entered production in 1953 and produced 10,000 pounds of thrust, marking the first engine in that class. ![Cannular combustor on a Pratt & Whitney JT9D turbofan][float-right] The J57's eight cannular flame tubes improved flame stability and fuel-air mixing through interconnected diffusers, addressing challenges like pressure losses and uneven temperature profiles in earlier can designs, which were prone to hotspots and reduced turbine life. These combustors operated with axial airflow, incorporating multiple fuel nozzles per can for better atomization and combustion completeness at varying throttle settings, essential for military applications like the Boeing B-52 bomber. Parallel developments in annular combustors, building on wartime prototypes like the Westinghouse J30's early annular setup, focused on minimizing wakes between cans to enhance radial uniformity, though can-annular remained dominant in U.S. engines for its balance of performance and serviceability. By the late 1950s, combustor advancements supported the rise of , with designs like the -derived adapting can-annular liners to handle core airflow in bypass configurations, achieving lower pressure drops and improved efficiency at subsonic speeds. Enhanced cooling via convection through perforated liners and dilution air jets allowed inlet temperatures to rise toward 1,200–1,400°C, extending material limits with and reducing carbon buildup from standardized post-1944. These innovations prioritized operational reliability over emissions, reflecting the era's emphasis on thrust-to-weight ratios amid demands.

Post-2000 Innovations

In the early 2000s, combustor designs for aero engines advanced toward staged lean-premixed combustion to minimize NOx formation by controlling peak flame temperatures below 1800 K, addressing and emerging standards. 's (TAPS) combustor, building on dual-annular concepts, features inner and outer premixing zones for sequential fuel staging, reducing NOx by up to 55% compared to prior single-annular designs while limiting CO and unburned hydrocarbons. This technology matured for production in the engine, which achieved FAA certification in 2008 and entered service on the in 2011, enabling operation at overall pressure ratios exceeding 40:1. Pratt & Whitney incorporated axially-staged lean-burn architectures in its PW1000G geared turbofan series, introduced with the PW1100G-JM variant certified in 2014 for the , emphasizing fuel-efficient mixing via axial fuel staging to achieve single-digit NOx margins under landing-takeoff cycles. To mitigate liner durability issues from high thermal loads, P&W deployed a redesigned ceramic-coated combustor liner in 2017, extending component life by optimizing cooling film distribution and reducing oxidation in hydrogen-enriched fuels. Rolls-Royce advanced similar radial-staging in its Trent XWB combustor, certified in 2013, which uses rich-lean zoning to cut NOx by 40% over predecessors through precise swirler-induced turbulence for uniform air-fuel premixing. For stationary gas turbines, dry low-NOx (DLN) systems evolved with multi-nozzle arrays for premixed operation at equivalence ratios near 0.5, as in GE's DLN 2.6+ combustor deployed from 2008 onward, achieving NOx below 9 ppm at base load in F-class units by suppressing hot spots via rapid mixing. Mitsubishi Heavy Industries integrated steam-enhanced combustion in J-series turbines from 2011, injecting water at ratios up to 1.5:1 to dilute flames and lower NOx by 30% without wet controls, supporting 1600°C inlet temperatures. NASA-supported efforts, such as GE's N+2 program completed in 2016, targeted 75% NOx reductions via trapped vortex and micromix injectors, informing hybrid architectures for 2030s hybrid-electric propulsion. These innovations prioritize empirical validation through sector rig tests, revealing trade-offs like increased pressure dynamics managed via active damping.

Design Elements

Core Components

The liner forms the primary containment for the combustion process within the combustor, separating the high-temperature flame zone from the outer casing while allowing controlled cooling via effusion holes, film cooling slots, or convective passages using approximately 25% of the compressor discharge air. This cooling maintains liner wall temperatures below material limits, typically under 1500°F, despite flame temperatures exceeding 2000°F. Fuel injectors, or nozzles, introduce atomized fuel into the primary zone, employing pressure atomizers for smaller droplets (around 10 μm achievable in 0.1 ms) or air-blast designs for larger engines where droplets up to 200 μm vaporize within 40 ms residence times. These injectors angle the fuel spray to promote mixing with swirled air, establishing fuel-rich conditions (equivalence ratio φ > 1.0) for stable ignition and reduced formation in some designs. Swirlers and the dome assembly at the combustor head generate aerodynamic recirculation s by imparting tangential velocity to inlet air, creating a low-velocity primary with swirl numbers exceeding 0.6 to the and ensure continuous auto-ignition of incoming . Primary and secondary air jets downstream bifurcate flow to enclose this , transitioning to leaner mixtures (φ ≈ 0.8) for CO oxidation before dilution holes admit additional air to lower exit temperatures to turbine-compatible levels (φ ≈ 0.3). Igniters, typically spark plugs or glow plugs, provide the initial energy input for startup and relight under off-design conditions, ensuring reliable establishment with minimal pressure loss (5-7% total across the combustor). The outer casing encloses these elements, maintaining structural integrity and routing , while ducts connect the combustor exit to the inlet, often incorporating variable geometry in advanced designs for pattern factor control.

Materials and Construction Techniques

Combustor liners, domes, and transition pieces are predominantly fabricated from nickel-based superalloys, such as Haynes 230 (Ni-22Cr-14W-2Mo) and Hastelloy X (Ni-22Cr-1.5Co-9Mo-18Fe), which provide high-temperature strength, creep resistance, and oxidation tolerance up to 1100°C. These alloys are selected over iron- or cobalt-based alternatives for their superior performance in oxidative and sulfidizing environments typical of combustion gases. To mitigate thermal fatigue and extend service life, metallic substrates are coated with thermal barrier coatings (TBCs), typically 100-500 μm thick layers of yttria-stabilized zirconia (7-8 wt% Y₂O₃-ZrO₂) applied via air plasma spraying, which reduce surface temperatures by 100-300°C through low thermal conductivity (≈1 W/m·K) and high thermal expansion mismatch tolerance. Construction begins with wrought or cast sheets or segments formed into cylindrical liners via stamping, , or to achieve wall thicknesses of 0.5-2 mm for optimal and weight reduction. Cooling features are integrated through or electro-discharge to create holes (0.3-1 mm ) spaced at densities up to 100 holes/cm², enabling film cooling with 20-30% of air to maintain liner metal temperatures below 900°C. For annular or can-annular designs, segments are joined using at 1100-1200°C under or inert atmosphere, or transient liquid-phase with nickel-based fillers, ensuring leak-tight seams without filler dilution that could degrade high-temperature properties. Advanced techniques incorporate double-wall constructions, where inner perforated liners are spaced 1-3 mm from outer pressure-containing walls via brazed spacers, promoting convective cooling channels that handle heat fluxes exceeding 2 MW/. Transition pieces connecting combustors to turbines employ similar superalloys with helical fins or quarl rings for accommodation, often repaired in-service via plasma-sprayed refurbishment or sheet replacement welding. Emerging methods, such as of nickel superalloys like IN718, enable monolithic fabrication of complex internal cooling passages, reducing part count by 50% in prototypes tested since 2019, though certification lags due to microstructure variability concerns.

Flow Dynamics

Air and Fuel Paths

Air discharged from the enters the combustor casing at and , typically around 1005–1574°F, before flowing radially inward through perforations, holes, and slots in the combustion liner. This is partitioned into primary, secondary (or ), and dilution to enable , complete oxidation, and gas reduction for turbine compatibility. Approximately 8–30% of the total compressor air supports stoichiometric in the primary zone, with the remainder used for dilution and cooling. In the primary zone, 20–25% of the total airflow enters via swirler vanes surrounding the fuel nozzle and primary liner holes, generating intense and a recirculation vortex approximately one duct diameter downstream for effective fuel-air mixing and flame anchoring. Conditions here are fuel-rich, with an equivalence ratio exceeding 1.0, promoting rapid ignition and initial heat release while minimizing formation through localized . Secondary air, admitted through intermediate wall jets and slots, constitutes about 20–30% of the and bifurcates from primary jets to mix with products, reducing the to approximately 0.8 and facilitating oxidation to via extended residence time at elevated temperatures. This zone completes bulk before downstream cooling. Dilution air, the largest fraction (50–70%), enters via larger downstream holes in the liner, rapidly the hot gases to inlet temperatures of 1700–2900°F while achieving a of about 0.3 and 15% excess oxygen by volume for uniform profile and protection. Portions of the air also follow cooling paths, forming protective films along liner walls or convecting heat away to prevent thermal damage. Fuel delivery begins upstream in the engine fuel system, where pumps pressurize liquid or gaseous fuel and route it through manifolds to nozzles at the combustor dome or head end. Injection occurs axially or at shallow angles into the primary zone, with liquid fuels atomized into fine droplets (evaporating in ~2 milliseconds) via pressure, airblast, or hybrid mechanisms, while gaseous fuels mix directly. This non-premixed diffusion process yields partially stratified fuel-air ratios, enhanced by turbulent eddies for microscale homogeneity, though advanced systems may incorporate staging or premixing for low-emissions operation.

Combustion Zoning and Mixing

Combustion zoning in combustors divides the chamber into distinct regions—typically primary, secondary, and dilution zones—to sequentially manage fuel-air mixing, ignition, reaction completion, and gas cooling. The primary zone, receiving approximately 10-25% of air, establishes a fuel-rich environment with an equivalence ratio of 1.2-1.8 to promote stability and ignition under high-velocity inflow conditions exceeding flame speeds by factors of 10-50. This zone relies on rapid initial mixing via fuel nozzle atomization and recirculation vortices induced by swirlers or dome geometry, ensuring sufficient residence time for initiation despite turbulent flows. In the secondary zone, additional air (20-30% of total) is introduced through sidewall holes or jets to complete oxidation of unburned , transitioning the mixture toward stoichiometric or conditions while controlling peak temperatures to mitigate formation. Mixing here involves shear-layer interactions between primary zone effluents and secondary air streams, with large-scale turbulent eddies dominating control over . The dilution zone then injects the bulk remaining air (50-70%) via downstream perforations to rapidly quench temperatures to 1200-1600 K for compatibility, relying on cross-stream mixing to achieve uniformity within milliseconds. Fuel-air mixing techniques prioritize , , and to minimize spatial nonuniformities that cause hot spots or incomplete . In diffusion-flame designs, fuel is injected centrally with air staged peripherally, fostering gradient-driven mixing, whereas lean-premixed systems employ upstream fuel-air blending in dedicated channels—primary for bulk fuel, secondary and tertiary for fine-tuning—to achieve homogeneity before ignition, reducing by operating below stoichiometric temperatures. Wall-cooling jets and quenches in rich-quench- (RQL) architectures further enhance mixing by entraining primary products into regimes, with air injection rates calibrated to ratios of 0.6-0.8 post-quench for thermal management. Empirical data from sector rig tests confirm that optimized hole geometries and swirl numbers yield mixing efficiencies exceeding 95% uniformity at exit, correlating with reduced pressure oscillations and emissions.

Combustor Configurations

Can-Type Combustors

Can-type combustors consist of multiple independent cylindrical chambers, typically numbering 6 to 18, arranged circumferentially within the casing and connected to the discharge and inlet via individual ducts. Each chamber, or "can," features a perforated liner that admits dilution and cooling air while containing the , a for atomized , and often a dedicated igniter for startup. Primary air enters through swirl vanes or holes near the dome end, mixing with to establish a stable zone, followed by dilution air injection downstream to moderate exit temperatures. Cross-fire tubes interconnect adjacent cans to propagate ignition sequentially during start. This configuration originated in early axial-flow designs, such as those developed during , where simplicity in fabrication and testing proved advantageous for . Individual cans can be bench-tested in isolation, facilitating iterative improvements in flame stability and efficiency without full engine disassembly. Pressure losses remain higher, often 5-7% of inlet total pressure, due to flow disruptions in the diffusers and cross-fire tubes, compared to integrated designs. However, the modular setup allows for straightforward , as a faulty can can be replaced without disturbing others, reducing downtime in industrial applications. Key operational challenges include achieving uniform circumferential temperature profiles at the turbine inlet, as variations between cans can arise from uneven air distribution or manufacturing tolerances, potentially leading to hot spots and accelerated turbine blade wear. Empirical data from simulated-altitude tests indicate that can-type combustors exhibit consistent performance across a range of inlet conditions but suffer from greater sensitivity to fuel-air ratio fluctuations than annular variants. Despite these drawbacks, their robustness in handling high combustion pressures—up to 20-30 bar in modern derivatives—sustains use in select stationary gas turbines, where reliability trumps compactness. Transition ducts from each can to the turbine must accommodate thermal expansion, often incorporating flexible seals to mitigate vibration-induced fatigue.

Can-Annular Combustors

Can-annular combustors feature multiple discrete cylindrical combustion chambers, or cans, arranged in a circular within a shared annular casing that connects to the discharge and inlet. Each can operates as an independent combustion zone with its own inner and outer liners, fuel nozzles, and igniters, while the outer casing distributes uniformly to all cans via interconnecting tubes or cross-fire tubes for flame propagation. This hybrid design evolved from early can-type systems to accommodate the annular airflow patterns of larger axial in mid-20th-century gas turbines. The configuration allows for easier individual testing and development of cans compared to fully annular combustors, where the entire ring must be evaluated as a , reducing experimental complexity and costs during engine maturation. Modular also enhances , as faulty cans can be replaced without disassembling the full assembly, a key advantage in high-thrust applications. However, the design incurs higher drops from air routing between cans and increased wetted surface area, potentially leading to greater heat losses and less uniform exit temperature profiles than seamless annular types. Notable implementations include the low-bypass engine, certified in 1963, which employed a nine-can annular to support ratings up to 17,000 lbf for commercial airliners like the 727. General Electric's early jet engines, such as the J47 from the 1940s, initially used separate cans before transitioning toward annular variants, highlighting can-annular as a transitional for scaling power output in and industrial gas turbines. These systems achieved combustion efficiencies around 99% under cruise conditions but faced challenges with durability in high-temperature environments, prompting material advancements like ceramic coatings by the 1970s. Despite their prevalence in engines from the to , can-annular combustors have largely been supplanted in modern designs by single annular combustors (SACs) for improved , reduced weight, and lower emissions, though legacy fleets continue to rely on them for proven reliability in retrofit and overhaul contexts. Empirical data from operational fleets indicate pattern factors (temperature nonuniformity) typically between 12-15%, balancing stability against the efficiency gains of annular alternatives.

Annular Combustors

Annular combustors feature a single, continuous ring-shaped formed by concentric inner and outer liners within an annular casing, enabling fuel-air mixing and across the entire annulus. This integrates multiple fuel nozzles around the , promoting without discrete cans. Design emphasizes stable through precise nozzle and baffle orientations to achieve even mixing and holding, often with swirl injectors for enhanced airflow patterns. Compared to can-annular designs, which arrange separate cylindrical liners in a ring for modular testing and servicing, annular combustors eliminate inter-can linkages, reducing weight and length by approximately 25% relative to tubo-annular variants while minimizing surface area for heat loss. Advantages include lower , higher , and improved exhaust temperature uniformity, which benefits downstream durability. However, development challenges arise from the complexity of ensuring ignition reliability and stability across the full annulus, as malfunctions can propagate circumferentially more readily than in isolated cans. Annular combustors evolved from early axial-flow engine designs as a compact alternative to can-types, gaining prevalence in modern aero-engines for their reduced axial footprint and enhanced performance. Notable implementations include the CFM International CFM56 series, which employs a single annular combustor for efficient operation in high-bypass turbofans. Advanced variants, such as double annular combustors (DAC) introduced on the CFM56-5B in 1995, incorporate dual concentric rings to optimize emissions and lean-burn stability, achieving service entry with airlines like Swissair. These designs support higher overall pressure ratios in contemporary engines by facilitating better airflow distribution and reduced cooling air requirements.

Advanced Variants

Advanced combustors incorporate staged combustion, premixing, and optimized zoning to minimize () emissions while enhancing efficiency and stability, addressing limitations of traditional diffusion-flame designs. These variants emerged in response to stricter environmental regulations, with development accelerating in the 1980s for stationary applications and the 1990s for aero-engines. Rich-Quench-Lean (RQL) combustors operate via a fuel-rich primary for stable ignition, followed by rapid air injection in a quench to dilute and cool the mixture, and a final to complete combustion at lower temperatures, achieving reductions of up to 50% compared to conventional designs. Introduced in 1980, RQL systems were tested in full-scale F-class turbines with combustor exit temperatures of 2550°F (1399°C), demonstrating viability for low-heating-value fuels. evaluations in confirmed RQL performance at elevated pressures up to 30 atm and inlet temperatures of 600 K, though challenges include soot formation in the rich and mixing uniformity in the quench. Lean premixed (LPM) combustors, often termed Dry Low (DLN) in commercial systems, premix fuel and air upstream to maintain ratios below 0.6, suppressing thermal formation by avoiding local hot spots; ratios are precisely controlled via swirlers and baffles for uniform mixing. Deployed widely in gas turbines since the , GE's DLN 2.6 upgrade for F-class units enables single-digit emissions (under 10 ) without diluents, supporting fuel flexibility including blends up to 100% in validation tests completed by 2025. Operability issues, such as combustion instability from acoustic-heat release , are mitigated through tuned premixers, with limits predicted via models showing up to 20 Hz oscillations. In aero-engines, GE's Twin Annular Premixing Swirler (TAPS) represents a dual-zone evolution, featuring inner and outer premixers that stage fuel delivery for takeoff (richer inner) and cruise (leaner outer), reducing NOx by over 50% versus single-annular predecessors while meeting CAEP/8 standards. First entering service on the GEnx engine in 2011 for Boeing 787 and 747-8 aircraft, TAPS III variants in the GE9X achieve premixing of over 70% of compressor air, with sector tests confirming emissions below 20 g/kN at 3000 K exit temperatures. Lean Premixed Prevaporized (LPP) concepts, researched by NASA since 1977, extend this by vaporizing fuel prior to mixing, further lowering emissions but requiring advanced injectors to prevent autoignition at high pressures. These designs trade increased complexity for durability, with pattern factors under 0.2 ensuring even turbine inlet profiles.

Performance and Reliability

Efficiency and Stability Factors

in combustors refers to the fraction of converted to in the exhaust gases, typically exceeding 99% across operating conditions in modern designs. This high arises from optimized air- mixing, sufficient for completion, and favorable thermodynamic conditions, with primary occurring in fuel-rich zones where and carbon convert to and H₂O, followed by secondary oxidation of to CO₂. Key factors influencing include combustor pressure, which inversely affects such that lower pressures (e.g., below 7 lb/sq in. abs) reduce due to diminished rates; air , where decreases (e.g., from 620 °R to 500 °R) impair and mixing; and reference velocity, with higher velocities (e.g., above 105 ft/s) shortening and lowering . Air- ratio also plays a critical role, peaking at optimal ratios (around 0.014) before declining in or extremes due to incomplete . Stability in combustors encompasses the maintenance of continuous without (blowout) or destructive oscillations, essential for reliable across varying loads and conditions. Lean blowout (LBO), a primary limit, occurs when the drops too low (typically φ ≈ 0.3-0.8 in dilution zones), failing to sustain anchoring, influenced by factors such as insufficient -air mixing, reduced contact time relative to and ignition delay, and diminished swirl strength (requiring swirl numbers >0.6 for recirculation zones). Higher velocities exacerbate LBO by straining stabilization, while and reductions at altitude or off-design points narrow the operable range. instabilities, manifesting as oscillations from unsteady heat release, stem from acoustic between dynamics and combustor acoustics, often mitigated by primary zone -rich conditions (φ >1.0) to anchor flames via aerodynamic recirculation. Overall indirectly impacts by altering conditions, with higher ratios demanding enhanced mixing to prevent dynamic instabilities.

Thermal Management and Durability

Thermal management in gas turbine combustors is essential due to combustion temperatures routinely exceeding 1800–2200 K, which surpass the melting points of structural materials by factors of 1.5 or more, necessitating active cooling to limit metal surface temperatures to 900–1200°C for operational longevity. Primary heat transfer modes include radiation (dominant internally, accounting for up to 60% of heat load), convection, and conduction, with cooling air comprising 20–30% of compressor discharge to mitigate these effects without excessively penalizing engine efficiency. Key cooling techniques encompass film cooling, which injects compressor bleed air through slots or holes to create a protective gaseous barrier reducing hot gas contact; effusion cooling, utilizing dense arrays of small-diameter perforations (e.g., 0.5–1 mm) for uniform transpiration-like coverage and enhanced durability under high-pressure ratios; and convection cooling via internal passages or fins to convect heat away from walls. Impingement cooling targets high-heat-flux zones like diluter holes, while advanced variants integrate these with backside convection for overall effectiveness factors exceeding 1.5 in modern designs. These methods must balance cooling air usage against performance losses, as excessive bleed reduces turbine inlet temperature potential by 50–100 K per percent increase. Durability challenges arise from prolonged exposure to thermo-mechanical stresses, including creep deformation under sustained loads above 800°C, where nickel-based superalloys exhibit rupture lives dropping by orders of magnitude per 50 K increment; oxidation forming adherent scales like Al2O3 or Cr2O3 that, if spalled, accelerate degradation; and thermal fatigue from cyclic operations inducing cracking at cooling hole edges. Dirt ingestion exacerbates erosion and insulating deposits, reducing cooling efficiency by 20–50% and halving liner life in dusty environments. Manufacturing variability in hole geometry or wall thickness can amplify local hot spots, with probabilistic models showing 10–20% scatter in predicted life. Enhancements for include thermal barrier coatings (TBCs) of (8–20 μm thick) providing 100–200 K and oxidation up to 1200°C, alongside single-crystal superalloys like CMSX-4 offering superior via gamma-prime precipitates stable to 1100°C. Liner segmentation and advanced designs mitigate stress concentrations, achieving 20,000–30,000-hour lives in aero engines, though empirical data from field inspections reveal oxidation-driven failures as primary in 40% of cases under hydrogen-rich fuels. Experimental validation via and gauges confirms these strategies extend , but trade-offs persist, as intensified cooling correlates with higher via incomplete mixing.

Emissions Analysis

Emission Mechanisms and Species

In gas turbine combustors, pollutant emissions primarily consist of nitrogen oxides (), carbon monoxide (), unburned hydrocarbons (UHC), and (PM), including , generated through incomplete combustion, high-temperature reactions, and localized stoichiometry variations. These species form due to the interplay of , profiles exceeding 2000 K in primary zones, and residence times on the order of milliseconds, with NOx dominating at high power settings and CO/UHC peaking at low loads or startup. arises mainly in fuel-rich pockets, contributing to radiative but also visible smoke emissions. NOx formation encompasses three primary pathways: thermal, prompt, and fuel-bound. Thermal NOx, the dominant mechanism in lean, high-temperature , proceeds via the Zeldovich mechanism, where atmospheric N2 dissociates at temperatures above approximately 1800 K and reacts with O atoms to form NO, with rates exponentially dependent on temperature (Arrhenius form, ~380 kJ/mol for key steps). Prompt NOx occurs in fuel-rich fronts, where radicals (e.g., , C2H) attack N2 to produce cyanogen intermediates like , rapidly converting to NO; this pathway contributes significantly under low-temperature, rich conditions or in turbulent diffusion flames, accounting for up to 20-50% of total in some regimes. Fuel NOx stems from organic nitrogen in the fuel (typically <0.1% in jet fuels), which oxidizes to NO during pyrolysis and combustion, though its contribution is minor in low-nitrogen fuels like natural gas or kerosene. CO emissions result from kinetically limited oxidation of CO to CO2, prevalent in near-stoichiometric or lean-quench zones where flame temperatures drop below 1500-1700 K or residence times are insufficient for complete burnout (reaction rate constants indicate half-life >1 ms at 1200 K). High CO levels, often exceeding 100 at , arise from poor fuel-air mixing, wall , or in dilute regions, with concentrations inversely scaling with combustor pressure and load. UHC emissions, including alkanes, alkenes, and aromatics, form via at cold walls (boundary layers <1 mm thick), incomplete vaporization of fuel droplets, or local in lean premixed flows, leading to slip hydrocarbons that evade post-flame oxidation; these peak during acceleration transients, comprising <10 at cruise but up to 1000 at startup. Soot particles, primarily elemental carbon aggregates (20-100 nm diameter), nucleate in fuel-rich pyrolysis zones (equivalence ratio >1.5-2.0) through (C2H2) , surface growth, and , with formation rates peaking at 1500-1800 K before oxidation by radicals in later stages. Suppression occurs via overall lean operation or staged air addition, reducing peak soot yields by factors of 10-100, though residual emissions include sulfates and metals from fuel impurities. These mechanisms are validated through detailed chemical kinetic models (e.g., GRI-Mech 3.0 for gas-phase) coupled with CFD simulations of combustor flowfields.

Reduction Strategies and Technologies

Reduction strategies for combustor emissions primarily target nitrogen oxides (), the dominant pollutant formed via , prompt, and fuel-bound mechanisms, by mitigating peak flame temperatures exceeding 1700 , optimizing equivalence ratios, and minimizing residence times in high-temperature zones. Combustion modifications, categorized as wet and dry techniques, alter the reaction environment directly within the combustor, while post-combustion methods like (SCR) treat exhaust gases downstream. These approaches achieve NOx levels as low as 2-25 ppmvd at 15% , depending on the system, though trade-offs include increased (CO) at lean conditions and operational stability challenges. Wet low-NOx technologies inject or into the combustor to act as a , reducing and thermal formation rates, which follow an exponential dependence on . In combustors, water-to-fuel ratios of 1.0 can reduce to 42 ppmvd at 15% , representing a 90% drop from uncontrolled baselines around 160-200 ppmvd on . injection similarly attains 25-65 ppmvd, but both methods incur efficiency penalties of 2-5% due to evaporative cooling and increased mass flow, alongside higher emissions from incomplete ; they remain viable for retrofits on smaller turbines like GE's MS7001E series. Dry low-NOx (DLN) systems, predominant in modern designs, premix fuel and air upstream to achieve lean equivalence ratios (φ ≈ 0.5), distributing heat release and suppressing peak temperatures without diluents, thereby avoiding efficiency losses. Lean premixed (LPM) variants stage for flame stability, yielding 9-25 ppmvd , as in GE's DLN upgrades for F-class turbines. Rich-quench-lean (RQL) configurations, introduced in 1980, operate a fuel-rich primary zone (φ >1) to convert fuel-bound to N2 rather than , followed by rapid air to prevent and a lean secondary zone for completion; this reduces fuel by up to 95% from sources like in , limiting total emissions to ~50 ppmvd even with 1000 ppm fuel . Lean direct injection (LDI) further minimizes mixing times, achieving ~1 ppmv at combustor exit temperatures of 600-1000°F and pressures up to 13.6 atm. Post-combustion SCR employs or over vanadium-titanium catalysts at 600-750°F to decompose to N2 and , attaining 2 vd with 2-5 ppm ammonia slip when integrated with DLN. This secondary control adds a 4 in.w.c. but enables ultra-low emissions in combined-cycle plants, often as lowest achievable emission rate (LAER) technology. Emerging variants like micromix or trapped vortex combustors enhance DLN stability for blends, but all dry methods risk combustion dynamics and spikes below 50% load, necessitating advanced monitoring and tuning.

Trade-offs, Regulations, and Empirical Outcomes

Emission reduction strategies in combustors, particularly dry low- (DLN) technologies employing lean premixed , necessitate trade-offs with operational performance. Lowering formation through reduced flame temperatures and enhanced air-fuel premixing decreases thermal but risks instability, elevated emissions at low loads, and potential penalties from incomplete mixing or extinction events. These designs often exhibit higher pressure drops and sensitivity to fuel composition variations, complicating fuel flexibility while aiming to balance suppression against power output and . In combustors, via or rich-lean-quench architectures can marginally increase specific fuel consumption, as optimized emission control diverts design priorities from pure thermodynamic . Regulatory standards enforce these trade-offs by mandating emission thresholds that drive technological adoption. For stationary gas turbines in the United States, the EPA's New Source Performance Standards (NSPS) under 40 CFR Part 60, Subpart KKKK, establish NOx limits varying by turbine size, fuel, and operation; for instance, simple-cycle units greater than 250 MMBtu/h heat input face corrected NOx concentrations of 15-42 ppmvd at 15% O2 depending on subcategory, with a November 2024 proposal to tighten limits for new, modified, and reconstructed fossil fuel-fired turbines using dry ultra-low-NOx burners as best system of emission reduction. Internationally and for aviation, ICAO's Committee on Aviation Environmental Protection (CAEP) sets engine certification standards in Annex 16, Volume II, regulating NOx (e.g., via correlation equations scaling with pressure and temperature), CO, unburned hydrocarbons, smoke, and non-volatile particulate matter for turbofan and turbojet engines above 26.7 kN thrust, with CAEP/10 adopting further stringency effective 2020 onward. Field deployments of advanced low- combustors yield empirical reductions aligning with regulatory goals, though outcomes vary by fuel and conditions. DLN retrofits on mature industrial gas turbines have demonstrated levels of 5-15 ppm on at , compared to over 100 ppm in legacy diffusion-flame systems, enabling without wet controls but requiring operational tuning to mitigate spikes. In , certified engines per CAEP standards exhibit margins 20-50% below limits in the ICAO emissions databank, correlating with fleet-wide reductions; for example, post-CAEP/6 implementations cut average by up to 15% relative to prior cycles. EPA analyses project that proposed NSPS tightening will avert 198 tons of in 2027 rising to 2,659 tons by 2032 from sources, yielding up to $340 million in net benefits from and environmental improvements, predicated on achievable DLN technologies. These outcomes underscore causal links between design innovations and emission declines, tempered by site-specific factors like ambient conditions and maintenance.

Applications

Aero-Engine Integration

In aero-engines, the combustor is positioned immediately downstream of the high-pressure and upstream of the high-pressure , forming a critical in the core flow path of , , and configurations. Compressed air enters the combustor at elevated temperatures typically between 400°C and 600°C and pressures ranging from 20 to 40 , depending on the engine's overall and . This preheated, high-pressure airflow, which constitutes the majority of mass flow, must be precisely managed to ensure efficient fuel-air mixing and while minimizing total losses, which are generally limited to 3-7% to preserve cycle . Fuel injection systems, often employing airblast atomizers or pressure-swirl nozzles, integrate directly into the combustor dome to achieve fine droplet sizes for rapid and mixing with the incoming swirl. Approximately 20-30% of the discharge is diverted for liner cooling and dilution to protect structural components from the temperatures exceeding 1700°C, while the remainder supports primary zones for stabilization. The combustor design must deliver a uniform exhaust profile to the inlet to avoid thermal hotspots that could reduce life, with modern annular or can-annular configurations favored for their compact radial dimensions and improved uniformity in high-bypass turbofans like the . Aerodynamic coupling with the upstream exit diffuser and downstream nozzle guide vanes demands careful flow matching to mitigate instabilities such as or over-temperature. Integration challenges include thermoacoustic instabilities arising from interactions between combustion dynamics and acoustic modes in the combustor-turbine annulus, exacerbated by operations for emissions reduction. These require advanced features like acoustic liners or variable geometry elements, though the latter remain limited to experimental engines due to mechanical complexity and weight penalties. Empirical testing in sector rigs and full annular setups verifies integration performance under varying flight conditions, from ground idle to at altitudes above 10 km, ensuring stable operation across the engine's range. Material selections, such as nickel-based superalloys for liners, must withstand cyclic thermal loads while maintaining seals at compressor-combustor and combustor-turbine interfaces to prevent hot gas leakage.

Stationary Gas Turbines

Combustors in stationary gas turbines, used primarily for generation and mechanical drive applications, are engineered for continuous, high-load under stable inlet conditions, prioritizing long-term durability, , and reduced pollutant emissions over the transient performance demands of aero-engines. These systems typically operate at discharge pressures of 10-30 bar and temperatures exceeding 1,200°C, converting and —predominantly —into hot gases to drive the . Heavy-frame designs, common in base-load plants, feature lower pressure ratios (around 15-20:1) compared to aeroderivative units, allowing for robust combustor geometries that accommodate higher flows and extended intervals exceeding 25,000 hours. The predominant combustor configurations in gas turbines include can-annular and annular types, with or frame-type variants in larger units where chambers are mounted externally for easier maintenance. Can-annular designs, consisting of multiple cylindrical cans arranged annularly around the turbine axis, provide for individual replacement and uniform distribution, as employed in Energy's SGT-400 and certain GE heavy-duty models. Annular combustors, by contrast, integrate a single continuous chamber within a shorter casing, reducing losses and weight while enabling compact packaging, as seen in Siemens' SGT-800 with its twin-shaft architecture. These geometries facilitate precise air-fuel mixing to achieve stable flame holding across a wide , essential for load-following in grid applications. Emission control has driven the shift to dry low-NOx (DLN) combustors since the 1990s, which employ lean premixed combustion—premixing fuel with 75% or more of the air at fuel-lean equivalence ratios (φ ≈ 0.5-0.6)—to suppress thermal NOx formation by maintaining peak flame temperatures below 1,800 K, achieving NOx levels as low as 9-25 ppm at 15% O2 without diluents like water injection. GE's DLN 2.6+ system, introduced for F-class turbines in 2015, uses multi-stage fuel injection and advanced swirlers for staged premixing, enabling operation on natural gas with up to 30% hydrogen blends while limiting CO to under 10 ppm. Siemens' distributed combustion systems similarly distribute fuel via multiple nozzles to avoid hot spots, supporting hydrogen-ready upgrades in models like the SGT-800, though challenges persist in managing combustion dynamics and flashback risks at high hydrogen contents. Thermal management in these combustors relies on film cooling, effusion cooling, and ceramic thermal barrier coatings to withstand turbine inlet temperatures up to 1,600°C, with liner lives extended through computational fluid dynamics-optimized airflow patterns that minimize hot streaks. Fuel flexibility upgrades, tested in facilities like those at NETL, allow co-firing with or biofuels, but empirical data indicate trade-offs in efficiency (1-2% drop) and increased maintenance for diffusion-flame pilots required for . Reliability metrics from field deployments show rates above 98% for DLN-equipped units, contingent on rigorous hot gas path inspections every 8,000-16,000 hours.

Specialized Propulsion Systems

In engines, the operates without mechanical compression, relying on ram compression from high-speed to achieve velocities. Flame stabilization is achieved through flameholders such as V-gutters or that create recirculation zones for continuous ignition, addressing the challenge of short fuel-air mixing times at velocities up to 3. Fuel injectors are positioned upstream to promote rapid and , with efficiencies typically exceeding 95% in operational designs due to the high temperatures and pressures generated by inlet diffusion. Scramjet combustors, in contrast, sustain in supersonic airflow ( 4-8), where the for fuel-air reaction is limited to approximately 1 , necessitating advanced mixing and ignition strategies. Cavity-based flameholders or injectors generate shock-induced recirculation to anchor flames, while fuel such as or hydrocarbons is injected transversely or at angles to enhance and for near-instantaneous . These designs prioritize minimal total loss, with experimental s demonstrating specific impulses up to 2000 seconds under hypersonic conditions, though thermal management via wall cooling remains critical to prevent material failure at temperatures exceeding 2000 K. Rocket engine combustion chambers differ fundamentally from air-breathing combustors by utilizing storable s without atmospheric oxygen, operating at chamber pressures of 50-300 bar and temperatures around 3000-3500 K. Coaxial or injectors ensure propellant mixing through impingement or shear, with via propellant circulation preventing in high-thrust designs like those in the Main Engine, which achieved mixture ratios of 6:1 oxygen to . Combustion instabilities, driven by acoustic coupling between chamber oscillations and heat release, are mitigated through baffles or acoustic absorbers, as evidenced in scaling studies showing instability thresholds scaling with chamber length and frequency. Unlike combustors, rocket chambers emphasize high handling (up to 100 MW/m²) over emissions, with efficiencies approaching 99% in steady-state operation.

Emerging Developments

Low-NOx and Sustainable Fuel Designs

Low-NOx combustor designs primarily achieve emission reductions through strategies that minimize peak flame temperatures, a primary driver of thermal formation via the Zeldovich mechanism, where and oxygen react at high temperatures above 1800 K. Dry low emissions (DLE) systems, also known as premixed combustion, introduce fuel and air in a premixed state at ratios below 0.6 to maintain conditions, suppressing by limiting local temperatures to under 1700 K while ensuring stable combustion across load ranges. These systems have demonstrated levels as low as 25 in operational gas turbines without water or steam injection, outperforming diffusion flame combustors by reducing both and through uniform fuel-air distribution. Rich-quench-lean (RQL) architectures divide combustion into a fuel-rich primary zone (equivalence ratio ~1.5-2.0) for , a rapid quench zone with excess air to drop temperatures below formation thresholds, and a lean secondary zone for burnout, converting fuel-bound nitrogen primarily to N2 rather than . Introduced in the , RQL designs achieve reductions of up to 70% relative to conventional combustors at conditions of 10-20 and 1500-1900 K, with empirical tests showing emissions below 50 in heavy-duty applications. Staged and variable geometry further enhance turndown ratios, enabling operation from 50% to full load without exceeding regulatory limits like California's 9 standard. Adaptations for sustainable fuels such as and integrate these low- principles with fuel-specific modifications to address instabilities and altered kinetics. 's high laminar (up to 2.3 m/s versus 0.4 m/s for ) necessitates advanced micromix injectors and swirl-stabilized premixers in DLE combustors to prevent flashback, achieving below 10 at 100% operation in a 65 kW recuperated tested in 2024, where preheated air inlet at 500 K improved stability and efficiency to 35%. , zero-carbon but prone to fuel- from its 82% content, employs rich-lean staging in non-premixed designs; a 2024 study modeled a staged combustor reducing to under 50 at 10 by maintaining rich-zone ratios above 1.2, converting ammonia-N to N2 before lean dilution, though unburned NH3 slip requires catalytic abatement. These designs prioritize empirical validation over simulations, revealing trade-offs like increased pressure drops (5-10% in premixers) but enabling net-zero compatibility without diluents.

Detonation-Based and Alternative Combustion

Detonation-based combustion represents a departure from conventional deflagrative processes in combustors, where subsonic flame fronts propagate through premixed reactants, by employing supersonic waves that achieve near-constant-volume combustion and inherent pressure gain across the reaction zone. This mode leverages the Chapman-Jouguet detonation theory, yielding velocities of 1,500–2,500 m/s depending on fuel-air mixtures, compared to deflagration speeds below 100 m/s, enabling higher thermodynamic efficiencies potentially 20–30% above traditional limits due to reduced generation and positive pressure rise (typically 15–25 times initial ). Empirical tests in subscale rigs have demonstrated stable detonation operation with natural gas-air mixtures, though full-scale integration into gas turbines remains challenged by wave and thermal management. Pulse detonation engines (PDEs), an early detonation-based variant, operate via cyclic filling, ignition, deflagration-to-detonation transition (), and exhaust purging, typically at 20–100 Hz frequencies, offering simplicity with fewer moving parts and scalability from to hypersonic regimes. Advantages include specific consumption reductions of up to 15% over deflagrative counterparts in theoretical models and zero-speed thrust capability, as validated in ground tests by organizations like the since the 1990s. However, challenges persist, including DDT initiation requiring 10–20% of tube length for reliable transition, high from peaks exceeding 20 atm, and levels necessitating advanced , limiting practical deployment despite prototypes achieving thrust-to-weight ratios superior to turbojets in simulations. Rotating detonation engines (RDEs) or combustors (RDCs) advance this paradigm with continuous azimuthal waves propagating at 1,000–2,000 m/s in an annular chamber, eliminating cyclic and enabling steady-state gain of 10–20% over inlet conditions, as demonstrated in U.S. Department of Energy-funded tests at NETL since 2016. Conceptualized in the by researchers like Nicholls but revitalized post-2000 with aiding wave mode prediction, RDEs have shown operational stability with hydrogen-oxygen fuels in hot-fire tests yielding specific impulses 10–15% above conventional rockets, and applications in rigs achieving efficiencies near 80%. Integration hurdles include injector choking from backpressure, heterogeneous mixing in liquid-fueled variants leading to mode extinction, and material endurance under 3,000 K peaks, though advancements like 3D-printed nozzles have enabled multi-wave modes for broader operability. Alternative combustion approaches, such as wave rotor topping , augment detonative concepts by incorporating unsteady prior to , achieving overall efficiencies up to 50% in hybrid gas turbine configurations per simplified models, though empirical validation lags due to rotor sealing and synchronization demands. These methods prioritize causal pressure recovery over steady , aligning with principles but avoiding full-wave to mitigate structural loads, with potential applications in aero-engines for 5–10% fuel savings as explored in ASME analyses. Ongoing research emphasizes hybrid - transitions for robust , underscoring -based systems' promise for next-generation amid empirical evidence of superior utilization despite unresolved scalability.

Integration with Additive Manufacturing and Digital Tools

Additive manufacturing (AM), also known as 3D printing, has enabled the production of combustor components with intricate geometries that enhance fuel-air mixing, cooling efficiency, and structural integrity, which are challenging or impossible with conventional casting or machining. For instance, in 2018, Siemens Energy successfully 3D-printed and engine-tested a dry low-emission (DLE) pre-mixer for the SGT-A05 aeroderivative gas turbine, demonstrating reduced manufacturing lead times and improved performance under operational conditions. Similarly, AM techniques like selective laser melting have been applied to fabricate burner tips and liners, allowing for optimized internal channels that minimize weight by up to 20-30% while maintaining thermal resistance. These advancements support higher turbine inlet temperatures and compatibility with low-carbon fuels, such as hydrogen blends, by enabling precise microstructures for flame stabilization. Integration of AM with combustor design often involves to create lattice structures or conformal cooling passages, reducing cooling air requirements by 10-15% in prototypes and thereby boosting overall . has developed AM processes for complex hot-section parts, including combustor swirlers, leveraging powder-bed fusion to achieve near-net-shape components with densities exceeding 99%. However, challenges persist, such as ensuring material anisotropy and fatigue resistance in high-temperature environments, necessitating post-processing like . Empirical testing confirms that AM combustor parts can endure thousands of cycles, with Siemens reporting successful field deployments in industrial turbines as of 2021. Digital tools, particularly (CFD) simulations, are integral to AM combustor development, allowing virtual iteration of designs to predict dynamics, pollutant formation, and thermoacoustic before physical prototyping. Software suites like Fluent and CFX model turbulent reacting flows in combustors, incorporating large eddy simulations for accurate flame front resolution and emission profiles. Adjoint-based optimization within CFD frameworks has been used to redesign fuel injectors for AM, targeting minimal and uniform mixing, as demonstrated in research optimizing micro combustors for efficiencies above 30%. These tools reduce physical testing iterations by 50-70%, enabling rapid validation against empirical data from rig tests. The synergy of AM and digital tools facilitates a design-manufacture-test cycle measured in weeks rather than months, with CFD guiding AM parameter selection to mitigate defects like . For example, employs CFD-driven digital workflows to repair and upgrade combustor burners, cutting downtime from weeks to days while preserving original performance specifications. Ongoing developments include enhancements to CFD for real-time design feedback, though validation against ground-truth experiments remains essential to counter simulation uncertainties in multiphase . This integration promises scalable production of next-generation combustors tailored for sustainable and power generation.

References

  1. [1]
    Combustor - Burner
    A combustor, or burner, is where fuel combines with high-pressure air and burns, located between the compressor and power turbine, and is where combustion ...
  2. [2]
    [PDF] Conventional Type Combustion 3.2.1.1-1 Introduction
    The role of the combustor in a gas turbine engine is two-fold. First, the combustor transforms the chemical energy resident in the fuel into thermal energy.
  3. [3]
    Gas Turbine Combustors - an overview | ScienceDirect Topics
    A gas turbine combustor is defined as a component where high-pressure air from the compressor is mixed with fuel and combusted to add heat, enabling the ...
  4. [4]
    How Gas Turbine Combustion Works - GE Vernova
    The COMPRESSOR is what takes in air from outside of the turbine and increases its pressure at hundreds of MPH, then feeds it to the combustion chamber. · The ...
  5. [5]
    Combustor - Burner - NASA Glenn Research Center
    There are three main types of combustors, and all three designs are found in modern gas turbines: The burner at the left is an annular combustor with the liner ...
  6. [6]
    Combustor - an overview | ScienceDirect Topics
    A combustor is defined as a critical component that converts the chemical energy of fuel into thermal energy to drive a turbine, operating as a constant ...
  7. [7]
    [PDF] AP-42, Vol. I, 3.1: Stationary Gas Turbines - EPA
    A gas turbine is an internal combustion engine that operates with rotary rather than reciprocating motion. Gas turbines are essentially composed of three major ...
  8. [8]
    Burner Basics - Purdue College of Engineering
    The burner recieves flow from the compressor, seapartes some of the flow, mixes it with fuel and ignites it, remixes the flow, and delivers it to the turbine.
  9. [9]
    How Gas Turbine Power Plants Work - Department of Energy
    The combustion produces a high temperature, high pressure gas stream that enters and expands through the turbine section. The turbine is an intricate array of ...
  10. [10]
    [PDF] Aircraft Engines - Federal Aviation Administration
    Turbine Engine Operating Principles. The principle used by a gas turbine engine as it provides force to move an airplane is based on Newton's Third Law. This ...<|separator|>
  11. [11]
    [PDF] Introduction to Gas Turbines for Non- Engineers - ASME
    The gas turbine is an internal combustion (IC) engine employing a continuous combustion process. This differs from the intermittent combustion occurring in ...Missing: definition | Show results with:definition
  12. [12]
    [PDF] 3.2.1.4.2-1 Introduction Low Swirl Combustion Robert K. Cheng
    It is formed when the swirl intensities are deliberately low such that vortex breakdown, a precursor to the formation of flow reversal and recirculation, does.
  13. [13]
    [PDF] Aircraft Propulsion Combustion chambers and Turbines
    This types of combustion chamber is used on centrifugal compressor type engines. it has several cans disposed around the engine. Page 2. Aircraft Propulsion.
  14. [14]
    Solved The stoichiometric air/fuel ratio for combustion of | Chegg.com
    Apr 30, 2024 · The stoichiometric air/fuel ratio for combustion of kerosene is approximately 15:1.However, in aircraft gas turbine engines the actual air/fuel ratio may vary ...
  15. [15]
    Combustion in the future: The importance of chemistry - PMC
    Sep 25, 2020 · Combustion involves chemical reactions that are often highly exothermic. Combustion systems utilize the energy of chemical compounds released during this ...
  16. [16]
    How do jet engines cool the gas in the combustion zone?
    Oct 13, 2021 · The adiabatic flame temperature of kerosene in air is 2093 C (source). Modern jet engines have a turbine inlet temperature (TIT) of 1700 to 1800 ...
  17. [17]
    Turbine Engine Thermodynamic Cycle - Brayton Cycle
    The combustion process in the burner occurs at constant pressure from station 3 to station 4. The temperature increase depends on the type of fuel used and the ...Missing: fundamentals | Show results with:fundamentals
  18. [18]
    Burner Thermodynamics
    It would appear that we can make the temperature ratio and resulting thrust as large as we want by just increasing the fuel flow rate and the fuel/air ratio.
  19. [19]
    The History of Gas Turbine Engines - The Inventors
    These form the basis for modern propulsion theory. 1791 - John Barber received the first patent for a basic turbine engine. His design was planned to use as a ...
  20. [20]
    The History of the Gas Turbine - Rochem Fyrewash
    Oct 21, 2019 · It was not until 1791 however, that a patent was awarded for the first true gas turbine. John Barber, an English inventor, designed a turbine ...<|separator|>
  21. [21]
    Engines - NASA Glenn Research Center
    It was Frank Whittle, a British pilot, who designed and patented the first turbo jet engine in 1930. The Whittle engine first flew successfully in May, 1941.
  22. [22]
    The First Patent - Sir Frank Whittle - inventor of the jet engine
    Frank Whittle filed this, his first patent for a gas turbine to propel an aircraft directly by its exhaust on 16th January 1930.
  23. [23]
    Pratt & Whitney J57-P-4 Turbojet Engine | Smithsonian Institution
    Type: Turbojet. Thrust: 66,720 N (15,000 lb) at 8,200 rpm. Compressor: 9-stage high pressure and 6-stage low pressure axial. Combustor: Cannular. Turbine ...
  24. [24]
    Combustor - Burner - NASA Glenn Research Center
    May 7, 2021 · The advantage to the can-annular design is that the individual cans are more easily designed, tested, and serviced.
  25. [25]
    Pratt & Whitney J57 Turbojet - Air Force Museum
    The J57 turbojet was the first production jet engine to produce 10,000 pounds of thrust. The J57 featured a dual-rotor axial-flow compressor, which lowered fuel ...Missing: combustor type
  26. [26]
    [PDF] TAPS II Combustor Final Report - Federal Aviation Administration
    Figure 6 shows the lean burn combustion process. TAPS combustor development started in 1995 as a GE/NASA emissions reduction technology program. The TAPS ...
  27. [27]
    [PDF] N+2 Advanced Low NOx Combustor Technology Final Report
    The development of N+2 combustor technology proceeded from CFD-based conceptual design through single cup evaluation rigs and finally the design and fabrication ...
  28. [28]
    Pratt & Whitney Advances Combustor Technologies with NASA ...
    Nov 10, 2022 · Pratt & Whitney was selected by NASA to develop advanced engine technologies to reduce fuel consumption and emissions for next generation single-aisle aircraft.Missing: post- | Show results with:post-
  29. [29]
    New P&W GTF Combustor-lining Design Will Aid Middle East ...
    Oct 29, 2017 · Pratt & Whitney developed a new combustor-lining design as a result of lining-degradation issues experienced by airlines operating Airbus ...
  30. [30]
    [PDF] Rolls-Royce CLEEN II Low NOx Combustor Final Report
    Under the CLEEN II program, Rolls-Royce initiated a program to develop combustion technology capable of achieving 65% margin to ICAO CAEP/8 NOx standards and ...
  31. [31]
    Dry-Low Emission Gas Turbine Technology: Recent Trends and ...
    The table shows that the three combustion technologies, COSTAIR, NanoSTAR and DLE, share an essential feature of very low COx and NOx emission. However, only ...
  32. [32]
    Recent Innovations from Gas Turbine and HRSG OEMs
    Jun 1, 2014 · The method is similar to steam injection but adds far more water to the combustion process. MHPS has been developing the technology since 2000.
  33. [33]
    [PDF] Aircraft Combustors
    Combustors: Requirements. • Convert chemical energy (fuel) to thermal energy (T) with. – high combustion (conversion) efficiency. – low pressure losses.
  34. [34]
    Several Modern Wrought Superalloys for Gas Turbine Applications
    Jan 27, 2015 · HAYNES® 230™ alloy (Ni-22Cr-14W-2Mo) is used for combustors and transition pieces to replace an old combustor alloy, HASTELLOY® X alloy (Ni-22Cr ...
  35. [35]
    (PDF) Nickel Based Super Alloys For Gas turbine Applications
    Nickel based Waspaloy is an age hardening super alloys with excellent high temperature strength and good corrosion resistance notable to oxidation.<|separator|>
  36. [36]
    Recent developments in nickel-based superalloys for gas turbine ...
    Nov 10, 2023 · Nickel-based superalloys are used in gas turbines due to their mechanical properties at high temperatures.
  37. [37]
    Thermal Barrier Coatings (TBCs) And Its Role | Oerlikon Metco
    Thermal Barrier Coatings, commonly referred to as TBCs, are advanced protective layers applied onto the critical components of gas turbine engines.
  38. [38]
    [PDF] History of Thermal Barrier Coatings for Gas Turbine Engines
    “The durability of thermal barrier coatings is limited by degradation of adhesion by environmental interactions rather than by mechanical stress per se.” P.A. ...
  39. [39]
    Gas Turbine Thermal Barrier Coating in Combustion Liners
    This two layer system, applied using the plasma spray process, protects the metal of the combustion liner from hot gas corrosion, erosion and thermal cycle ...
  40. [40]
    US6434821B1 - Method of making a combustion chamber liner
    A method of fabricating an annular liner for a combustion chamber of a gas turbine engine. The combustion chamber has a dome including a fuel nozzle for ...
  41. [41]
    US20080172876A1 - Methods for repairing combustor liners
    This may include forming the sheet of material into a cylindrical, semi-cylindrical, spherical, semi-spherical, or any required shape. Openings are formed into ...Missing: construction | Show results with:construction
  42. [42]
    Combustor Wall Cooling Concepts for Dirt Mitigation - Ascent
    As double-wall cooling designs for combustors continue to evolve, it is important to assess the likelihood of dirt deposition. Modern gas turbine engines ...
  43. [43]
    Best practice in combustor transition piece design - Gas Turbine World
    Free delivery over $1,000 30-day returnsNov 7, 2018 · Best practices when it comes to TP design – focusing on the selection of a cooling concept, optimal support design, and stringent airflow ...
  44. [44]
    Novel Nickel alloy combustion chamber - Airborne Engineering
    Nov 19, 2019 · The project aims to demonstrate that OxMet's new Nickel based superalloy, ABD®-900AM, has better performance than industry incumbent IN718 ...
  45. [45]
    [PDF] Presented at the - NASA Technical Reports Server (NTRS)
    In the secondary zone, the remaining air is mixed with the primary zone combustion products to complete combustion of the fuel, and cool the gas stream to the ...
  46. [46]
    [PDF] Gas Turbine Technology Lecture 05 - ACS College of Engineering
    Air flows in around the fuel nozzle and through the first row of combustion air holes in the liner. •. Fuel is introduced in the front of the combustor by fuel ...<|separator|>
  47. [47]
    [PDF] Application of Mixing-Controlled Combustion Models to Gas Turbine ...
    Combustion takes place in the rich and lean zones at temperatures well below the stoichiometric flame temperature, so NO, production is minimized. The lean, ...
  48. [48]
    Turbulent mixing and NOx formation in gas turbine combustors
    An order of magnitude analysis is presented which shows that large-scale mixing in gas turbine combustors controls the stoichiometry of the combustion ...<|separator|>
  49. [49]
    [PDF] Combustion, Fuels and Emissions for Industrial Gas Turbines
    A lean pre-mix combustor design comprises 4 main features: • Fuel / air injection device. • Stability device. • Pre-mixing zone. • Flame stabilization zone.
  50. [50]
    Industrial Trent Combustor—Combustion Noise Characteristics
    The combustor consists of three premixing channels, which are respectively referred to as the primary, secondary, and tertiary premixers. The primary premix ...
  51. [51]
    [PDF] Mix, Lean Burn (RQL) Combustor 3.2.1.3-1 Introduction
    The RQL concept is predicated on the premise that the primary zone of a gas turbine combustor operates most effectively with rich mixture ratios. (Figure 1).
  52. [52]
    [PDF] Spectral Flame Radiance From a Tubular-Can - Combustor
    The primary-zone, secondary-zone, and tertiary-zone optical ports were located 6, 21, and 33 centimeters, respectively, downstream of the fuel nozzle.
  53. [53]
    Classification of Combustion Chamber - Propulsion 1
    Can type combustors were most widely used in early gas turbine engines ... Cannular type combustor can be called as can annular type combustor. Like the ...
  54. [54]
    [PDF] RESEARCH MEMORANDUM
    A comparison of tbe simulated-altitude performance of United. States annular- and can-type combustors and a German can-type com- bustor with inlet-air ...
  55. [55]
    Annular Combustor - an overview | ScienceDirect Topics
    In the annular combustor an annular flame tube is placed within the cylindrical liner or casing. The annular combustor has a lower pressure loss and is more ...
  56. [56]
  57. [57]
    Why would a jet engine use cannular combustors if the annular ...
    May 19, 2016 · Modern jet engines do not use can or cannular combustors. GE made the switch to annular combustors by at least the J85, which had a first engine to test in the ...What is the difference between single and dual annular combustor?How is combustion flame maintained in the combustion chamber ...More results from aviation.stackexchange.com
  58. [58]
    Annular Combustion Chamber - an overview | ScienceDirect Topics
    It is a direct development of the early type of Whittle combustion chamber. The major difference is that the Whittle chamber had a reverse flow as illustrated ...
  59. [59]
    CFM'S Advanced Double Annular Combustor Technology
    Sep 6, 1998 · CFM56 Double Annular Combustor (DAC) development was initiated in 1989 in response to growing airline concerns over future planned reductions ...Missing: invention | Show results with:invention
  60. [60]
    Understanding Gas Turbine Combustors: Types and Key Features
    It creates a toroidal flow reversal, that entrains and recirculates a part of hot combustion product to provide a continuous source of ignition to the incoming ...
  61. [61]
    [PDF] Lean Premixed/Prevaporized Combustion
    Lean premixed/prevaporized combustion involves carefully controlling fuel vaporization, fuel-air mixing, and fuel/air proportions in the combustion process.
  62. [62]
    The GE Rich-Quench-Lean Gas Turbine Combustor
    A full-scale, F-class (2550°F combustor exit temperature), rich-quench-lean (RQL) gas turbine combustor, designated RQL2, for low heating value (LHV) fuel.
  63. [63]
    [PDF] Performance of a Model Rich Burn-Quick Mix-Lean Burn Combustor ...
    The two main goals of this research are to characterize the performance of the RQL combustor at elevated inlet temperatures and pressures and to. NASA/CR. 2002- ...
  64. [64]
    Analysis of soot formation in a lab-scale Rich-Quench-Lean ...
    RQL combustors are characterized by an initial fuel-rich zone, followed by quick mixing (facilitated through air dilution jets) leading to a subsequent lean ...
  65. [65]
    Modeling of lean premixed combustion in stationary gas turbines
    Abstract. Lean premixed combustion (LPC) of natural gas is of considerable interest in land-based gas turbines for power generation.
  66. [66]
    [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 ...
  67. [67]
    Dry Low NOx (DLN 2.6) Combustion Upgrade - GE Vernova
    Dry Low NOx combustor upgrades. Our Dry Low NOx (DLN) upgrades enhance your plant flexibility and help your F-class turbine do more. Download fact sheet PDF.
  68. [68]
    A Mechanism of Combustion Instability in Lean Premixed Gas ...
    Combustion instability in lean premixed combustors is likely due to a feedback process involving heat release, acoustic pressure, and ϕ oscillations.Introduction · Mechanism of LP Combustion... · Comparisons with... · Final Remarks
  69. [69]
    The Lean Blowout Prediction Techniques in Lean Premixed Gas ...
    The lean premixed (LPM) combustion has significantly reduced NOx emission while keeping high efficiency in power production. On the contrary, lean premixed ...
  70. [70]
    GE9X's New TAPS Combustor to Maintain Its Cool Under Fire
    Nov 20, 2014 · "Since 2011, GE engineers have been working the GE9X TAPS III combustor. The TAPS III combustor will feature fuel nozzle tips manufactured using ...Missing: date | Show results with:date
  71. [71]
    Development of the GE Aviation Low Emissions TAPS ... - AIAA ARC
    Nov 6, 2012 · Development of the GE Aviation Low Emissions TAPS Combustor for Next Generation Aircraft Engines ... GE TAPS Injector Configuration · Yolanda ...
  72. [72]
    [PDF] CLEEN Phase II GE TAPS - Federal Aviation Administration
    September 30, 2020. Publication No. XXXXX. Deliverable In Response to Government OTA No. DTFAWA-15-A-80013. Levent Ileri, Program Manager.
  73. [73]
    [PDF] Combustion Efficiency
    Gas Turbine Combustion Short Course. Dr Vishal Sethi. Centre for Propulsion ... • Typical values: • > 99% for all operating conditions. • 75% - 80% for ...
  74. [74]
    None
    Summary of each segment:
  75. [75]
  76. [76]
    An FV-EE model to predict lean blowout limits for gas turbine ...
    The critical condition for LBO is that the contact time is equal to the sum of the fuel evaporation time and the ignition delay time. The model has been ...
  77. [77]
    Effusion-Cooling Performance at Gas Turbine Combustor ...
    Jul 9, 2013 · The thermal management of aero gas turbine engine combustion systems commonly employs effusion-cooling in combination with various cold-side ...
  78. [78]
    Combustor technology of high temperature rise for aero engine
    Jul 1, 2023 · Another technical approach is to pursue the use of materials with greater resistance to higher temperatures, more advanced cooling design ...
  79. [79]
    [PDF] Combustor Heat Transfer and Cooling
    Combustor heat transfer includes internal and external radiation, internal and external convection, and film cooling. Internal radiation is the largest heat ...
  80. [80]
    Advanced liner-cooling techniques for gas turbine combustors
    A basic reverse-flow combustor geometry was being maintained while different advanced liner wall cooling techniques were investigated.Missing: durability | Show results with:durability
  81. [81]
    Gas Turbine Combustor Effusion Cooling Heat Transfer
    The thermal characteristics are critical in assessing the cooling performance of effusion cooling designs and validating CFD models of combustor gas dynamics.
  82. [82]
    Evaluation of Flow and Heat Transfer Inside Lean Pre-Mixed ...
    This will support the development of more effective cooling schemes to maintain and improve combustor durability, leading to reduced maintenance costs for ...Missing: techniques | Show results with:techniques
  83. [83]
    High-Temperature Alloys: The Backbone of Extreme Environment ...
    Sep 18, 2025 · Their outstanding creep, fatigue, and corrosion resistance at temperatures up to 1200°C (2192°F) makes them the undisputed choice for critical ...
  84. [84]
    Segmented approach with advanced cooling techniques
    Jul 1, 1981 · Durability characteristics of current combustor liners severely limit liner life requirements of advanced gas turbine engines.
  85. [85]
    Impact of Manufacturing Variability on Combustor Liner Durability
    This paper presents a probability-based systems-level approach for assessing the impact of manufacturing variability on combustor liner durability.
  86. [86]
    High-temperature Alloys - Haynes International
    This new alloy has excellent creep strength in the temperature range of 1200 to 1700°F (649° to 927°C), surpassing that of Waspaloy alloy and approaching that ...<|control11|><|separator|>
  87. [87]
    Hot Section Materials and Durability | Reacting Flow Dynamics ...
    Increasing temperature can also result in durability issues for the combustor hardware, including surface burning, thermal stresses, and accelerated component ...
  88. [88]
    Thermal Crystal Temperature Characterization in a High ...
    Experimental research to study combustor liner thermal management requires temperature and heat flux measurements at relevant operating conditions, which is ...
  89. [89]
    [PDF] GER-4211 - Gas Turbine Emissions and Control - GE Vernova
    The major species (CO2, N2,. H2O, and O2) are present in percent concen- trations. The minor species (or pollutants) such as CO, UHC, NOx, SOx, and particulates.
  90. [90]
    [PDF] Methodology for the Numerical Prediction of Pollutant Formation in ...
    The combustion inside gas turbine burners generates emissions of the greenhouse gas CO2, as well as pollutants such as NOx, CO, soot particles and UHC ...<|separator|>
  91. [91]
    Soot Particle Emissions: Formation and Suppression Mechanisms in ...
    Any combustion equipment fueled with hydrocarbons generates not only gaseous emissions (SOx; NOx, CO; UHC; PAH), but also variable quantities of soot particles ...
  92. [92]
    [PDF] 3.2-1 Introduction 3.2-2 NOx Formation Combustion Strategies for ...
    Thermal NOx is formed by oxidation of nitrogen in air and requires sufficient temperature and time to produce NOx. A rule of thumb is that below approximately ...
  93. [93]
    What are the mechanisms for Prompt NOx formation during ... - IFRF
    Jan 20, 2003 · Prompt NO, also called the [GLOSS]Fenimore NO[/GLOSS], is produced rapidly in the flame front, before there would be time to form NO by the thermal mechanism.
  94. [94]
    Prompt NOx, fuel NOx and thermal NOx: The DLE strategy
    Feb 24, 2020 · The primary mechanisms of combustion-produced NOx include Thermal NOx, Prompt NOx, and Fuel NOx. The reactions are highly temperature dependent ...
  95. [95]
    [PDF] Gas Turbine Combustor - NASA Technical Reports Server (NTRS)
    carbon monoxide and unburned hydrocarbons. The carbon monoxide emissions were modelled using finite rate chemical kinetics in a plug flow scheme. The combustor ...
  96. [96]
    A Model for the Prediction of Thermal, Prompt, and Fuel NOx ...
    A model has been developed for the prediction of NOx emissions from combustion turbines. Thermal, prompt, and fuel NO are all treated and are all assumed.
  97. [97]
    [PDF] Combustion Turbine NOx Control Technology Memo - EPA
    Jan 1, 2022 · DLN and ULN technologies are applicable to both combined cycle and simple cycle units, and are a developing technology that is being included in ...
  98. [98]
    Combustion performance of a low NOx gas turbine combustor using ...
    Mar 15, 2021 · However, it can easily lead to combustion efficiency losses, and the trade-off between emission reduction benefit and performance penalty must ...
  99. [99]
    Effect of Natural Gas Composition on Low NOx Burners Operation in ...
    Nov 5, 2019 · Modifications to the burner geometry and fuel injection optimization have shown to be able to reach a good trade-off while keeping low NOx ...
  100. [100]
    On the trade-off between aviation NOx and energy efficiency
    Jan 1, 2017 · Zhang et al. [16] introduced a novel double-vortex combustor for gas turbine engines burning kerosene and lower emissions of CO, NOx and UHC ...
  101. [101]
    Trading off Aircraft Fuel Burn and NOx Emissions for Optimal ...
    Feb 8, 2018 · Conversely, combustor modifications to reduce NOx may increase CO2. Hence, a technology trade-off exists, which also translates to a trade-off ...
  102. [102]
    Stationary Gas and Combustion Turbines: New Source Performance ...
    On November 22, 2024, the EPA proposed to strengthen limits on emissions of nitrogen oxides (NOx) from most new, modified, and reconstructed fossil fuel-fired ...
  103. [103]
    Review of New Source Performance Standards for Stationary ...
    Dec 13, 2024 · Specifically, in subpart KKKK, the EPA identifies 14 subcategories of stationary combustion turbines and establishes NOX emission limits for ...What outreach and... · How did the EPA consider... · Proposed Determinations of...
  104. [104]
    CAEP-WG3 - ICAO
    The main aim of Working Group 3 (WG3, Emissions Technical) is to keep ICAO engine emissions and aeroplane CO2 emissions certification standards (Annex 16, ...
  105. [105]
    Control of Air Pollution From Aircraft Engines: Emission Standards ...
    Nov 23, 2022 · When developing new emission standards, ICAO/CAEP seeks to capture the technological advances made in the control of emissions through the ...
  106. [106]
    ICAO Aircraft Engine Emissions Databank | EASA - European Union
    The databank covers engine types which emissions are regulated, namely turbojet and turbofan engines with a static thrust greater than 26.7 kilonewtons. The ...Missing: CAEP | Show results with:CAEP
  107. [107]
    EPA proposes tightening NOx limits for new gas-fired power plants
    Nov 25, 2024 · EPA estimates its proposal would cut NOx emissions by 198 tons in 2027 and 2,659 tons in 2032, providing net benefits of up to $340 million. The ...
  108. [108]
    EPA Proposes Tighter Limits on Harmful NOx Emissions from New ...
    Nov 22, 2024 · EPA Proposes Tighter Limits on Harmful NOx Emissions from New Stationary Combustion Turbines to Better Protect Nearby Communities ; Large ...
  109. [109]
    How does a turbofan engine work? – The structure - AEROREPORT
    Jan 29, 2024 · Inside the combustor, the compressed air flowing into the chamber is mixed with fuel, where it burns at a temperature of about 1,700 degrees ...
  110. [110]
    ELI5 why pressure doesnt increase in the combustor when fuel is ...
    Sep 21, 2022 · Generally, the combustor flow area is designed to result in about 3-7% pressure drop. This pressure drop is required to allow control of the ...<|control11|><|separator|>
  111. [111]
    The aerodynamic challenges of aeroengine gas-turbine combustion ...
    The components of an aeroengine gas-turbine combustor have to perform multiple tasks – control of external and internal air distribution, fuel injector feed, ...
  112. [112]
  113. [113]
    SGT-400 Industrial gas turbine - Siemens Energy
    The Siemens Energy SGT-400 is a durable twin-shaft gas turbine suitable for power generation and mechanical drive applications in the 10 – 15 MW power band.
  114. [114]
    SGT-800 Combustion Turbine for Combined Cycle and ... - YouTube
    Apr 28, 2016 · The SGT-800 gas turbine is a small 50MW machine featuring a reduction gear-driven four-pole generator. The gas turbine utilizes an annular ...
  115. [115]
    Ramjet Propulsion
    The combustion that produces thrust in the ramjet occurs at a subsonic speed in the combustor. For a vehicle traveling supersonically, the air entering the ...
  116. [116]
    Ramjets and Scramjets - Purdue College of Engineering
    A ramjet uses the ram effect, compression in an inlet, for all of the compression in an engine. Hence, it has no rotor or blades.
  117. [117]
    Scramjet Propulsion
    The combustion process in a ramjet occurs at subsonic speeds in the combustor. For a vehicle traveling supersonically the air entering the engine must be slowed ...
  118. [118]
    [PDF] History of Ramjet and Scramjet Propulsion Development for U.S. ...
    These include surface- and air-launched subsonic combustion ramjets, supersonic combustion ramjets (scramjets), and mixed-cycle ramjet/scramjet/rocket.
  119. [119]
    [PDF] Scaling of Performance in Liquid Propellant Rocket Engine ...
    This paper discusses scaling of combustion and combustion performance in liquid propellant rocket engine combustion devices.
  120. [120]
    Prediction and control of combustion instabilities in real engines
    This paper presents recent progress in the field of thermoacoustic combustion instabilities in propulsion engines such as rockets or gas turbines.
  121. [121]
    Dry Low Emission - an overview | ScienceDirect Topics
    Dry Low Emission (DLE) combustors operate without water or steam, minimizing pollutants by burning fuel at cool, fuel-lean conditions.
  122. [122]
    Dry Low Emissions Combustor Development - ASME Digital Collection
    Dec 23, 2014 · This technology permits the operator to run with reduced emissions of NOx as well as CO and UHC over a wide load setting. Emissions goals of 25 ...
  123. [123]
    Towards Low NOx Emissions Performance of a 65kW Recuperated ...
    Aug 28, 2024 · Abstract. This work supports the development of a low NOx emission 65 kW natural gas turbine capable of operating on 100% hydrogen.
  124. [124]
    Modelling and development of ammonia-air non-premixed low NOX ...
    Oct 28, 2024 · This paper aims to develop a novel combustion strategy and describe the design of a low-NOX gas turbine combustor based on the rich-lean staged ...
  125. [125]
    Development and Testing of a Low NOX Hydrogen Combustion ...
    Aug 5, 2025 · A new premixing fuel injector for high-hydrogen fuels was designed to balance reliable, flashback-free operation, reasonable pressure drop, and low emissions.<|separator|>
  126. [126]
    [PDF] A Simplified Model for Detonation Based Pressure-Gain Combustors
    A time-dependent model is presented which simulates the essential physics of a detonative or otherwise constant volume, pressure-gain combustor for gas turbine ...
  127. [127]
    Thermodynamic analysis of a gas turbine engine with a rotating ...
    Jun 1, 2017 · A rotating detonation combustor is a form of “pressure gain combustion” where one or more detonations continuously travel around an annular ...
  128. [128]
    NETL Advances Rotating Detonation-Wave Combustor Technology
    Mar 23, 2021 · Rotating detonating engines create controlled, continuous detonation waves that rotate inside a modified gas turbine combustion chamber.
  129. [129]
    [PDF] Operability of a Natural Gas-Air Rotating Detonation Engine - OSTI.gov
    The present study seeks to understand how the dynamic injection, mixing, and chemical kinetic processes that occur in RDWCs affect engine operation for the ...<|separator|>
  130. [130]
    Pulse detonation propulsion: challenges, current status, and future ...
    The advantages of PDE for air-breathing propulsion are simplicity and easy scaling, reduced fuel consumption, and intrinsic capability of operation from zero ...
  131. [131]
    [PDF] Summary of Recent Research on Detonation Wave Engines at UTA
    A brief historical review of the early pulse detonation engine (PDE) research at. UTA is provided to lay the background for the development of a large-scale PDE ...
  132. [132]
    Pulse Detonation Engines: Advantages and Limitations - SpringerLink
    This paper reviews the efforts made over the years in adapting detonations for propulsion applications, and highlights new challenges in studying detonation<|separator|>
  133. [133]
    [PDF] Rotating Detonation Combustion for Gas Turbines
    To advance combustion turbine technologies for combined cycle applications… …by integrating a Rotating Detonation Engine (RDE), pressure gain combustion system ...
  134. [134]
    (PDF) Rotating Detonation: History, Results, Problems - ResearchGate
    The idea of using the phenomenon of rotating detonation to propulsion has its roots in fifties of the last century in works of Adamson et al. and Nicholls et al ...
  135. [135]
    Liquid fuels in rotating detonation engines: Advances and challenges
    Dec 20, 2024 · However, key challenges, such as multiple ignitions and the rapid and effective formation of detonation waves, lead to intermittent thrust and ...
  136. [136]
    [PDF] NASA's Rotating Detonation Rocket Engine Development
    The Rotating Detonation Rocket Engine has maintained steady development at NASA with many staggering performance advantages demonstrated to date over the state ...Missing: history | Show results with:history
  137. [137]
    Detonation Gas Turbines - ASME Digital Collection
    This article focuses on various technical and functional aspects of detonation gas turbines. Detonation combustion involves a supersonic flow.
  138. [138]
    Investigation of rotating detonation gas turbine cycle with different ...
    Rotating detonation combustion offers pressure gain characteristic, high thermal release rate, low pollution, and a wide working range [1], making it suitable ...
  139. [139]
    Combustor Design for Additive Manufacturing
    These operational issues, including flame static stability and thermoacoustic oscillations, are first-order drivers of improved engine performance, particularly ...Missing: aero | Show results with:aero
  140. [140]
    Design, simulation, and validation of additively manufactured high ...
    Nov 15, 2021 · This work provides guidance on the incorporation of additively manufactured features in combustors for any gas turbine application.
  141. [141]
    Siemens achieves breakthrough with 3D-printed combustion ...
    Aug 8, 2018 · The company has successfully 3D-printed and engine tested a dry low emission (DLE) pre-mixer for the SGT-A05 aeroderivative gas turbine.
  142. [142]
    Combustor Components - ADDere Additive Manufacturing
    Nov 1, 2022 · Additive manufacturing can offer several advantages for engine combustor design and fabrication, such as reducing weight, improving fuel ...
  143. [143]
    [PDF] additive manufacture and the gas turbine combustor - -ORCA
    Sep 10, 2021 · Advances in gas turbine (GT) combustion are enabled by metal additive manufacturing (AM) using selective laser melting (SLM) and other methods.<|separator|>
  144. [144]
    Challenges and Opportunities to Enable Low-Carbon Fuel Flexibility
    The additive manufacturing (AM) technique enables the fabrication of advanced burner components to enhance the hydrogen capability of the existing gas turbines ...
  145. [145]
    [PDF] Development of Metal AM Technology for Gas Turbine Components
    Mitsubishi Heavy Industries, Ltd. (MHI) Group has been developing additive manufacturing. (AM) as a method that can manufacture parts with complicated shapes ...
  146. [146]
    [PDF] ADDITIVE MANUFACTURING FOR HOT GAS PATH PARTS
    Additive Manufacturing (AM) offers a vast potential for the manufacturing of gas turbine components. Especially the powder bed based method utilizing a laser.
  147. [147]
    Printing parts for gas turbines is about to get easy. - Siemens Energy
    May 10, 2021 · 3D-printed parts for gas turbines under real life conditions possible. They help run turbines more efficiently, lower emissions and, if so desired, even extend ...<|separator|>
  148. [148]
    Combustion simulation and consultancy using CFD
    Computational fluid dynamics (CFD) simulation of your combustion processes gives you an overview, financial savings and reduces downtime at the plant.
  149. [149]
    CFD Software: Fluid Dynamics Simulation Software - Ansys
    High performance CFD software recognized for outstanding accuracy, robustness and speed with turbomachinery applications. Streamlined turbo setups; Blade design ...Ansys Fluent · CFD Software · Ansys CFX · Ansys Rocky
  150. [150]
    A Computational Fluid Dynamics-Based Small-Scale Combustor ...
    Oct 25, 2023 · The CFD is based on steady, incompressible, three-dimensional simulations with k–ω SST RANS and flamelet/progress variable combustion model.2 Modeling And Simulation... · 2.2 Combustor Parametric... · 4.2 Combustor Design...
  151. [151]
    BB-Agema Combustor Design - Success Story - Volupe - Volupe.com
    Combustor design relies on CFD simulation to give insight into flames, mixing, temperatures, and emissions. In this webinar, B&B-AGEMA presents results from ...
  152. [152]
    3D Printing to Enhance Industrial Gas Turbines - EOS GmbH
    Innovation for Maintenance: EOS additive manufacturing technology shortens repair times and reduces maintenance costs for industrial gas turbines.
  153. [153]
    Siemens Energy Uses 3D Printing to Speed Up Turbine Repairs
    Sep 9, 2025 · Siemens Energy is using EOS metal 3D printing to speed turbine repairs, modernize older components, and reduce maintenance time and ...
  154. [154]
    Advanced Computational Tools for Combustion Analysis and ...
    Argonne has developed a suite of five advanced computational tools for addressing complex challenges related to combustion analysis and engine design.
  155. [155]
    Current and future topics in additive manufacturing for gas turbine ...
    Particularly in the field of gas turbine combustion, AM offers potential in manufacturing components with improved mechanical properties and increasingly ...