Fact-checked by Grok 2 weeks ago

Stagnation temperature

Stagnation temperature, also known as total temperature, is the achieved by a flowing when it is brought to rest isentropically, converting its into without or irreversibilities. This concept arises in the steady flow energy for compressible fluids, where the stagnation T_0 relates to the static T and flow velocity V by the T_0 = T + \frac{V^2}{2 c_p}, with c_p being the specific heat at constant pressure for an . For real gases, deviations from behavior must be accounted for, particularly at high speeds where significant rises occur. In and gas dynamics, stagnation is a critical for characterizing conditions in high-speed environments, such as those encountered in engines, wind tunnels, and atmospheric reentry vehicles. It remains constant along streamlines in adiabatic flows without work, making it useful for performance analysis and . of stagnation typically involves total temperature probes that decelerate the adiabatically, though corrections are needed for non-ideal effects like conduction and losses. The is especially vital in hypersonic flows, where it can exceed thousands of , influencing and material limits.

Fundamentals

Definition and Physical Interpretation

Stagnation temperature, denoted as T_0, is defined as the temperature that a flowing attains when it is brought isentropically to rest, meaning through a reversible with no external work or . This concept arises in and , particularly in compressible flows, where it represents the state at a —such as the leading edge of an object in a stream—where the fluid velocity drops to zero. At this point, the fluid's fully converts into , elevating the temperature beyond the static temperature measured in the moving flow. Physically, stagnation temperature embodies the total content of the , combining its static with the dynamic from bulk motion. It signifies the maximum achievable by decelerating the to zero in an ideal, frictionless process, providing a in adiabatic duct flows without shaft work. This interpretation is crucial for understanding energy partitioning in high-speed flows, where the conversion of ordered into random occurs without increase in the isentropic idealization. Intuitively, the relation between stagnation and static temperatures follows from conservation of energy in steady flow: the total enthalpy remains constant, so for a calorically perfect gas, h_0 = h + \frac{v^2}{2}, where h_0 and h are stagnation and static enthalpies, and v is velocity. Since enthalpy h = c_p T for an ideal gas, this yields the basic equation T_0 = T + \frac{v^2}{2 c_p}, with T as static temperature and c_p as specific heat at constant pressure, illustrating how kinetic energy contributes to thermal energy.

Relation to Static and Total Temperatures

In compressible flows, the static temperature T represents the thermodynamic temperature of the fluid, corresponding to the random thermal motion of its molecules in the undisturbed, moving state. The stagnation temperature T_0, commonly synonymous with the total temperature T_t for ideal gases under isentropic conditions, is the temperature achieved when the flow is decelerated to rest through an isentropic process, incorporating both thermal and kinetic energy components. Within isentropic flows, stagnation establishes a relative to static , where T \leq T_0, with the stagnation value serving as an upper bound that remains along streamlines in steady, adiabatic conditions without external work. This conservation property arises from the steady , ensuring that the total enthalpy—and thus total —is constant for such processes. For ideal gases, the explicit relation between stagnation and static temperatures is \frac{T_0}{T} = 1 + \frac{\gamma - 1}{2} M^2 where \gamma denotes the specific heat ratio and M is the . This equation quantifies the temperature rise due to effects and underpins the design of tables and charts, which tabulate ratios like T_0 / T and p_0 / p as functions of Mach number to facilitate rapid property evaluation in aerodynamic analyses.

Theoretical Derivation

Adiabatic Flow Case

In the adiabatic case, the derivation of stagnation temperature assumes steady, one-dimensional, inviscid, and adiabatic of a perfect gas with constant specific heats. The process begins with the first law of applied to a steady energy , where the total remains constant along a streamline due to the absence of transfer and work. For a , this yields h + \frac{v^2}{2} = h_0, where h is the static , v is the , and h_0 is the stagnation corresponding to zero . For a perfect gas, is a function of only, given by h = c_p T, where c_p is the specific heat at constant pressure and T is the static . Substituting this relation produces c_p T + \frac{v^2}{2} = c_p T_0, which rearranges to the stagnation T_0 = T + \frac{v^2}{2 c_p}. To express this in dimensionless form, introduce the M = \frac{v}{a}, where a = \sqrt{\gamma [R](/page/R) T} is the , \gamma is the ratio of specific heats, and [R](/page/R) is the . Since c_p = \frac{\gamma [R](/page/R)}{\gamma - 1}, the velocity term becomes \frac{v^2}{2 c_p} = \frac{(\gamma - 1) M^2 T}{2}. Thus, the stagnation temperature is T_0 = T \left( 1 + \frac{\gamma - 1}{2} M^2 \right), relating the stagnation temperature directly to the static temperature and flow Mach number. This relation holds for calorically perfect gases under the stated assumptions but breaks down in hypersonic flows where real gas effects become significant.

Flow with Heat Addition

In Rayleigh flow, which models frictionless, steady, one-dimensional of an in a constant-area duct with heat addition or removal, the stagnation temperature varies along the duct due to the energy input from . This contrasts with the adiabatic flow case, where stagnation temperature remains constant in the absence of . The model assumes constant specific heats and is particularly relevant for analyzing processes like in confined spaces, where increases primarily from heat addition. The derivation begins with the steady-flow energy equation applied between two stations in the duct: c_p T_2 + \frac{v_2^2}{2} = c_p T_1 + \frac{v_1^2}{2} + q where q is the added per unit mass, T is the static temperature, v is the flow velocity, and c_p is the specific at constant . Expressing this in terms of stagnation temperatures T_0, the equation simplifies to show that the change in stagnation temperature directly results from the heat addition: T_{0_2} = T_{0_1} + \frac{q}{c_p} This relation holds regardless of the flow regime, indicating that heat addition always increases the stagnation temperature by an amount proportional to q / c_p. For integration along the duct in subsonic or supersonic flows, the momentum and continuity equations are combined with the energy equation to yield property ratios relative to sonic conditions (M = 1), such as the stagnation temperature ratio T_0 / T_{0^*} = \frac{2 (\gamma + 1) M^2 \left(1 + \frac{\gamma - 1}{2} M^2 \right) }{ (1 + \gamma M^2)^2 }, where \gamma is the specific heat ratio and * denotes sonic reference values; these allow computation of how T_0 evolves with progressive heat addition. The increase in stagnation temperature with heat addition has significant implications for flow behavior: in subsonic flow (M < 1), it accelerates the flow by increasing the Mach number toward unity while raising velocity; in supersonic flow (M > 1), it decelerates the flow by decreasing the Mach number toward unity while reducing velocity. This tendency to drive the flow toward sonic conditions at the throat can lead to if sufficient heat is added, limiting the maximum before the flow reaches M = 1. The model was developed and applied in the context of combustion analysis during the 1940s, as part of early efforts by organizations like NACA to understand heat addition in high-speed propulsion systems.

Measurement and Practical Considerations

Experimental Determination

Stagnation temperature is experimentally determined using specialized total temperature probes that facilitate isentropic deceleration of the flow to capture the at zero . These instruments typically employ , such as iron-constantan or platinum-rhodium types, housed within vented shields to minimize radiative and conductive losses. In settings, thermocouple rakes—arrays of multiple probes—enable spatial mapping of stagnation conditions across the flow field, as demonstrated in hypersonic facilities operating at numbers up to 10. The measurement procedure involves positioning the in the stream, where the gas is slowed through a duct-like to approximate isentropic while reducing viscous dissipation. The indicated T_{\text{probe}} is then adjusted via the \eta, defined as \eta = \frac{T_{\text{probe}} - T}{T_0 - T}, where T is the static and T_0 is the true stagnation ; typical values of \eta from 0.8 to 0.99, achieving near-unity (e.g., 0.998) at high Reynolds numbers but declining with lower speeds or higher due to incomplete . Calibration curves relate probe readings to parameters like and , often derived from tunnel surveys or free-jet methods to ensure accuracy within 1-2%. Key challenges in these measurements include heat conduction errors along probe leads and supports, which can cause significant under-reading (up to 6% error at low Reynolds numbers) by transferring away from the sensing , necessitating designs with high length-to-diameter ratios (e.g., >100) and low-conductivity materials like silica shields. In high-enthalpy environments, such as arc-jet tunnels simulating reentry conditions, calibration demands advanced techniques like (CFD) modeling with multi-species air chemistry to quantify uncertainties, often dominated by bias in and measurements that indirectly inform stagnation temperature. Adherence to established standards ensures measurement reliability, with ASME PTC 19.3 providing guidelines for thermocouple-based determination in flows, including corrections for conduction and installation effects in compressible regimes. For hypersonic applications, post-2000 protocols from AIAA and emphasize in high-temperature testing, such as through methods in arc-jet facilities, targeting errors below 5% for stagnation conditions.

Effects of Real Gases and Dissociation

In real gases, particularly under high-temperature conditions encountered in hypersonic flows, the ideal gas assumptions of constant specific heat at constant pressure (c_p) and constant specific heat ratio (\gamma = c_p / (c_p - R)) no longer hold, leading to significant deviations in stagnation temperature calculations. Vibrational excitation of diatomic molecules like N_2 and O_2 begins around 3–5, causing c_p to increase with temperature and \gamma to decrease from its room-temperature value of 1.4 toward 1.3 or lower, which alters the isentropic relations used to derive stagnation temperature from static conditions. These effects make the gas calorically imperfect, requiring iterative solutions to the energy equation where total h_0 = h + \frac{1}{2} V^2 is conserved, but the corresponding stagnation temperature T_0 is lower than the ideal gas prediction for the same enthalpy due to the higher effective c_p. To account for these deviations, equations of state beyond the are employed, such as the , which introduces corrections for molecular volume and intermolecular attractions: \left(P + \frac{a}{V_m^2}\right)(V_m - b) = RT, where a and b are substance-specific constants, V_m is , and this modifies pressure and volume in the thermodynamic relations for stagnation properties. For air in hypersonic applications, more advanced real-gas models, including virial expansions or tabular data from thermochemical equilibrium codes, are preferred over van der Waals for accuracy at elevated temperatures, as they better capture variable c_p and \gamma through polynomial fits or . Charts of real-gas isentropic exponents and enthalpy-temperature relations, derived from such models, allow engineers to correct ideal stagnation temperature estimates by up to 20–30% at Mach numbers above 5. At temperatures exceeding 2000 K, common in re-entry or high-speed environments, and introduce additional non-idealities by absorbing thermal energy through endothermic chemical reactions, further reducing the effective stagnation temperature rise compared to ideal predictions. For instance, oxygen (O_2 \rightarrow 2O) becomes significant above 2500 K, followed by (N_2 \rightarrow 2N) around 4000 K, and (e.g., O → O^+ + e^-) above 6000 K, all of which increase the number of and effectively raise c_p, diverting from sensible heating to energy. In re-entry flows, such as those around blunt bodies at –25, this dissociative cooling in the shock layer lowers post-shock stagnation temperatures by 10–25% relative to frozen-flow assumptions, mitigating heat loads but complicating boundary-layer predictions. These processes make air a reacting , often modeled with 5–11 (N_2, O_2, NO, N, O, etc.), where compositions are computed to determine the temperature corresponding to the total . Stagnation temperature in such regimes is calculated using frozen-flow or equilibrium-flow models to distinguish reaction rates from flow timescales. In frozen flow, chemical composition remains fixed (no dissociation/ionization), yielding higher T_0 as all kinetic energy converts to thermal energy without chemical absorption; this approximates rapid expansions or low-temperature limits. Conversely, equilibrium flow assumes instantaneous reactions, where dissociation lowers T_0 by distributing enthalpy across chemical modes; tools like NASA's Chemical Equilibrium with Applications (CEA) code solve for species mole fractions and thermodynamic properties at specified h_0 and pressure. For example, in air at Mach 10 and 100 km altitude (static T \approx 200 K, velocity \approx 3 km/s, total enthalpy h_0 \approx 4.5 MJ/kg), CEA equilibrium calculations yield T_0 \approx 3500–4000 K, compared to the ideal gas estimate of \approx 4200 K, due to \approx 20\% O_2 dissociation and vibrational excitation. Recent advancements in hypersonic since 2010 have incorporated quantum effects into real-gas models to improve accuracy in non-equilibrium flows, particularly for rates and radiative heating. The quantum-kinetic (Q-K) chemistry model, validated against (DSMC) for air , uses quantum mechanical cross-sections to predict dissociation and ionization probabilities more precisely than classical total collision energy models, reducing errors in stagnation region predictions by up to 15% in CFD simulations of re-entry vehicles. In 2024, the Q-K model was extended to include an anharmonic oscillator representation for more accurate reproduction of experimental measurements in non-equilibrium flows. Post-2010 validations, including Ames and AIAA studies on bow-shock experiments, demonstrate that Q-K enhancements in Navier-Stokes solvers like US3D better capture quantum tunneling in vibrational-dissociation coupling at temperatures 3000–8000 , essential for next-generation hypersonic design.

Applications

Aerodynamics and High-Speed Flows

In high-speed external flows, stagnation temperature plays a critical role in the around vehicles, particularly at the where the oncoming flow decelerates to zero , leading to maximum heating in the . For blunt bodies, such as the nose regions of hypersonic vehicles, the at the is governed by relations like the Fay-Riddell equation, which approximates q \propto \sqrt{\frac{\rho}{R}} (H_0 - H_w), where q is the , \rho is the , R is the nose radius of curvature, H_0 is the , and H_w is the wall enthalpy. This heating arises from the conversion of into across the shock layer and , with the stagnation temperature determining the driving potential for convective to the surface. Practical applications of stagnation temperature concepts are evident in the design of for missiles and high-speed , where the stagnation region experiences intense thermal loads that dictate material choices and shaping to minimize heating. For instance, in hypersonic missiles, the blunt geometry balances drag reduction with management, relying on stagnation temperature predictions to ensure structural integrity under flight conditions exceeding 5. Similarly, in testing of aerodynamic models, the recovery temperature—closely related to stagnation temperature—serves as a key metric to validate flow conditions and calibrate , providing insights into real-flight boundary layer behavior without direct of total temperature. The importance of stagnation temperature in predicting thermal loads is underscored by its use in selecting heat-resistant materials for vehicle surfaces, as it quantifies the difference driving or requirements. A notable case is the during atmospheric reentry, where peak temperatures reached approximately 1600 , necessitating advanced thermal protection systems like reinforced carbon-carbon for the nose cap to withstand the intense convective heating. This analysis not only informed the Shuttle's design but also highlighted the need for accurate external modeling to extend vehicle reusability in high-speed regimes.

Propulsion and Engine Performance

In propulsion systems such as turbojets and ramjets, stagnation temperature is integral to the Brayton thermodynamic cycle that governs engine operation. The cycle's thermal efficiency depends on the compressor inlet stagnation temperature T_{01}, which represents the total temperature of the airflow entering the compressor after deceleration in the inlet; higher inlet stagnation temperatures from ram compression increase the temperature ratio across the compressor, thereby enhancing overall efficiency up to the limits imposed by material constraints. For ideal conditions, this efficiency is expressed as \eta_B = 1 - \frac{1}{r_p^{(\gamma-1)/\gamma}}, where r_p is the pressure ratio influenced by the inlet stagnation conditions. Ram recovery in the engine further elevates , calculated as T_{02} = T_0 \left(1 + \frac{\gamma - 1}{2} M^2 \right), where T_0 is the static , \gamma is the specific , and M is the flight ; this assumes an adiabatic, isentropic diffuser that preserves total while converting to . In practice, total remains constant through the for adiabatic flow, but pressure recovery inefficiencies in supersonic inlets can indirectly affect downstream stagnation conditions by altering mass flow and compression. Key performance metrics, including (TSFC), are constrained by the inlet stagnation temperature T_{04}, typically limited to around 1700 K by the melting points and resistance of nickel-based superalloys used in blades. Elevating T_{04} reduces TSFC by allowing greater in the , but it demands sophisticated cooling to prevent material failure, directly impacting engine and operational range. In military turbojet applications, afterburners inject fuel downstream of the turbine to raise the exhaust stagnation temperature, often by 800–1200 K, enabling supersonic thrust augmentation for maneuvers in aircraft like the F-15 Eagle. This heat addition increases exhaust velocity and thrust at the cost of higher fuel use, but it is limited by nozzle materials to avoid thermal shock. For hypersonic scramjets, stagnation temperatures surpass 3000 K at 8+, posing dissociation risks to air molecules and requiring active cooling to sustain combustion stability and structural integrity. Post-2020 advancements in additive manufacturing have facilitated higher-temperature designs by enabling intricate internal cooling geometries in superalloys, improving cycle in next-generation engines.

Solar Thermal Collectors

In solar thermal collectors, stagnation refers to the maximum reached by the absorber or when there is no heat extraction, such as during pump failure or excess solar input, where absorbed solar radiation balances thermal losses to the . This condition can lead to overheating, posing risks to system integrity and performance. The stagnation is calculated using the relation T_{\text{stag}} = T_{\text{amb}} + \frac{G \cdot (\tau \alpha)}{U_L}, where T_{\text{amb}} is the ambient , G is the solar irradiance, (\tau \alpha) is the transmittance-absorptance product representing optical , and U_L is the overall heat loss coefficient. For typical conditions with G = 1000 W/m², flat-plate collectors achieve stagnation temperatures of approximately 180–210°C with selective coatings. Design implications of stagnation temperature are significant, as high values limit material choices and necessitate protective features. For instance, polymers used in gaskets or seals often degrade above 150°C, requiring alternatives like metals or silicones that withstand higher temperatures. To mitigate damage, systems incorporate venting mechanisms or expansion vessels that release pressure from vaporized fluids, preventing bursts or corrosion. Different collector types exhibit varying stagnation temperatures due to their optical and thermal designs. Evacuated tube collectors, with vacuum insulation reducing losses, reach 220–300°C under stagnation, higher than non-evacuated flat-plate types but still manageable with standard fluids. In contrast, concentrating systems like parabolic trough collectors, which focus sunlight to ratios of 20–80, can attain stagnation temperatures up to 500°C, demanding high-temperature synthetics oils or molten salts as fluids. Testing standards such as ISO 9806 (updated 2017) specify procedures for determining stagnation temperature, including dry stagnation tests at standard irradiance to ensure reliability and safety across these types.

References

  1. [1]
    Stagnation Temperature - Real Gas Effects
    If the moving flow is isentropically brought to a halt on the body, we measure the stagnation temperature. The stagnation temperature is important because it is ...
  2. [2]
    Steady Flow Energy Equation - MIT
    The stagnation temperature and stagnation pressure are the conditions the fluid would reach if it were brought to zero speed relative to some reference frame, ...Missing: dynamics | Show results with:dynamics
  3. [3]
    [PDF] Stagnation Properties and Mach Number Compressible p and ...
    • Stagnation Temperature. – from energy conservation: no work but flow ... – always defined in the flow's reference frame still air moving air still ...
  4. [4]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics ...
    Dec 15, 2021 · (5) Note that the stagnation temperature is greater than the flow temperature since when the flow is decelerated to zero velocity, the ...
  5. [5]
    [PDF] AME 436 - Paul D. Ronney
    ➢ Define stagnation temperature Tt = temperature of gas when decelerated ... ➢ T = static temperature - T measured by thermometer moving with flow.Missing: fluid dynamics
  6. [6]
    Thermodynamic Foundations – Introduction to Aerospace Flight ...
    The stagnation temperature, by contrast, is defined as the temperature the fluid would reach if it were decelerated to zero velocity in an isentropic manner.
  7. [7]
    HISTORY OF BOUNDARY LA YER THEORY - Annual Reviews
    The boundary-layer theory began with Ludwig Prandtl's paper On the motion of a fluid with very small viscosity, which was presented at the Third ...
  8. [8]
    Lecture T10: Stagnation Quantities General comments - MIT
    If location (2) is a place where the flow is stagnated (so that c2=0) then we call T2 the stagnation temperature and give it the general symbol TT. which stands ...
  9. [9]
    Steady Flow Energy Equation - MIT
    Note that for any steady, adiabatic flow with no external work, the stagnation temperature is constant.
  10. [10]
    [PDF] STEADY FLOW ENERGY EQUATION
    Note that for any steady, adiabatic flow with no external work, the stagnation temperature is constant.
  11. [11]
    [PDF] Equations, Tables and Charts for Compressible Flow
    The tables present useful dimensionless ratios for continuous one-dimen- sional flow and for normal shock waves as functions of Mach number for air considered ...Missing: stagnation | Show results with:stagnation
  12. [12]
  13. [13]
    [PDF] Rayleigh Flow – compressible flow with heat transfer
    Use the ratio equation for stagnation temperature (inversely) to calculate Ma2: ... Use the remaining Rayleigh flow equations, ideal gas law, speed of sound ...
  14. [14]
    Flows with heat transfer (Rayleigh flows) — Gas Dynamics notes
    First, we need to find the Mach number and stagnation temperature at the inlet. We can also identify the change in stagnation temperature using ...
  15. [15]
    [PDF] Rayleigh & Fanno Flows
    This equation clearly shows that the heat addition directly changes the total temperature of the flow. • Heat addition → T0 increases. • Heat removal → T0 ...Missing: stagnation | Show results with:stagnation
  16. [16]
    A Century of Ramjet Propulsion Technology Evolution - AIAA ARC
    The ramjet engine began to receive attention during the second- half of the 1940s and reached a relative peak during the 1950s with a number of operational ...
  17. [17]
    [PDF] STAGNATION TEMPERATURE PROBES FOR USE AT HIGH ... - DTIC
    For boundary layer studies in a wind tunnel a small probe, of the order of 1-mm height at the air entrance, is desirable in order to obtain a sufficient number ...
  18. [18]
    [PDF] Total Temperature Probes for High-Temperature Hypersonic ...
    This paper presents design details and test results of two types of total temperature probes that were used for hypersonic boundary-layer measurements.
  19. [19]
    Thermocouple Recovery Factor for Temperature Measurements in ...
    Thermocouple Recovery Factor for Temperature Measurements in Turbomachinery Test Facilities. Mark Fernelius and; Steven E. Gorrell. Mark Fernelius.Missing: probe | Show results with:probe
  20. [20]
    [PDF] Calibration Probe Uncertainty and Validation for the Hypersonic ...
    In addition to the surface pressure and heat flux, conditions in the arc jet are measured, including plenum total pressure downstream of the arc heater, prior ...
  21. [21]
    Temperature Measurement - PTC 19.3 - ASME
    In stockThe purpose of this ASME Performance Test Code (PTC) is to give instructions and guidance for the accurate determination of temperature values.<|control11|><|separator|>
  22. [22]
    Real Gas Effects
    Because the total temperature does not change through a shock wave, the stagnation temperature and and the total temperature have the same value at a stagnation ...
  23. [23]
    [PDF] A METHOD TO CALCULATE THE REAL GAS STAGNATION ...
    The real gas effect influence the total properties, which is evidenced by the total temperature distribution. It is recognized that for inviscid flows, total ...
  24. [24]
    [PDF] FUNDAMENTAL CONCEPTS OF REAL GASDYNAMICS
    We can define macroscopic fluid properties by taking averages over the ... Using the definitions of stagnation temperature (2.31) and stagnation pressure.
  25. [25]
  26. [26]
    High-Temperature Air Effects in High-Speed Flows | IntechOpen
    The dependencies of these properties on flow quantities become more complex in high-temperature air when processes of dissociation and ionization are apparent.
  27. [27]
    [PDF] STAGNATION POINT HEAT TRANSFER FOR HYPERSONIC FLOW
    The heat transfer equation incor- porates real gas effects through use of dissociated air properties and a fitted curve for the stagnation point velocity ...
  28. [28]
    [PDF] hypersonic shock-heated flow parameters for velocities to 46000 ...
    Thermochemical equilibrium-flow properties are given for the inviscid normal shock and stagnation point, and both equilibrium and frozen-flow parameters ...
  29. [29]
    [PDF] NASA Chemical Equilibrium with Applications (CEA) Tutorial
    May 1, 2024 · CEA, or Chemical Equilibrium with Applications, is a robust chemical equilibrium solver with numerous applications, used in government, ...Missing: stagnation | Show results with:stagnation
  30. [30]
    Open-Source Direct Simulation Monte Carlo Chemistry Modeling for ...
    This paper presents an open-source chemistry modeling for direct simulation Monte Carlo (DSMC) using the quantum-kinetic method in dsmcFoam for hypersonic ...
  31. [31]
    [PDF] Continuum Simulation of Hypersonic Flows using the Quantum ...
    The results show that in thermal equilibrium the reaction rates between these two models are comparable. The QK model predicts greater rates for some chemical ...Missing: post- | Show results with:post-
  32. [32]
    None
    ### Summary of Stagnation Point Heating for Blunt Bodies
  33. [33]
    None
    Insufficient relevant content. The provided URL (https://ntrs.nasa.gov/api/citations/19650024663/downloads/19650024663.pdf) returned a "Not Found" status (404), indicating the document is unavailable. No content could be extracted or summarized regarding the Fay-Riddell equation for stagnation point heat transfer.
  34. [34]
    [PDF] Aerodynamic Heating - DTIC
    Aerodynamic heating is the conversion of kinetic energy into heat energy as the result of the relative motion between a body and a fluid, and.
  35. [35]
    Wind-Tunnel Tests for Temperature Recovery Factors at Supersonic ...
    The value of the temperature recovery factor is 88.0 per cent for these experiments. This is a little cooler than Squire's theoretical value of 89.42 per cent ...
  36. [36]
    [PDF] SPACE SHUTTLE ORBITER ENTRY HEATING AND TPS RESPONSE
    In this paper, representative aerothermodynamic and TPS thermal response data obtained on the first atmospheric entry test flight (STS-1) of the Space Shuttle ...
  37. [37]
    3.7 Brayton Cycle - MIT
    The Brayton cycle thermal efficiency contains the ratio of the compressor exit temperature to atmospheric temperature, so that the ratio is not based on the ...
  38. [38]
    11.6 Performance of Jet Engines - MIT
    The ratios of stagnation to static pressure at exit and at inlet are the same, with the consequence that the inlet and exit Mach numbers are also the same. $\ ...Missing: recovery | Show results with:recovery
  39. [39]
    Inlet Performance
    Inlet performance is measured by pressure recovery, and a good inlet has high recovery, low spillage drag, and low distortion. Total temperature is constant.Missing: stagnation | Show results with:stagnation
  40. [40]
    (PDF) EFFECT OF COMPRESSION RATIO AND TURBINE INLET ...
    Dec 13, 2020 · Considering that the compressor compression ranges from 5 to 35 and the turbine inlet temperature ranges from 1100 K to 1700 K, surface graphs ...
  41. [41]
    Combustor technology of high temperature rise for aero engine
    Jul 1, 2023 · In terms of engine cycle performance, increasing the turbine inlet temperature (TIT) is the most direct way to increase the thrust-to-weight ...
  42. [42]
    Afterburner - an overview | ScienceDirect Topics
    The increased stagnation temperature permits the generation of higher exit velocity V7 and thereby increases the thrust capability of the engine. Sign in to ...
  43. [43]
    Turbojet Engines – Introduction to Aerospace Flight Vehicles
    Nevertheless, the turbine vanes and blades require internal cooling passages to maintain acceptable material temperatures. 5. Exhaust stage through a nozzle.
  44. [44]
    [PDF] High Temperature Additive Architectures for 65% Efficiency
    Nov 10, 2021 · Develop a feasible conceptual design for advanced additive turbine inlet components that enable 65% CC efficiency through analytical methods and ...
  45. [45]
    Leveraging Additive Manufacturing to Fabricate High Temperature ...
    Exciting developments in employing powder-based directed energy deposition (DED) to manufacture functionally graded materials (FGMs) with material compositions ...Missing: post- | Show results with:post-
  46. [46]
  47. [47]
    [PDF] Stagnation behaviour of solar thermal systems - IEA SHC
    Then, the temperature of the absorber rises rapidly and reaches the so-called stagnation temperature, which is from. 180 to 210 °C for selective coated ...Missing: formula | Show results with:formula
  48. [48]
    Potential for Mitigation of Solar Collector Overheating ... - JSDEWES
    The temperature of stagnation of the most basic flat plate solar collector design is high and regularly exceeds 150 °C. This limits the selection of materials ...
  49. [49]
    [PDF] Overheating prevention and stagnation handling in solar process ...
    Jan 15, 2015 · As explained in detail in this report, stagnation describes the state of a solar thermal system in which the flow in the collector loop is ...
  50. [50]
    [PDF] GUIDE TO STANDARD ISO 9806:2017 A Resource for ... - IEA SHC
    Oct 26, 2017 · The following chart shows exemplary temperature ranges of reachable standard stagnation temperatures depending on the collector technology.