Fact-checked by Grok 2 weeks ago

Rayleigh flow

Rayleigh flow refers to the steady, one-dimensional, frictionless flow of a compressible through a constant-area duct with addition or removal, serving as a fundamental model in aerothermodynamics for analyzing non-adiabatic processes. This flow is characterized by the absence of wall , allowing to dominate changes in flow properties such as temperature, pressure, velocity, and , while the duct's constant cross-section maintains . The analysis of Rayleigh flow relies on conservation of mass, , and , assuming a calorically perfect gas with constant specific heats and no body forces or viscous effects. Key relations include the linking static pressures and dynamic pressures across sections, and the incorporating heat input as q = c_p (T_{02} - T_{01}), where T_0 is the . These yield ratios such as stagnation pressure \frac{P_{02}}{P_{01}} = \frac{1 + k Ma_1^2}{1 + k Ma_2^2} \left( \frac{1 + \frac{k-1}{2} Ma_2^2}{1 + \frac{k-1}{2} Ma_1^2} \right)^{k/(k-1)}, highlighting how alters thermodynamic states. A central concept is the Rayleigh line, plotted on a temperature-entropy (T-s) or enthalpy-entropy (h-s) diagram, which traces possible states from subsonic to supersonic regimes, with the sonic point (Mach number Ma = 1) as a reference. Heat addition drives subsonic flows toward sonic conditions by increasing Mach number, velocity, and temperature while decreasing density and pressure; conversely, in supersonic flows, it decelerates the flow toward Ma = 1, raising temperature and pressure but lowering velocity and density. This process can lead to thermal choking at the sonic point, limiting maximum heat input to q_{\max} = c_p (T_{0*} - T_{01}), beyond which mass flow rate decreases. Unlike Fanno flow, which models frictional effects without heat transfer, Rayleigh flow emphasizes heat's role in flow acceleration or deceleration. Rayleigh flow models critical components in systems, including combustors, afterburners, and inlets, where heat addition influences performance and efficiency. It also applies to heat exchangers and nozzles with significant thermal effects, providing insights into phenomena and stagnation pressure losses, which are essential for designing high-speed systems.

Fundamentals

Definition and Historical Context

Rayleigh flow refers to the frictionless, steady, one-dimensional flow of an through a constant-area duct, where addition or removal is the dominant process affecting the flow properties. This model isolates the effects of exchange on behavior, without considering frictional losses or area changes, making it a fundamental tool for analyzing non-adiabatic processes in gas dynamics. The concept originated with Lord Rayleigh, whose full name was John William Strutt (1842–1919), a British physicist renowned for contributions to acoustics and . In his late 19th-century work on gas dynamics (circa 1885), Rayleigh analyzed gas flows involving heat exchange, particularly the interplay between mechanical and thermal energy during flow discontinuities or chemical reactions. This laid the groundwork for understanding how influences flow regimes, including the phenomenon of thermal choking, where excessive heat addition limits further mass flow. Rayleigh flow was formalized as a standard model in modern gas dynamics during the mid-20th century, notably through the comprehensive treatments in texts like Ascher H. Shapiro's The Dynamics and Thermodynamics of Compressible Fluid Flow (Volume 1, 1953), which integrated it into the broader framework of compressible flow theory. Applicable to both subsonic and supersonic regimes, it distinguishes itself from adiabatic flows by emphasizing heat transfer's role in altering velocity, pressure, and temperature profiles. As a complementary model to Fanno flow, which incorporates frictional effects instead of heat transfer, Rayleigh flow aids in studying thermal limits in engineering applications such as combustors and nozzles.

Assumptions of the Model

The Rayleigh flow model relies on a set of simplifying assumptions to analyze the effects of heat addition or removal on in a duct. The flow is steady, meaning all properties are time-independent at any given location. It is also one-dimensional, with flow properties uniform across the duct cross-section and varying only in the axial direction, neglecting radial or circumferential variations. The duct has a constant cross-sectional area, ensuring no geometric changes influence the momentum balance. Friction is absent, treating the flow as inviscid with no wall or viscous dissipation. Heat transfer is the sole external interaction, occurring without any mechanical work input or output to the fluid. The fluid is modeled as a perfect gas that is calorically perfect, obeying the PV = RT and exhibiting constant specific heats c_p and c_v, with the ratio of specific heats \gamma = c_p / c_v remaining fixed regardless of . This assumption holds even in scenarios involving , where gas composition changes are ignored to maintain constant properties like the R. Diffusive effects, such as longitudinal conduction and viscous stresses in the flow direction, are neglected to focus solely on convective transport and heat addition. These idealizations exclude real-world phenomena like growth, multi-dimensional flow structures, and property variations due to or at high temperatures, which can significantly alter flow behavior. By omitting and , the model emphasizes transfer's role in driving changes in and thermodynamic states, making it suitable for approximating conditions in high-speed internal flows—such as those in afterburners or supersonic combustors—where frictional effects are secondary to thermal ones. In practice, however, the assumptions permit closed-form analytical solutions but often require empirical corrections for s and gas properties to align with experimental data.

Governing Equations

Derivation from Conservation Principles

Rayleigh flow is modeled as a steady, one-dimensional in a constant-area duct under inviscid conditions with addition or removal, allowing the application of integral principles along the flow path. The for steady in a duct of constant cross-sectional area A yields the relation \rho u A = \dot{m} = \text{constant}, where \rho is the , u is the , and \dot{m} is the . Since A is constant, this simplifies to \rho u = \text{constant}, implying that any increase in velocity must be accompanied by a corresponding decrease in to maintain . Conservation of , in the absence of wall friction and body , integrates along the duct to give the Rayleigh line p + \rho u^2 = \text{constant}, where p is the . This relation arises from the differential form dp + \rho u \, du = 0, obtained by considering the balance over a : the net balances the change in , leading to a linear relationship in the p-\rho when combined with the . The accounts for and states that the h_0 = h + \frac{u^2}{2} varies along the flow due to addition, where h is the static . For an with constant specific heat, the added per unit mass q relates to the change in as q = c_p (T_{02} - T_{01}), with T_0 denoting ; thus, integrating the yields h_2 + \frac{u_2^2}{2} = h_1 + \frac{u_1^2}{2} + q. Combining these conservation laws—mass (\rho u = \text{constant}), momentum (p + \rho u^2 = \text{constant}), and energy (h + \frac{u^2}{2} = h_0, with h_0 varying via q)—provides a framework to relate the thermodynamic state variables p, \rho, u, and temperature T (via h = c_p T) across sections of the duct without initially introducing the explicit heat transfer term, establishing the foundational integral relations for analyzing property changes induced by heating or cooling.

Key Differential Relations

The key differential relation in Rayleigh flow governs the evolution of the Mach number M as heat is added or removed, providing the analytical foundation for predicting how the flow accelerates or decelerates toward sonic conditions. This relation is derived from the energy equation, which equates the differential change in stagnation enthalpy to the heat addition per unit mass: dh_0 = dq = c_p dT_0, where c_p is the constant-pressure specific heat and T_0 is the stagnation temperature. Combining this with the definition of the speed of sound a = \sqrt{\gamma R T} and isentropic relations for an ideal gas, the momentum and continuity equations in differential form are incorporated to express changes in velocity, temperature, and pressure. The derivation proceeds by applying the differential form of the energy equation dh_0 = c_p dT_0 alongside the , which implies du/u = -d\rho/\rho, and the equation yielding dp/p = -\gamma M^2 (du/u). These are normalized relative to sonic reference conditions (denoted by *), where M = 1, to obtain the relation for the change in squared: \frac{dM^2}{M^2} = \frac{2(1 - M^2)}{1 + (\gamma - 1)M^2} \frac{dT_0}{T_0}. The significance of this differential relation lies in its demonstration of the flow's behavior under heat : for (M < 1), the term (1 - M^2) > 0 results in dM^2 > 0, accelerating the flow toward M = 1; conversely, for supersonic flow (M > 1), the negative (1 - M^2) term yields dM^2 < 0, decelerating the flow toward M = 1. This convergence to sonic conditions at the throat underscores the potential for thermal choking in ducted flows with heat transfer.

Properties of Rayleigh Flow

Thermodynamic Property Variations

In Rayleigh flow, the static thermodynamic properties—pressure, temperature, density, and velocity—vary algebraically with the local Mach number M for an ideal gas with constant specific heat ratio \gamma. These relations are obtained by integrating the governing conservation equations along the duct and normalizing all properties to the sonic reference state (denoted by *), which corresponds to the hypothetical location where M = 1. This normalization facilitates quantitative analysis of how heat addition or removal alters the flow state relative to the choking condition. The static pressure ratio is expressed as \frac{p}{p^*} = \frac{\gamma + 1}{1 + \gamma M^2}, where p^* is the pressure at M = 1. This formula indicates that pressure is highest for low-Mach-number flows and decreases as M increases in either subsonic or supersonic regimes. The static temperature ratio follows from the energy equation and is given by \frac{T}{T^*} = \frac{M^2 (\gamma + 1)^2}{(1 + \gamma M^2)^2}, with T^* denoting the temperature at the sonic point. Temperature exhibits a minimum near M = 1/\sqrt{\gamma} (approximately 0.845 for \gamma = 1.4) and rises toward stagnation-like values as M \to 0 or as M \to \infty. The static density ratio, derived from the ideal gas law and the above relations, is \frac{\rho}{\rho^*} = \frac{1 + \gamma M^2}{(\gamma + 1) M^2}, where \rho^* is the density at M = 1. Density becomes very large as M \to 0 due to the low temperature, and it decreases toward zero as M \to \infty. Since mass flow rate is conserved in the constant-area duct, the velocity ratio is the reciprocal: \frac{u}{u^*} = \frac{(\gamma + 1) M^2}{1 + \gamma M^2}, with u^* the velocity at sonic conditions; velocity starts at zero for M = 0 and approaches an asymptotic maximum for large M. These property ratios highlight the qualitative behavior under heat addition. For subsonic flow (M < 1), heat addition reduces while increasing velocity (and toward 1). For supersonic flow (M > 1), the trends reverse: heat addition raises while reducing velocity (and toward 1). Such variations underscore the role of in modulating flow or deceleration without changing the duct area.

Stagnation Parameters and Entropy

In Rayleigh flow, the stagnation temperature ratio relative to the sonic reference state (denoted by *) is given by \frac{T_0}{T_0^*} = \frac{2(\gamma + 1)M^2 \left(1 + \frac{\gamma - 1}{2}M^2\right)}{\left(1 + \gamma M^2\right)^2}, where \gamma is the specific heat ratio and M is the . This relation arises from combining the equation with the static temperature ratio across the flow, reflecting how heat addition alters the total thermal energy available for conversion to . For subsonic flows (M < 1), the ratio T_0 / T_0^* increases toward unity as heat is added, reaching its reference value at the choking condition (M = 1); in supersonic flows (M > 1), heat addition decreases the ratio below unity. The stagnation pressure ratio, which quantifies the loss in total pressure due to the irreversible process, is expressed as \frac{p_0}{p_0^*} = \frac{\gamma + 1}{1 + \gamma M^2} \left[ \frac{2}{\gamma + 1} \left(1 + \frac{\gamma - 1}{2}M^2\right) \right]^{\gamma / (\gamma - 1)}. This formula incorporates the ratio and the isentropic relations at the local and sonic conditions. Unlike isentropic flow, where remains constant, in Rayleigh flow the ratio exceeds 1 in subsonic regimes and decreases monotonically toward 1 as is added (increasing M toward 1); in supersonic regimes, it is less than 1 and increases toward 1 with addition (decreasing M toward 1). Overall, the ratio is maximized near M \to 0 and minimized as M \to \infty. For example, with \gamma = 1.4 and M = 0.5, the ratio is approximately 1.11, indicating a modest total recovery compared to the sonic reference. Entropy changes in Rayleigh flow highlight the inherent irreversibility of frictional at constant area. The dimensionless entropy change relative to the state is \frac{\Delta s}{c_p} = \ln \left[ M^2 \left( \frac{\gamma + 1}{1 + \gamma M^2} \right)^{(\gamma + 1)/\gamma} \right], where c_p is the specific heat at constant . This expression, derived from the second applied to the relations for and , shows that reaches its maximum at M = 1, corresponding to the point where addition is maximized for a given . addition invariably increases across both subsonic and supersonic branches, as the process involves irreversible mixing of with the , leading to a net rise in disorder until the limit is approached. In contrast, removal decreases , but the flow remains non-isentropic due to the underlying thermal gradients.

Flow Behavior and Limits

Impact of Heat Addition and Removal

In Rayleigh flow, the addition of heat to a flow (Mach number M < 1) results in an increase in the and flow velocity, approaching sonic conditions at M = 1, while the static temperature rises and the static pressure and density decrease. This behavior arises from the conservation of mass and momentum in a constant-area duct, where the energy input expands the gas, accelerating the flow despite the fixed cross-section. Conversely, heat removal in flow reverses these trends, decreasing the and velocity while increasing pressure and density. For supersonic flow (M > 1), heat addition leads to a decrease in the and flow velocity, again driving the flow toward the sonic limit at M = 1, with an accompanying increase in static , , and . The compression-like effects on and stem from the slowing of the high-speed flow, which compresses the gas molecules closer together under the influence of added . Heat removal in the supersonic regime inverts these changes, increasing the and velocity while decreasing , , and . These reversal points highlight the thermal phenomenon, where subsonic flows accelerate to and supersonic flows decelerate to ; however, transitioning from supersonic to subsonic conditions requires an external , as heat transfer alone cannot cross the sonic barrier without it. The qualitative variations in flow properties can be illustrated using property ratios relative to sonic reference conditions, such as those for and , which underscore the divergent behaviors between and supersonic branches. On a -entropy (T-s) , the process traces the Rayleigh line, a locus of states satisfying and for varying addition; heat input moves the state point upward along this line, increasing in both regimes and converging toward the point of maximum at M = 1.

Choking Condition and Maximum Heat Transfer

In Rayleigh flow, the choking condition arises when heat addition drives the Mach number to unity (M = 1) at the duct exit, establishing a throat that fixes the for given inlet conditions. Beyond this point, any additional heat input cannot increase the mass flow; instead, it triggers an upstream adjustment to maintain the choked state, effectively limiting the duct's throughput. This phenomenon is derived from the conservation equations, where the and balances constrain the flow such that the sonic condition represents the maximum state along the Rayleigh line. The maximum heat addition, denoted as q_{\max}, occurs precisely at this choking point for subsonic inlet flows and is given by q_{\max} = \dot{m} c_p (T_{0}^* - T_{0,1}), where \dot{m} is the , c_p is the specific heat at constant pressure, T_{0,1} is the inlet , and T_{0}^* is the at the sonic reference state (M = 1). This limit stems from the Rayleigh flow relations, where the ratio T_0 / T_0^* approaches unity as M reaches 1, marking the peak heat input before the flow chokes. For supersonic inlet flows, choking requires heat addition to reduce M to 1. Heating drives the flow to sonic conditions along the Rayleigh line. In Rayleigh flow, heat addition progressively accelerates the flow toward M = 1, while in supersonic flow, heating decelerates it to the same limit, with potential for a post-choke if heating exceeds the threshold in supersonic cases. These behaviors highlight the asymmetric response to thermal loading across the divide. Practically, the condition imposes strict limits on design in systems, such as ramjets or afterburners, where stagnation temperatures are capped near 2000 K to prevent thermal choking and ensure stable operation without upstream flow disruption.

Comparisons and Extensions

Comparison to Fanno and Isentropic Flows

Rayleigh flow and Fanno flow both describe steady, one-dimensional flows in constant-area ducts for an ideal gas with constant specific heats, but they differ fundamentally in their physical mechanisms. Rayleigh flow assumes frictionless conditions with heat addition or rejection, leading to variations in stagnation temperature T_0 while stagnation pressure p_0 decreases due to irreversibilities from heat transfer. In contrast, Fanno flow is adiabatic with wall friction but no heat transfer, maintaining constant T_0 while p_0 also decreases from frictional losses. Both processes increase entropy, but Rayleigh flow models heat-dominated phenomena like combustion, whereas Fanno flow captures friction-dominated effects in long ducts. On the Mollier (enthalpy-entropy, h-s) diagram, and Fanno lines originate from the same initial state and intersect at two points, representing the pre- and post-normal conditions for the same and . For example, starting from an initial supersonic of 3, the intersection yields a post- of approximately 0.475, illustrating how a normal satisfies both (heat-like) and Fanno (friction-like) constraints simultaneously. These intersections are crucial for analyzing combined effects in duct flows, such as in engines where both addition and occur. Compared to isentropic flow, Rayleigh flow is inherently non-isentropic, with rising (\Delta s > 0) due to across a finite difference, and it occurs in constant-area ducts without nozzle-like . Isentropic flow, by , is reversible and adiabatic (\Delta s = 0), typically involving area changes to accelerate or decelerate the while preserving both T_0 and p_0. Thus, Rayleigh flow deviates from the ideal isentropic path on the h-s diagram by following a curved Rayleigh line of constant momentum flux, enabling analysis of thermal effects absent in frictionless, adiabatic expansions.

Extensions to Non-Ideal Conditions

In real-world applications, the ideal Rayleigh flow model assumes constant specific heats and a calorically perfect gas, but deviations arise when gases exhibit variable specific heats due to vibrational excitation or other molecular effects at high temperatures. For calorically imperfect gases, the specific heats c_p and c_v vary with temperature, often modeled using for diatomic gases where vibrational modes contribute. These modifications require of the governing differential equations rather than closed-form algebraic ratios, as the energy equation involves integrating variable c_p(T). Seminal work by Eggers analyzed one-dimensional flows of imperfect diatomic gases, primarily focusing on shock and isentropic cases. Caloric imperfections affect stagnation properties and heat addition limits compared to the perfect gas case. For flows involving wall shear, the frictionless assumption of ideal Rayleigh flow is extended by combining it with effects, resulting in coupled models that account for both and frictional losses in constant-area ducts. These combined Rayleigh-Fanno models solve the with a friction term \frac{dp}{dx} + \rho V \frac{dV}{dx} + \frac{f \rho V^2}{2D} = 0 alongside the energy for addition dq = c_p dT + V dV, where f is the and D the . A proportionality function r(M) = \frac{n - k M^2}{1 - M^2} (with n = r + (k - r) M^2 and k the specific ) parameterizes the relative influence of and on evolution, enabling numerical solutions via methods like for integrating property differentials. For cases with small , techniques approximate the solution by treating as a small \epsilon, expanding variables around the frictionless Rayleigh solution and solving successive orders of the perturbed s, which is particularly useful in short ducts where wall shear is minor but non-negligible. Assumptions include a perfect gas, for , and transverse , as detailed in foundational analyses. The one-dimensional assumption in Rayleigh flow neglects multi-dimensional effects such as non-uniform heating and development, which become significant in short ducts or with asymmetric heat addition. (CFD) extensions model these by solving the full Navier-Stokes equations with conjugate boundary conditions, capturing radial temperature gradients and velocity profiles that deviate from uniform 1D predictions. For instance, in non-uniform heating scenarios, CFD reveals enhanced mixing and due to secondary flows, with boundary layers thickening near heated walls. The 1D limitation is most pronounced in short ducts (length-to-diameter < 20), where entrance effects and viscous invalidate uniformity, necessitating multi-dimensional simulations for accurate recovery and distributions. In hypersonic applications, Rayleigh flow extensions incorporate high-temperature effects like molecular dissociation and ionization, treating the gas as chemically reacting with variable composition. For dissociating flows, the equation of state becomes p = \rho R T (1 + \alpha) (where \alpha is the dissociation degree), and enthalpy includes dissociation energy h = c_p T + \alpha D, leading to a modified sound speed based on the effective specific heat ratio. These require integrating species continuity equations alongside momentum and energy, often using equilibrium assumptions for rapid reactions, which shift the Rayleigh line toward lower Mach numbers and increase thermal choking sensitivity to heat addition. In hypersonic ducts, such as scramjet combustors, dissociation reduces effective \gamma from 1.4 to below 1.2 at temperatures over 2500 K. Lighthill's dynamics of dissociating gases provide the foundational framework for these non-equilibrium effects.

Applications

Combustion Chambers and Propulsion

In turbojet and ramjet engines, the combustor is typically modeled as a Rayleigh flow , where is added through in a constant-area duct without significant effects. This model captures the flow entering the combustor at numbers of approximately 0.2 to 0.4 to prevent premature , allowing for controlled release that raises the to 1500–2000 K. Afterburners in engines extend this Rayleigh flow analysis to post-turbine sections, where additional ignites in the exhaust stream, often at or mildly supersonic conditions, to boost during high-demand maneuvers. In scramjets, designed for at and above, Rayleigh flow principles predict the effects of supersonic , including the formation of trains due to heat addition and thermal limits that constrain maximum heat input before occurs. Design of these combustors relies on Rayleigh line diagrams in the temperature-entropy plane to determine the maximum allowable heat addition without reaching the choking condition at Mach 1, ensuring stable operation and avoiding upstream flow disruptions. Engineers often intersect lines with Fanno lines (accounting for in downstream nozzles) to match combustor exit conditions with or requirements, optimizing overall engine performance. In modern applications, such as high-speed civil concepts and vehicles, Rayleigh flow models inform the analysis of variable specific heat ratio (γ) effects in hypersonic propulsion, with numerical simulations validating designs for combustors under real-gas conditions. For instance, one-dimensional Rayleigh-based simulations predict performance deficits from addition in scramjet inlets, aiding the development of reusable hypersonic systems.

Heat Exchangers and Industrial Uses

Rayleigh flow provides a fundamental model for analyzing in gas-to-gas heat exchangers, particularly in constant-area configurations such as uninsulated or regenerators, where frictional effects are minimal and addition or removal governs the behavior. This frictionless, one-dimensional approach predicts axial profiles by linking input to changes in , , and , thereby establishing limits for exchange without violating constraints. For instance, in subsonic air entering at 220 and 0.41, addition can elevate the exit while increasing the , illustrating the model's utility in sizing exchanger ducts for optimal performance. In industrial furnaces and boilers, Rayleigh flow analysis evaluates gas dynamics in constant-area combustion zones, where heat release from must be managed to prevent thermal choking in downstream hot gas or lines. By applying , , and , engineers can determine permissible loads that avoid conditions at the duct exit, ensuring safe operation and maximizing throughput in high-temperature processes like reheating or power generation. This approach highlights the trade-off between rates and flow acceleration, guiding the design of liners to sustain velocities under varying inputs. Emerging applications extend Rayleigh flow principles to microscale in devices, where the model predicts changes and limits in micronozzles subjected to wall heating, influencing and in compact actuators. In contemporary hypersonic research as of 2025, Rayleigh flow models are integrated with (CFD) for designing advanced combustors in reusable launch vehicles, accounting for non-ideal gas effects and variable geometry to enhance performance in sustained .

References

  1. [1]
    [PDF] Rayleigh Flow – compressible flow with heat transfer
    • Introduce Rayleigh flow: flow in a duct with heat addition (or removal) but no friction. • Discuss Rayleigh flow qualitatively and quantitatively. • Do an ...
  2. [2]
    [PDF] AME 436 - Paul D. Ronney
    AME 436 - Spring 2019 - Lecture 12 - 1D Compressible Flow. Heat addition at const. Area (Rayleigh flow). ➢ What if neither the initial state (1) nor final ...
  3. [3]
    [PDF] AA210A Fundamentals of Compressible Flow - Stanford University
    Nov 1, 2020 · The Rayleigh Line refers to the H (or T) versus entropy line determined by eliminating the Mach number between temperature and entropy. s−s∗. C".
  4. [4]
    [PDF] Modeling of Compressible Flow with Friction and Heat Transfer ...
    On the other hand, Raleigh flow provides analytical solution of frictionless compressible flow with heat transfer where incoming subsonic flow can be choked at ...
  5. [5]
    None
    Below is a merged response that consolidates all the information from the provided summaries into a single, comprehensive overview of Rayleigh Flow's definition and historical context. To maximize detail and clarity, I’ve organized the information into a table in CSV format, followed by a narrative summary that integrates the key points. This approach ensures all details are retained while maintaining readability.
  6. [6]
    [PDF] The Dynamics And Thermodynamics Of Compressible Fluid Flow
    Beginning wire a brief review of the foundation concepts of fluid dynamics and thermodynamics and an introduction to the concepts of compressible flow, this ...
  7. [7]
  8. [8]
    [PDF] LECTURE NOTES ON GAS DYNAMICS - University of Notre Dame
    4.6 Flow with heat transfer–Rayleigh flow. Flow with heat transfer is commonly known as Rayleigh flow. To isolate the effect of heat transfer, the following ...
  9. [9]
    [PDF] Fanno Flow and Rayleigh Flow Calculations for Bleed Flow through ...
    Rayleigh flow refers to adiabatic flow through a constant area duct where the effect of heat addition or rejection is considered. Compressibility effects often ...
  10. [10]
    [PDF] THERMODYNAMICS OF GAS FLOW - DTIC
    The following equations can be written for simple T flow: ££ = ä-^ ^ dT p J* T (state). dT ^ 2 _ d>£ dT0. "^T T. jC,jfa^' «2 Ä. T (total temp definition). ^?-f ...
  11. [11]
    Flows with heat transfer (Rayleigh flows) — Gas Dynamics notes
    Rayleigh flow is the specific case of frictionless flow in a constant-area duct, with heat transfer. Rayleigh flow applies to constant-area heat exchangers and ...
  12. [12]
    [PDF] Rayleigh Flow
    Rayleigh Flow -1. School of Aerospace Engineering. Copyright © 2001-2002 by ... • Method 1: direct solution of integrated equations (X.28-34). – requires ...Missing: conservation | Show results with:conservation
  13. [13]
    Thermodynamic bounds for existence of normal shock in ... - SciELO
    Rayleigh described the geometric loci of constant linear momentum in the Mollier diagram as a curve that became known as the Rayleigh curve. Fanno (Shapiro 1953) ...
  14. [14]
    Fanno flow - Wikipedia
    In a nozzle, the converging or diverging area is modeled with isentropic flow, while the constant area section afterwards is modeled with Fanno flow.Theory · Additional Fanno flow relations · Applications
  15. [15]
    [PDF] Equations, Tables, and Charts for Compressible Flow - DTIC
    If the gas is calorically perfect—that is, the specific heats ... flow properties is possible if the gas exhibits both thermal and caloric imperfections.
  16. [16]
  17. [17]
    [PDF] Combined fanno-line, rayleigh-line heat transfer and friction ... - CORE
    Heat is transferred continuously and completely hut only transver- sely throughout the passage. The conventional variables ( pressure, temperature, density, and ...
  18. [18]
    [PDF] thermal-flow code for modeling gas dynamics and heat transfer in ...
    65%. Table 2. The Mach Number Predicted by. SFLOW for the Rayleigh Flow Test Case. 20 cells. 200 cells. 1,000 cells. M,,,. 0.1954. 0.1987. 0.1991 error in M,,.Missing: multi- non-
  19. [19]
    [PDF] Generalized Dissociating Gas Flow - Scholarly Commons
    This flow is analogous to the classical one-dimensional Rayleigh flow CJlow with heat addition), Fanno flow (flow with fric- tion), and shock wave flow. * ...Missing: impact | Show results with:impact
  20. [20]
    [PDF] Supersonic Free-Jet Combustion in a Ramjet Burner
    Figure 1(a) shows a cross-sectional view of a generic scramjet engine. Processes that govern scramjet efficiency are inlet momentum losses, Rayleigh losses due ...
  21. [21]
    [PDF] Rayleigh & Fanno Flows
    This curve is known as the Rayleigh Curve. • Heat addition always drives the Mach numbers toward 1. It decelerates a supersonic flow and accelerates a subsonic ...
  22. [22]
    [PDF] Control-Relevant Modeling, Analysis, and Design for Scramjet ...
    The simple 1D Rayleigh flow engine model discussed within7, 19, 26, 29 will be used in the current paper. Hypersonic Vehicle Control Issues. Within this ...
  23. [23]
    Heat addition with variable area: Methodology for preliminary design ...
    Apr 4, 2023 · The one-dimensional Rayleigh flow theory can be applied to estimate the behavior of thermodynamic properties and velocities when the combustion ...<|control11|><|separator|>
  24. [24]
    Demonstration and Analysis of Filtered Rayleigh Scattering Flow ...
    Aug 1, 1996 · Filtered Rayleigh Scattering (FRS) is a diagnostic technique which measures velocity, temperature, and pressure by determining Doppler shift ...
  25. [25]
    Advancements on the use of Filtered Rayleigh Scattering (FRS) with ...
    This study demonstrates the further development of the FRS technique, with the ultimate goal of inlet flow distortion measurements for in-flight environments.
  26. [26]
    Analysis of a Rayleigh scattering measurement system in a ...
    Aug 22, 2012 · Analysis of a Rayleigh scattering measurement system in a hypersonic wind tunnel ; Charles Tyler. USAF, Wright Lab., Wright-Patterson AFB, OH.Missing: diagnostics non-