Fact-checked by Grok 2 weeks ago

Transonic

Transonic aerodynamics describes the behavior of fluid flow around an object, such as an aircraft, when the object's speed is close to the speed of sound, typically in the Mach number range of approximately 0.8 to 1.2. In this regime, the airflow exhibits a complex mixture of subsonic and supersonic regions, with local accelerations over surfaces leading to supersonic pockets terminated by shock waves. This mixed flow pattern introduces nonlinear effects that significantly alter aerodynamic forces, including a sharp rise in drag known as the transonic drag divergence. The transonic regime poses unique challenges in aircraft design and performance, as the formation of shock waves can cause , buffet, and control difficulties, necessitating specialized shaping techniques like the to minimize . Most modern commercial jet transports operate their cruise in the transonic , typically at numbers around 0.8, to balance and speed while avoiding the instabilities near Mach 1. Advances in and testing have been crucial for predicting and mitigating these phenomena, enabling safer and more efficient high-subsonic flight.

Fundamentals

Definition and Speed Range

The transonic regime describes the flight condition in which the airflow over an or other body begins to exhibit a mix of and supersonic characteristics locally, even as the freestream flow remains near the . This transitional phase typically occurs at numbers ranging from approximately 0.8 to 1.2, marking the onset of significant effects as the aircraft approaches speeds. The Mach number (M), a dimensionless quantity central to this regime, is defined as the ratio of the local flow speed (v) to the speed of sound (a) in the surrounding medium: M = v / a. The speed of sound itself depends primarily on the temperature of the air, which decreases with altitude in the troposphere, leading to a corresponding variation in a—typically around 340 m/s at sea level but dropping to about 295 m/s at 11 km altitude under standard conditions. As a result, the true airspeed required to reach a given Mach number increases at higher altitudes, influencing the practical boundaries of transonic flight. This regime is distinguished from purely flight (where M < 0.8 and airflow remains entirely below sonic speeds) and supersonic flight (where M > 1.2 and the is predominantly faster than sound). A key threshold within the transonic range is the (M_\text{crit}), defined as the lowest freestream Mach number at which sonic conditions (M = 1) are first reached at any point on the surface, such as over a , initiating local supersonic pockets. The precise limits of the transonic range can vary based on factors including the aircraft's geometry (e.g., wing thickness and sweep), flight altitude, and atmospheric conditions like and , which alter the onset of these local transitions.

Key Flow Characteristics

Transonic flow is distinguished by the coexistence of and supersonic regions within a predominantly flowfield, typically at Mach numbers ranging from approximately 0.8 to 1.2. This mixed regime arises as the airflow accelerates over curved surfaces, such as airfoils or wings, forming localized supersonic pockets that are abruptly terminated by embedded shock waves, which decelerate the flow back to speeds. These embedded shocks introduce significant complexities, as they propagate and strengthen with increasing Mach number, altering pressure distributions and flow patterns across the body. The behavior of transonic flow is inherently nonlinear, stemming from the fundamental role of the as a barrier that governs effects. As the flow approaches and crosses Mach 1 locally—often due to over the curvature of lifting surfaces—the governing equations shift from elliptic (subsonic) to hyperbolic (supersonic) characteristics, leading to discontinuous solutions and sensitivity to small perturbations in geometry or conditions. This nonlinearity manifests in rapid changes in flow properties, such as velocity and pressure, particularly near critical points where local Mach numbers reach unity, complicating predictive modeling and requiring specialized numerical approaches. Viscosity and development play critical roles in transonic regimes, influencing the interaction between the outer and the near-surface layer where shear stresses dominate. The thickens downstream of embedded shocks due to adverse gradients, increasing the risk of , especially on at higher angles of attack or numbers, as seen in cases like the RAE 2822 airfoil where separation bubbles form near the shock foot. This separation can lead to unsteady phenomena and reduced aerodynamic performance, necessitating accurate viscous modeling to capture these effects. In comparing inviscid and viscous effects, inviscid models, such as full potential equations, adequately approximate the outer flow but overestimate strengths and positions while neglecting growth and separation risks inherent to transonic conditions. Viscous simulations, incorporating Navier-Stokes equations, reveal more realistic interactions, such as displacement that displaces shocks forward and promotes separation under strong adverse gradients, highlighting the limitations of inviscid assumptions for practical design. These differences underscore the need for coupled viscous-inviscid approaches to achieve reliable predictions in transonic flows.

Aerodynamic Principles

Compressibility Effects

In transonic flows, which occur near the , air can no longer be treated as incompressible, as variations become significant due to the finite speed at which pressure disturbances propagate through the medium. This transition marks a departure from low-speed , where is assumed constant, to regimes where alters flow behavior fundamentally. Under isentropic conditions, typical for inviscid compressible flows without shocks, the ρ varies with p according to the relation ρ ∝ p^{1/γ}, where γ is the specific heat ratio, equal to 1.4 for diatomic air at standard conditions. This proportionality arises from the conservation of in reversible adiabatic processes, leading to local increases in density and pressure as accelerates toward speeds. The , defined as a = √(γ p / ρ), plays a critical role in , limiting the upstream influence of downstream disturbances in accelerating flows and causing about the to travel only at finite speeds relative to the . In transonic conditions, this results in regions where local numbers exceed 1, even if the freestream is , amplifying effects. These density changes impact aerodynamic coefficients, particularly for thin airfoils, where the Prandtl-Glauert transformation provides a correction for compressible effects on incompressible solutions. The transformation yields the pressure coefficient in compressible flow as C_{p, \text{compressible}} \approx \frac{C_{p, \text{incompressible}}}{\sqrt{1 - M^2}}, where M is the freestream Mach number, effectively scaling pressures upward as M approaches 1. Similar corrections apply to lift coefficients, predicting increases until the transonic regime introduces nonlinearities. This linear approximation, derived from potential flow theory, highlights how compressibility enhances lift but foreshadows limitations near sonic speeds. The recognition of these compressibility effects emerged in the 1930s through experiments at the (NACA), where researchers like Lyman J. Briggs and Hugh L. Dryden observed sudden changes in lift and drag at high speeds, termed "compressibility burble." These findings, building on earlier theoretical work by and Hermann Glauert in the , established the need to account for air's elastic properties in high-speed design.

Shock Waves and Drag Rise

In transonic flows, where the freestream approaches unity, waves form as a consequence of local supersonic regions decelerating to speeds, leading to abrupt changes in properties. Normal waves occur perpendicular to the direction, typically across minimum area sections like airfoil throats, while oblique waves arise at inclined surfaces such as leading edges or compression ramps, deflecting the and producing a weaker pressure jump. These shocks emerge during the transition at M ≈ 1 due to effects that accelerate over curved surfaces beyond the . The jump conditions across these shocks are governed by the Rankine-Hugoniot relations, derived from , , and . For a normal shock, the static pressure ratio is given by \frac{p_2}{p_1} = 1 + \frac{2\gamma}{\gamma + 1} (M_1^2 - 1), where p_2 and p_1 are the post- and pre-shock pressures, \gamma is the specific heat ratio (typically 1.4 for air), and M_1 is the upstream normal to the shock. This relation quantifies the sudden pressure rise, with entropy increasing across the discontinuity, marking the irreversible nature of the shock. For oblique shocks, similar relations apply using the normal component of M_1, resulting in less severe jumps but still significant flow deflection. Transonic drag divergence refers to the rapid increase in the wave C_{D_{wave}} near M = 1, primarily due to the formation of these shocks and subsequent separation. The drag rise is characterized by C_{D_{wave}} \approx 20 (M - M_{crit})^4 for M > M_{crit}, where M_{crit} is the , leading to a sharp escalation as shocks strengthen and induce , thickening the and increasing pressure drag. In typical transonic , wave can constitute around 10% of the total drag at divergence, though it peaks higher in unoptimized designs due to shock- interactions. Wave drag differs from parasitic drag, which arises from skin friction and form effects in incompressible flow, and induced drag, which stems from lift generation via trailing vortices; in transonic regimes, wave drag dominates the rise while parasitic and induced components form a baseline "bucket" in drag polars. Airfoil polars for supercritical sections, for example, exhibit a wide low-drag bucket at moderate lift coefficients, where C_D remains minimized until shock onset elevates C_{D_{wave}}, contrasting with narrower buckets in conventional airfoils prone to early separation. To mitigate this, area ruling distributes the aircraft's cross-sectional area smoothly along the longitudinal axis, reducing shock strength by minimizing local Mach number gradients and delaying drag rise by up to 60% near M = 1.

Historical Development

Early Observations and Discoveries

Early theoretical insights into high-speed aerodynamic challenges emerged in the through the work of , who published his first paper on supersonic flow in 1912, highlighting potential difficulties in airflow at speeds approaching and exceeding the . These warnings laid the groundwork for later empirical validations, emphasizing the nonlinear effects in regimes. In the pre-1940s era, empirical observations of transonic phenomena were advanced by the (NACA) under John Stack's leadership. During the and , Stack's tests at NACA's laboratory revealed a sharp drag rise in airfoils at high subsonic speeds, attributed to the onset of compressibility effects where local supersonic flow regions formed over the wing. These experiments, conducted in facilities like the Variable Density Tunnel, demonstrated that drag coefficients could increase significantly, often doubling, near Mach 0.7-0.8, marking the initial recognition of the transonic drag divergence. The 1940s brought direct flight test evidence of transonic issues during World War II, as high-performance aircraft pushed subsonic limits. The Messerschmitt Me 209, a record-breaking racer achieving speeds over 750 km/h (approximately Mach 0.75 at altitude), experienced handling challenges at high speeds. Similarly, British Supermarine Spitfire evaluations in dives approaching Mach 0.8 revealed high-speed control issues, including aileron reversal due to shock wave formation on the wing, which could reduce control effectiveness. World War II accelerated transonic research as military demands exposed these phenomena more frequently, particularly through propeller-driven aircraft where tip speeds routinely approached sonic velocities. Propeller tips on fighters like the Spitfire and often exceeded 1 locally at high power settings, generating shock waves that increased noise, vibration, and efficiency losses, prompting urgent investigations into mitigation. These wartime experiences, combining data with in-flight anomalies, solidified the empirical foundation for understanding transonic flight challenges.

Evolution in Aircraft Design

Following , aircraft designers addressed transonic drag rise by incorporating swept wings to delay the onset of shock waves. Swept wings reduce the component of freestream velocity normal to the leading edge, thereby lowering the effective on the wing and postponing supercritical flow conditions. The , introduced in 1947, exemplified this adaptation as the first operational swept-wing jet fighter, achieving superior transonic performance during the through its 35-degree wing sweep. In the 1950s, NACA aerodynamicist Richard T. Whitcomb developed the to further minimize transonic by ensuring a smooth distribution of cross-sectional area along the aircraft's length, often resulting in fuselage "waist" designs that integrated the wing and body seamlessly. This principle reduced drag rise by up to 60% near Mach 1, as validated in tests. Applied to the in 1953, the redesign enabled the interceptor to exceed Mach 1, transforming it from a failure to a supersonic success. During the 1960s and 1970s, advanced transonic technology with supercritical airfoils, pioneered by Whitcomb, featuring a flattened upper surface, large leading-edge radius, and to promote isentropic recompression and weaken shock waves. These airfoils raised the () by approximately 0.1 compared to conventional NACA 6-series profiles, allowing efficient cruise at 0.8-0.85 while maintaining low drag and good low-speed lift. Examples include the SC(2)-0710 , which achieved values up to 0.82 at typical lift coefficients. In modern aircraft, transonic design principles continue to evolve, as seen in the 's composite wings, which leverage for optimized sweep and thickness to enhance transonic efficiency and reduce induced . This configuration contributes to a 20% reduction in fuel burn relative to similarly sized predecessors like the .

Theoretical and Mathematical Foundations

The theoretical foundations of transonic flow analysis emerged from the need to address the breakdown of classical linear perturbation theories near 1, where the flow transitions between and supersonic regimes, leading to mixed-type partial differential equations that exhibit both elliptic and behaviors. Linear theories, such as the Prandtl-Glauert transformation for flows or Ackeret's for supersonic flows, fail at exactly M=1 because the parameter β = √(1 - M²) approaches zero, rendering the equations degenerate and unable to capture the nonlinear interactions essential for formation and rise. This limitation necessitated the development of nonlinear approximations to the full equations derived from the inviscid Euler equations under irrotational assumptions. A key advancement was the transonic small disturbance (TSD) theory, which provides a simplified model for weakly nonlinear flows by perturbing the steady, inviscid, compressible Navier-Stokes equations around a uniform , assuming small perturbations in velocity potential φ such that the disturbance velocities are much smaller than the freestream speed. The resulting TSD equation in the takes the form: \frac{\partial^2 \phi}{\partial x^2} + \beta^2 \left( \frac{\partial^2 \phi}{\partial y^2} + \frac{\partial^2 \phi}{\partial z^2} \right) = 0, where β = √(1 - M²) is the Prandtl-Glauert parameter, x is the streamwise direction, and y, z are transverse coordinates; this elliptic equation for flow (β > 0) becomes in local supersonic regions (β imaginary), highlighting the mixed-type nature of transonic flows. However, to accurately model shocks and nonlinear effects, the full TSD includes quadratic terms in the , making it quasi-linear and solvable via type-dependent finite-difference schemes. This framework, pioneered in the mid-1950s, enabled similarity rules and scaling laws for transonic similarity, allowing solutions for arbitrary numbers near 1 by adjusting thickness and angle-of-attack parameters. In transonic flows, particularly around sharp edges or convex corners in local supersonic pockets, Prandtl-Meyer expansion fans arise as isentropic wave structures that smoothly turn the flow and accelerate it, in contrast to abrupt, entropy-increasing shock waves that occur at concave deflections. These centered expansion fans, consisting of infinite simple waves propagating along characteristics, allow the to increase across the fan while preserving total pressure, providing a for flow adjustment at transonic boundaries without the losses associated with shocks; this distinction is crucial for modeling embedded supersonic regions in otherwise flows. The evolution of numerical methods for solving these equations began in the 1950s with the , which exploited the hyperbolic nature of supersonic subregions to march solutions along lines, as applied to transonic problems by researchers like Guderley and Yoshihara for axisymmetric flows. By the , relaxation techniques and finite-difference schemes, such as Murman and Cole's type-dependent differencing for the nonlinear TSD , enabled iterative solutions over mixed grids, treating subsonic regions elliptically and supersonic regions . This progressed in the to full potential solvers using multigrid acceleration and, subsequently, inviscid Euler -based (CFD) methods, exemplified by Jameson's finite-volume schemes with flux limiters for shock capturing, which provided higher fidelity for three-dimensional transonic flows without the small-disturbance assumption. These approaches underscored the shift from analytical perturbations to robust numerical frameworks capable of handling the nonlinearities inherent at M ≈ 1.

Observable Phenomena

Condensation Clouds

In transonic flows around , condensation clouds form due to the rapid pressure drop across shock waves and expansion regions, leading to adiabatic cooling that lowers the local temperature below the of atmospheric . This process causes the vapor to condense into visible droplets, creating transient clouds that highlight regions of local supersonic . For instance, at wingtips, fans accelerate air to supersonic speeds, producing characteristic conical clouds. These clouds typically require high ambient relative humidity near and occur at flight altitudes around 8–12 km (approximately 26,000–39,000 feet), where local cooling favors . The phenomenon is brief, lasting only seconds during dynamic maneuvers such as or high-angle-of-attack turns, as the clouds dissipate once the recovers and the air reheats across the terminating . Notable examples include observations on F/A-18 Hornet jets during high-G turns, where vapor cones envelop the , confirming the presence of transonic shock structures. Similarly, F-16 Fighting Falcon aircraft in level flight at Mach 0.9 exhibit wingtip condensation under humid conditions, visualizing local supersonic pockets over the wings. These sightings, such as during U.S. Navy flight demonstrations, provide direct evidence of transonic flow regimes without . Shock-induced condensation clouds differ from lift-induced aerodynamic condensation, which arises solely from the steady low-pressure regions over lifting surfaces in flight without supersonic acceleration or shocks. The former is tied to transonic compressibility effects, producing structured, conical formations bounded by oblique shocks, whereas the latter forms diffuse, persistent trails independent of transitions.

Flow Visualization Techniques

Flow visualization techniques are essential for experimentally observing transonic flows in controlled environments such as wind tunnels, where density gradients, shock waves, and surface streamlines must be captured to validate aerodynamic models. These methods allow researchers to qualitatively and quantitatively assess flow structures that are otherwise invisible, providing insights into effects and behaviors at numbers near 1.0. Early techniques, developed primarily in the 1930s, focused on optical and surface-based approaches, while modern advancements incorporate diagnostics for precise measurements. Schlieren and shadowgraphy represent foundational optical techniques for visualizing density gradients and shock waves in transonic tests. , reinvented by August Toepler in the mid-19th century but refined for aerodynamic applications in , uses a knife-edge setup to detect variations caused by density changes, rendering shock waves as bright or dark lines against a background. Shadowgraphy, a simpler variant, captures the second spatial derivative of density (Laplacian) by projecting defocused shadows, effectively highlighting abrupt disturbances like shocks without the need for precise alignment. Both methods gained prominence during for studies, with contributions from researchers like H. Schardin, who advanced background-distortion concepts applicable to transonic shock visualization. In transonic tunnels, these techniques reveal lambda shocks and expansion fans, aiding the analysis of wave patterns and flow disturbances at numbers from 0.7 to 1.1. Surface flow visualization methods, such as oil flow and tuft techniques, provide critical data on behavior and separation lines in transonic tests. The colored-oil technique involves applying a of oil, paint, and to the model surface before operation; shear forces streak the oil into patterns that trace streamlines, revealing regions of flow attachment, separation, and reattachment. Developed for high-speed applications in the , this method was employed in facilities like the NACA 16-Foot Transonic to map surface flows on wing-fuselage models, identifying shock-induced separation bubbles. Tuft methods complement this by attaching lightweight filaments (e.g., fluorescent minitufts) to the surface; under transonic conditions, tufts align with local flow direction or flutter to indicate separation, offering real-time qualitative insights during blowdown tests. These approaches are particularly effective for low-Reynolds transonic simulations, where they highlight vortex formation and transition without optical interference. Modern laser-based methods, including (PIV), enable quantitative mapping of two-dimensional velocity fields in transonic flows, surpassing the qualitative limits of earlier techniques. PIV seeds the flow with micron-sized tracer particles illuminated by laser sheets, capturing particle displacements via double-frame imaging to compute velocity vectors and quantify positions through strong gradients in the supersonic pockets. In transonic tests (e.g., NACA 0012 at 0.75), PIV visualizes instantaneous flow fields, identifying locations by velocity discontinuities and adjusting for particle lag near shocks using image shifting. Challenges in transonic applications include ensuring particle fidelity in high-gradient regions, often addressed by combining PIV with for hybrid validation. This method has become standard in contemporary wind tunnels for precise shock-boundary layer interaction studies. Transonic flow visualization relies on specialized wind tunnels designed to replicate flight conditions by scaling Reynolds and numbers. The NACA's 16-Foot Transonic Tunnel, operational by the early at Langley Field, Virginia (initially as the High-Speed Tunnel in 1941 and converted for transonic testing in 1947–48), used a slotted test section with 8–10 longitudinal slots (12% open-area ratio) to minimize wall interference for models like slender bodies and wing-fuselage combinations, enabling schlieren-based observations and oil flow surface mapping that matched free-air data with minimal corrections (e.g., Δp/q ≈ 0.03). Such tunnels were pivotal in transonic research, supporting developments like the aircraft by providing interference-free visualizations up to 1.1.

Applications Beyond Aviation

Transonic Flows in Astrophysics

In stellar winds, transonic flows describe the acceleration of from speeds near the stellar surface to supersonic velocities at large distances, a process essential for understanding mass loss in stars like . The seminal model, introduced in 1958, provides the theoretical framework for this transition in spherically symmetric, isothermal winds driven by thermal pressure against gravity. In this model, the flow reaches the sonic point—where the speed equals the local sound speed—at a critical radius r_s \approx \frac{GM}{c_s^2}, with G the , M the , and c_s the isothermal sound speed; beyond this radius, the flow expands supersonically, carrying away and . This transonic structure resolves the apparent of a static by predicting a continuous outflow, validated by in-situ measurements from like in 1962. Transonic flows also govern accretion processes in astrophysical environments, particularly the infall of gas onto compact objects such as black holes within accretion disks. The model, formulated in 1952, analyzes steady, spherically symmetric inflow from a uniform medium at rest at , where the transonic transition occurs at the Bondi radius r_B = \frac{GM}{c_s^2}, analogous to the sonic radius in winds but for inward motion. At this point, overcomes thermal pressure, accelerating the flow to supersonic speeds closer to the accretor; the model predicts the accretion rate \dot{M} \propto \frac{M^2 c_s^3}{G^3}, setting a baseline for more complex disk geometries where and modify the transonic behavior. This framework applies to phenomena like the fueling of supermassive black holes in galactic centers. Observational signatures of transonic flows in include shock structures in radio jets emanating from active galactic nuclei (AGN), where relativistic outflows interact with the , producing bright emission knots indicative of transonic transitions and internal shocks. These shocks, observed via (VLBI) at radio wavelengths, arise from pressure mismatches across the jet boundary or velocity gradients within the flow, mirroring aerodynamic shocks but scaled to lengths and relativistic speeds. For instance, in sources like M87, such structures reveal recollimation shocks where the jet narrows and expands through sonic points, enhancing and particle acceleration. Astrophysical transonic flows differ fundamentally from terrestrial aerodynamic cases due to relativistic effects and the conducting nature of the involved. Near black holes, alters the metric, shifting sonic points and introducing that can create multiple critical surfaces in rotating flows. Moreover, these environments feature highly magnetized, conductive s where magnetic fields enforce frozen-in flux, leading to magnetohydrodynamic (MHD) instabilities and enhanced dissipation at transonic boundaries, unlike the neutral gases in .

Industrial and Engineering Contexts

In industrial applications, transonic flows are prevalent in the stages of jet engines, where tips often experience local numbers exceeding 1, leading to shock- interactions that can precipitate and . These interactions occur as the relative accelerates over the suction surface, forming passage shocks that impinge on the , increasing losses and reducing efficiency, particularly in high-bypass engines designed for . For instance, in modern turbofans like the GE90 or CFM56 series, tip speeds approach 400 m/s, pushing the flow into the transonic regime and amplifying unsteady aerodynamic loading. Beyond aviation compressors, transonic effects influence , where bullets or shells encounter a sharp rise upon decelerating through the transonic range, typically approximately 270–410 m/s ( 0.8–1.2) at standard conditions (15°C), which alters and reduces . This drag divergence stems from the formation of shock waves around the as effects intensify, causing yaw and tumbling if the bullet's factor drops below critical thresholds. A representative example is the cartridge, with a of approximately 900 m/s, which experiences significant increase and potential destabilization around 300-400 meters downrange, impacting long-range accuracy in applications. In steam turbines, nozzle flows frequently reach sonic speeds due to rapid expansion, resulting in transonic conditions that are analyzed using airfoil-like modeling to predict shock formation and loss generation. The low-pressure stages of large power plants, such as those in nuclear or coal-fired facilities, feature long blades where tip sections operate transonically, with steam velocities approaching Mach 1 and inducing similar boundary layer disruptions as in gas turbines. Experimental cascades have shown that these flows incur higher losses in the transonic regime compared to subsonic, necessitating precise nozzle profiling to minimize entropy rise. To mitigate these transonic challenges in industrial designs, techniques such as blade twisting and porous materials are employed to manage strength and behavior. Forward-swept or twisted profiles reduce shock- interaction intensity by altering the incidence and delaying separation, improving stall margin in compressors by up to 10-15% in tested configurations. Porous surfaces or bleed slots, often integrated into the casing or walls, facilitate suction or injection, attenuating waves and preventing separation; for example, porous treatments in transonic cascades have demonstrated loss reductions of 20-30% through controlled extraction. These methods draw from the same rise principles observed in external but are adapted for internal, rotating machinery to enhance and durability.

References

  1. [1]
    [PDF] 7. Transonic Aerodynamics of Airfoils and Wings
    Mar 10, 2006 · Transonic flow occurs when there is mixed sub- and supersonic local flow in the same flowfield (typically with freestream Mach numbers from M = ...
  2. [2]
    Mach Number
    As the speed of the object approaches the speed of sound, the flight Mach number is nearly equal to one, M = 1, and the flow is said to be transonic. At ...<|control11|><|separator|>
  3. [3]
    [PDF] advanced transonic aerodynamic technology
    It is generally recognized that the transonic speed range is that in which the flow about an aerodynamic configuration is an interacting mixture of sub- sonic ...
  4. [4]
    [PDF] Transonic Unsteady Aerodynamic and Aeroelastic Calculations ...
    For example, large conrmercial aircraft cruise most effectively in the transonic flight regime and computational fluid dynamics (CFD) provides a new tool, which.Missing: aviation | Show results with:aviation
  5. [5]
    Critical Mach Number - an overview | ScienceDirect Topics
    Transonic flight conditions are assumed when the free stream Mach number lies in the range 0.8<M0<1.2. The greatest degree of aerodynamic unpredictability is ...
  6. [6]
    Transonic Flow - Centennial of Flight
    Additionally, transonic flow pertains to the range of speeds in which flow patterns change from subsonic to supersonic or vice versa, about Mach 0.8 to 1.2.
  7. [7]
    Speed of Sound
    The speed of sound depends on the temperature, and temperature changes with altitude in a very complex way.
  8. [8]
    Mach Number
    The Mach number depends on the speed of sound in a gas. The speed of sound in air depends on the temperature which, in turn, depends on the altitude in a rather ...
  9. [9]
    Critical Mach Number | SKYbrary Aviation Safety
    In aerodynamics, the critical Mach Number (Mcr or Mcrit) of an aircraft is the lowest Mach number at which the airflow over any part of the aircraft reaches the ...
  10. [10]
    [PDF] some recent developments in numerical methods for t~nso~c flows
    (1) The equations of gas dynamics are nonlinear. (2) Solutions in the transonic range will ordinarily be discontinuous. Thus we may expect to find shock waves.
  11. [11]
    [PDF] Analysis of Viscous Transonic Flow Over Airfoil Sections
    The boundary layer i s the arclength of the grid line measured from the airfoil surface to the node identified as having the velocity, Ue.
  12. [12]
    Investigations related to the inviscid-viscous interaction in transonic ...
    Jan 1, 1977 · Inviscid small-disturbance theory has been shown to predict three-dimensional transonic flows about finite wings reasonably well as long as ...
  13. [13]
    Isentropic Flow Equations
    We can determine the value of the constant by defining total conditions to be the pressure and density when the flow is brought to rest isentropically. The "t" ...Missing: 1.4 | Show results with:1.4
  14. [14]
    The effect of compressibility on the lift of an aerofoil - Journals
    An attempt has now been made to estimate theoretically the effect of compressibility on the lift of an aerofoil in two-dimensional motion.
  15. [15]
    Research in Supersonic Flight and the Breaking of the Sound Barrier
    He concluded that the successful development of such transonic wind tunnels was a problem of Herculean proportions, and was far into the future. In order to ...<|control11|><|separator|>
  16. [16]
    Oblique Shocks Interactive Simulator | Glenn Research Center | NASA
    ### Summary of Shock Waves in Transonic Flows
  17. [17]
    [PDF] Equations, Tables and Charts for Compressible Flow
    Many relations for normal shock waves are conveniently expressed in terms of either upstream Mach number M, or the static-pressure ratio across the shock έ=p2/ ...
  18. [18]
    [PDF] Empirically Based - Transonic Aircraft Drag Buildup Technique
    contribution is in the order of 10 percent of total drag. In evaluating the ... be combined to define a systematic procedure for the total drag buildup of.
  19. [19]
    The Whitcomb Area Rule: NACA Aerodynamics Research ... - NASA
    The area rule revolutionized how engineers looked at high-speed drag and impacted the design of virtually every transonic and supersonic aircraft ever built.Missing: mitigation | Show results with:mitigation
  20. [20]
    [PDF] THEODORE VON KARMAN, distinguished aeronautical engi
    Von Karman early acquired an interest in the theory of fluid flow at high subsonic, transonic, and supersonic speeds. His first published paper on ...
  21. [21]
    The Transonic Wind Tunnel and the NACA Technical Culture - NASA
    Since a propeller is an airfoil, a complex, precise aerodynamic shape like a wing, this made transonic aerodynamics a practical aeronautical research issue, ...
  22. [22]
  23. [23]
    Messerschmitt Me 209 - Plane-Encyclopedia
    Feb 19, 2023 · Some of these included the aircraft's tendency to abruptly dive in mid-flight, the controls being heavy and hard to work with either in the air ...
  24. [24]
    [PDF] The Aerodynamics of the Spitfire - Royal Aeronautical Society
    The Spitfire's aerodynamics include a semi-elliptical wing, thinner aerofoil for lower drag, and high subsonic Mach number performance. Drag comparison with ...Missing: buzz | Show results with:buzz
  25. [25]
    [PDF] 19740023386 Progress In Aircraft Design Since 1903
    NORTH AMERICAN F-86F "SABRE". NORTH AMERICAN F-86F: The "Sabre" was blooded in combat over Korea. This country's first swept-wing jet fighter will be ...
  26. [26]
    [PDF] NASA Supercritical Airfoils
    This dis- tinctive airfoil shape, based on the concept of local supersonic flow with isentropic recompression, was characterized by a large leading-edge radius, ...Missing: M_crit | Show results with:M_crit
  27. [27]
    787 Dreamliner By Design
    ### Summary of Boeing 787 Composite Wings and Fuel Efficiency
  28. [28]
    Review of unsteady transonic aerodynamics: Theory and applications
    The author dedicates this paper to the memory of Professor Julian Cole, whose pioneering research on transonic flow and expertly taught courses on perturbation ...Missing: origin | Show results with:origin
  29. [29]
    [PDF] Transonic flow computations using nonlinear potential methods
    This report discusses transonic flow computations using nonlinear potential methods, including the full potential equation, and transonic small disturbance ( ...Missing: barrier | Show results with:barrier
  30. [30]
    Prandtl-Meyer Angle - NASA Glenn Research Center
    Through an expansion fan, the Mach number increases, the static pressure decreases and the total pressure remains constant. Expansion fans are isentropic. The ...
  31. [31]
    [PDF] AN HISTORICAL PERSPECTIVE - Aerospace Computing Lab
    Multigrid was applied to the transonic small-disturbance equation by South & Brandt [57], and to the full potential equation by. Jameson [21]. Just as ...Missing: derivation | Show results with:derivation
  32. [32]
    [PDF] On the possible origin of a vapour cone occurring during the ...
    Such cloud is a product of an aircraft travelling at transonic speed in moist and warm air. The air which is in the relative motion with respect to an aircraft ...Missing: mechanism | Show results with:mechanism
  33. [33]
    Aerodynamic Contrails: Phenomenology and Flow Physics in
    Feb 1, 2009 · Condensation in transonic flows (i.e., subsonic flow becomes supersonic over the wing and other upward curved surfaces, such as the pilot ...
  34. [34]
    [PDF] A review of recent developments in schlieren and shadowgraph ...
    Feb 15, 2017 · Schlieren and shadowgraph techniques are used for imaging and measuring phenomena in transparent media. Recent developments include digital ...
  35. [35]
    [PDF] Transonic Wind Tunnel Testing - DTIC
    Transonic wind tunnel testing is difficult because the flow changes from subsonic to mixed-type, requiring experiments for aerodynamic information.
  36. [36]
    [PDF] Flow-Visualization Techniques Used at High Speed by ...
    The colored-oil technique is used for observ- ing surface flow features and was applied by the Con- figuration Aerodynamics. Wind-Tunnel-Test. Team in a number ...
  37. [37]
    [PDF] Feasibility of Particle Image Velocimetry (PIV) in High Speed Flow ...
    Section 2 discusses the feasibility of PIV in supersonic flow tunnels, section 3 about transonic flows, section 4 regarding shock tubes and shock tunnels,.
  38. [38]
    The Transonic Wind Tunnel and the NACA Technical Culture - NASA
    The Variable-Density Tunnel, or VDT, could, with fairly good success for the time, counteract scale effects—the skewing of test data inherent in testing scale ...
  39. [39]
    Dynamics of the Interplanetary Gas and Magnetic Fields. - ADS
    Parker, E. N.. Abstract. We consider the dynamical consequences of Biermann's suggestion that gas is often streaming outward in all directions from the sun with ...
  40. [40]
    On Spherically Symmetrical Accretion - Oxford Academic
    The special accretion problem is investigated in which the motion is steady and spherically symmetrical, the gas being at rest at infinity.Missing: original | Show results with:original
  41. [41]
    Shocks in relativistic transverse stratified jets
    By establishing a relation between the transverse stratification of the jets, the internal shock properties, and the multiple observed AGN jet morphologies and ...Missing: transonic | Show results with:transonic
  42. [42]
    General relativistic hydrodynamic simulations of perturbed transonic ...
    We performed three-dimensional general relativistic hydrodynamic simulations of rotating transonic flows with velocity perturbations injected from a spherical ...
  43. [43]
    Numerical Simulation of Radiatively Driven Transonic Relativistic Jets
    We show that starting from a very low injection velocity at the base, jets can be accelerated to relativistic terminal speeds when traveling through the ...
  44. [44]
    Flow in Transonic Compressors - Aerospace Research Central
    Usually the flow is diffused to subsonic velocities in the flow passages formed by the rotating blades, generating part of the desired pressure rise. A further.
  45. [45]
    The Present Challenge of Transonic Compressor Blade Design
    Aug 6, 2025 · In addition, the interaction of these shock waves with the blade boundary layer inherently leads to a very unsteady flow behavior. The aim of ...
  46. [46]
    Passage shock wave/boundary layer interaction control for transonic ...
    Jul 5, 2021 · The investigated conclusion implies a potential promise for future practical applications of SCB in both transonic and supersonic compressors.
  47. [47]
    The Tuning of a CFD Model for External Ballistics, Followed ... - MDPI
    This contribution focuses on the analysis of drag force acting on a .223 rem caliber projectile in both subsonic and supersonic regimes.
  48. [48]
    External Ballistics Summary | Close Focus Research
    May 3, 2009 · Some of the Lapua-provided drag coefficient data shows drastic increases in the measured drag around or below the Mach 1 flight velocity region.<|control11|><|separator|>
  49. [49]
    Practical Thoughts About Transonic Bullet Stability and Accuracy
    Sep 22, 2014 · What we observed was a rise in BC at extreme range, and round bullet holes. This indicates stability of the bullet traversing the transonic zone ...
  50. [50]
    Transonic and Low Supersonic Flow Losses of Two Steam Turbine ...
    Mar 11, 2005 · A linear cascade experiment was conducted to investigate transonic and low supersonic flow losses of two nozzle blades for the steam ...Missing: sonic modeling
  51. [51]
    [PDF] Investigation-of-transonic-and-supersonic-flow-in-the-rotor-tip ...
    ABSTRACT. Design of a new generation of the large power steam turbines leads to very long last rotor blades with tip sections operating often in supersonic ...
  52. [52]
    [PDF] Experimental Tests of Transonic Nozzles in Linear Cascades ...
    The results presented in the paper refer to a steam turbine nozzle cascade (shown in Fig 2) ... Statistical Approach to the Experimental. Evaluation of Transonic ...
  53. [53]
    Blade-end treatment to improve the performance of axial compressors
    Benini and Biollo [32] showed that the use of swept blades effectively reduced shock losses and improved the aerodynamic behavior of the transonic rotors.
  54. [54]
    Effect of Rotor Blade Twist on Axial Transonic Compressor ...
    Increasing the pressure in gas turbine cycle is done by Compressor which is one of the most important components of the Gas Turbines.
  55. [55]
    Effects of Number of Bleed Holes on Shock-Wave/Boundary-Layer ...
    Jan 10, 2024 · An extensive numerical investigation is conducted to characterize the flow separation control in a transonic compressor cascade with a porous bleed.Missing: materials | Show results with:materials
  56. [56]
    Experimental Identification of a New Secondary Wave Pattern in ...
    The current paper refers to experimental findings on a transonic linear cascade specifically designed to mitigate shock waves using porous walls on the blades.