Fact-checked by Grok 2 weeks ago

Skin friction drag

Skin friction drag, a primary component of aerodynamic , arises from the viscous interaction between a —such as air—and of a moving through it, resulting in stresses that oppose the motion. This is fundamentally tied to the , a thin region near where the velocity transitions from zero at the wall (due to the ) to the free-stream velocity, with viscous effects dominating within this layer. In practical terms, it manifests as aerodynamic , where air molecules adhere to and are slowed by , creating that scales with factors like , , and flow speed. For and other vehicles, skin friction constitutes a significant portion of total , often around half in high-speed flight, influencing and design choices such as surface polishing or control. The 's nature—laminar or —profoundly affects the magnitude of skin friction . In a , flow is smooth and orderly, producing lower stresses and thus reduced drag, though it is prone to separation under adverse gradients. Conversely, a involves eddies that increase momentum transfer to the surface, elevating shear stresses and drag by factors of several times compared to laminar flow, but it resists separation better, aiding generation on wings. The between these states depends on the , a dimensionless defined as Re = \frac{\rho V L}{\mu} (where \rho is fluid density, V is velocity, L is , and \mu is dynamic ), with higher values promoting turbulence. Mitigating skin friction drag is a key focus in , achieved through strategies like maintaining smooth, clean surfaces to minimize roughness-induced and using materials or coatings that reduce effective interactions. For instance, flush riveting and waxing on skins lower drag by preserving longer along the surface. In broader applications, such as marine hulls or blades, similar principles apply to optimize performance by quantifying drag via skin friction coefficients, often derived from empirical correlations like the Blasius solution for flat plates in . Overall, understanding and reducing skin friction drag enhances vehicle efficiency, reduces energy consumption, and informs models for predicting aerodynamic behavior.

Fundamentals of Skin Friction Drag

Definition and Physical Mechanism

Skin friction drag, also known as frictional drag, represents the component of the total aerodynamic drag on an object moving through a that arises from the tangential stresses exerted by the viscous on the object's surface. This , often denoted as wall shear stress τ_w, is generated primarily within the , a thin region adjacent to the surface where viscous effects are dominant. The forms due to the at the solid surface, where the fluid velocity is zero, creating steep velocity gradients as the flow accelerates to match the free-stream velocity outside this layer; these gradients are the direct cause of the frictional forces. At its core, the physical mechanism of skin friction drag involves the viscous transfer of momentum between adjacent fluid layers parallel to the surface, which opposes the object's motion and dissipates as heat through internal friction. In , molecular governs this momentum exchange via direct collisions between fluid molecules, while in turbulent flow, random fluctuations from turbulent eddies enhance the transfer, effectively behaving as an augmented eddy that amplifies the beyond what molecular alone would produce. This understanding of skin friction drag was first rigorously quantified by in his seminal 1904 boundary layer theory, which provided a framework for analyzing viscous effects near surfaces and enabled practical predictions of drag in aerodynamic applications. In contrast to pressure drag (or form drag), which originates from unbalanced normal pressure forces across the body typically linked to , skin friction drag is exclusively tangential and viscous in nature, often comprising the majority of drag on streamlined shapes.

Boundary Layer Development

The in fluid flow over a originates from the , which dictates that the fluid velocity at the wall is zero, while the free-stream velocity U_\infty remains constant outside this region. This velocity disparity establishes a steep near the wall, where the is maximized, fundamentally shaping the viscous effects central to skin friction. introduced this concept in his seminal work, resolving the limitations of theories by isolating a thin region near the surface where dominates. From the of the surface, the \delta—defined as the from the wall where the reaches 99% of U_\infty—begins at zero and grows progressively with downstream x. This arises as diffuses perpendicular to the surface through viscous , with the rate of growth depending inversely on U_\infty (faster free-stream speeds compress the layer) and directly on the fluid's kinematic viscosity \nu (more viscous fluids allow thicker layers). In typical external flows, such as over a flat plate, \delta scales with the of x, reflecting the diffusive nature of the process, though exact profiles vary with surface and conditions. To quantify the boundary layer's influence on the outer , integral measures like displacement thickness \delta^* and momentum thickness \theta are employed. The displacement thickness \delta^* indicates the outward shift of the streamline due to reduced velocity within the layer, effectively reducing the cross-sectional area available to the external flow. Meanwhile, the thickness \theta captures the deficit in flow caused by the , serving as a key parameter for assessing overall viscous losses. These thicknesses are derived from integrating the velocity profile across \delta, providing conserved quantities useful in approximate analyses without resolving the full profile. Certain flow conditions, particularly adverse pressure gradients where pressure increases in the streamwise direction, can accelerate boundary layer growth by promoting instability and hastening transition to turbulence. In such gradients, the decelerating external flow exacerbates momentum loss near the wall, thickening \delta more rapidly than in zero or favorable pressure gradient cases, as observed in experimental studies of decelerating boundary layers. This enhanced growth underscores the interplay between pressure forces and viscous diffusion in boundary layer evolution. Skin friction drag emerges from the wall driven by these velocity gradients within the developing .

Flow Regimes and Skin Friction

Laminar Flow Characteristics

In over a surface, particles move in smooth, parallel layers with no mixing across streamlines, resulting in an orderly velocity profile that minimizes momentum transfer perpendicular to the flow direction. This layered structure contrasts with more chaotic regimes and leads to lower skin friction drag compared to turbulent flow, as viscous shear occurs primarily within adjacent layers rather than through intense . The stability of laminar boundary layers is governed by linear stability analysis, which identifies Tollmien-Schlichting waves as the primary instabilities initiating the transition to turbulence. These waves arise from small-amplitude disturbances in the flow and grow when the local exceeds critical values, amplifying perturbations that the laminar . The , defined as Re_x = \frac{U_\infty x}{\nu}, where U_\infty is the , x is the distance from the , and \nu is the , determines the maintenance of laminarity; for a flat plate, typically persists up to Re_x \approx 5 \times 10^5. In this regime, the skin friction coefficient decreases with increasing Re_x, reflecting the thinning of the relative to the flow scale. Despite these theoretical characteristics, fully laminar flow is rare in practical high-speed applications due to environmental disturbances that trigger early transition. Surface roughness introduces localized perturbations that amplify instabilities, while freestream turbulence provides external energy to bypass the stable laminar regime, often reducing the effective transition Reynolds number by orders of magnitude. Achieving sustained laminar flow thus requires meticulous control of these factors, such as polished surfaces and low-turbulence wind tunnels, to realize the associated drag benefits.

Turbulent Flow Characteristics

In turbulent boundary layers, the flow exhibits chaotic, irregular motion characterized by intermittent eddies and bursts that originate near the wall and extend outward, facilitating rapid momentum transfer from the freestream to the surface. These coherent structures, as described in Townsend's attached eddy model, enhance mixing across the layer, effectively increasing the momentum diffusivity far beyond molecular viscosity and thereby elevating skin friction drag compared to laminar flows. The turbulent is stratified into distinct regions based on distance from : the viscous sublayer, where viscous forces dominate and the flow remains nearly laminar; the adjacent buffer layer, a transitional zone where turbulent fluctuations begin to compete with viscous effects; and the outer logarithmic layer, where inertial forces prevail and the velocity profile follows a self-similar form. This structure arises from the balance between and production, as outlined in classical analyses. A key feature is the universal velocity profile expressed in wall units, defined as u^+ = \frac{u}{u_\tau} and y^+ = \frac{y u_\tau}{\nu}, where u is the streamwise velocity, y is the wall-normal distance, u_\tau = \sqrt{\tau_w / \rho} is the friction velocity, \tau_w is the wall shear stress, \rho is the fluid density, and \nu is the kinematic viscosity. In the logarithmic region, the law of the wall relates these as u^+ = \frac{1}{\kappa} \ln y^+ + B, with von Kármán constant \kappa \approx 0.41 and additive constant B \approx 5.0, providing a collapse of profiles across different flows when scaled appropriately. This formulation, originally derived by von Kármán, underscores the wall's universal influence on near-wall turbulence. The intense vortical motions in turbulent boundary layers lead to higher rates of energy dissipation through viscous effects, resulting in elevated skin friction coefficients that can be several times larger than in laminar regimes. However, this enhanced mixing can delay on bluff bodies, reducing overall ; for instance, dimples on golf balls promote early transition to , shifting the separation point downstream and lowering the total by up to 50% at typical flight Reynolds numbers.

Transitional Flow Phenomena

Transitional flow in boundary layers represents the dynamic process where the initially breaks down into , characterized by the formation and propagation of turbulent spots amid predominantly laminar regions. This arises as disturbances amplify, leading to localized turbulent patches that expand and coalesce downstream. The significantly influences skin friction drag, as the intermittent turbulent regions enhance momentum transfer compared to pure . Two primary mechanisms drive this transition: instability-driven and bypass. Instability-driven transition occurs through receptivity of the to external disturbances, such as or surface vibrations, which excite Tollmien-Schlichting () waves—viscous instability modes that amplify nonlinearly over a range of frequencies and wavelengths. These waves, initially damped or neutral at low Reynolds numbers, grow exponentially in the unstable regime, eventually leading to the breakdown into turbulent spots. In contrast, transition is triggered by high levels of or other strong external perturbations that overwhelm the process, causing abrupt formation of turbulent spots without significant TS wave amplification. This mode dominates in environments with intensity greater than 1%, such as behind grids or in practical engineering flows like turbine blades. The extent of transition is quantified by the intermittency factor, γ, defined as the of time the at a given point is turbulent, ranging from 0 in fully laminar regions to 1 in fully turbulent ones. In the transitional zone, turbulent spots occupy a probabilistic of the , with γ increasing monotonically downstream as spots grow at an of approximately 10–15 degrees to the and merge. This reflects the nature of spot generation and evolution, making it a key parameter for modeling the gradual shift in structure. Transition typically initiates at a local based on distance from the , Rex, in the range of 5 × 105 to 3 × 106 for a flat plate in low-turbulence conditions, though this critical range varies with flow parameters. Surface lowers the transition Rex by generating disturbances that promote early spot formation. Adverse pressure gradients destabilize the by reducing stability margins, accelerating transition, while favorable gradients enhance stability and delay it. Predicting transitional flow remains challenging due to its acute sensitivity to initial and environmental conditions, including turbulence levels, , and surface imperfections, which can shift the transition location by orders of magnitude in Rex. Linear stability theory captures amplification rates but fails to pinpoint exact onset without accounting for nonlinear receptivity and spot dynamics, complicating reliable forecasting in complex geometries. In this regime, the average skin friction exceeds that of pure but remains below fully turbulent levels, reflecting the partial turbulent character.

Skin Friction Coefficient

Mathematical Definition

The local skin friction coefficient, denoted as c_f, quantifies the at the wall relative to the freestream dynamic pressure and is defined as c_f = \frac{\tau_w}{\frac{1}{2} \rho_\infty U_\infty^2}, where \tau_w is the wall , \rho_\infty is the , and U_\infty is the . The average skin friction C_f over a surface of L in the streamwise direction is the spatial average of the local coefficient, given by C_f = \frac{1}{L} \int_0^L c_f \, dx. This represents the overall frictional contribution across the surface. The total skin friction drag force D_f acting on a body with wetted surface area S is related to the average coefficient through D_f = \frac{1}{2} \rho_\infty U_\infty^2 \, S \, C_f. This expression links the coefficient directly to the net drag due to surface friction. The magnitude of the skin friction coefficient depends primarily on the Re, defined as Re = \rho_\infty U_\infty L / \mu (with \mu as the dynamic viscosity), which governs the balance between inertial and viscous effects, as well as on surface conditions like roughness height relative to the boundary layer thickness. The value of the skin friction coefficient varies with flow regime, generally being lower in compared to turbulent flow.

Laminar Flow Coefficients

In laminar boundary layers over a flat plate, the skin friction arises from the viscous at the wall, characterized by specific coefficients derived from similarity solutions. The local skin friction coefficient, c_{f,x}, which represents the dimensionless wall at a x from the , is given by c_{f,x} = \frac{0.664}{\sqrt{\mathrm{Re}_x}}, where \mathrm{Re}_x = \frac{U_\infty x}{\nu} is the local based on free-stream velocity U_\infty and kinematic \nu. This expression indicates that skin friction decreases with the of the , reflecting the thinning of the downstream. For the average skin friction coefficient over the entire plate of length L, C_f, the value is obtained by integrating the local coefficient: C_f = \frac{1.328}{\sqrt{\mathrm{Re}_L}}, with \mathrm{Re}_L = \frac{U_\infty L}{\nu}. This average quantifies the total due to per unit wetted area and is twice the local value at the trailing edge, consistent with the self-similar growth of the laminar . These coefficients stem from the exact similarity solution for the equations presented by Blasius in 1908. The formulas apply under key assumptions of , zero (as in uniform free-stream conditions over a flat plate), and a surface, where viscous effects dominate without separation or . They hold for local Reynolds numbers \mathrm{Re}_x up to approximately $5 \times 10^5, beyond which to typically occurs, rendering effects or negligible in this regime.

Turbulent Flow Coefficients

In turbulent boundary layers over flat plates, skin friction coefficients are typically derived from empirical correlations that approximate the velocity profile using power-law forms, such as Prandtl's 1/7th power law, which assumes a velocity distribution u/U_\infty = (y/\delta)^{1/7} near the wall. This leads to a local skin friction coefficient expressed as c_{f,x} \approx 0.0576 \, \mathrm{Re}_x^{-1/5}, where \mathrm{Re}_x = U_\infty x / \nu is the local based on distance x from the leading edge. These power-law approximations stem from mixing length theory, providing a simple framework for estimating shear stress in fully developed turbulent flows without pressure gradients. For the average skin friction coefficient over a plate of length L, integration of the local value yields the Schlichting formula C_f \approx 0.074 \, \mathrm{Re}_L^{-1/5} for smooth surfaces, where \mathrm{Re}_L = U_\infty L / \nu. This correlation, developed from experimental data and theoretical integration, offers a practical estimate for overall drag contributions in engineering applications. Surface roughness significantly alters these coefficients by disrupting the near-wall flow structure, requiring adjustments via the equivalent sand-grain roughness height k_s, which represents irregular surfaces as uniform sand grains producing equivalent drag. In the logarithmic region of the turbulent boundary layer, the velocity profile follows the law of the wall: u^+ = (1/\kappa) \ln y^+ + B - \Delta B(k_s^+), where \kappa \approx 0.41 is von Kármán's constant, B \approx 5.0 for smooth walls, and \Delta B is the roughness function dependent on the roughness Reynolds number k_s^+ = u_\tau k_s / \nu. This shifts the skin friction downward in the fully rough regime (k_s^+ > 70), increasing drag compared to smooth cases, as validated by Nikuradse's pipe flow experiments extended to boundary layers. These power-law formulas are applicable for Reynolds numbers $10^6 < \mathrm{Re}_L < 10^9 on smooth plates under zero-pressure-gradient conditions, beyond which logarithmic laws or direct numerical simulations provide higher fidelity, especially for complex geometries where (CFD) resolves three-dimensional effects and . Limitations arise at very high Reynolds numbers or with strong pressure gradients, where the assumptions of layers break down.

Analytical and Computational Methods

Blasius Boundary Layer Solution

The Blasius boundary layer solution provides the exact analytical treatment for the development of a laminar boundary layer over a flat plate in steady, two-dimensional, incompressible flow with zero pressure gradient. The governing equations are the continuity equation, \frac{\partial u}{\partial x} + \frac{\partial v}{\partial y} = 0, and the momentum equation in the x-direction, u \frac{\partial u}{\partial x} + v \frac{\partial u}{\partial y} = \nu \frac{\partial^2 u}{\partial y^2}, where u and v are the velocity components, x is the streamwise coordinate, y is the wall-normal coordinate, and \nu is the kinematic viscosity. To solve these partial differential equations, Blasius introduced a that reduces the problem to an . The similarity variable is defined as \eta = y \sqrt{\frac{U_\infty}{\nu x}}, where U_\infty is the free-stream , and the \psi is \psi = \sqrt{\nu x U_\infty} \, f(\eta), leading to the velocity components u = U_\infty f'(\eta) and v = \frac{1}{2} \sqrt{\frac{\nu U_\infty}{x}} \left( \eta f'(\eta) - f(\eta) \right). Substituting these into the boundary layer equations yields the nonlinear third-order f''' + \frac{1}{2} f f'' = 0, subject to the boundary conditions f(0) = 0, f'(0) = 0, and f'(\infty) = 1. This equation has no closed-form solution and requires numerical integration. The numerical evaluation provides the key value f''(0) \approx 0.332, which determines the velocity gradient at the wall. From this, the wall shear stress is \tau_w = \mu \left( \frac{\partial u}{\partial y} \right)_{y=0} = 0.332 \rho U_\infty^2 / \sqrt{\mathrm{Re}_x}, where \mu is the dynamic viscosity, \rho is the fluid density, and \mathrm{Re}_x = U_\infty x / \nu is the local Reynolds number. The local skin friction coefficient then follows as c_f = \frac{\tau_w}{\frac{1}{2} \rho U_\infty^2} = \frac{0.664}{\sqrt{\mathrm{Re}_x}}. Additionally, the boundary layer thickness, defined at 99% of the free-stream velocity, is approximately \delta \approx 5 x / \sqrt{\mathrm{Re}_x}. These results from the Blasius solution form the foundation for predicting laminar skin friction in low-Reynolds-number flows over flat plates.

Prandtl's Mixing Length Approximations

Prandtl introduced the mixing length in as a semi-empirical approach to model turbulent transfer in shear flows, including boundary layers, by analogizing turbulent eddies to molecules in kinetic theory. In this framework, coherent lumps of displace transversely by a characteristic distance, termed the mixing length l, before exchanging with surrounding due to differences. The resulting Reynolds shear is approximated as -\rho \overline{u'v'} = \rho l^2 \left( \frac{du}{dy} \right)^2, where the ensures the correct sign, leading to a turbulent eddy \nu_t = l^2 \left| \frac{du}{dy} \right|. This allows the turbulent stress to be incorporated into the mean equation as an effective , facilitating analytical approximations for profiles and drag in boundary layers. For wall-bounded turbulent flows, von Kármán extended Prandtl's concept in 1930 by proposing that the mixing length scales linearly with distance from the wall, l = \kappa y, where \kappa \approx 0.41 is the von Kármán constant derived from similarity arguments in the inertial region. Substituting this into the eddy viscosity expression and assuming constant total shear stress in the near-wall region yields the logarithmic velocity law, u^+ = \frac{1}{\kappa} \ln y^+ + B, with B \approx 5.0, valid in the inertial sublayer away from both the wall and the outer edge. To simplify calculations across the full boundary layer thickness \delta, Prandtl approximated the mean velocity profile with a 1/7th power law, \frac{u}{U_\infty} = \left( \frac{y}{\delta} \right)^{1/7}, which provides a reasonable fit to experimental data for moderate Reynolds numbers and captures the steeper gradient near the wall compared to the log law. From this profile, the local skin friction coefficient is derived as c_f \approx 0.045 \, \mathrm{Re}_\delta^{-1/4}, where \mathrm{Re}_\delta = U_\infty \delta / \nu, emphasizing the weak dependence on Reynolds number typical of turbulent drag. These approximations are applied via the integral form of the momentum equation for growth, originally formulated by von Kármán in 1921: \frac{d\theta}{dx} = \frac{c_f}{2}, where \theta is the momentum thickness. Inserting the power-law profile to express \theta and c_f in terms of \delta results in a solvable for the development, yielding \delta / x \approx 0.37 \, \mathrm{Re}_x^{-1/5} and c_f \approx 0.059 \, \mathrm{Re}_x^{-1/5} based on the streamwise distance x, aligning well with flat-plate experiments for $10^6 < \mathrm{Re}_x < 10^9. Despite its utility, the mixing length model with l = \kappa y fails in the viscous sublayer, where y^+ < 5, as it predicts vanishing \nu_t at the wall while the actual profile is linear due to molecular dominance, overestimating the near-wall gradient; enhancements involve damping the mixing length near the wall or superimposing the log law onto a linear inner profile for better accuracy.

Modern Computational Techniques

Direct Numerical Simulation (DNS) resolves all relevant turbulence scales in the Navier-Stokes equations without modeling subgrid phenomena, enabling precise predictions of skin friction drag in complex flows, including transitional regimes. In transitional boundary layers, DNS combined with the Computational Preston Tube Method (CPM) calibrates virtual probes to extract local skin friction coefficients from near-wall velocity data, achieving high accuracy in capturing intermittency effects. For instance, DNS studies of supersonic turbulent boundary layers at Mach 2.5 have demonstrated skin friction reductions up to 7% due to surface modifications like riblets, providing benchmark data for drag mitigation strategies. Reynolds-Averaged Navier-Stokes (RANS) simulations, augmented with turbulence models such as the k-ω model, offer computationally efficient predictions of the skin (c_f) for practical applications. The k-ω SST model blends k-ω near the wall and k-ε in the free stream, improving accuracy in adverse pressure gradients and separation-prone flows, with typical errors in c_f predictions below 10% for attached boundary layers when calibrated appropriately. These models are widely adopted for vehicle design, where they validate against empirical approximations like Prandtl's mixing length by reproducing integral drag contributions within 5-15% of experimental values. Large Eddy Simulation (LES) explicitly resolves large-scale turbulent structures while modeling subgrid-scale (SGS) effects, proving effective for high-Reynolds-number (high-Re) flows where skin friction dominates total . SGS models, such as the wall-adapting local eddy-viscosity (WALE) or dynamic Smagorinsky, capture near-wall , yielding skin friction predictions accurate to within 5% of DNS in flows at Re_τ up to 5200. In high-Re wall-bounded flows, LES with explicit algebraic models has shown superior performance over RANS in predicting c_f variations due to gradients. Recent advancements integrate machine learning (ML) surrogates to accelerate DNS for hypersonic vehicle applications, where skin friction under extreme conditions contributes over 50% to total drag. Physics-informed neural networks serve as surrogates for chemical reactions or wall modeling in DNS, reducing computational costs by factors of 10-100 while maintaining c_f accuracy within 6% relative error in hypersonic boundary layers. These ML-enhanced approaches, applied to transition-continuum predictions, enable rapid prototyping of drag reduction for reentry vehicles. Additionally, as of 2024-2025, machine learning models, including neural networks, have been developed to predict the evolution of skin friction in forced turbulent channel flows and wall aerodynamic quantities using Euler equation embeddings, improving predictive capabilities for complex aerodynamic scenarios.

Interrelations with Transport Phenomena

Reynolds Analogy to Heat Transfer

The Reynolds analogy establishes a fundamental relationship between skin friction drag and convective heat transfer in turbulent boundary layers by assuming that the eddy diffusivities for momentum and heat are equal, leading to similar transport mechanisms for shear stress and heat flux. This analogy, originally proposed by Osborne Reynolds, links the skin friction coefficient c_f, which quantifies momentum transfer at the surface, to the thermal transport via the dimensionless Stanton number St. The core equation of the Reynolds analogy is St = \frac{c_f}{2}, where the Stanton number is defined as St = \frac{Nu}{Re \, Pr} = \frac{q_w}{\rho_\infty U_\infty c_p (T_w - T_\infty)}, with Nu the , Re the , Pr the , q_w the wall , \rho_\infty the freestream density, U_\infty the freestream , c_p the specific , and T_w, T_\infty the wall and freestream temperatures, respectively. This relation implies that regions of high skin , such as near leading edges or in accelerating flows, experience correspondingly high convective rates due to the coupled of and thermal layers. The analogy holds under specific assumptions: the Pr \approx 1 (as for air, where Pr \approx 0.71), the flow is fully developed turbulent with dominant over , and conduction within the fluid is negligible compared to . These conditions ensure that the velocity and temperature profiles scale similarly, allowing the wall \tau_w = \frac{1}{2} \rho_\infty U_\infty^2 c_f to directly inform the . For fluids with Pr \neq 1, such as (Pr \approx 7) or oils, the basic overpredicts or underpredicts ; an extension known as the Colburn analogy (or Chilton-Colburn modification) accounts for this by incorporating a correction: St \, Pr^{2/3} = \frac{c_f}{2}. This form empirically adjusts for differences in molecular diffusivities of and while retaining the turbulent similarity. In practical applications, the is particularly valuable for cooling in engines, where elevated skin in high-velocity passages correlates directly with intensified from hot gases, guiding the placement and efficiency of cooling holes to maintain blade integrity without excessive cooling air usage. For instance, in turbine cascades, deviations from the due to freestream can increase by up to 20-30% relative to predictions, informing refined cooling designs.

Implications for Mass Transfer and Combustion

The Chilton-Colburn analogy extends the between skin friction and to phenomena, providing a framework for predicting mass transport rates from transfer data in turbulent layers. This empirical correlation equates the Colburn j-factor for to half the skin friction coefficient, expressed as j_m = \frac{\mathrm{Sh}}{\mathrm{Re} \, \mathrm{Sc}^{2/3}} = \frac{c_f}{2}, where \mathrm{Sh} is the , \mathrm{Re} is the , \mathrm{Sc} is the , and c_f is the skin friction coefficient. The analogy holds well for turbulent flows with moderate Schmidt numbers (typically 0.6 < Sc < 60), enabling estimation of species diffusion across boundary layers based on measured friction. In combustion processes, the turbulent friction underlying the promotes enhanced mixing of fuel and oxidizer within the by intensifying momentum transfer and , which is critical for achieving uniform fuel-air blending and stable reaction zones. However, this heightened also correlates with increased convective to the walls via the , resulting in greater heat losses that can reduce overall and necessitate advanced cooling strategies. In engines, skin friction plays a pivotal role in during hypersonic flows, directly influencing ignition delay and holding by affecting fuel penetration and reactivity, as evidenced in 2020s research on enhancement. The exhibits limitations in reacting flows with non-unity Lewis numbers (Le = Sc / Pr ≠ 1), where differential between heat and mass transport disrupts the assumed similarity, leading to inaccuracies in predicting species profiles and structures.

Practical Effects

Contribution to Total Aerodynamic Drag

Skin friction drag represents a significant portion of the total aerodynamic in various vehicles, particularly in flight regimes where viscous effects dominate the . In , it typically accounts for 45-50% of the total during cruise conditions. A seminal 1974 technical memorandum on reduction concepts reported that skin friction contributes approximately 45% of the total for commercial transports, underscoring its role as the primary viscous component. More recent analyses, including a 2021 review of turbulent reduction techniques, confirm this share remains around 50% in modern applications, reflecting ongoing relevance despite advancements in surface treatments. The contribution varies with vehicle configuration and operating conditions, influenced by factors such as wetted surface area and flow regime. Vehicles with high wetted areas relative to , like gliders optimized for low-speed , exhibit elevated skin friction shares due to minimized induced drag components. In contrast, see a reduced skin friction fraction of about 30%, as from shock waves overtakes viscous effects in the drag polar. A 2023 study on hypersonic layers highlighted that even in high-speed regimes, turbulent skin friction persists at roughly 30% of total , emphasizing its enduring impact across numbers. Measurement of skin friction's contribution relies on both wind tunnel experiments and full-scale flight tests, with necessary corrections for discrepancies arising from scale effects. Wind tunnel data, obtained from scaled models, underpredict skin friction at lower Reynolds numbers compared to flight conditions, requiring adjustments via methods like the reference temperature approach to extrapolate to full-scale performance. A 2002 brief on high-Reynolds-number skin friction estimation stressed the importance of such corrections for accounting for Reynolds effects on transition. Flight tests provide direct validation but are costlier, often confirming wind tunnel predictions within 5% after corrections. These methods apply to various designs, benefiting from precise instrumentation like embedded sensors.

Impacts on Vehicle Performance and Design

Skin friction drag imposes a notable penalty on the fuel efficiency of aircraft, as it constitutes approximately 50% of the total aerodynamic drag during cruise conditions for typical long-range jets. Reducing this component of drag directly improves operational range and reduces fuel consumption; for instance, a 1% reduction in total drag yields approximately a 1% increase in range for jet aircraft, stemming from the proportional relationship in the Breguet range equation where range scales with the lift-to-drag ratio. In vehicle design, minimizing skin friction often involves trade-offs between aerodynamic performance and structural durability. Smoother surfaces, achieved through flush riveting or polished finishes, lower friction by promoting and reducing , but they are more susceptible to from environmental factors like accumulation, , or paint wear, necessitating frequent . Similarly, laminar flow wings extend regions of low-friction laminar boundary layers, potentially reducing overall aircraft drag by 15-20% through improvements in compared to fully turbulent designs, yet their to surface demands specialized protocols and limits applicability in rugged operational environments. At high speeds, particularly in regimes near 0.8, effects exacerbate skin friction drag through shock- interactions that thicken the boundary layer, promote premature transition to , and elevate stresses. This increased friction contributes to the overall rise, influencing design choices such as shaping to delay shock formation and mitigate these penalties. The environmental implications of skin friction drag are profound, as lower friction translates to reduced fuel burn and thus fewer emissions; a 1% drag reduction can save 2-3 tonnes of CO2 per flight on civil . In alignment with the European Union's Flightpath 2050 initiative, which targets a 75% reduction in CO2 emissions per passenger-kilometer by 2050 relative to 2000 levels, aerodynamic improvements including skin friction reductions are essential to meeting these goals and curbing aviation's climate impact.

Drag Reduction Strategies

Passive Reduction Techniques

Passive reduction techniques for skin friction drag involve fixed surface modifications that do not require external energy input, primarily targeting the to minimize without altering flow dynamics actively. These methods leverage geometric or material alterations to disrupt production or promote slip conditions, achieving drag reductions typically in the range of 5-20% under controlled conditions. Seminal has focused on bio-inspired and engineered patterns that interact with near-wall coherent structures in turbulent flows. Riblets consist of streamwise-aligned micro-grooves on the surface, inspired by natural microstructures, which reduce skin friction by limiting lateral fluid motion and cross-stream vorticity in the turbulent . Optimal riblet dimensions, with spacing around 15-20 wall units, can yield drag reductions of 2-8% in flat-plate turbulent flows, as demonstrated in numerous studies. Notably, applied riblets to the A320 aircraft in the 1980s and 1990s, with full-airframe tests confirming approximately 2% total drag reduction, influencing experimental designs in . Biomimetic surfaces draw from biological adaptations, such as or , to replicate low-drag textures that modulate turbulence. -inspired surfaces, featuring micro-scale ridges and cusps, have shown reductions of up to 31% in recent experimental studies by altering vortex formation and momentum transfer near the wall. Similarly, -inspired flexible skins promote wave propagation that delays transition and reduces friction by up to 15%, as validated in hydrodynamic tests on scaled models. These approaches highlight the role of compliant or hierarchical structures in enhancing passive control. Dimples and controlled roughness patterns introduce shallow concavities or protrusions that generate localized low-drag wakes, countering the high in turbulent layers through secondary flows and redistribution. Laboratory tests on flat plates with optimized arrays have reported skin friction reductions of 5-15%, particularly effective at Reynolds numbers typical of applications, by promoting embedded vortices that inhibit near-wall bursts. A 2023 study using staggered circular cavities confirmed these benefits, achieving up to 10% net savings without increasing form drag significantly. Such patterns are advantageous for their simplicity in and integration on curved surfaces. Superhydrophobic coatings create Cassie-Baxter states with trapped air layers, inducing slip at the liquid-solid interface and reducing effective wetted area in submerged or high-humidity flows. These coatings can lower skin friction by up to 20% in low-speed turbulent conditions, as measured in experiments, by minimizing viscous through the plastron . However, reviews as of 2021 and 2025 emphasize durability challenges, including air layer collapse under or , limiting practical longevity to short-term applications despite advances in robust nanoparticle-infused formulations. Laminar flow control through passive leading-edge designs extends the region by optimizing contours to delay without , thereby reducing overall skin friction compared to fully turbulent flows. Natural (NLF) achieve 50-80% lower skin friction in the laminar portion, with surface slots in passive configurations further stabilizing the layer via geometric diffusion control. Recent design studies confirm these techniques enable up to 15-20% total drag savings on wings by promoting favorable gradients at the .

Active Reduction Methods

Active reduction methods for skin friction drag involve the input of external energy to dynamically manipulate the turbulent in , enabling that surpasses the limitations of fixed passive designs. These techniques target the near-wall region to disrupt coherent structures, suppress production, or remove low-momentum fluid, often achieving significant local drag reductions through actuators, blowing, or oscillatory mechanisms. Recent advancements in the have focused on integrating sensors and systems for optimized performance in high-Reynolds-number flows, with applications in and marine vehicles. Boundary layer suction and injection represent foundational active control strategies, where low-momentum fluid is actively removed via suction slots or replenished through injection to stabilize the and reduce . Uniform suction can delay to and lower skin friction by 10-30% in turbulent flows, as demonstrated in tests evaluating multiple suction configurations on zero-pressure-gradient surfaces. Injection, such as micro-blowing through arrays of small orifices, similarly attenuates near-wall streaks, yielding up to 20% local reduction in direct numerical simulations (DNS) of flat-plate s, with effects persisting downstream due to staggered hole arrangements. These methods require precise of mass flow rates to balance energy input against net drag savings, particularly in high-speed applications like hypersonic vehicles where cooling benefits are also realized. Spanwise traveling employ oscillatory motion generated by actuators to impose periodic spanwise perturbations, effectively modulating the near- and weakening turbulent . In DNS studies of flows at moderate to high s, these achieve 15-20% skin friction reductions by accelerating spanwise motions that suppress sweep and ejection events, with optimal wave speeds around 15-20% of the friction . Recent 2024 investigations confirm up to 24.5% drag cuts in fully developed turbulent boundary layers using blowing-suction distributions to mimic traveling , highlighting dependence where efficacy diminishes slightly at higher but remains viable for regimes. Plasma actuators, particularly dielectric barrier discharge (DBD) types, utilize ionized air from high-voltage electrodes to create virtual wall shaping or induced flows without mechanical parts, offering lightweight solutions for real-time . In high-speed turbulent boundary layers, opposite-charged DBD configurations generate counter-rotating vortices that disrupt near-wall , resulting in 5-10% skin friction reductions as reported in 2025 AIAA studies on annular actuators for supersonic flows. Pulsed-DC variants enhance by producing spanwise oscillations analogous to traveling waves, with net power savings observed in experiments where local drops by up to 8-10% at highway-equivalent speeds, scalable to . Vibration and compliant walls draw from bionic principles, such as dolphin skin, where active flexibility or microvibrations via piezoelectric actuators mimic longitudinal ridges to manage wave propagation and reduce pressure gradients. 2024 bionic research on flexible surfaces inspired by cetacean demonstrates up to 28% drag reductions in channel tests, attributed to enhanced stability from downstream-traveling ultrasonic perturbations that attenuate intensity. These systems actively adjust compliance based on , outperforming rigid analogs by 5-6% in maximum reduction rates, with applications in submerged vehicles where combined and hydrophobicity amplify effects. Large Eddy Break-Up (LEBU) devices, when augmented with feedback control, actively disrupt outer-layer eddies using adjustable flat plates or flaps positioned in the to break up large-scale structures. Traditional LEBU yields mixed results with local skin friction reductions of 10-20%, but recent trials incorporating AI-driven feedback for have improved efficacy by optimizing placement in , achieving more consistent 15% net drag cuts in turbulent channels as per 2022 large-eddy simulations. These advancements address historical limitations like downstream drag penalties, enabling adaptive operation in variable flow conditions.

References

  1. [1]
    What is Drag? | Glenn Research Center - NASA
    Jul 21, 2022 · We can think of drag as aerodynamic friction, and one of the sources of drag is the skin friction between the molecules of the air and the ...Drag · Aerodynamic Friction · Form Drag
  2. [2]
    Boundary Layer Flows – Introduction to Aerospace Flight Vehicles
    This latter result is significant because the boundary layer is the primary source of shear stress drag, also known as skin friction drag, on a flight vehicle.
  3. [3]
    [PDF] Chapter 5: Aerodynamics of Flight - Federal Aviation Administration
    Skin Friction Drag. Skin friction drag is the aerodynamic resistance due to the contact of moving air with the surface of an aircraft. Every surface, no matter ...
  4. [4]
    Skin Friction - an overview | ScienceDirect Topics
    Skin friction is defined as the drag generated by shear stresses acting tangentially on a body's surface, resulting from viscosity, and opposing the ...
  5. [5]
    Boundary Layer
    A thin layer of fluid near the surface in which the velocity changes from zero at the surface to the free stream value away from the surface.
  6. [6]
    Turbulent Flows – Introduction to Aerospace Flight Vehicles
    The concept of “eddy” viscosity augments the laminar molecular viscosity by accounting for the effects of eddies and mixing caused by turbulence. The Boussinesq ...
  7. [7]
    [PDF] Introductory Lectures on Turbulence: Physics, Mathematics and ...
    we give up the ability to predict skin friction drag with this model. ... models incorporate the notion that eddy viscosity “acts like” molecular viscosity—the ...
  8. [8]
    [PDF] Ludwig Prandtl's Boundary Layer
    boundary-layer theory occurred in the late 1920s when de- signers began to use the theory's results to predict skin- friction drag on airships and airplanes.
  9. [9]
    Drag of Blunt Bodies and Streamlined Bodies
    Frictional drag comes from friction between the fluid and the surfaces over which it is flowing. This friction is associated with the development of boundary ...Missing: definition | Show results with:definition
  10. [10]
    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 International ...
  11. [11]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Dec 15, 2021 · Over the back half the pressure gradient is adverse. (1) Boundary layer separation requires an adverse pressure gradient; however, an adverse ...Missing: growth | Show results with:growth
  12. [12]
    [PDF] Boundary Layers - Purdue Engineering
    Dec 15, 2021 · The momentum thickness, δM , is the thickness of a stagnant layer that has the same momentum deficit, relative to the outer flow, as the actual ...
  13. [13]
    Momentum Thickness - an overview | ScienceDirect Topics
    Momentum thickness is defined in relation to the momentum flow rate within the boundary layer. This rate is less than the rate that would occur if no boundary ...
  14. [14]
    [PDF] Simulation of a Turbulent Flow Subjected to Favorable and Adverse ...
    The boundary layer growth, indicated by the increases in 𝛿, 𝛿∗ and 𝜃 in the upstream adverse pressure gradient region, is accompanied by a corresponding drop in ...
  15. [15]
    [PDF] Boundary Layers with Pressure Gradients
    If dP/dx > 0, dU/dx < 0, the flow is decelerating. This is an unfavorable pressure gradient (also called an adverse pressure gradient).
  16. [16]
    [PDF] Introduction to Aerospace Engineering - TU Delft OpenCourseWare
    Laminar boundary layer: thin, low skin friction drag. Turbulent boundary layer: thick, high skin friction drag. Transition “point”. V. Laminar. B.L.. Page 28 ...
  17. [17]
    Tollmien-Schlichting Wave - an overview | ScienceDirect Topics
    Tollmien-Schlichting waves are defined as primary instabilities that lead to the transition from laminar to turbulent flow in two-dimensional and ...
  18. [18]
    Laminar/Turbulent Flow - Rose-Hulman
    Here are some generally accepted numbers for the laminar to turbulent transition: Internal flow: Recr = 2,300. External forced flow (Flat Plate): Recr = 5 X 10.
  19. [19]
    Smooth Flat Plate - an overview | ScienceDirect Topics
    The transition Reynolds number Rxtr depends partly upon the turbulence in the free stream; Rxtr may be as low as 5 × 104 or as high as 5 × 106. For laminar flow ...
  20. [20]
    Toward Practical Laminar Flow Control— Remaining Challenges
    Within certain limits on freestream turbulence and 3-D surface roughness, they are more sensitive to freestream sound than to freestream turbulence [3] ...
  21. [21]
    Experiments on Discrete Roughness Element Technology for Swept ...
    Freestream turbulence is particularly important when dealing with the crossflow instability, mainly because a low freestream-turbulence environment is dominated ...
  22. [22]
    Attached Eddy Model of Wall Turbulence - Annual Reviews
    Jan 5, 2019 · Modeling wall turbulence remains a major challenge, as a sufficient physical understanding of these flows is still lacking.
  23. [23]
    Modification of burst events in the near-wall region of turbulent ...
    Burst events in the near-wall region of turbulent boundary layers are the main contributors to skin friction drag in wall flows.
  24. [24]
    Boundary-Layer Theory | SpringerLink
    This new edition of the near-legendary textbook by Schlichting and revised by Gersten presents a comprehensive overview of boundary-layer theory.Missing: sublayer | Show results with:sublayer
  25. [25]
    [PDF] By Th. v. Karman Reprint from Nachrichten von der Gesellschaft der ...
    of the velocity distribution? method of attack in this sense, and endeavors an attempt to nake tbe laws of turbulent flow in grooves amenable for calculation.
  26. [26]
    Golf Ball Aerodynamics | Aeronautical Quarterly | Cambridge Core
    Jun 7, 2016 · A wind tunnel technique has been developed to measure the aerodynamic forces acting on golf balls over a wide range of Reynolds number and spin rate.
  27. [27]
    [PDF] A review of factors affecting boundary-layer transition
    A brief review is made of the current state of the art of boundary-layer transition. Discussed, in various degrees of detail, are experimentally determined ...
  28. [28]
    [PDF] Boundary-Layer Linear Stability Theory
    Boundary-Layer Linear Stability Theory includes elements of stability theory, and covers incompressible and compressible stability theory.
  29. [29]
    [PDF] Bypass Transition in Boundary Layers Including Curvature and ...
    stability theory. There are no standard criteria or parameters for defining bypass transition, but it is known to be the mode of transition when the flow is ...
  30. [30]
    A rational method for determining intermittency in the transitional ...
    Dec 3, 2019 · The value of the intermittency factor varies from 0 to 1, where a zero value represents a fully laminar region and the value of one represents a ...
  31. [31]
    [PDF] local skin friction coefficients and boundary-layer profiles obtained in ...
    Boundary-layer and local skin friction data for Mach numbers up to 2. 5 and. Reynolds numbers up to 3. 6 X 108 were obtained in flight at three locations on.
  32. [32]
    [PDF] NATIONAL ADVISORY COMMITTEE FOR AERONAUTICS
    The average skin-friction coefficient for a surface is the inte- grated ... The second term inside the integral on the right-hand side of equa- tion (9) ...
  33. [33]
    [PDF] THE PRACTICAL CALCULATION OF THE AERODYNAMIC ...
    The skin friction coefficient is defined as the skin friction drag of a plane surface divided by the dynamic pressure and the wetted area of the surface.
  34. [34]
    [PDF] Skin Friction at Very High Reynolds Numbers in the National ...
    Aug 1, 2006 · Increased accuracy in the prediction of high Reynolds number skin friction coefficients can translate into ... felt that surface roughness ...
  35. [35]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics ...
    Dec 15, 2021 · The friction coefficient for a laminar boundary layer is given by, . (1). Re-arranging to solve for τw in terms of x gives, . (2). Thus, as x ...
  36. [36]
    Skin-Friction Coefficient - an overview | ScienceDirect Topics
    The skin friction coefficient is defined as a measure of the frictional resistance experienced by a fluid over a surface, calculated as the average of the ...
  37. [37]
    [PDF] Blasius_Paper_NACA-TM-1256.pdf
    This paper by H. Blasius discusses boundary layers in fluids with little friction, based on a translation of a 1908 work.
  38. [38]
    Skin friction coefficient - CFD Online
    Jan 14, 2016 · The skin friction coefficient, C_f , is defined by: C_f \equiv \frac{\tau_w}{\frac{1}{. Where \tau_w is the local wall shear stress, \rho ...
  39. [39]
    [PDF] Turbulent Boundary Layers 4 - 1 David Apsley h δ* δ* h+ h 4 ...
    skin-friction coefficient cf is finite, the momentum thickness must be continuous at the transition point. Irrespective of any intermediate development, the ...
  40. [40]
    [PDF] Predicting the Drag of Rough Surfaces - Naval Academy
    In summary, the offset d is a parameter determined by the flow (like the equivalent sand-grain roughness, ks; Section 2.2.2). Geometrical definitions (e.g., ...
  41. [41]
    A new equivalent sand grain roughness relation for two-dimensional ...
    Apr 8, 2022 · A new equivalent sand grain roughness relation for two-dimensional rough wall turbulent boundary layers - Volume 940.
  42. [42]
    [PDF] prediction of very high reynolds number compressible skin friction
    Flat plate skin friction calculations over a range of. Mach numbers from 0.4 to 3.5 at Reynolds numbers from 16 million to 492 million using a Navier Stokes.
  43. [43]
    [PDF] Blasius Solution for a Flat Plate Boundary Layer
    The first exact solution to the laminar boundary layer equations, discovered by Blasius (1908), was for a simple constant value of U(s) and pertains to the ...
  44. [44]
    [PDF] REPORT ONINVESTIGATION O,FDEVELOPED TURBULENCE By L ...
    Translation of “Bericht fiber Untersuchungen zur ausgebildeten. Turbulent.” Zeitschrift fiir angewand~e Mathematik und. Mechanik, vol. 5, no. 2, April 1925.Missing: Zusammenarbeiten | Show results with:Zusammenarbeiten
  45. [45]
    Aerodynamic Theory Vol-iii : Durand William Frederick
    Jan 19, 2017 · PDF download · download 1 file · PDF WITH TEXT download · download 1 file · SINGLE PAGE PROCESSED JP2 ZIP download · download 1 file · TORRENT ...Missing: Prandtl | Show results with:Prandtl
  46. [46]
    [PDF] Prandtl's Mixing Length Hypothesis - Sites
    Equation (12) clearly shows that the kinetic energy of turbulence is convected, diffused, produced, and dissipated. Modeling k-Equation. To close the k-equation ...Missing: skin friction<|control11|><|separator|>
  47. [47]
    Measurement of transitional boundary layer on a flat plate using a ...
    A computational Preston tube method (CPM) originally proposed by Nitsche et al. (1983) was adopted and refined to measure the skin friction coefficients in ...
  48. [48]
    [PDF] Direct Numerical Simulations of High-Speed Turbulent Boundary ...
    Section III presents results pertaining to the effects of riblets on skin friction at high speeds, including the drag-reduction curve and drag-reduction ...Missing: seminal | Show results with:seminal
  49. [49]
    Direct numerical simulations of a supersonic turbulent boundary ...
    Apr 6, 2021 · I. INTRODUCTION. Pursuing lower turbulent friction drag is a permanent and important goal for aeronautical vehicles and it has attracted ...Missing: seminal | Show results with:seminal
  50. [50]
    Calibration method of the k-ω SST turbulence model for wind turbine ...
    Jan 15, 2024 · The present study intends to enhance the accuracy of the k–ω SST turbulence model for numerical wind turbine simulation in stall condition.2. Computational Method · 2.1. Turbulence Model · 3. Results And Discussion
  51. [51]
    [PDF] PERFORMANCE OF SST k-ω TURBULENCE MODEL FOR ...
    From this table it may be concluded that the SST k-ω turbulence model computes more accurate drag coefficient for slender bodies. Fully turbulent flow is ...
  52. [52]
    Large Eddy Simulation of High-Reynolds-Number Free and Wall ...
    Large Eddy Simulation of High-Reynolds-Number Free and Wall-Bounded Flows ... skin friction and other bulk flow variables in two-dimensional rectangular duct flow.
  53. [53]
    Taking large-eddy simulation of wall-bounded flows to higher ...
    Mar 3, 2017 · Large-eddy simulations of a fully developed channel flow are performed with friction Reynolds numbers of 550, 2000, and 5200 by using the ...
  54. [54]
    Detailed investigation of subgrid scale models in large-eddy ...
    Nov 16, 2021 · In this paper, to investigate the effect of the definition of a filter width on the prediction accuracy of an SGS model, we report on a priori tests of several ...
  55. [55]
    Accelerating hypersonic reentry simulations using deep learning ...
    Sep 27, 2022 · This paper accelerates hypersonic reentry simulations by coupling a traditional fluid solver with a neural network for chemical reactions, ...<|control11|><|separator|>
  56. [56]
    Examination of Machine Learning for the Modeling of Hypersonic ...
    This work uses machine learning to reconstruct velocity gradient on the wall in hypersonic boundary layers, improving relative error from 50% to within 6%.
  57. [57]
    (PDF) Physics-Based Machine Learning Closures and Wall Models ...
    Jul 15, 2025 · We develop a physics-constrained machine learning framework that augments transport models and boundary conditions to extend the applicability ...
  58. [58]
    17.1 The Reynolds Analogy - MIT
    In a turbulent boundary layer, the dominant mechanisms of shear stress and heat transfer change in nature as one moves away from the wall. Figure 17.3: Velocity ...<|separator|>
  59. [59]
    A Critical Assessment of Reynolds Analogy for Turbine Flows
    Freestream turbulence has the opposite effect of increasing heat transfer more than skin friction, thus the Reynolds analogy factor increases with turbulence ...
  60. [60]
    Reynolds Analogy - an overview | ScienceDirect Topics
    3.1 Skin friction, surface temperature and scalar concentration. Relation between skin friction and heat transfer is generally known as the Reynolds analogy.Missing: drag | Show results with:drag
  61. [61]
    REYNOLDS ANALOGY - Thermopedia
    A modified Reynolds analogy, also known as the Chilton-Colburn analogy, is found to be applicable under these conditions: (2). (3). Equations (2) and (3) are ...Missing: H. 1933<|control11|><|separator|>
  62. [62]
    Mass Transfer (Absorption) Coefficients Prediction from Data on ...
    article November 1, 1934. Mass Transfer (Absorption) Coefficients Prediction from Data on Heat Transfer and Fluid Friction. Click to copy article linkArticle ...Missing: original | Show results with:original
  63. [63]
    An Experimental Study of Chilton–Colburn Analogy Between ...
    The Chilton–Colburn analogy is demonstrated to be valid for turbulent flow and heat transfer of supercritical kerosene through horizontal straight circular ...
  64. [64]
    Hypersonic Turbulent Boundary-Layer Fuel Injection and Combustion
    Immediately above the wall surface is the viscous sublayer, inside which turbulent eddy structures are smaller than the Kolmogorov scale.A. Boundary Layer Model · Iii. Analysis · B. Skin Friction And Fuel...
  65. [65]
    Numerical simulation and performance evaluation of skin friction ...
    Jul 5, 2025 · Key findings reveal that fuel injection demonstrates superior skin friction reduction efficacy compared to inert gases, especially hydrogen, ...
  66. [66]
    Numerical simulation on the combustion characteristics of scramjet ...
    It works by compressing the hypersonic incoming air through the diffuser to a suitable temperature and pressure, and then passing through the isolator to enter ...
  67. [67]
    Influence of Lewis number on strain rate effects in turbulent ...
    Aug 8, 2025 · For non-unity Lewis numbers it is shown that the variations of temperature and scalar gradient in response to tangential strain rate on a given ...
  68. [68]
    [PDF] A GENERAL REVIEW OF CONCEPTS FOR REDUCING SKIN ...
    Application of suction to an airfoil may produce two favorable effects: (1) a reduction in skin-friction drag by delaying or preventing boundary-layer ...
  69. [69]
    [PDF] A review of turbulent skin-friction drag reduction by near-wall ...
    In civil aviation, skin-friction drag accounts for around 50% of the total drag in cruise conditions, thus being a preferential target for research. With ...<|control11|><|separator|>
  70. [70]
    On the drag reduction mechanism of hypersonic turbulent boundary ...
    Feb 21, 2023 · Even under flight conditions dominated by shock drag in a supersonic/hypersonic flow, the turbulent friction drag still accounts for about 30% ...Missing: percentage | Show results with:percentage
  71. [71]
    [PDF] Skin-friction estimation at high Reynolds numbers and Reynolds ...
    The purpose of this research brief is to conduct a brief review of skin-friction estimation over a range of Reynolds numbers, as this is one of the key ...
  72. [72]
    Drag model for extremely fast estimation of pressure and friction ...
    The total drag coefficient is then calculated by combining the various sources, using Eq. ... drag, will help guide AI-built UAV designs to make reasonable ...
  73. [73]
    [PDF] Innovative Flow Control Concepts for Drag Reduction
    Jan 4, 2016 · Since skin friction. (or viscous drag) contributes to about 50% of the total drag of a transport aircraft,3,4 its reduction has the potential of.
  74. [74]
    A review of turbulent skin-friction drag reduction by near-wall ...
    May 1, 2021 · The present article provides a near-exhaustive review of research into the response of turbulent near-wall layers to the imposition of unsteady and wavy ...Missing: seminal | Show results with:seminal<|separator|>
  75. [75]
    [PDF] Breguet Range Equation - MIT OpenCourseWare
    An excellent estimate of the range (maximum distance) that an aircraft can fly (under some important assumptions) is provided by the Breguet Range Equation. The ...
  76. [76]
    [PDF] Aircraft Drag Prediction and Reduction - DTIC
    The various sources and relative contributions of aircraft drag are described including skin friction drag, pressure drag, interference drag and lift induced ...
  77. [77]
    Transonic Flow - Centennial of Flight
    Up to a free-stream Mach number of about 0.7 to 0.8, compressibility effects have only minor effects on the flow pattern and drag. The flow is subsonic ...<|control11|><|separator|>
  78. [78]
    [PDF] NATURAL LAMINAR FLOW AIRFOIL ANALYSIS AND TRADE ...
    The theoretical possibility of achieving laminar airflow over airplane wings and realizing the performance benefits of the resultant drag reduction has been ...Missing: durability | Show results with:durability
  79. [79]
    Skin Friction Drag - an overview | ScienceDirect Topics
    The skin friction drag is the result of stresses produced within the boundary layer. Initial flow within the boundary layer is laminar (regular, continuous ...
  80. [80]
    Aircraft Lift and Drag Decomposition in Transonic Flows | Journal of ...
    May 10, 2017 · The problem is further complicated when compressibility effects, such as shock waves, are considered.Missing: skin | Show results with:skin
  81. [81]
    [PDF] Time for change | Acare
    The ratified Green Deal objectives demand that the European aviation sector achieves drastically reduced emissions by 2030 and climate neutral aviation by 2050.
  82. [82]
    Riblets - an overview | ScienceDirect Topics
    Riblet surfaces are made of grooves aligned with the flow (Fig. 23). The role of these surfaces is to reduce the friction drag of a turbulent boundary layer.
  83. [83]
    (PDF) Smart Flow Control with Riblets - ResearchGate
    Numerous studies (Walsh, 1990;Behert et al., 1997; Choi, 2013) have reported that riblets can reduce the friction drag of a flat-plate turbulent boundary layer ...
  84. [84]
    [PDF] Evaluation of Skin Friction Drag Reduction in the Turbulent ...
    Nov 29, 2019 · The turbulent statistics, such as the turbulent scales and intensities, in the boundary layer were identified based on the freestream velocity ...<|control11|><|separator|>
  85. [85]
    Experimental Studies of Bioinspired Shark Denticles for Drag ...
    Sep 27, 2024 · This review highlights the past 15 years of manufacturing techniques and experimental measurements of drag over denticle-inspired surface ...Missing: dolphin | Show results with:dolphin
  86. [86]
    A novel aerodynamic drag-reduction mechanism using dolphin ...
    Apr 21, 2025 · Here we introduce a novel strategy to reduce drag while enhancing lift-to-drag ratio by utilizing dolphin skin-inspired downstream-traveling longitudinal micro ...
  87. [87]
    Dimples for Skin-Friction Drag Reduction: Status and Perspectives
    Jul 13, 2022 · Dimples have been proposed as a roughness pattern that is capable of reducing the turbulent drag of a flat plate by providing a reduction of skin friction.Missing: 2023 | Show results with:2023
  88. [88]
    Drag reduction by means of an array of staggered circular cavities at ...
    Passive skin friction reduction techniques such as cavities and dimples are preferred to wall-normal blowing and spanwise wall oscillations as they do not ...
  89. [89]
    Superhydrophobic drag reduction in turbulent flows: a critical review
    Oct 16, 2021 · This review is aimed to help guide the design and application of SHPo surfaces for drag reduction in the large-scale turbulent flows of field conditions.<|separator|>
  90. [90]
    Durable superhydrophobic coating with energy-saving drag ...
    Jul 15, 2025 · The coating shows excellent drag reduction of up to 94 %. At −10 °C, the superhydrophobic coating was able to delay icing for about 758 s, ...
  91. [91]
    Laminar Flow Control - an overview | ScienceDirect Topics
    Laminar flow control has been shown to provide skin friction reduction as high as 75–80% when compared to the skin friction of a solid flat plate. [2,3] ...
  92. [92]
    [PDF] Low Drag Airfoil Design Utilizing Passive Laminar Flow and ... - DTIC
    These three design areas define the passive laminar flow stabilization region, transition stabilization zone, and trailing edge diffusion. Two design areas ...