Fact-checked by Grok 2 weeks ago

Adverse pressure gradient

An refers to a condition in where the increases in the direction of the flow (dp/dx > 0), causing deceleration of fluid particles, especially near solid surfaces within the . This opposes the inertial forces of the flow and is a key factor in many contexts, such as and , where it arises from geometric features like curved surfaces or diverging channels. The primary effect of an adverse pressure gradient is the potential for separation, where the near-wall flow reverses direction and detaches from the surface, forming a recirculation and wake. This separation occurs when the wall approaches zero, often preceded by an in the velocity profile, and is more likely in laminar layers due to their lower compared to turbulent ones. In aerodynamic applications, such as over the rear of an , it leads to increased form , reduced , and stall at high angles of attack, significantly impacting vehicle performance and efficiency. Adverse pressure gradients also thicken the , elevate levels in the outer , and reduce skin friction downstream, with losses up to 50% or more in severe cases. They are commonly induced in diffusers, compressors, and vortex flows, where factors like swirl intensity exacerbate and lower critical Reynolds numbers for . Understanding and mitigating these gradients through design optimizations, such as surface curvature control or control devices, is essential for maintaining attached and minimizing energy losses in practical systems.

Definition and Fundamentals

Definition

An in is a spatial variation of in which the increases in the direction of the , qualitatively described as rising downstream. This condition typically arises in decelerating external flows, such as those over the aft portions of airfoils or in diffusers, where the surrounding imposes a rising on the . In distinction, a favorable pressure gradient involves a decrease in downstream, which accelerates the flow and thins the . The physical intuition behind an adverse pressure gradient lies in its opposition to the flow's : it decelerates the by converting into pressure energy, thereby retarding motion especially near the surface where viscous effects are prominent. The concept originated in Ludwig Prandtl's 1904 boundary layer theory, introduced as a critical element in understanding viscous flows adjacent to surfaces and their susceptibility to disruption. Adverse pressure gradients can precipitate separation as a potential outcome.

Mathematical Formulation

In , an adverse pressure gradient is quantitatively defined as a condition in which the P increases along the streamwise x, mathematically expressed as \frac{dP}{dx} > 0. This formulation assumes a one-dimensional aligned with the coordinate , where the positive x- corresponds to the primary . For more general flows, including those in non-orthogonal coordinates or with velocity not strictly aligned with the axes, the condition generalizes to the vector form \nabla P \cdot \mathbf{u} > 0, where \nabla P is the pressure gradient vector and \mathbf{u} is the local velocity vector. This dot product indicates that the pressure rise opposes the local flow direction, leading to deceleration. The implications for flow behavior can be understood through the inviscid Euler equation for steady flow along a streamline in the outer flow region, where the streamwise velocity gradient relates to the pressure gradient as \frac{dU}{dx} = -\frac{1}{\rho} \frac{dP}{dx}, with \rho denoting fluid density and U the outer flow velocity. For an adverse pressure gradient (\frac{dP}{dx} > 0), this yields \frac{dU}{dx} < 0, signifying deceleration of the inviscid flow outside the boundary layer. A key dimensionless parameter characterizing the strength of the pressure gradient in similarity-based analyses is the Falkner-Skan parameter \beta = \frac{2m}{m+1}, where m relates to the external velocity variation U \propto x^m. Adverse conditions correspond to \beta < 0 (or equivalently m < 0), which applies in frameworks like the for wedge flows.

Physical Principles

Relation to Bernoulli's Principle

Bernoulli's equation provides the foundational inviscid relation linking pressure and velocity in fluid flows. For steady, incompressible, inviscid flow along a streamline, it states: P + \frac{1}{2} \rho U^2 + \rho g h = \constant where P is static pressure, \rho is fluid density, U is flow speed, g is gravitational acceleration, and h is elevation. Neglecting gravitational effects in horizontal flows, an increase in P must be balanced by a decrease in U, demonstrating the inverse relationship between pressure and velocity. This principle directly underlies adverse pressure gradients, where the streamwise pressure gradient \frac{dP}{dx} > 0 implies flow deceleration \frac{dU}{dx} < 0. In regions of flow diffusion, such as the rear portion of an airfoil or a diverging duct, the inviscid outer flow experiences pressure recovery as velocity diminishes, converting kinetic energy into pressure energy along streamlines. For instance, on an airfoil's upper surface beyond the maximum thickness, the pressure rises toward the trailing edge, creating an adverse gradient that slows the flow. The equation assumes inviscid, steady, and incompressible conditions, which idealize the outer flow but overlook viscous dissipation within boundary layers; real viscous effects can exacerbate deceleration and lead to flow separation as a consequence of this energy shift. A practical example is a diffuser, where an increase in cross-sectional area enforces velocity reduction via the continuity equation, thereby raising pressure and imposing an adverse gradient; ideal pressure recovery follows C_{P} = 1 - \frac{1}{(\text{AR})^2}, where AR is the area ratio, though viscous losses reduce efficiency.

Imposition in External Flows

In external flows, adverse pressure gradients are primarily imposed by the inviscid outer flow field, which dictates the pressure distribution along the surface through solutions to the potential flow equations. For instance, in the potential flow around an airfoil, the outer flow accelerates over the forward curved surface and decelerates toward the rear, creating regions of increasing pressure that act on the boundary layer. Similarly, in wakes behind bluff bodies, the inviscid flow solution leads to a recovery of pressure downstream, imposing an adverse gradient on any adjacent shear layers. Geometrically, these gradients arise from convex surface curvatures or diffusing flow paths that cause the external flow to slow down. On the rear portion of an airfoil, the convex curvature forces streamlines to diverge, reducing the external velocity and thereby increasing the static pressure in the upstream direction. In diffusers or expanding sections of external flow geometries, such as the aft regions of vehicle underbodies, the area increase similarly decelerates the flow, generating an upstream pressure rise that challenges the near-wall flow. The profile of the external velocity U_e directly governs the adverse pressure gradient, as derived from the inviscid Bernoulli equation along the edge of the : \frac{dP}{dx} = -\rho U_e \frac{dU_e}{dx} An adverse gradient occurs when \frac{dU_e}{dx} < 0, leading to \frac{dP}{dx} > 0 as U_e decreases along the surface. This relationship highlights how the outer flow's deceleration imposes a decelerating force on the . These phenomena were first systematically observed in experiments during the 1920s by and his collaborators, who linked adverse pressure gradients to the abrupt rise in drag on airfoils at higher angles of attack through separation.

Effects on Boundary Layer Flow

Flow Deceleration and Separation

In an adverse pressure gradient, where the static pressure increases in the streamwise direction (\frac{\partial P}{\partial x} > 0), the pressure force acts opposite to the flow momentum, causing a progressive deceleration of the . This effect is most pronounced near the wall, where fluid particles have the lowest and are least able to overcome the retarding pressure rise. As a result, the streamwise velocity component u diminishes, particularly at the wall (y=0), potentially reducing to zero and inducing reversal if the gradient is sufficiently strong. The fundamental mechanism governing this deceleration is captured by the boundary layer momentum equation: \frac{\partial u}{\partial t} + u \frac{\partial u}{\partial x} + v \frac{\partial u}{\partial y} = -\frac{1}{\rho} \frac{\partial P}{\partial x} + \nu \frac{\partial^2 u}{\partial y^2}, in which the adverse pressure term -\frac{1}{\rho} \frac{\partial P}{\partial x} (positive for adverse gradients) dominates the near-wall balance, overpowering convective and viscous diffusion effects to slow the flow. This leads to a thickening boundary layer and a flattening velocity profile, reducing the momentum transport from the outer flow. Flow separation occurs at the point where the wall shear stress \tau_w = \mu \left( \frac{\partial u}{\partial y} \right)_{y=0} reaches , marking the onset of reverse flow and the formation of a recirculation characterized by streamwise . At this criterion, \left( \frac{\partial u}{\partial y} \right)_{y=0} = [0](/page/0), the velocity profile develops an , and the detaches from the surface, transitioning to a wake-like structure. Associated with this process, the skin friction coefficient C_f = \frac{2 \tau_w}{\rho U_\infty^2} steadily decreases under the adverse gradient as the near-wall velocity gradient weakens, ultimately dropping to zero precisely at the separation point. This decline reflects the diminishing interaction between the fluid and surface, contributing to a shift from skin friction-dominated drag to pressure drag in separated flows. While the deceleration physics is universal, turbulent boundary layers exhibit greater resistance to separation than laminar ones due to enhanced mixing that sustains near-wall momentum.

Laminar vs. Turbulent Responses

In laminar boundary layers, momentum transfer occurs primarily through , resulting in a thinner layer with lower near the wall. This limited renewal of momentum makes laminar flows particularly susceptible to separation under even mild adverse pressure gradients, as the near-wall decelerates rapidly without sufficient of high-speed from the outer . In contrast, turbulent boundary layers exhibit enhanced mixing due to the action of eddies, which effectively transport high-momentum fluid from the outer toward the wall. This mechanism increases the near-wall velocity and , allowing turbulent layers to resist stronger adverse pressure gradients and delay separation compared to their laminar counterparts. The from laminar to turbulent plays a critical role in response to adverse pressure gradients, with boundary layers typically transitioning at Reynolds numbers Re_x > 5 \times 10^5 based on from the , promoting that delays separation. Experimentally, adverse pressure gradients amplify Tollmien-Schlichting waves in laminar boundary layers, leading to more regular wave patterns and increased amplification rates that accelerate the to . The Falkner-Skan \beta, which quantifies strength (negative for adverse), further influences laminar by promoting unstable disturbances as \beta decreases.

Theoretical Frameworks

Falkner-Skan Similarity Solutions

The Falkner-Skan similarity solutions provide an exact analytical framework for solving the Prandtl boundary layer equations in two-dimensional, incompressible, steady flows where the external varies as a along the surface, U_e(x) \propto x^m. This approach assumes a self-similar form for the profile, enabling reduction of the partial differential equations to a single (ODE). These solutions are particularly insightful for understanding adverse s, corresponding to negative values of the \beta < 0, which model decelerating flows such as those over the rear portion of a wedge. The derivation begins with the boundary layer continuity and momentum equations: \frac{\partial u}{\partial x} + \frac{\partial v}{\partial y} = 0, \quad u \frac{\partial u}{\partial x} + v \frac{\partial u}{\partial y} = U_e \frac{d U_e}{dx} + \nu \frac{\partial^2 u}{\partial y^2}, with boundary conditions u(x, y \to \infty) = U_e(x), u(x,0) = 0, and v(x,0) = 0. To achieve similarity, introduce the transformed coordinate \eta = y \sqrt{\frac{U_e}{\nu x}} and stream function \psi = \sqrt{\nu x U_e} \, f(\eta), where the velocity components are u = U_e f'(\eta) and v = \frac{1}{2} \sqrt{\frac{\nu U_e}{x}} \left( \eta f' - f \right). Substituting these into the boundary layer equations yields the : f''' + f f'' + \beta \left( 1 - (f')^2 \right) = 0, subject to the boundary conditions f(0) = f'(0) = 0 and f'(\infty) = 1. Here, the primes denote derivatives with respect to \eta, and the pressure gradient parameter is defined as \beta = \frac{2m}{m+1}, linking the flow exponent m to the imposed pressure gradient via Bernoulli's principle, \frac{dP}{dx} = -\rho U_e \frac{dU_e}{dx}. For adverse pressure gradients, m < 0 implies \beta < 0. Solutions to the Falkner-Skan equation exist for \beta > -0.1988, with the wall shear stress proportional to f''(0) > 0. At the critical value \beta \approx -0.1988, corresponding to m \approx -0.0904, the solution exhibits separation, marked by zero wall shear f''(0) = 0, beyond which no physically meaningful attached flow solutions are possible. This threshold was determined through numerical integration of the ODE, highlighting the onset of reverse flow in adverse gradient boundary layers. For \beta < 0, the velocity profiles show inflection points and reduced skin friction compared to the zero-pressure-gradient Blasius case (\beta = 0), underscoring the destabilizing effect of adverse gradients on laminar boundary layers.

Momentum Integral Methods

Momentum integral methods offer approximate solutions to the boundary layer equations by integrating the momentum equation across the layer thickness, providing a practical framework for predicting boundary layer behavior under adverse pressure gradients without solving the full partial differential equations. These techniques assume a form for the velocity profile and use integral quantities like momentum and displacement thicknesses to model flow development. The approach is particularly useful for engineering calculations where exact solutions are intractable, balancing simplicity with reasonable accuracy for separation onset predictions. The von Kármán momentum integral equation forms the core of these methods, derived by integrating the streamwise boundary layer momentum equation from the wall to the edge of the layer. It expresses the rate of change of momentum thickness θ along the streamwise direction x as: \frac{d\theta}{dx} + \left(2 + \frac{\delta^*}{\theta}\right) \frac{\theta}{U_e} \frac{dU_e}{dx} = \frac{C_f}{2}, where δ* is the displacement thickness, U_e the external flow velocity, and C_f the skin friction coefficient. In the presence of an adverse pressure gradient (dU_e/dx < 0), the second term on the left-hand side promotes growth in θ, reflecting deceleration of the near-wall flow and increased likelihood of separation. This equation requires closure through assumptions on the velocity profile to relate C_f, δ*, and θ. A key diagnostic in these analyses is the shape factor H = δ*/θ, which characterizes the fullness of the velocity profile; for zero pressure gradient flows, H ≈ 2.59, but it increases under adverse gradients as the profile becomes more skewed toward the wall. In the Pohlhausen method, the shape factor H increases under adverse gradients, reaching approximately 4 at separation (when Λ = -12), indicating zero wall shear stress. To achieve closure, the Pohlhausen method employs a quartic polynomial approximation for the dimensionless velocity profile u/U_e across the normalized coordinate η = y/δ: \frac{u}{U_e} = a \eta + b \eta^2 + c \eta^3 + d \eta^4, with coefficients determined by boundary conditions including no-slip at the wall, matching external velocity and zero shear at the edge, and a curvature condition at the wall derived from the boundary layer equation. The pressure gradient influence is captured by the parameter Λ = (δ^2 / ν) (dU_e / dx), which is negative for adverse gradients; separation is predicted when Λ = -12, corresponding to zero wall shear. This polynomial family allows computation of integral thicknesses and skin friction as functions of Λ, enabling numerical integration of the momentum equation to find separation locations. These integral methods excel in their computational efficiency for preliminary design, often yielding separation predictions within 10-20% accuracy compared to exact solutions for laminar flows. The Falkner-Skan similarity solutions provide exact benchmarks against which these approximations can be validated for specific pressure gradient exponents.

Engineering Applications

Aerodynamic Contexts

In external aerodynamic flows, adverse pressure gradients significantly influence the performance of airfoils, particularly on the upper surface aft of the maximum thickness location where flow deceleration imposes a rising pressure that challenges boundary layer attachment. At high angles of attack, this gradient intensifies, causing the boundary layer to separate, which triggers stall and a sudden loss of lift. For conventional airfoils, the steep adverse gradient following the suction peak exacerbates separation vulnerability, limiting the maximum angle of attack before stall onset. Post-separation, the formation of a low-pressure wake behind the airfoil substantially increases pressure drag, as the separated flow creates a pressure imbalance between the forward stagnation region and the rearward wake. This contributes to a sharp rise in the total drag coefficient C_D, often dominating over skin friction drag in stalled conditions; for instance, at separation extending 0.2 chord lengths forward of the trailing edge, pressure drag can comprise up to 69% of the total drag. Turbulent boundary layers provide greater resistance to such gradients, helping to delay separation and stall compared to laminar ones. High-lift devices like flaps modify the pressure distribution around airfoils to mitigate adverse gradients and postpone separation, thereby enhancing lift for takeoff and landing. Slotted flaps, in particular, energize the boundary layer through the slot gap, reducing the gradient severity on the main airfoil's aft section and allowing higher maximum lift coefficients C_{L\max}. Historical NACA tests in the 1930s, conducted in variable-density wind tunnels, systematically quantified these gradient effects on C_{L\max} for various airfoil-flap combinations, demonstrating increases of up to 90% in lift with slotted configurations over plain airfoils.

Internal Flow Scenarios

In internal flow scenarios, such as those in pipes, ducts, and channels, adverse pressure gradients typically arise from geometric expansions or directional changes that decelerate the , converting kinetic energy to but risking boundary layer separation and reduced system efficiency. These gradients, where static increases in the flow direction (dP/dx > 0), can lead to recirculation zones and energy dissipation through , particularly in pressure-driven systems like HVAC ducts or industrial . Unlike external flows, the confined nature amplifies the effects, as separated can cause significant blockage and losses, impacting overall performance metrics like head loss or flow uniformity. Diffuser flows exemplify this phenomenon, where an increasing cross-sectional area inherently imposes an to recover from decelerating . This deceleration thickens the , and if the divergence angle exceeds 7-10 degrees, often occurs, leading to blockage and . For instance, in conical or two-dimensional diffusers, separation initiates near due to the inability of low-momentum to overcome the rising , resulting in eddy formation and reduced effective flow area. Such separation is more pronounced in laminar regimes but can be delayed in turbulent flows, though excessive angles still cause and up to 50% loss in efficiency. In pipe expansions, sudden enlargements create a severe adverse as the flow transitions from high-velocity in the smaller section to low-velocity in the larger one, prompting the main to continue downstream while separates into turbulent eddies. This separation forms recirculating zones immediately after the expansion, dissipating and generating losses equivalent to the excess of the incoming flow. The effect, observed as the jet contracts slightly before expanding, exacerbates the gradient, with head losses calculated as h_L = (V_1 - V_2)^2 / (2g), where V_1 and V_2 are upstream and downstream velocities. These losses can account for 20-30% of total in multi-component systems, underscoring the need for gradual transitions to minimize inefficiency. Heat exchanger applications further highlight the role of adverse pressure gradients in bends and valves, where sharp curvatures or restrictions induce localized deceleration, promoting and recirculation that disrupts uniform distribution. In pipe bends, the gradient causes secondary flows and separation bubbles on the inner wall, while valves introduce abrupt expansions or contractions that amplify the effect, leading to formation and eddy shedding. This separation reduces efficiency by up to 25% through uneven profiles and increased drops, as turbulent mixing in recirculation zones, though enhancing local transfer near reattachment, overall elevates pumping costs and lowers effectiveness. Studies show peak coefficients 4 times higher than fully developed occur 1.25-2.5 diameters downstream of separation points, but at the expense of substantial penalties. A key performance metric for evaluating these effects in internal flows, particularly diffusers, is the pressure recovery coefficient C_p, defined as
C_p = \frac{P_\text{out} - P_\text{in}}{P_0 - P_\text{in}}
where P_out and P_in are static pressures at outlet and inlet, and P_0 is the total ( at inlet. Strong adverse pressure gradients that induce separation significantly lower C_p, often to below 0.5 for angles exceeding optimal limits, reflecting incomplete kinetic-to-static energy conversion and heightened losses from blockage. Ideal C_p approaches 1 - 1/AR (area ratio) without separation, but real values drop due to gradient-induced nonuniformities.

Mitigation Strategies

Delay of Separation Techniques

Boundary layer control techniques aim to delay separation by energizing the low-momentum fluid near the wall, thereby enhancing its ability to resist the decelerating effects of an . This principle involves introducing additional into the to counteract the caused by the , preventing of near . Such methods are grounded in the recognition that separation occurs when the near-wall fluid lacks sufficient to overcome the imposed increase. Active control through or blowing represents a primary approach to achieve this energization. removes low-speed from the adjacent to the wall, thinning the layer and increasing its overall , which allows it to better withstand adverse gradients. Conversely, blowing injects high-speed tangentially at the wall, directly adding to the low-energy regions and re-energizing the to postpone separation. These techniques have been demonstrated to extend attached significantly, with often requiring lower energy input for laminar layers while blowing proves effective in turbulent conditions. Plasma actuators provide a non-mechanical alternative for separation delay by generating electromagnetic body forces that induce flow without moving parts. These devices, typically (DBD) actuators, create ionized air () through high-voltage AC excitation, producing a localized that accelerates neutral air molecules near the surface. This induces a virtual aerodynamic shape that modifies the local , effectively reducing the adverse pressure rise and promoting momentum transfer into the . Experimental studies have shown plasma actuators capable of reattaching separated flows under strong gradients, with pulsed operation often yielding higher efficiency than steady modes. The Stratford criterion offers a theoretical for predicting incipient separation under adverse gradients, serving as a for evaluating effectiveness. Developed for turbulent boundary layers, it defines a distribution where the skin friction approaches zero, marking the onset of separation based on inner and outer layer approximations. This , expressed through a relation involving the local and recovery factor, helps quantify the gradient strength at which interventions must act to prevent . Passive devices such as vortex generators can complement these active methods by mixing high-momentum outer flow into the boundary layer.

Practical Engineering Examples

In aviation, vortex generators are small, low-profile fins mounted on aircraft wings to counteract adverse pressure gradients that lead to boundary layer separation. By inducing streamwise vortices, these devices mix high-momentum free-stream air into the low-momentum boundary layer near the surface, thereby delaying flow separation and enhancing lift at high angles of attack. This results in an increase in the maximum lift coefficient (C_Lmax) by approximately 10-20%, depending on the wing configuration and placement, allowing for improved stall characteristics and safer low-speed handling. The dimpled surface of a exemplifies passive flow control in , where the indentations promote an early transition from laminar to turbulent s. This turbulent layer better withstands the adverse pressure gradient on the ball's surface, reducing the wake size and pressure compared to a smooth . Consequently, is lowered by up to 50% at typical flight Reynolds numbers, enabling the ball to travel farther with less energy loss. This benefit arises from the superior tolerance of turbulent flows to adverse pressure gradients, as noted in boundary layer theory. Serrated trailing edges on blades serve as a bio-inspired modification to manage adverse pressure gradients in the blade wakes, where pressure recovery can induce separation and turbulence. The serrations disrupt large-scale coherent structures in the , mitigating broadband aerodynamic noise generation while preserving or slightly enhancing lift-to-drag ratios. Studies show noise reductions of 4-6 dB(A) in key frequency bands, with minimal impact on overall output, though some optimized designs report slight gains of less than 1%. In , ground-effect diffusers on race cars, such as those in Formula 1, are contoured underbody components that exploit proximity to the road to control adverse pressure gradients. By gradually expanding the high-speed underbody flow, the diffuser recovers pressure without inducing separation, generating significant —often equivalent to 2-3 times the car's weight at high speeds—while minimizing penalties. This design prevents boundary layer detachment in the diffuser ramp, ensuring stable aerodynamic grip during cornering.

References

  1. [1]
    Adverse Pressure Gradient - an overview | ScienceDirect Topics
    Adverse pressure gradient is defined as a condition in fluid flow where the pressure increases in the direction of the flow, which can lead to flow separation.
  2. [2]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Dec 15, 2021 · In going from point B to C, however, the fluid particle experiences an adverse pressure gradient that acts to decelerate the fluid particle. If ...
  3. [3]
    Boundary Layer Flows – Introduction to Aerospace Flight Vehicles
    Beyond the minimum pressure point, an adverse pressure gradient develops over the airfoil, leading to boundary-layer transition to turbulence.
  4. [4]
    [PDF] Hermann Schlichting (Deceased) Klaus Gersten Ninth Edition
    However a new fifth section on numerical methods in boundary–layer theory has been added. The partition into chapters had to be somewhat modified in order to.
  5. [5]
    [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). Figures from Çengel and ...<|control11|><|separator|>
  6. [6]
    Dynamic Pressure
    r * u * du/dx = - dp/dx. where r is the density of the gas, p is the pressure, x is the direction of the flow, and u is the velocity in the x direction.
  7. [7]
    [PDF] Falkner-Skan Solutions to Laminar Boundary Layer Equations
    We note that m < 0 corresponds to a locally decelerating external flow corresponding to an “adverse pressure gradient” whereas m > 0 ... 2(m + 1)F d. 2. F dη.
  8. [8]
    [PDF] Ludwig Prandtl's Boundary Layer
    The type of external inviscid flow that promotes. boundary-layer separation is a flow that produces an ad- verse pressure gradient—in other words, an ...
  9. [9]
    Boundary layer separation and pressure drag - Galileo
    Separation is bound to occur in a sufficiently large adverse pressure gradient. On the other hand, boundary layers like decreasing pressure gradients, which ...
  10. [10]
    [PDF] 7.3. Separation - Aerospace Computing Lab
    The presence of an adverse pressure gradient (increasing pressure) causes a deceleration of the fluid. ... Where the constant S is 0.35 when d p/dx < 0 (concave ...
  11. [11]
    [PDF] Transition and Turbulent Boundary Layers AA200b Lecture 9 ...
    Feb 3, 2005 · For low Reynolds numbers (typically Re < 3 × 105), and for a flat plate at zero pressure gradient, the disturbances are damped and the flow ...
  12. [12]
    Effects of Adverse Pressure Gradients on the Nature and Length of ...
    For an adverse pressure gradient, the Tollmien–Schlichting waves appear more regularly with a well-defined spectral peak. As the adverse pressure gradient is ...
  13. [13]
    The behaviour of turbulence production in adverse pressure ...
    May 7, 2025 · This paper investigates the behaviour of turbulence production in adverse pressure gradient (APG) turbulent boundary layers (TBLs).
  14. [14]
  15. [15]
    Momentum-Integral Equation - an overview | ScienceDirect Topics
    M in = ∫ 0 δ ρ u 2 d y . This is the momentum-integral equation first derived by von Karman. Since no assumption is made at this stage about the relationship ...
  16. [16]
    Über laminare und turbulente Reibung - Kármán - Wiley Online Library
    Über laminare und turbulente Reibung. Th. V. Kármán,. Th. V. Kármán. Aachen ... First published: 1921. https://doi.org/10.1002/zamm.19210010401. Citations ...
  17. [17]
    [PDF] jen92a2.tmp
    Realizing the limitations of the Von Karman momentum equation, these investigators summed the mean- flow terms and the wall shear stress and then subtracted ...
  18. [18]
    Zur näherungsweisen Integration der Differentialgleichung der ...
    Pohlhausen,. K. Pohlhausen. Aachen. Search for more papers by this author. First published: 1921. https://doi.org/10.1002/zamm.19210010402. Citations: 347.
  19. [19]
    [PDF] NATIONAL FOR AERONAUTICS
    In some of these methods, for e-le, the Pohlhausen method (reference 3) and the. Wieghardt method (reference 4), a functional form is chosen for the veloc- ity ...
  20. [20]
    [PDF] unsteady airfoil stall
    ... stall. As the pressure gradient grows more and more adverse (with increasing a) the spots grow in number and size, start to coalesce, and finally ...
  21. [21]
    [PDF] dynamic stall characteristics of drooped leading edge airfoils
    The conventional airfoil has a high suction peak, and a rather steep adverse pressure gradient and is more prone to stall. c) At higher angles of attack, (ct = ...
  22. [22]
    [PDF] NASA Technical Memorandum 102722
    At an angle of attack of 18°,where the separation has reached 0.2 chord ahead or the trailing edge, the pressure drag has reached 69 percent of the total.
  23. [23]
    [PDF] Modification of the MML Turbulence Model for Adverse Pressure ...
    for an airfoil near stall,. S modifications are needed to make it applicable to general boundary layer flows. In order to make MML applicable over a range of ...
  24. [24]
    [PDF] c"_ /._- Design Methodology for Multi-Element High-Lift Systems on ...
    Slotted flaps offer improved efficiency over plain flaps due to a delay in flow separation. ... adverse pressure gradients than thick ones before separating.
  25. [25]
    [PDF] High-Lift Systems on Commercial Subsonic Airliners
    ... flap slot reduces the adverse pressure gradient on the aft portion of the main wing. The deployment and strength of the trailing-edge flaps also have an.
  26. [26]
    [PDF] N88-23739 - NASA Technical Reports Server (NTRS)
    A review of the. NACA and. NASA low-drag airfoil research is presented with particular emphasis given to the development of mechanical high-lift flap systems.
  27. [27]
    [PDF] NATIONAL TECHNICAL INFORMATION SERVICE
    Pressure-distribution tests of an N. A. C. A. 23012 airfoil equipped with a slotted flap and with a plain flap were made in the 7- by lO-foot wind tunnel. A ...
  28. [28]
    [PDF] Performance Prediction of Straight Two•Dimensional Diffusers
    COMPARISON OF IDEAL, CALCULATED, AND MEASURED C. We follow the standard definition of pressure recovery coefficient, Cp. Pe - pi. pC=p- where pe and pl are ...Missing: formula | Show results with:formula
  29. [29]
    [PDF] MASTER - OSTI.GOV
    Diffusion at angles greater than these produces flow instabilities; witp angles greater than 3 degrees, random flow separation and · reattachment was observed. ...
  30. [30]
    Losses Due to Sudden Enlargement - NPTEL Archive
    The fluid flows against an adverse pressure gradient. · The fluid particles near the wall due to their low kinetic energy cannot overcome the adverse pressure ...
  31. [31]
    Heat Transfer to Separation Flow in Heat Exchangers - IntechOpen
    Oct 31, 2012 · Adverse pressure gradient and viscosity are the two governing factors of flow separation. By changing or maintaining the structure of viscosity ...
  32. [32]
    Pressure Recovery Coefficient - an overview | ScienceDirect Topics
    The pressure recovery coefficient is defined as the ratio of the total pressure in the test section to the sum of the total pressures at the afterburner and ...
  33. [33]
    [PDF] Boundary-Layer Control - MIT
    Increasing the angle of attack increases the adverse gradient. Such gradients are destabilizing, so that the amount of laminar flow over a structure decreases ...
  34. [34]
    Boundary Layer Suction - an overview | ScienceDirect Topics
    Boundary layer suction refers to the method of removing low-energy air from the boundary layer near a surface to prevent flow separation, thus allowing for ...
  35. [35]
    [PDF] Active control of Boundary Layer Separation & Flow Distortion in ...
    Executive Summary. Inlets to aircraft propulsion systems must supply flow to the compressor with minimal pressure loss, flow distortion or unsteadiness.
  36. [36]
    Turbulent boundary-layer control with plasma actuators - Journals
    Apr 13, 2011 · This paper reviews turbulent boundary-layer control strategies for skin-friction reduction of aerodynamic bodies. The focus is placed on the ...
  37. [37]
    [PDF] INVESTIGATING ACTIVE VORTEX GENERATORS AS A NOVEL ...
    The induced energy stabilizes the boundary layer, keeps the flow attached and delays separation to higher angles of attack. ... increase in CLmax of 14%, a rise ...
  38. [38]
    Mechanism of drag reduction by dimples on a sphere - AIP Publishing
    Apr 17, 2006 · There are two most interesting flow phenomena about golf-ball dimples.1 One is that dimples reduce drag on a sphere as much as 50% as compared ...
  39. [39]
    [PDF] Survey of Techniques for Reduction of Wind Turbine Blade Trailing ...
    The serrations resulted in noise reductions of 4-6 dB(A) at 500 Hz, 2-4 dB(A) at 1000 Hz, and marginal reduction at 2000 Hz. While experiments have indicated ...
  40. [40]
    [PDF] Integrating machine learning and CFD for enhanced trailing edge ...
    Oct 9, 2025 · One effective way to improve the aerodynamic efficiency of wind turbine blade airfoils is the addition of serrated trailing edges, which can ...<|separator|>
  41. [41]
    A Review of Ground-Effect Diffuser Aerodynamics | J. Fluids Eng.
    The ground-effect diffuser generates downforce by creating a suction effect underneath a racing car. When the high-velocity airflow travels underneath the floor ...
  42. [42]
    [PDF] Aerodynamics and Experimental Optimisation of Automotive ...
    Therefore, as the diffuser angle is increased, so is the area ratio and the resulting adverse pressure gradient, causing flow separation at a higher ride ...<|control11|><|separator|>