Fact-checked by Grok 2 weeks ago

Blasius boundary layer

The Blasius boundary layer describes the steady, two-dimensional, incompressible laminar flow over a semi-infinite flat plate aligned with a uniform free-stream velocity U, where viscous effects create a thin layer of retarded fluid near the plate surface. This boundary layer forms due to the no-slip condition at the plate, with the flow transitioning from zero velocity at the surface to the free-stream value asymptotically far from it, under the assumptions of constant viscosity, negligible pressure gradient, and high Reynolds number to justify the boundary layer approximation. The solution, derived via a similarity transformation, reduces the partial differential equations of the Navier-Stokes system to a single nonlinear ordinary differential equation, providing the first exact analytical-numerical profile for such a configuration. Named after Heinrich Blasius, who published the seminal work in 1908 as part of Ludwig Prandtl's theory, the addressed the limitations of full Navier-Stokes equations for high-Reynolds-number flows by focusing on the thin viscous region near solid boundaries. Blasius's approach built on Prandtl's 1904 concept of s, applying mechanical similitude to flows with small , and was motivated by problems like drag on plates and around bodies such as cylinders. The work involved using series expansions and asymptotic matching, as no closed-form exists, establishing a for laminar behavior. Mathematically, the equations for and are \frac{\partial u}{\partial x} + \frac{\partial v}{\partial y} = 0 and u \frac{\partial u}{\partial x} + v \frac{\partial u}{\partial y} = \nu \frac{\partial^2 u}{\partial y^2}, with conditions u = v = 0 at y = 0, u \to U as y \to \infty, and u = U at x = 0. Introducing the \psi such that u = \frac{\partial \psi}{\partial y} and v = -\frac{\partial \psi}{\partial x}, and the similarity variable \eta = y \sqrt{\frac{U}{ \nu x}} (or scaled variants), yields the Blasius f''' + \frac{1}{2} f f'' = 0, where f(\eta) is a dimensionless satisfying f(0) = f'(0) = 0 and f'(\infty) = 1. The profile is \frac{u}{U} = f'(\eta), solved numerically with f''(0) \approx 0.332. Key results include the boundary layer thickness \delta \approx 5.0 \sqrt{\frac{\nu x}{U}} (defined at 99% of U), displacement thickness \delta_1 = 1.721 \sqrt{\frac{\nu x}{U}}, and momentum thickness \delta_2 = 0.664 \sqrt{\frac{\nu x}{U}}. The local skin friction coefficient is c_f = \frac{0.664}{\sqrt{\mathrm{Re}_x}}, where \mathrm{Re}_x = \frac{U x}{\nu}, leading to average drag coefficient C_D = \frac{1.328}{\sqrt{\mathrm{Re}_L}} for a plate of length L. These scalings highlight the diffusive growth of the layer proportional to \sqrt{x}. The Blasius solution remains foundational in , serving as a reference for validating computational methods, approximating more complex flows, and understanding transition to in zero-pressure-gradient boundary layers. It underpins applications in , such as design and , and extends to similar problems like Falkner-Skan flows with pressure gradients.

Boundary Layer Fundamentals

Prandtl's Boundary Layer Equations

The concept, foundational to understanding viscous flows at high Reynolds numbers, was introduced by in his 1904 presentation at the Third in . Prandtl recognized that in fluids with low viscosity relative to inertial forces—characterized by large Reynolds numbers—the effects of friction are not uniformly distributed but are instead confined to a thin region adjacent to the solid surface. This accommodates the at the wall while allowing the flow to transition smoothly to an inviscid outer flow, resolving the between ideal fluid predictions and experimental drag observations. The Prandtl boundary layer approximation relies on several key assumptions for high-Reynolds-number flows: the δ is much smaller than the streamwise length scale L (δ ≪ L), making the layer geometrically thin; streamwise of is negligible compared to transverse near the wall; and inertial terms dominate viscous terms in the outer part of the layer, while viscous effects balance inertia close to the surface. These assumptions hold because the large Re = UL/ν implies that viscous forces are significant only over short transverse distances, leading to a disparity in diffusion scales. To derive the equations, begin with the two-dimensional, steady, incompressible Navier-Stokes equations in Cartesian coordinates, where u and v are the streamwise and transverse components, respectively, p is , ρ is , and ν is kinematic :
  • Continuity equation: \frac{\partial u}{\partial x} + \frac{\partial v}{\partial y} = 0
  • Streamwise momentum equation: u \frac{\partial u}{\partial x} + v \frac{\partial u}{\partial y} = -\frac{1}{\rho} \frac{\partial p}{\partial x} + \nu \left( \frac{\partial^2 u}{\partial x^2} + \frac{\partial^2 u}{\partial y^2} \right)
  • Transverse momentum equation: u \frac{\partial v}{\partial x} + v \frac{\partial v}{\partial y} = -\frac{1}{\rho} \frac{\partial p}{\partial y} + \nu \left( \frac{\partial^2 v}{\partial x^2} + \frac{\partial^2 v}{\partial y^2} \right)
Under the boundary layer assumptions, the transverse pressure gradient is negligible (\partial p / \partial y \approx 0), so pressure varies only in the streamwise direction and is determined by the outer inviscid flow via Bernoulli's principle: U_\infty \, dU_\infty / dx = - (1/\rho) \, dp/dx, where U_\infty(x) is the outer flow velocity. Additionally, streamwise viscous diffusion terms (\partial^2 u / \partial x^2 and \partial^2 v / \partial x^2) are small compared to transverse terms (\partial^2 u / \partial y^2), as their ratio scales with δ²/L² ∼ 1/Re ≪ 1. The resulting Prandtl boundary layer equations simplify to:
  • Continuity: \frac{\partial u}{\partial x} + \frac{\partial v}{\partial y} = 0
  • Streamwise : u \frac{\partial u}{\partial x} + v \frac{\partial u}{\partial y} = U_\infty \frac{d U_\infty}{dx} + \nu \frac{\partial^2 u}{\partial y^2}
The transverse equation reduces to hydrostatic balance. These approximations are rigorously justified by nondimensionalizing the Navier-Stokes equations using scales x ∼ L, y ∼ δ ∼ L / √, u ∼ U, v ∼ U δ / L ∼ U / √, and p ∼ ρ U². In the streamwise equation, the inertial and transverse viscous terms remain O(1), while the streamwise viscous diffusion becomes O(1/) and can be neglected for ≫ 1; the transverse equation similarly shows that is constant across the layer to leading order. This framework applies generally to external flows with imposed gradients, including the zero--gradient case over a flat plate where dU_\infty / dx = 0.

Flat-Plate Boundary Layer Setup

The Blasius boundary layer problem considers the steady, two-dimensional of an incompressible over a semi-infinite flat plate aligned at zero incidence to the oncoming flow. The plate extends from the at x=0 to x=∞ in the streamwise direction, with a uniform free-stream U_∞ parallel to the plate surface. The flow is governed by Prandtl's boundary layer equations, which approximate the Navier-Stokes equations under the thin-layer assumption. This setup models the development of viscous effects near the plate while the outer flow remains essentially inviscid. The essential boundary conditions are the at the wall (y=0), where the streamwise velocity u=0 and the normal velocity v=0, and matching to the free stream as y→∞, where u→U_∞ and v→0. At the (x=0), the boundary layer thickness is zero, and it develops downstream as diffuses perpendicularly from the wall into the free stream. The local Reynolds number, defined as Re_x = U_∞ x / ν with ν the kinematic , serves as the characteristic scale for the flow regime; the laminar approximation holds for Re_x ≲ 5×10^5, after which transition to typically occurs due to amplification. Key physical quantities include the skin friction, represented by the wall shear stress τ_w = μ (∂u/∂y){y=0} where μ is the dynamic , which quantifies the tangential force per unit area exerted by the fluid on the plate. The thickness δ^* = ∫0^∞ (1 - u/U∞) dy measures the effective outward shift of the inviscid streamlines due to deficit in the . The δ, conventionally taken as the distance where u reaches 99% of U∞, grows proportionally as δ ∼ √(ν x / U_∞), reflecting the diffusive nature of viscous along the plate.

The Blasius Solution

Derivation of the Blasius Equation

The derivation of the Blasius equation begins with the motivation to find self-similar solutions to the Prandtl boundary layer equations for over a flat plate, where the velocity profiles are expected to scale with the local inverse square root of the , Re_x^{-1/2}, rendering them independent of the streamwise position x. This similarity assumption is justified by the growing as δ(x) ≈ √(νx / U_∞), where ν is the kinematic and U_∞ is the free-stream , allowing a dimensionless transverse coordinate to collapse the profiles. For a flat plate aligned with the free stream, the zero (dp/dx = 0) is crucial, as it ensures the outer imposes constant pressure across the , eliminating explicit x-dependence in the governing equations and enabling exact similarity transformation. To achieve this, Blasius introduced a similarity variable η and a ψ defined as \eta = y \sqrt{\frac{U_\infty}{\nu x}}, \quad \psi = \sqrt{\nu x U_\infty} \, f(\eta), where y is the transverse coordinate perpendicular to the plate, and f(η) is a dimensionless function to be determined. The velocity components are then expressed using the stream function properties u = ∂ψ/∂y and v = -∂ψ/∂x, yielding u = U_\infty f'(\eta), \quad v = \frac{1}{2} \sqrt{\frac{\nu U_\infty}{x}} \left( \eta f'(\eta) - f(\eta) \right). These expressions satisfy the continuity equation ∂u/∂x + ∂v/∂y = 0 automatically due to the form of ψ. Substituting these into the streamwise momentum equation for the boundary layer, u ∂u/∂x + v ∂u/∂y = ν ∂²u/∂y² (with dp/dx = 0), involves computing the partial derivatives. First, ∂u/∂x = - (U_∞ η f''(η)) / (2x) and ∂u/∂y = U_∞ f''(η) √(U_∞ / (ν x)), while ∂²u/∂y² = U_∞ f'''(η) (U_∞ / (ν x)). Plugging in and simplifying, the left-hand side becomes - (U_∞² / (2x)) f f''(η), and the right-hand side is (U_∞² / x) f'''(η). Equating and multiplying through by 2x / U_∞² yields f'''(η) + (1/2) f(η) f''(η) = 0. This third-order nonlinear governs the similarity function f(η). The boundary conditions in the similarity variables reflect the physical constraints at the wall and far from it: at the no-slip wall (η = 0), f(0) = 0 (zero normal ) and f'(0) = 0 (zero tangential ); asymptotically, as η → ∞, f'(∞) = 1 to match the free-stream U_∞. These conditions, combined with the zero , ensure the solution is uniquely determined and applicable to the entire flat-plate downstream of the .

Similarity Transformation and Solution

The Blasius equation, obtained through the of Prandtl's boundary layer equations for steady, over a flat plate, is a third-order nonlinear given by f'''(\eta) + \frac{1}{2} f(\eta) f''(\eta) = 0, subject to the boundary conditions f(0) = 0, f'(0) = 0, and f'(\infty) = 1, where \eta = y \sqrt{U_\infty / (\nu x)} is the similarity variable, f(\eta) is the dimensionless , and primes denote differentiation with respect to \eta. This lacks a closed-form analytical , so Blasius originally employed a expansion around \eta = 0 to approximate the behavior near , assuming f(\eta) = \sum_{n=0}^\infty a_n \eta^n and recursively determining coefficients from and initial conditions. More accurate results require numerical methods; Howarth integrated step-by-step from outward, adjusting the initial guess for f''(0) iteratively to satisfy the far-field condition, a precursor to modern shooting techniques that convert the into initial value problems solved via Runge-Kutta or similar integrators. The velocity profile emerges as the derivative u / U_\infty = f'(\eta), which approaches 1 asymptotically. The boundary layer thickness \delta, conventionally defined at the 99% velocity point where u = 0.99 U_\infty, corresponds to \eta \approx 5, yielding \delta / x \approx 5 / \sqrt{\mathrm{Re}_x} with \mathrm{Re}_x = U_\infty x / \nu. The wall shear stress is \tau_w = \mu \left( \frac{\partial u}{\partial y} \right)_{y=0} = 0.332 \mu U_\infty^{3/2} (\nu x)^{-1/2}, leading to the local skin friction coefficient c_f = \tau_w / (\frac{1}{2} \rho U_\infty^2) = 0.664 / \sqrt{\mathrm{Re}_x}, where the constant 0.332 arises from the numerical value f''(0) \approx 0.332. From the similarity solution, the momentum thickness is \theta / x = 0.664 / \sqrt{\mathrm{[Re](/page/Re)}_x}, obtained by integrating \theta = \int_0^\infty (u/U_\infty) (1 - u/U_\infty) \, dy = \left[ \int_0^\infty f' (1 - f') \, d\eta \right] / \sqrt{\mathrm{[Re](/page/Re)}_x / x}. For a flat plate of length [L](/page/L'), the average skin friction coefficient over one side is the integral of the local value, giving C_f = 1.328 / \sqrt{\mathrm{[Re](/page/Re)}_L}, which determines the total drag force D = \frac{1}{2} \rho U_\infty^2 b L C_f for plate width b. Far from the wall, the solution exhibits asymptotic behavior f(\eta) \sim \eta - \beta as \eta \to \infty, where \beta \approx 1.721 quantifies the displacement effect, corresponding to the displacement thickness \delta^* / x = \beta / \sqrt{\mathrm{Re}_x}.

Uniqueness and Properties of the Solution

The and of the solution to the , satisfying the conditions f(0) = f'(0) = 0 and f'(\infty) = 1, were first rigorously established by Weyl using techniques from analytic function theory, including a transformation to demonstrate a unique shooting parameter f''(0). A simpler proof was later provided by Serrin, employing the von Mises transformation to reduce the third-order to a second-order form, followed by application of the to show that no two distinct solutions can coexist while satisfying the conditions. These results confirm that there is precisely one physical describing the steady laminar over a flat plate. The solution exhibits monotonicity properties essential to its physical interpretation: the dimensionless stream function f(\eta) is strictly increasing with \eta, while the velocity component f'(\eta) monotonically increases from 0 at the wall to 1 in the freestream, reflecting the smooth acceleration from no-slip to free-stream conditions. The wall shear derivative f''(\eta) remains positive for all \eta \geq 0, starting at f''(0) \approx 0.332 and asymptotically approaching 0, which ensures the velocity profile is concave upward everywhere without an . This absence of an aligns with Rayleigh's theorem, implying inviscid stability of the profile to certain disturbances, though viscous effects introduce at higher Reynolds numbers. Linear stability analysis of the Blasius profile reveals vulnerability to Tollmien-Schlichting (TS) waves, the primary mechanism for laminar-turbulent transition in boundary layers. The Orr-Sommerfeld equation, derived from a normal-mode decomposition of small perturbations, yields a neutral curve with a critical Re_{\delta^*} \approx 520 (based on displacement thickness), below which all disturbances decay and above which TS waves with specific and amplify. Seminal computations by Tollmien and Schlichting identified the least stable mode at a critical \alpha \approx 0.038 and frequency parameter F \approx 2.3 \times 10^{-3}, marking the onset of . The lack of an precludes inviscid modes but permits these viscous TS instabilities, which drive transition when amplification reaches order-one amplitudes. Experimental validations, such as those by Schubauer and Skramstad using hot-wire anemometry on a flat plate, confirm the theoretical velocity profiles u/U_e = f'(\eta) with close agreement in the inner layer (up to \eta \approx 5) and capture the growth of waves consistent with linear theory, including the critical for observable oscillations. These measurements underscore the practical relevance of the Blasius solution, showing deviations only near due to nonlinear effects or external disturbances.

Higher-Order and Perturbation Extensions

Second-Order Boundary Layer Corrections

The second-order boundary layer corrections extend the Prandtl by accounting for higher-order effects, primarily the streamwise viscous diffusion term \partial^2 u / \partial x^2 in the momentum , which is of order $1/\mathrm{Re} relative to the leading terms. The velocity field is expanded as a series u = u_1 + \epsilon u_2 + O(\epsilon^2), where u_1 is the Blasius solution and \epsilon \approx \mathrm{Re}_x^{-1/2} represents the small parameter based on the scaling. Similarly, expansions are applied to v, p, and the \psi, with the second-order terms satisfying boundary conditions at the wall and matching the outer flow. This approach improves accuracy for finite Reynolds numbers, particularly near the where the breaks down more noticeably. For the semi-infinite flat plate with zero , the second-order theory, as developed by researchers like , yields no correction to the classical Blasius skin friction coefficient, as the displacement thickness does not induce a and streamwise diffusion effects balance without altering the wall shear to this order. The governing equations for the second-order quantities arise from substituting the expansions into the full Navier-Stokes equations and collecting terms at O(\epsilon^2). The remains zero to second order, and the solution preserves a self-similar structure away from the , though modified near x=0 due to the singular nature of the . Numerical methods, such as series expansions and asymptotic matching, confirm that velocity profiles u_2(\eta) contribute negligibly to wall shear for high \mathrm{Re}_x. These extensions provide insight into the validity of the approximation but do not change drag predictions at this order for the semi-infinite case.

Third-Order Boundary Layer Corrections

The third-order boundary layer corrections refine the Blasius solution by extending the asymptotic to include terms of order \epsilon^2, where \epsilon = \mathrm{Re}_x^{-1/2} represents the small parameter based on the local \mathrm{Re}_x. This approach approximates the full Navier-Stokes equations for the over a semi-infinite flat plate, capturing viscous effects neglected in the leading-order Prandtl equations, such as streamwise of and the normal across the layer. The streamwise velocity component is expanded as u = u_1(\eta) + \epsilon u_2(\eta) + \epsilon^2 u_3(\eta) + \cdots, where \eta = y \sqrt{U_\infty / (\nu x)} is the , u_1(\eta) is the classical Blasius profile satisfying the zero-pressure-gradient , and u_2(\eta) = 0 for the semi-infinite plate due to the absence of a second-order effect. The function u_3(\eta) arises from substituting the expansion into the and at order \epsilon^2, yielding a linear inhomogeneous that incorporates higher-order viscous diffusion terms and interactions with the leading-order solution. This third-order equation includes contributions from the streamwise \partial^2 u / \partial x^2, which introduces transverse effects relative to the scaling, as well as higher moments of the profile that account for non-similarity in the streamwise direction. Analytical progress is limited, but numerical solutions for u_3(\eta) have been computed using series truncation methods on the full Navier-Stokes formulation in , confirming the series converges for moderate Reynolds numbers and providing the correction profile across the . The resulting corrections modify the at the wall, refining the local skin friction coefficient C_f by terms of order O(\mathrm{Re}_x^{-3/2}) beyond the leading-order Blasius value of $0.664 \mathrm{Re}_x^{-1/2}. These adjustments enhance predictions of total in laminar boundary layers, particularly for flows at finite Reynolds numbers where higher-precision modeling is needed to assess proximity to or low-Reynolds-number regimes.

Variations with Mass Transfer

Blasius Boundary Layer with Suction

The Blasius boundary layer with modifies the classical flat-plate problem by introducing wall modeled through a constant similarity parameter in the transformation, leading to a normal at the wall v_w = -\frac{1}{2} f_w \sqrt{\frac{U_\infty \nu}{x}} < 0, where f_w = f(0) > 0 is the parameter, while the transverse v \to 0 as y \to \infty. This setup applies to flows over porous surfaces where extraction influences development, with the decreasing as x^{-1/2} to maintain self-similarity. The streamwise free-stream remains U_\infty, and the flow is steady, incompressible, and laminar. The employed is identical to the standard Blasius case, with the \psi = \sqrt{\nu x U_\infty} \, f(\eta) and \eta = y \sqrt{U_\infty / (\nu x)}. The boundary conditions are altered to account for : f(0) = f_w > 0, f'(0) = 0, and f'(\infty) = 1. These conditions reflect the at the wall while preserving no-slip in the streamwise direction. The governing retains the Blasius form, f''' + \frac{1}{2} f f'' = 0, but the nonzero wall boundary condition on f(0) results in a family of solutions parameterized by the strength f_w, solved numerically via methods or series expansions. This adjustment confines low-momentum near the wall, yielding profiles that approach the free stream more rapidly with increasing f_w. significantly reduces the compared to the no- case, where \delta \sim 5 \sqrt{\nu x / U_\infty}; for moderate , the thickness scales as a reduced multiple of this value. This thinning effect enhances momentum transfer and elevates skin friction. Moreover, the fuller profiles stabilize the flow, increasing the critical for to from approximately 520 in the Blasius case to up to around 41,000 for strong , thereby delaying instability onset. In practical applications, the Blasius boundary layer with serves as a foundational model for active flow control, particularly in to mitigate drag on wings and fuselages via porous suction panels, which can reduce overall skin-friction drag by 20-30% while preventing separation. Such techniques have been explored since the mid-20th century for high-lift devices and laminar flow maintenance.

Von Mises Transformation for Suction

The Von Mises transformation, originally developed by in 1927 for analyzing general two-dimensional incompressible flows, replaces the transverse coordinate y with the \psi as an independent variable, while retaining s = x as the streamwise coordinate. This yields the relation \partial u / \partial \psi = (\partial u / \partial y) / u, effectively recasting the velocity gradient in terms of the stream function. For boundary layers involving constant wall suction with normal velocity v_w < 0, the stream function is adjusted to \psi = \int_0^y u \, dy - v_w x to satisfy the boundary condition at the wall, preserving \partial \psi / \partial y = u. Substituting into the Prandtl momentum equation produces the transformed form u \frac{\partial u}{\partial s} = \nu \frac{\partial}{\partial \psi} \left( u \frac{\partial u}{\partial \psi} \right) + v_w \frac{\partial u}{\partial \psi}, with boundary conditions u(s, 0) = 0 and u(s, \infty) = U_\infty. This formulation offers key advantages by eliminating the explicit appearance of the transverse velocity v from the governing equation, reducing the system to a single partial differential equation in u(s, \psi). It facilitates efficient numerical solutions through streamwise marching procedures, where the \psi-derivatives can be integrated accurately as an initial-value problem at each s-step, improving convergence and precision near the wall where velocity gradients are steep. In the context of the Blasius flat-plate boundary layer with uniform suction, the transformation enables numerical integration that aligns with similarity principles for moderate suction rates, linking solutions to a reduced ordinary differential equation in a similarity variable \eta related to \psi / \sqrt{\nu s}, adjusted by the suction parameter -v_w \sqrt{s / \nu U_\infty}. For vanishing suction (v_w = 0), it precisely recovers the classical u / U_\infty = f'(\eta).

Asymptotic Suction Profile

In the strong suction limit, where the wall suction velocity v_0 significantly exceeds the characteristic transverse velocity scale of the boundary layer without suction, specifically v_0 \gg \sqrt{\nu U_\infty / x}, the boundary layer develops a fully established velocity profile that becomes independent of the streamwise distance x. This regime occurs far downstream or under sufficiently intense uniform suction, preventing the boundary layer from growing and leading to a constant-thickness profile that satisfies the boundary layer equations exactly. The exact velocity solution in this asymptotic state is given by \frac{u}{U_\infty} = 1 - \exp\left( -\frac{v_0 y}{\nu} \right), where u(y) is the streamwise velocity, U_\infty is the free-stream velocity, y is the wall-normal coordinate, \nu is the kinematic viscosity, and v_0 > 0 denotes the magnitude of the inward velocity. This profile arises from integrating the momentum equation under constant suction, resulting in a constant throughout the layer and no streamwise diffusion. The \delta, defined as the distance where u/U_\infty reaches 99% of its free-stream value, is constant and given by \delta \approx 4.5 \nu / v_0, representing the minimal possible thickness for a suction-controlled layer. Similarly, the thickness is \delta^* = \nu / v_0. The wall skin friction \tau_w in this profile is \tau_w = \rho U_\infty v_0, where \rho is the fluid density, and it remains independent of , highlighting the dominance of suction over viscous . Due to its fuller velocity profile and the damping effect of suction, the asymptotic suction boundary layer exhibits high stability, with linear stability analysis yielding a critical Reynolds number (based on displacement thickness) of approximately 54,382—far exceeding the value of about 520 for the standard Blasius profile. Experimental studies confirm no transition to turbulence occurs even at high Reynolds numbers under practical suction conditions, as disturbances decay rapidly without amplification.

Compressible Extensions

Compressible Blasius Boundary Layer

The extension of the Blasius boundary layer solution to flows addresses high-speed where and variations significantly influence the structure, particularly in the presence of effects. The governing equations are the compressible Prandtl boundary layer equations, comprising the \partial(\rho u)/\partial x + \partial(\rho v)/\partial y = 0, the momentum equation \rho(u \partial u / \partial x + v \partial u / \partial y) = -\partial p / \partial x + \partial / \partial y (\mu \partial u / \partial y), and the energy equation \rho c_p (u \partial T / \partial x + v \partial T / \partial y) = \partial / \partial y (\lambda \partial T / \partial y) + \mu (\partial u / \partial y)^2. These incorporate variable \rho(T, p) via the , temperature-dependent viscosity \mu(T) often modeled by Sutherland's law, and effects through in the energy dissipation term and recovery . For a flat plate with zero , a is applied using the variable \eta = y \sqrt{U_\infty / (\nu_\infty x)}, where the is \psi = \sqrt{\nu_\infty x U_\infty} \, f(\eta) and is \theta(\eta) = (T - T_w)/(T_{aw} - T_w). Unlike the incompressible case, the ordinary differential equations now couple the dimensionless f(\eta) with the profile \theta(\eta), resulting in a system solved numerically, as originally formulated by Illingworth. The momentum equation becomes f''' + \frac{1}{2} f f'' = 0 in transformed variables (e.g., Howarth-Dorodnitsyn), but the energy equation introduces additional coupling dependent on the and wall ratio. The Crocco relation holds for Pr=1, relating velocity and linearly; for general Pr ≠1, fully coupled numerical solutions are required. When the Prandtl number Pr = 1, the Crocco integral simplifies the solution by relating velocity and temperature through linear total enthalpy profiles: the total enthalpy H = c_p T + u^2 / 2 is constant across the boundary layer, yielding T = T_w + (T_{aw} - T_w) (u / U_\infty). Here, the adiabatic wall temperature T_{aw} is defined as T_{aw} = T_\infty \left[1 + r \frac{\gamma - 1}{2} M_\infty^2 \right], where r \approx \sqrt{\Pr} is the recovery factor (approximately 0.85 for air in laminar flow with Pr=0.72) and \gamma is the specific heat ratio. This relation, derived by Crocco, decouples the equations and allows direct computation of thermal profiles from velocity data. The dimensionless velocity profile u / U_\infty = f'(\eta) remains qualitatively similar to the incompressible Blasius solution but is scaled by the local density ratio \rho_\infty / \rho to account for compressibility-induced thickening or thinning of the layer, with the \delta \propto \sqrt{\nu x / U_\infty} \cdot (\rho_\infty / \rho_w)^{1/2}. Numerical solutions, such as those by Howarth, confirm that for moderate numbers, the profile shape is preserved but shifted due to temperature gradients. The local skin friction coefficient for the compressible case is c_f = \frac{0.664}{\sqrt{\Rey_x}} \left( \frac{T_w}{T_\infty} \right)^{-1/2}, where \Rey_x = U_\infty x / \nu_\infty, for the case Pr=1 and \mu \propto T. For Sutherland's law (\mu \sim T^{0.76}), numerical corrections are applied, but the similarity solution uses the -1/2 power, validated against experimental data for numbers up to 5. This correction arises from the wall \tau_w = \mu_w (\partial u / \partial y)_w, scaled by and ratios at the wall.

Howarth Transformation

The Howarth transformation provides a method to convert the governing equations of a steady, two-dimensional compressible with zero into an equivalent incompressible form, allowing the use of known solutions like the Blasius profile for the velocity field. Developed by Leslie Howarth in 1948, this approach builds on foundational ideas from Illingworth's work on boundary layer transformations. The transformation defines new coordinates as \bar{x} = x and \bar{y} = \int_0^y \frac{\rho}{\rho_\infty} \, dy', where \rho is density and \rho_\infty is the density at the outer edge of the boundary layer. This density-weighted scaling adjusts the normal coordinate to account for compressibility effects. An approximation often used, particularly when viscosity \mu varies inversely with density (as in \mu \propto T for an ideal gas with Pr=), is the Illingworth transformation \bar{y} \approx y \sqrt{\frac{\rho}{\rho_\infty}}. In the transformed variables, the and equations reduce to the incompressible form, recovering the Blasius f''' + \frac{1}{2} f f'' = 0, where f(\eta) is the dimensionless with similarity \eta \propto \bar{y} / \sqrt{\nu_\infty x / U_\infty}, \nu_\infty is the kinematic at the edge, and U_\infty is the free-stream . This mapping is exact for certain viscosity laws, such as when \mu \propto \rho^{-1} and the is unity, as it eliminates property terms in the . The transformation assumes steady, two-dimensional flow with zero pressure gradient and neglects viscous dissipation in the energy equation for density evaluation. While analytical for the flat-plate Blasius case under these conditions, general compressible profiles require numerical integration of the coupled energy equation to determine density variations before applying the transformation. The Howarth uses the integral form, while the Illingworth approximation simplifies for \mu \propto T.

Alternative Coordinate Solutions

Blasius Solution in Parabolic Coordinates

The Blasius boundary layer solution can be generalized using parabolic coordinates (\xi, \eta), where \xi measures distance along the surface and \eta is perpendicular to it, with scale factors h_\xi = h_\eta = \sqrt{\xi^2 + \eta^2}. This curvilinear system aligns naturally with the parabolic character of the boundary layer equations, facilitating analysis of flows near sharp leading edges or over bodies like parabolic cylinders, where Cartesian coordinates introduce singularities at the origin. To achieve similarity, the transformation employs the variable \zeta = \eta / \sqrt{\xi} and the \psi = \sqrt{\nu \xi U_\infty} \, f(\zeta), where \nu is kinematic and U_\infty is the free-stream . These substitutions collapse the partial differential equations into an . For the flat-plate case in (resolving leading-edge effects), it yields the Blasius f''' + f f'' = 0. This is solved numerically, yielding the u / U_\infty = f'(\zeta), with f''(0) \approx 0.332. For flows over parabolic bodies, where the external potential flow varies as U \sim \sqrt{x} (corresponding to Falkner-Skan parameter \beta = 2/3), the equation becomes the Falkner-Skan form f''' + f f'' + \frac{2}{3} \left(1 - (f')^2 \right) = 0, with f''(0) \approx 0.4696. Boundary conditions are tailored to the configuration: at the surface (\zeta = 0), f(0) = 0 (impermeability) and f'(0) = 0 (no-slip); asymptotically, f'(\infty) = 1 (matching free-stream). For wedge-like flows or stagnation regions in parabolic setups, these incorporate pressure gradients via \beta. Applications include exact similarity solutions for laminar boundary layers near leading edges of flat plates (recovering \beta = 0) or over semi-infinite parabolic cylinders (\beta = 2/3). This approach, developed in mid-20th-century analyses (e.g., by ), connects to the for non-zero pressure gradients.

References

  1. [1]
    [PDF] Blasius Solution for a Flat Plate Boundary Layer
    The Blasius solution is best presented as an example of a similarity solution to the non-linear, partial differential equation (Bjd4). In a similarity solution ...
  2. [2]
    [PDF] 1 Introduction. 2 Boundary Layer Governing Equations. - MIT
    4.1 Quantities for the Blasius Boundary Layer Solution. For the Blasius similarity solution for a two-dimensional boundary layer given by equa- tion (3.47) ...
  3. [3]
  4. [4]
    HISTORY OF BOUNDARY LA YER THEORY - Annual Reviews
    In 1908 there appeared two papers on boundary layers, one by Blasius and the other by Boltze, both prepared as dissertations at Gottingen under Prandtl's.<|control11|><|separator|>
  5. [5]
    [PDF] Prandtl's Boundary Layer Theory - UC Davis Math
    The Navier-Stokes equations are a singular perturbation of the Euler equations because they contain higher-order spatial derivatives. As a result, the Navier- ...
  6. [6]
    [PDF] Blasius_Paper_NACA-TM-1256.pdf
    They refer to the formation of boundary layers on solid bodies and the origin of separation of jets from these boundary layers suggested by Prandtl. The writer ...
  7. [7]
    [PDF] Blasius solution - Principles of Fluid Dynamics
    It is possible to find more accurate solutions of the boundary layer equations by a series expansion approach. But the starting point of the mathematical.
  8. [8]
    [PDF] NATIONAL ADVISORY COMMITTEE FOR AERONAUTICS
    These equations establish, to a certain extent, a basis for a special mechanics of boundary layers, since the outer flow enters only in “impressed” manner. 3.
  9. [9]
    [PDF] Flat-Plate Blasius Solution - Ansys Customer Center
    The core of Blasius' analysis centers upon transforming the partial differential equations (PDEs) which comprise the flat plate boundary layer equations, with ...
  10. [10]
    On the solution of the laminar boundary layer equations. - Journals
    An approximate method of solution of the boundary layer equations due to Kármán and Pohlhausen was thought to be reasonably accurate.
  11. [11]
    [PDF] A SURVEY OF HIGHER-ORDER BOUNDARY-LAYER THEORY
    If the surface of a body is curved, rather than flat, centrifugal forces yield pressure changes across the boundary layer that exert a second-order effect. This ...
  12. [12]
  13. [13]
    Laminar incompressible flow past a semi-infinite flat plate
    Mar 28, 2006 · 1957 Second approximation to the laminar boundary-layer flow over a flat plate J. ... Kuo, Y. H. 1953 On the flow of an incompressible ...
  14. [14]
    [PDF] Exact Solution of the Boundary-layer Equations Under Particular ...
    The ordinary Blasius profile is obtained by solving (2) with the conditions f(0) = 0, f' (0) = 0, f' (oo) = 2. If however we give f (0) a finite value, we ...
  15. [15]
    On the stability of a Blasius boundary layer subject to localised suction
    May 24, 2019 · In this study the origins of premature transition due to oversuction in boundary layers are studied. An infinite row of circular suction ...Missing: seminal | Show results with:seminal
  16. [16]
    None
    ### Summary of Von Mises Transformation for Boundary Layer with Suction
  17. [17]
    [PDF] The Asymptotic Theory of Boundary-layer Flow with Suction - AERADE
    At the leading edge the boundary layer is of zero thickness and has Blasius' profile, but ultimately it tends to the asymptotic suction profile, which is ...
  18. [18]
    [PDF] 20050031104.pdf - NASA Technical Reports Server (NTRS)
    The solutions for the laminar boundary layer with suction existing so far ... Iglisch about the asymptotic behavior of the plane stegnation point flow ...
  19. [19]
    On the disturbance growth in an asymptotic suction boundary layer
    May 13, 2003 · This comparison shows satisfactory agreement even though the stability of the asymptotic suction profile is somewhat overpredicted by the theory ...Missing: JFM | Show results with:JFM
  20. [20]
    [PDF] Investigations of the asymptotic suction boundary layer Jens H. M. ...
    This study investigates the effect of boundary layer suction on laminar-turbulent transition, using an asymptotic suction boundary layer established in a wind ...
  21. [21]
    Concerning the effect of compressibility on lam inar boundary layers ...
    Concerning the effect of compressibility on lam inar boundary layers and their separation. Leslie Howarth.
  22. [22]
    [PDF] General solution of the laminar compressible boundary layer in the ...
    A modified Howarth-Dorodnitsyn transformation has been used to write the laminar boundary-layer equations for axisymmetric flow in terms of transformed ...