Blasius boundary layer

Last updated

In physics and fluid mechanics, a Blasius boundary layer (named after Paul Richard Heinrich Blasius) describes the steady two-dimensional laminar boundary layer that forms on a semi-infinite plate which is held parallel to a constant unidirectional flow. Falkner and Skan later generalized Blasius' solution to wedge flow (Falkner–Skan boundary layer), i.e. flows in which the plate is not parallel to the flow.

Contents

Prandtl's boundary layer equations

A schematic diagram of the Blasius flow profile. The streamwise velocity component
u
(
e
)
/
U
{\displaystyle u(\eta )/U}
is shown, as a function of the similarity variable
e
{\displaystyle \eta }
. Blasius1.PNG
A schematic diagram of the Blasius flow profile. The streamwise velocity component is shown, as a function of the similarity variable .

Using scaling arguments, Ludwig Prandtl [1] argued that about half of the terms in the Navier-Stokes equations are negligible in boundary layer flows (except in a small region near the leading edge of the plate). This leads to a reduced set of equations known as the boundary layer equations. For steady incompressible flow with constant viscosity and density, these read:

Here the coordinate system is chosen with pointing parallel to the plate in the direction of the flow and the coordinate pointing normal to the plate, and are the and velocity components, is the pressure, is the density and is the kinematic viscosity.

A number of similarity solutions to this set of equations have been found for various types of flow, including flow on a thin flat-plate. The term similarity refers to the property that the velocity profiles at different positions in the flow are the same apart from scaling factors. Similarity scaling factors reduce the set of partial differential equations to a relatively easily solved set of non-linear ordinary differential equations. Paul Richard Heinrich Blasius, [2] one of Prandtl's students, developed the similarity model corresponding to the flow for the case where the pressure gradient, /, along a thin flat-plate is negligible compared to any pressure gradient in the boundary layer region.

Blasius equation - first-order boundary layer

Blasius showed that for the case where , the Prandtl -momentum equation has a self-similar solution. The self-similar solution exists because the equations and the boundary conditions are invariant under the transformation

where is any positive constant. He introduced the self-similar variables

Developing Blasius boundary layer (not to scale). The velocity profile
f
'
{\displaystyle f'}
is shown in red at selected positions along the plate. The blue lines represent, in top to bottom order, the 99% free stream velocity line (
d
99
%
,
e
[?]
5.29
{\displaystyle \delta _{99\%},\eta \approx 5.29}
), the displacement thickness (
d
*
,
e
[?]
1.79
{\displaystyle \delta _{*},\eta \approx 1.79}
) and
d
(
x
)
{\displaystyle \delta (x)}
(
e
=
1.51
{\displaystyle \eta =1.51}
). See Boundary layer thickness for a more detailed explanation. Blasius.png
Developing Blasius boundary layer (not to scale). The velocity profile is shown in red at selected positions along the plate. The blue lines represent, in top to bottom order, the 99% free stream velocity line (), the displacement thickness () and (). See Boundary layer thickness for a more detailed explanation.

where is the boundary layer thickness, is the free stream velocity, and is the stream function. The stream function is directly proportional to the normalized function, , which is only a function of the similarity thickness variable. This leads directly to the velocity components: [3] :136

Where the prime denotes derivation with respect to . Substitution into the -momentum equation gives the Blasius equation

The boundary conditions are the no-slip condition, the impermeability of the wall and the free stream velocity outside the boundary layer

This is a third-order non-linear ordinary differential equation which can be solved numerically, e.g. with the shooting method.

With the solution for and its derivatives in hand, the Prandtl -momentum equation can be non-dimensionalized and rearranged to obtain the -pressure gradient, /, as [4] :46

where is the Blasius displacement thickness.

The Blasius normal velocity and the -pressure gradient asymptotes to a value of 0.86 and 0.43, respectively, at large -values whereas asymptotes to the free stream velocity . As goes to zero, the scaled -pressure gradient goes to 0.16603.

A schematic depiction of the scaled velocities and
y
{\displaystyle y}
-pressure gradient for Blasius boundary layer flow versus the scaled normal height above the wall,
e
{\displaystyle \eta }
. The red line is the Blasius scaled velocity
u
(
x
,
y
)
/
U
{\displaystyle u(x,y)/U}
, the green line is the Blasius
v
(
x
,
y
)
/
n
U
/
x
{\textstyle v(x,y)/{\sqrt {\nu U/x}}}
and the blue line is the Blasius
1
r
d
P
d
y
/
n
U
3
/
x
3
{\displaystyle {\frac {1}{\rho }}{\frac {dP}{dy}}/{\sqrt {\nu U^{3}/x^{3}}}}
. Graph Blasius velocities and y-pressure gradient versus normal height above wall.png
A schematic depiction of the scaled velocities and -pressure gradient for Blasius boundary layer flow versus the scaled normal height above the wall, . The red line is the Blasius scaled velocity , the green line is the Blasius and the blue line is the Blasius .

The limiting form for small is

and the limiting form for large is [5]

The characteristic parameters for boundary layers are the two sigma viscous boundary layer thickness, [6] , the displacement thickness , the momentum thickness , the wall shear stress and the drag force acting on a length of the plate. For the Blasius solution, they are given by

The factor in the drag force formula is to account both sides of the plate.

The Von Kármán Momentum integral and the energy integral for Blasius profile reduce to

where is the wall shear stress, is the wall injection/suction velocity, is the energy dissipation rate, is the momentum thickness and is the energy thickness.

Uniqueness of Blasius solution

The Blasius solution is not unique from a mathematical perspective, [7] :131 as Ludwig Prandtl himself noted it in his transposition theorem and analyzed by series of researchers such as Keith Stewartson, Paul A. Libby. [8] To this solution, any one of the infinite discrete set of eigenfunctions can be added, each of which satisfies the linearly perturbed equation with homogeneous conditions and exponential decay at infinity. The first of these eigenfunctions turns out to be the derivative of the first order Blasius solution, which represents the uncertainty in the effective location of the origin.

Second-order boundary layer

This boundary layer approximation predicts a non-zero vertical velocity far away from the wall, which needs to be accounted in next order outer inviscid layer and the corresponding inner boundary layer solution, which in turn will predict a new vertical velocity and so on. The vertical velocity at infinity for the first order boundary layer problem from the Blasius equation is

The solution for second order boundary layer is zero. The solution for outer inviscid and inner boundary layer are [7] :134

Again as in the first order boundary problem, any one of the infinite set of eigensolution can be added to this solution. In all the solutions can be considered as a Reynolds number.

Third-order boundary layer

Since the second order inner problem is zero, the corresponding corrections to third order problem is null i.e., the third order outer problem is same as second order outer problem. [7] :139 The solution for third-order correction does not have an exact expression, but the inner boundary layer expansion is of the form,

where is the first eigensolution of the first order boundary layer solution (which is derivative of the first order Blasius solution) and solution for is nonunique and the problem is left with an undetermined constant.

Blasius boundary layer with suction

Suction is one of the common methods to postpone the boundary layer separation. [9] Consider a uniform suction velocity at the wall . Bryan Thwaites [10] showed that the solution for this problem is same as the Blasius solution without suction for distances very close to the leading edge. Introducing the transformation

into the boundary layer equations leads to

with boundary conditions,

Von Mises transformation

Iglisch obtained the complete numerical solution in 1944. [11] If further von Mises transformation [12] is introduced

then the equations become

with boundary conditions,

This parabolic partial differential equation can be marched starting from numerically.

Asymptotic suction profile

Since the convection due to suction and the diffusion due to the solid wall are acting in the opposite direction, the profile will reach steady solution at large distance, unlike the Blasius profile where boundary layer grows indefinitely. The solution was first obtained by Griffith and F.W. Meredith. [13] For distances from the leading edge of the plate , both the boundary layer thickness and the solution are independent of given by

Stewartson [14] studied matching of full solution to the asymptotic suction profile.

Compressible Blasius boundary layer

Here Blasius boundary layer with a specified specific enthalpy at the wall is studied. The density , viscosity and thermal conductivity are no longer constant here. The equation for conservation of mass, momentum and energy become

where is the Prandtl number with suffix representing properties evaluated at infinity. The boundary conditions become

Unlike the incompressible boundary layer, similarity solution exists only if the transformation

holds and this is possible only if .

Howarth transformation

Compressible Blasius boundary layer CompBlasius.jpg
Compressible Blasius boundary layer

Introducing the self-similar variables using Howarth–Dorodnitsyn transformation

the equations reduce to

where is the specific heat ratio and is the Mach number, where is the speed of sound. The equation can be solved once are specified. The boundary conditions are

The commonly used expressions for air are . If is constant, then . The temperature inside the boundary layer will increase even though the plate temperature is maintained at the same temperature as ambient, due to dissipative heating and of course, these dissipation effects are only pronounced when the Mach number is large.

First-order Blasius boundary layer in parabolic coordinates

Since the boundary layer equations are Parabolic partial differential equation, the natural coordinates for the problem is parabolic coordinates. [7] :142 The transformation from Cartesian coordinates to parabolic coordinates is given by

See also

Footnotes

  1. Prandtl, L. (1904). "Über Flüssigkeitsbewegung bei sehr kleiner Reibung". Verhandlinger 3. Int. Math. Kongr. Heidelberg: 484–491.
  2. Blasius, H. (1908). "Grenzschichten in Flüssigkeiten mit kleiner Reibung". Z. Angew. Math. Phys. 56: 1–37.
  3. Schlichting, H., (1979). Boundary-Layer Theory, 7th ed., McGraw-Hill, New York.
  4. Weyburne, David (2022). Aspects of Boundary Layer Theory. ISBN   978-0-578-98334-9.
  5. Boyd, J. (2008). "The Blasius function: computations before computers, the value of tricks, undergraduate projects, and open research problems". SIAM Rev. 50 (4): 791–804. Bibcode:2008SIAMR..50..791B. doi:10.1137/070681594.
  6. Weyburne, D. (2014). "New thickness and shape parameters for the boundary layer velocity profile". Experimental Thermal and Fluid Science. 54: 22–28. doi:10.1016/j.expthermflusci.2014.01.008.
  7. 1 2 3 4 Van Dyke, Milton (1975). Perturbation methods in fluid mechanics. Parabolic Press. ISBN   9780915760015.
  8. Libby, Paul A., and Herbert Fox. "Some perturbation solutions in laminar boundary-layer theory." Journal of Fluid Mechanics 17.3 (1963): 433-449.
  9. Rosenhead, Louis, ed. Laminar boundary layers. Clarendon Press, 1963.
  10. Thwaites, Bryan. On certain types of boundary-layer flow with continuous surface suction. HM Stationery Office, 1946.
  11. Iglisch, Rudolf. Exakte Berechnung der laminaren Grenzschicht an der längsangeströmten ebenen Platte mit homogener Absaugung. Oldenbourg, 1944.
  12. Von Mises, Richard. "Bemerkungen zur hydrodynamik." Z. Angew. Math. Mech 7 (1927): 425-429.
  13. Griffith, A. A., and F. W. Meredith. "The possible improvement in aircraft performance due to the use of boundary layer suction." Royal Aircraft Establishment Report No. E 3501 (1936): 12.
  14. Stewartson, K. "On asymptotic expansions in the theory of boundary layers." Studies in Applied Mathematics 36.1-4 (1957): 173-191.

Related Research Articles

<span class="mw-page-title-main">Boundary layer</span> Layer of fluid in the immediate vicinity of a bounding surface

In physics and fluid mechanics, a boundary layer is the thin layer of fluid in the immediate vicinity of a bounding surface formed by the fluid flowing along the surface. The fluid's interaction with the wall induces a no-slip boundary condition. The flow velocity then monotonically increases above the surface until it returns to the bulk flow velocity. The thin layer consisting of fluid whose velocity has not yet returned to the bulk flow velocity is called the velocity boundary layer.

The oceanic wind driven Ekman spiral is the result of a force balance created by a shear stress force, Coriolis force and the water drag. This force balance gives a resulting current of the water different from the winds. In the ocean, there are two places where the Ekman spiral can be observed. At the surface of the ocean, the shear stress force corresponds with the wind stress force. At the bottom of the ocean, the shear stress force is created by friction with the ocean floor. This phenomenon was first observed at the surface by the Norwegian oceanographer Fridtjof Nansen during his Fram expedition. He noticed that icebergs did not drift in the same direction as the wind. His student, the Swedish oceanographer Vagn Walfrid Ekman, was the first person to physically explain this process.

<span class="mw-page-title-main">Oblate spheroidal coordinates</span> Three-dimensional orthogonal coordinate system

Oblate spheroidal coordinates are a three-dimensional orthogonal coordinate system that results from rotating the two-dimensional elliptic coordinate system about the non-focal axis of the ellipse, i.e., the symmetry axis that separates the foci. Thus, the two foci are transformed into a ring of radius in the x-y plane. Oblate spheroidal coordinates can also be considered as a limiting case of ellipsoidal coordinates in which the two largest semi-axes are equal in length.

In physics, the Spalart–Allmaras model is a one-equation model that solves a modelled transport equation for the kinematic eddy turbulent viscosity. The Spalart–Allmaras model was designed specifically for aerospace applications involving wall-bounded flows and has been shown to give good results for boundary layers subjected to adverse pressure gradients. It is also gaining popularity in turbomachinery applications.

<span class="mw-page-title-main">Shallow water equations</span> Set of partial differential equations that describe the flow below a pressure surface in a fluid

The shallow-water equations (SWE) are a set of hyperbolic partial differential equations that describe the flow below a pressure surface in a fluid. The shallow-water equations in unidirectional form are also called Saint-Venant equations, after Adhémar Jean Claude Barré de Saint-Venant.

Acoustic streaming is a steady flow in a fluid driven by the absorption of high amplitude acoustic oscillations. This phenomenon can be observed near sound emitters, or in the standing waves within a Kundt's tube. Acoustic streaming was explained first by Lord Rayleigh in 1884. It is the less-known opposite of sound generation by a flow.

In mathematics, the spectral theory of ordinary differential equations is the part of spectral theory concerned with the determination of the spectrum and eigenfunction expansion associated with a linear ordinary differential equation. In his dissertation, Hermann Weyl generalized the classical Sturm–Liouville theory on a finite closed interval to second order differential operators with singularities at the endpoints of the interval, possibly semi-infinite or infinite. Unlike the classical case, the spectrum may no longer consist of just a countable set of eigenvalues, but may also contain a continuous part. In this case the eigenfunction expansion involves an integral over the continuous part with respect to a spectral measure, given by the Titchmarsh–Kodaira formula. The theory was put in its final simplified form for singular differential equations of even degree by Kodaira and others, using von Neumann's spectral theorem. It has had important applications in quantum mechanics, operator theory and harmonic analysis on semisimple Lie groups.

<span class="mw-page-title-main">Falkner–Skan boundary layer</span> Boundary Layer

In fluid dynamics, the Falkner–Skan boundary layer describes the steady two-dimensional laminar boundary layer that forms on a wedge, i.e. flows in which the plate is not parallel to the flow. It is also representative of flow on a flat plate with an imposed pressure gradient along the plate length, a situation often encountered in wind tunnel flow. It is a generalization of the flat plate Blasius boundary layer in which the pressure gradient along the plate is zero.

In fluid dynamics, the Burgers vortex or Burgers–Rott vortex is an exact solution to the Navier–Stokes equations governing viscous flow, named after Jan Burgers and Nicholas Rott. The Burgers vortex describes a stationary, self-similar flow. An inward, radial flow, tends to concentrate vorticity in a narrow column around the symmetry axis, while an axial stretching causes the vorticity to increase. At the same time, viscous diffusion tends to spread the vorticity. The stationary Burgers vortex arises when the three effects are in balance.

Von Kármán swirling flow is a flow created by a uniformly rotating infinitely long plane disk, named after Theodore von Kármán who solved the problem in 1921. The rotating disk acts as a fluid pump and is used as a model for centrifugal fans or compressors. This flow is classified under the category of steady flows in which vorticity generated at a solid surface is prevented from diffusing far away by an opposing convection, the other examples being the Blasius boundary layer with suction, stagnation point flow etc.

In fluid dynamics, a stagnation point flow refers to a fluid flow in the neighbourhood of a stagnation point or a stagnation line with which the stagnation point/line refers to a point/line where the velocity is zero in the inviscid approximation. The flow specifically considers a class of stagnation points known as saddle points wherein incoming streamlines gets deflected and directed outwards in a different direction; the streamline deflections are guided by separatrices. The flow in the neighborhood of the stagnation point or line can generally be described using potential flow theory, although viscous effects cannot be neglected if the stagnation point lies on a solid surface.

In fluid dynamics, Bickley jet is a steady two-dimensional laminar plane jet with large jet Reynolds number emerging into the fluid at rest, named after W. G. Bickley, who gave the analytical solution in 1937, to the problem derived by Schlichting in 1933 and the corresponding problem in axisymmetric coordinates is called as Schlichting jet. The solution is valid only for distances far away from the jet origin.

In fluid dynamics, Rayleigh problem also known as Stokes first problem is a problem of determining the flow created by a sudden movement of an infinitely long plate from rest, named after Lord Rayleigh and Sir George Stokes. This is considered as one of the simplest unsteady problem that have exact solution for the Navier-Stokes equations. The impulse movement of semi-infinite plate was studied by Keith Stewartson.

<span class="mw-page-title-main">Stokes problem</span>

In fluid dynamics, Stokes problem also known as Stokes second problem or sometimes referred to as Stokes boundary layer or Oscillating boundary layer is a problem of determining the flow created by an oscillating solid surface, named after Sir George Stokes. This is considered one of the simplest unsteady problems that has an exact solution for the Navier-Stokes equations. In turbulent flow, this is still named a Stokes boundary layer, but now one has to rely on experiments, numerical simulations or approximate methods in order to obtain useful information on the flow.

In combustion, a Burke–Schumann flame is a type of diffusion flame, established at the mouth of the two concentric ducts, by issuing fuel and oxidizer from the two region respectively. It is named after S.P. Burke and T.E.W. Schumann, who were able to predict the flame height and flame shape using their simple analysis of infinitely fast chemistry in 1928 at the First symposium on combustion.

In fluid dynamics, Berman flow is a steady flow created inside a rectangular channel with two equally porous walls. The concept is named after a scientist Abraham S. Berman who formulated the problem in 1953.

In astrophysics, Chandrasekhar's white dwarf equation is an initial value ordinary differential equation introduced by the Indian American astrophysicist Subrahmanyan Chandrasekhar, in his study of the gravitational potential of completely degenerate white dwarf stars. The equation reads as

In combustion, Emmons problem describes the flame structure which develops inside the boundary layer, created by a flowing oxidizer stream on flat fuel surfaces. The problem was first studied by Howard Wilson Emmons in 1956. The flame is of diffusion flame type because it separates fuel and oxygen by a flame sheet. The corresponding problem in a quiescent oxidizer environment is known as Clarke–Riley diffusion flame.

Schlichting jet is a steady, laminar, round jet, emerging into a stationary fluid of the same kind with very high Reynolds number. The problem was formulated and solved by Hermann Schlichting in 1933, who also formulated the corresponding planar Bickley jet problem in the same paper. The Landau-Squire jet from a point source is an exact solution of Navier-Stokes equations, which is valid for all Reynolds number, reduces to Schlichting jet solution at high Reynolds number, for distances far away from the jet origin.

In the study of partial differential equations, particularly in fluid dynamics, a self-similar solution is a form of solution which is similar to itself if the independent and dependent variables are appropriately scaled. Self-similar solutions appear whenever the problem lacks a characteristic length or time scale. These include, for example, the Blasius boundary layer or the Sedov–Taylor shell.

References