Sunteți pe pagina 1din 15

Boundary layer

In physics and fluid mechanics, a boundary layer is the layer of fluid in the immediate vicinity of a
bounding surface where the effects of viscosity are significant.

In the Earth's atmosphere, the atmospheric boundary layer is the air layer near the ground affected by
diurnal heat, moisture or momentum transfer to or from the surface. On an aircraft wing the boundary
layer is the part of the flow close to the wing, where viscous forces distort the surrounding non-viscous
flow.

Contents
Types of boundary layer
Aerodynamics
Boundary layer equations
Prandtl's transposition theorem
Von Kármán momentum integral
Energy integral
Von Mises transformation
Crocco's transformation
Turbulent boundary layers
Heat and mass transfer
Convective transfer constants from boundary layer analysis
Naval architecture
Boundary layer turbine
Predicting transient boundary layer thickness in a cylinder using dimensional analysis
Predicting convective flow conditions at the boundary layer in a cylinder using
dimensional analysis
Boundary layer ingestion
See also
References
External links

Types of boundary layer


Laminar boundary layers can be loosely classified according to their structure and the circumstances
under which they are created. The thin shear layer which develops on an oscillating body is an example of
a Stokes boundary layer, while the Blasius boundary layer refers to the well-known similarity solution
near an attached flat plate held in an oncoming unidirectional flow and Falkner–Skan boundary layer, a
generalization of Blasius profile. When a fluid rotates and viscous forces are balanced by the Coriolis
effect (rather than convective inertia), an
Ekman layer forms. In the theory of heat
transfer, a thermal boundary layer occurs. A
surface can have multiple types of boundary
layer simultaneously.

The viscous nature of airflow reduces the local


velocities on a surface and is responsible for
Boundary layer visualization, showing transition from
skin friction. The layer of air over the wing's
laminar to turbulent condition
surface that is slowed down or stopped by
viscosity, is the boundary layer. There are two
different types of boundary layer flow: laminar and turbulent.[1]

Laminar boundary layer flow

The laminar boundary is a very smooth flow, while the turbulent boundary layer contains swirls or
"eddies." The laminar flow creates less skin friction drag than the turbulent flow, but is less stable.
Boundary layer flow over a wing surface begins as a smooth laminar flow. As the flow continues back
from the leading edge, the laminar boundary layer increases in thickness.

Turbulent boundary layer flow

At some distance back from the leading edge, the smooth laminar flow breaks down and transitions to a
turbulent flow. From a drag standpoint, it is advisable to have the transition from laminar to turbulent flow
as far aft on the wing as possible, or have a large amount of the wing surface within the laminar portion of
the boundary layer. The low energy laminar flow, however, tends to break down more suddenly than the
turbulent layer.

Aerodynamics
The aerodynamic boundary layer was first defined by Ludwig
Prandtl in a paper presented on August 12, 1904 at the third
International Congress of Mathematicians in Heidelberg, Germany.
It simplifies the equations of fluid flow by dividing the flow field
into two areas: one inside the boundary layer, dominated by viscosity
and creating the majority of drag experienced by the boundary body;
and one outside the boundary layer, where viscosity can be neglected
without significant effects on the solution. This allows a closed-form
solution for the flow in both areas, a significant simplification of the
full Navier–Stokes equations. The majority of the heat transfer to
and from a body also takes place within the boundary layer, again
allowing the equations to be simplified in the flow field outside the
boundary layer. The pressure distribution throughout the boundary
layer in the direction normal to the surface (such as an airfoil)
remains constant throughout the boundary layer, and is the same as
Ludwig Prandtl
on the surface itself.
The thickness of the velocity boundary layer is
normally defined as the distance from the solid
body to the point at which the viscous flow
velocity is 99% of the freestream velocity (the
surface velocity of an inviscid flow).
Displacement thickness is an alternative definition
stating that the boundary layer represents a deficit
in mass flow compared to inviscid flow with slip Laminar boundary layer velocity profile
at the wall. It is the distance by which the wall
would have to be displaced in the inviscid case to
give the same total mass flow as the viscous case. The no-slip condition requires the flow velocity at the
surface of a solid object be zero and the fluid temperature be equal to the temperature of the surface. The
flow velocity will then increase rapidly within the boundary layer, governed by the boundary layer
equations, below.

The thermal boundary layer thickness is similarly the distance from the body at which the temperature is
99% of the freestream temperature. The ratio of the two thicknesses is governed by the Prandtl number. If
the Prandtl number is 1, the two boundary layers are the same thickness. If the Prandtl number is greater
than 1, the thermal boundary layer is thinner than the velocity boundary layer. If the Prandtl number is less
than 1, which is the case for air at standard conditions, the thermal boundary layer is thicker than the
velocity boundary layer.

In high-performance designs, such as gliders and commercial aircraft, much attention is paid to
controlling the behavior of the boundary layer to minimize drag. Two effects have to be considered. First,
the boundary layer adds to the effective thickness of the body, through the displacement thickness, hence
increasing the pressure drag. Secondly, the shear forces at the surface of the wing create skin friction drag.

At high Reynolds numbers, typical of full-sized aircraft, it is desirable to have a laminar boundary layer.
This results in a lower skin friction due to the characteristic velocity profile of laminar flow. However, the
boundary layer inevitably thickens and becomes less stable as the flow develops along the body, and
eventually becomes turbulent, the process known as boundary layer transition. One way of dealing with
this problem is to suck the boundary layer away through a porous surface (see Boundary layer suction).
This can reduce drag, but is usually impractical due to its mechanical complexity and the power required
to move the air and dispose of it. Natural laminar flow techniques push the boundary layer transition aft
by reshaping the airfoil or fuselage so that its thickest point is more aft and less thick. This reduces the
velocities in the leading part and the same Reynolds number is achieved with a greater length.

At lower Reynolds numbers, such as those seen with model aircraft, it is relatively easy to maintain
laminar flow. This gives low skin friction, which is desirable. However, the same velocity profile which
gives the laminar boundary layer its low skin friction also causes it to be badly affected by adverse
pressure gradients. As the pressure begins to recover over the rear part of the wing chord, a laminar
boundary layer will tend to separate from the surface. Such flow separation causes a large increase in the
pressure drag, since it greatly increases the effective size of the wing section. In these cases, it can be
advantageous to deliberately trip the boundary layer into turbulence at a point prior to the location of
laminar separation, using a turbulator. The fuller velocity profile of the turbulent boundary layer allows it
to sustain the adverse pressure gradient without separating. Thus, although the skin friction is increased,
overall drag is decreased. This is the principle behind the dimpling on golf balls, as well as vortex
generators on aircraft. Special wing sections have also been designed which tailor the pressure recovery so
laminar separation is reduced or even eliminated. This represents an optimum compromise between the
pressure drag from flow separation and skin friction from induced turbulence.

When using half-models in wind tunnels, a peniche is sometimes used to reduce or eliminate the effect of
the boundary layer.

Boundary layer equations


The deduction of the boundary layer equations was one of the most important advances in fluid
dynamics. Using an order of magnitude analysis, the well-known governing Navier–Stokes equations of
viscous fluid flow can be greatly simplified within the boundary layer. Notably, the characteristic of the
partial differential equations (PDE) becomes parabolic, rather than the elliptical form of the full Navier–
Stokes equations. This greatly simplifies the solution of the equations. By making the boundary layer
approximation, the flow is divided into an inviscid portion (which is easy to solve by a number of
methods) and the boundary layer, which is governed by an easier to solve PDE. The continuity and
Navier–Stokes equations for a two-dimensional steady incompressible flow in Cartesian coordinates are
given by

where and are the velocity components, is the density, is the pressure, and is the kinematic
viscosity of the fluid at a point.

The approximation states that, for a sufficiently high Reynolds number the flow over a surface can be
divided into an outer region of inviscid flow unaffected by viscosity (the majority of the flow), and a
region close to the surface where viscosity is important (the boundary layer). Let and be streamwise
and transverse (wall normal) velocities respectively inside the boundary layer. Using scale analysis, it can
be shown that the above equations of motion reduce within the boundary layer to become

and if the fluid is incompressible (as liquids are under standard conditions):
The order of magnitude analysis assumes the streamwise length scale significantly larger than the
transverse length scale inside the boundary layer. It follows that variations in properties in the streamwise
direction are generally much lower than those in the wall normal direction. Apply this to the continuity
equation shows that , the wall normal velocity, is small compared with the streamwise velocity.

Since the static pressure is independent of , then pressure at the edge of the boundary layer is the
pressure throughout the boundary layer at a given streamwise position. The external pressure may be
obtained through an application of Bernoulli's equation. Let be the fluid velocity outside the boundary
layer, where and are both parallel. This gives upon substituting for the following result

For a flow in which the static pressure also does not change in the direction of the flow

so remains constant.

Therefore, the equation of motion simplifies to become

These approximations are used in a variety of practical flow problems of scientific and engineering
interest. The above analysis is for any instantaneous laminar or turbulent boundary layer, but is used
mainly in laminar flow studies since the mean flow is also the instantaneous flow because there are no
velocity fluctuations present. This simplified equation is a parabolic PDE and can be solved using a
similarity solution often referred to as the Blasius boundary layer.

Prandtl's transposition theorem


Prandtl[2] observed that from any solution which satisfies the boundary layer
equations, further solution , which is also satisfying the boundary layer equations,
can be constructed by writing

where is arbitrary. Since the solution is not unique from mathematical perspective,[3] to the solution
can added any one of an infinite set of eigenfunctions as shown by Stewartson[4] and Paul A. Libby.[5][6]

Von Kármán momentum integral


von Kármán[7] derived the integral equation by integrating the boundary layer equation across the
boundary layer in 1921. The equation is
where

is the wall shear stress, is the suction/injection velocity at the wall, is the
displacement thickness and is the momentum thickness. Kármán–Pohlhausen
Approximation is derived from this equation.

Energy integral
The energy integral was derived by Wieghardt.[8][9]

where

is the energy dissipation rate due to viscosity across the boundary layer and is the
energy thickness.[10]

Von Mises transformation


For steady two-dimensional boundary layers, von Mises[11] introduced a transformation which takes
and (stream function) as independent variables instead of and and uses a dependent variable
instead of . The boundary layer equation then become

The original variables are recovered from

This transformation is later extended to compressible boundary layer by von Kármán and HS Tsien.[12]

Crocco's transformation
For steady two-dimensional compressible boundary layer, Luigi Crocco[13] introduced a transformation
which takes and as independent variables instead of and and uses a dependent variable
(shear stress) instead of . The boundary layer equation then becomes
The original coordinate is recovered from

Turbulent boundary layers


The treatment of turbulent boundary layers is far more difficult due to the time-dependent variation of the
flow properties. One of the most widely used techniques in which turbulent flows are tackled is to apply
Reynolds decomposition. Here the instantaneous flow properties are decomposed into a mean and
fluctuating component. Applying this technique to the boundary layer equations gives the full turbulent
boundary layer equations not often given in literature:

Using a similar order-of-magnitude analysis, the above equations can be reduced to leading order terms.
By choosing length scales for changes in the transverse-direction, and for changes in the streamwise-
direction, with , the x-momentum equation simplifies to:

This equation does not satisfy the no-slip condition at the wall. Like Prandtl did for his boundary layer
equations, a new, smaller length scale must be used to allow the viscous term to become leading order in
the momentum equation. By choosing as the y-scale, the leading order momentum equation for
this "inner boundary layer" is given by:

In the limit of infinite Reynolds number, the pressure gradient term can be shown to have no effect on the
inner region of the turbulent boundary layer. The new "inner length scale" is a viscous length scale, and
is of order , with being the velocity scale of the turbulent fluctuations, in this case a friction
velocity.
Unlike the laminar boundary layer equations, the presence of two regimes governed by different sets of
flow scales (i.e. the inner and outer scaling) has made finding a universal similarity solution for the
turbulent boundary layer difficult and controversial. To find a similarity solution that spans both regions of
the flow, it is necessary to asymptotically match the solutions from both regions of the flow. Such analysis
will yield either the so-called log-law or power-law.

The additional term in the turbulent boundary layer equations is known as the Reynolds shear stress
and is unknown a priori. The solution of the turbulent boundary layer equations therefore necessitates the
use of a turbulence model, which aims to express the Reynolds shear stress in terms of known flow
variables or derivatives. The lack of accuracy and generality of such models is a major obstacle in the
successful prediction of turbulent flow properties in modern fluid dynamics.

A constant stress layer exists in the near wall region. Due to the damping of the vertical velocity
fluctuations near the wall, the Reynolds stress term will become negligible and we find that a linear
velocity profile exists. This is only true for the very near wall region.

Heat and mass transfer


In 1928, the French engineer André Lévêque observed that convective heat transfer in a flowing fluid is
affected only by the velocity values very close to the surface.[14][15] For flows of large Prandtl number,
the temperature/mass transition from surface to freestream temperature takes place across a very thin
region close to the surface. Therefore, the most important fluid velocities are those inside this very thin
region in which the change in velocity can be considered linear with normal distance from the surface. In
this way, for

when , then

where θ is the tangent of the Poiseuille parabola intersecting the wall. Although Lévêque's solution was
specific to heat transfer into a Poiseuille flow, his insight helped lead other scientists to an exact solution
of the thermal boundary-layer problem.[16] Schuh observed that in a boundary-layer, u is again a linear
function of y, but that in this case, the wall tangent is a function of x.[17] He expressed this with a modified
version of Lévêque's profile,

This results in a very good approximation, even for low numbers, so that only liquid metals with
much less than 1 cannot be treated this way. [16] In 1962, Kestin and Persen published a paper describing
solutions for heat transfer when the thermal boundary layer is contained entirely within the momentum
layer and for various wall temperature distributions.[18] For the problem of a flat plate with a temperature
jump at , they propose a substitution that reduces the parabolic thermal boundary-layer equation to
an ordinary differential equation. The solution to this equation, the temperature at any point in the fluid,
can be expressed as an incomplete gamma function.[15] Schlichting proposed an equivalent substitution
that reduces the thermal boundary-layer equation to an ordinary differential equation whose solution is the
same incomplete gamma function.[19]

Convective transfer constants from boundary layer analysis


Paul Richard Heinrich Blasius derived an exact solution to the above laminar boundary layer
equations.[20] The thickness of the boundary layer is a function of the Reynolds number for laminar
flow.

= the thickness of the boundary layer: the region of flow where the velocity is less than
99% of the far field velocity ; is position along the semi-infinite plate, and is the
Reynolds Number given by ( density and dynamic viscosity).

The Blasius solution uses boundary conditions in a dimensionless form:

at

at and

Note that in many cases, the no-slip boundary condition holds that
, the fluid velocity at the surface of the plate equals the velocity
of the plate at all locations. If the plate is not moving, then
. A much more complicated derivation is required if fluid slip is
allowed.[21]

In fact, the Blasius solution for laminar velocity profile in the


boundary layer above a semi-infinite plate can be easily extended
to describe Thermal and Concentration boundary layers for heat
and mass transfer respectively. Rather than the differential x-
momentum balance (equation of motion), this uses a similarly
derived Energy and Mass balance:

Velocity Boundary Layer (Top,


Energy: orange) and Temperature Boundary
Layer (Bottom, green) share a
functional form due to similarity in the
Momentum/Energy Balances and
Mass:
boundary conditions.

For the momentum balance, kinematic viscosity can be


considered to be the momentum diffusivity. In the energy balance this is replaced by thermal diffusivity
, and by mass diffusivity in the mass balance. In thermal diffusivity of a substance, is
its thermal conductivity, is its density and is its heat capacity. Subscript AB denotes diffusivity of
species A diffusing into species B.
Under the assumption that , these equations become equivalent to the momentum balance.
Thus, for Prandtl number and Schmidt number the Blasius solution
applies directly.

Accordingly, this derivation uses a related form of the boundary conditions, replacing with or
(absolute temperature or concentration of species A). The subscript S denotes a surface condition.

at

at and

Using the streamline function Blasius obtained the following solution for the shear stress at the surface of
the plate.

And via the boundary conditions, it is known that

We are given the following relations for heat/mass flux out of the surface of the plate

So for

where are the regions of flow where and are less than 99% of their far field values.[22]

Because the Prandtl number of a particular fluid is not often unity, German engineer E. Polhausen who
worked with Ludwig Prandtl attempted to empirically extend these equations to apply for . His
results can be applied to as well.[23] He found that for Prandtl number greater than 0.6, the thermal
boundary layer thickness was approximately given by:

and therefore

From this solution, it is possible to characterize the convective heat/mass transfer constants based on the
region of boundary layer flow. Fourier's law of conduction and Newton's Law of Cooling are combined
with the flux term derived above and the boundary layer thickness.
This gives the local convective constant at one point on the
semi-infinite plane. Integrating over the length of the plate gives
an average

Plot showing the relative thickness in


the Thermal boundary layer versus
the Velocity boundary layer (in red)
Following the derivation with mass transfer terms ( = convective for various Prandtl Numbers. For
mass transfer constant, = diffusivity of species A into , the two are equal.
species B, ), the following solutions are obtained:

These solutions apply for laminar flow with a Prandtl/Schmidt number greater than 0.6.[22]

Naval architecture
Many of the principles that apply to aircraft also apply to ships, submarines, and offshore platforms.

For ships, unlike aircraft, one deals with incompressible flows, where change in water density is
negligible (a pressure rise close to 1000kPa leads to a change of only 2–3 kg/m3). This field of fluid
dynamics is called hydrodynamics. A ship engineer designs for hydrodynamics first, and for strength only
later. The boundary layer development, breakdown, and separation become critical because the high
viscosity of water produces high shear stresses. Another consequence of high viscosity is the slip stream
effect, in which the ship moves like a spear tearing through a sponge at high velocity.

Boundary layer turbine


This effect was exploited in the Tesla turbine, patented by Nikola Tesla in 1913. It is referred to as a
bladeless turbine because it uses the boundary layer effect and not a fluid impinging upon the blades as in
a conventional turbine. Boundary layer turbines are also known as cohesion-type turbine, bladeless
turbine, and Prandtl layer turbine (after Ludwig Prandtl).

Predicting transient boundary layer thickness in a cylinder using


dimensional analysis
By using the transient and viscous force equations for a cylindrical flow you can predict the transient
boundary layer thickness by finding the Womersley Number ( ).

Transient Force =
Viscous Force =

Setting them equal to each other gives:

Solving for delta gives:

In dimensionless form:

where = Womersley Number; = density; = velocity; ?; = length of transient boundary


layer; = viscosity; = characteristic length.

Predicting convective flow conditions at the boundary layer in a


cylinder using dimensional analysis
By using the convective and viscous force equations at the boundary layer for a cylindrical flow you can
predict the convective flow conditions at the boundary layer by finding the dimensionless Reynolds
Number ( ).

Convective force:

Viscous force:

Setting them equal to each other gives:

Solving for delta gives:

In dimensionless form:
where = Reynolds Number; = density; = velocity; = length of convective boundary layer; =
viscosity; = characteristic length.

Boundary layer ingestion


Boundary layer ingestion promises an increase in aircraft fuel efficiency with an aft-mounted propulsor
ingesting the slow fuselage boundary layer and re-energising the wake to reduce drag and improve
propulsive efficiency. To operate in distorted airflow, the fan is heavier and its efficiency is reduced, and
its integration is challenging. It is used in concepts like the Aurora D8 or the French research agency
Onera’s Nova, saving 5% in cruise by ingesting 40% of the fuselage boundary layer.[24]

Airbus presented the Nautilius concept at the ICAS congress in September 2018: to ingest all the fuselage
boundary layer, while minimizing the azimuthal flow distortion, the fuselage splits into two spindles with
13-18:1 bypass ratio fans. Propulsive efficiencies are up to 90% like counter-rotating open rotors with
smaller, lighter, less complex and noisy engines. It could lower fuel burn by over 10% compared to a
usual underwing 15:1 bypass ratio engine.[24]

See also
Boundary layer separation
Boundary-layer thickness
Thermal boundary layer thickness and shape
Boundary layer suction
Boundary layer control
Blasius boundary layer
Falkner–Skan boundary layer
Ekman layer
Planetary boundary layer
Logarithmic law of the wall
Shape factor (boundary layer flow)
Shear stress
Surface layer

References
1. Young, A.D. (1989). Boundary layers (1st publ. ed.). Washington, DC: American Institute of
Aeronautics and Astronautics. ISBN 0930403576.
2. Prandtl, L. "Zur berechnung der grenzschichten." ZAMM‐Journal of Applied Mathematics and
Mechanics/Zeitschrift für Angewandte Mathematik und Mechanik 18.1 (1938): 77–82.
3. Van Dyke, Milton. Perturbation methods in fluid mechanics. Parabolic Press, Incorporated,
1975.
4. Stewartson, K. "On asymptotic expansions in the theory of boundary layers." Studies in
Applied Mathematics 36.1-4 (1957): 173–191.
5. Libby, Paul A., and Herbert Fox. "Some perturbation solutions in laminar boundary-layer
theory." Journal of Fluid Mechanics 17.03 (1963): 433-449.
6. Fox, Herbert, and Paul A. Libby. "Some perturbation solutions in laminar boundary layer
theory Part 2. The energy equation." Journal of Fluid Mechanics 19.03 (1964): 433–451.
7. Kármán, Th V. "Über laminare und turbulente Reibung." ZAMM‐Journal of Applied
Mathematics and Mechanics/Zeitschrift für Angewandte Mathematik und Mechanik 1.4
(1921): 233–252.
8. Wieghardt, K. On an energy equation for the calculation of laminar boundary layers. Joint
Intelligence Objectives Agency, 1946.
9. Wieghardt, Karl. "über einen Energiesatz zur Berechnung laminarer Grenzschichten."
Ingenieur-Archiv 16.3–4 (1948): 231–242.
10. Rosenhead, Louis, ed. Laminar boundary layers. Clarendon Press, 1963.
11. Von Mises, Richard. "Bemerkungen zur hydrodynamik." Z. Angew. Math. Mech 7 (1927):
425–429
12. Karman. "T. von & Tsien, HS." J. Aero. Sci. 1938 (1938): 5–227.
13. Crocco, L. "A characteristic transformation of the equations of the boundary layer in gases."
ARC 4582 (1939): 1940.
14. Lévêque, A. (1928). "Les lois de la transmission de chaleur par convection". Annales des
Mines ou Recueil de Mémoires sur l'Exploitation des Mines et sur les Sciences et les Arts
qui s'y Rattachent, Mémoires (in French). XIII (13): 201–239.
15. Niall McMahon. "André Lévêque p285, a review of his velocity profile approximation" (https://
web.archive.org/web/20120604004223/http://www.computing.dcu.ie/~nmcmahon/biography/l
eveque/leveque_approximation.html). Archived from the original (http://www.computing.dcu.i
e/~nmcmahon/biography/leveque/leveque_approximation.html) on 2012-06-04.
16. Martin, H. (2002). "The generalized Lévêque equation and its practical use for the prediction
of heat and mass transfer rates from pressure drop". Chemical Engineering Science. 57
(16). pp. 3217–3223. doi:10.1016/S0009-2509(02)00194-X (https://doi.org/10.1016%2FS000
9-2509%2802%2900194-X).
17. Schuh, H. (1953). "On asymptotic solutions for the heat transfer at varying wall temperatures
in a laminar boundary layer with Hartree's velocity profiles". Jour. Aero. Sci. 20 (2). pp. 146–
147.
18. Kestin, J. & Persen, L.N. (1962). "The transfer of heat across a turbulent boundary layer at
very high prandtl numbers". Int. J. Heat Mass Transfer. 5: 355–371. doi:10.1016/0017-
9310(62)90026-1 (https://doi.org/10.1016%2F0017-9310%2862%2990026-1).
19. Schlichting, H. (1979). Boundary-Layer Theory (7 ed.). New York (USA): McGraw-Hill.
20. Blasius, H. (1908). "Grenzschichten in Flüssigkeiten mit kleiner Reibung". Z. Math. Phys. 56:
1–37. (English translation)
21. Martin, Michael J. Blasius boundary layer solution with slip flow conditions. AIP conference
proceedings 585.1 2001: 518-523. American Institute of Physics. 24 Apr 2013.
22. Geankoplis, Christie J. Transport Processes and Separation Process Principles: (includes
Unit Operations). Fourth ed. Upper Saddle River, NJ: Prentice Hall Professional Technical
Reference, 2003. Print.
23. Pohlhausen, E. (1921), Der Wärmeaustausch zwischen festen Körpern und Flüssigkeiten mit
kleiner reibung und kleiner Wärmeleitung. Z. angew. Math. Mech., 1: 115–121.
doi:10.1002/zamm.19210010205 (https://doi.org/10.1002%2Fzamm.19210010205)
24. Graham Warwick (Nov 19, 2018). "The Week In Technology, November 19-23, 2018" (http://
aviationweek.com/future-aerospace/week-technology-november-19-23-2018). Aviation Week
& Space Technology.

Chanson, H. (2009). Applied Hydrodynamics: An Introduction to Ideal and Real Fluid Flows
(http://espace.library.uq.edu.au/view/UQ:191112). CRC Press, Taylor & Francis Group,
Leiden, The Netherlands, 478 pages. ISBN 978-0-415-49271-3.
A.D. Polyanin and V.F. Zaitsev, Handbook of Nonlinear Partial Differential Equations,
Chapman & Hall/CRC Press, Boca Raton – London, 2004. ISBN 1-58488-355-3
A.D. Polyanin, A.M. Kutepov, A.V. Vyazmin, and D.A. Kazenin, Hydrodynamics, Mass and
Heat Transfer in Chemical Engineering, Taylor & Francis, London, 2002. ISBN 0-415-27237-
8
Hermann Schlichting, Klaus Gersten, E. Krause, H. Jr. Oertel, C. Mayes "Boundary-Layer
Theory" 8th edition Springer 2004 ISBN 3-540-66270-7
John D. Anderson, Jr., "Ludwig Prandtl's Boundary Layer" (http://aps.org/units/dfd/resources/
upload/prandtl_vol58no12p42_48.pdf), Physics Today, December 2005
Anderson, John (1992). Fundamentals of Aerodynamics (2nd ed.). Toronto: S.S.CHAND.
pp. 711–714. ISBN 0-07-001679-8.
H. Tennekes and J. L. Lumley, "A First Course in Turbulence", The MIT Press, (1972).
Lectures in Turbulence for the 21st Century by William K. George (http://www.turbulence-onli
ne.com/Publications/Lecture_Notes/Turbulence_Lille/TB_16January2013.pdf)

External links
National Science Digital Library – Boundary Layer (https://web.archive.org/web/2005042008
5336/http://www.nsdl.arm.gov/Library/glossary.shtml#boundary_layer)
Moore, Franklin K., "Displacement effect of a three-dimensional boundary layer (http://naca.l
arc.nasa.gov/reports/1953/naca-report-1124/)". NACA Report 1124, 1953.
Benson, Tom, "Boundary layer (http://www.grc.nasa.gov/WWW/K-
12/airplane/boundlay.html)". NASA Glenn Learning Technologies.
Boundary layer separation (https://web.archive.org/web/20041204054810/http://www.maths.
man.ac.uk/~ruban/blsep.html)
Boundary layer equations: Exact Solutions (http://eqworld.ipmnet.ru/en/solutions/npde/npde-t
oc5.pdf) – from EqWorld
Jones, T.V. BOUNDARY LAYER HEAT TRANSFER (http://www.thermopedia.com/content/59
6/)

Retrieved from "https://en.wikipedia.org/w/index.php?title=Boundary_layer&oldid=940642866"

This page was last edited on 13 February 2020, at 19:43 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using
this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia
Foundation, Inc., a non-profit organization.

S-ar putea să vă placă și