Stagnation point flow

Last updated

In fluid dynamics, a stagnation point flow refers to a fluid flow in the neighbourhood of a stagnation point (in two-dimensional flows) or a stagnation line (in three-dimensional flows) 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.

Contents

Stagnation point flow without solid surfaces

When two streams either of two-dimensional or axisymmetric nature impinge on each other, a stagnation plane is created, where the incoming streams are diverted tangentially outwards; thus on the stagnation plane, the velocity component normal to that plane is zero, whereas the tangential component is non-zero. In the neighborhood of the stagnation point, a local description for the velocity field can be described.

General three-dimensional velocity field

The stagnation point flow corresponds to a linear dependence on the coordinates, that can be described in the Cartesian coordinates with velocity components as follows

where are constants (or time-dependent functions) referred as the strain rates; the three strain rates are not completely arbitrary since the continuity equation requires , that is to say, only two of the three constants are independent. We shall assume so that flow is towards the stagnation point in the direction and away from the stagnation point in the direction. Without loss of generality, one can assume that . The flow field can be categorized into different types based on a single parameter [1]

Planar stagnation-point flow

The two-dimensional stagnation-point flow belongs to the case . The flow field is described as follows

where we let . This flow field is investigated as early as 1934 by G. I. Taylor. [2] In the laboratory, this flow field is created using a four-mill apparatus, although these flow fields are ubiquitous in turbulent flows.

Axisymmetric stagnation-point flow

The axisymmetric stagnation point flow corresponds to . The flow field can be simply described in cylindrical coordinate system with velocity components as follows

where we let .

Radial stagnation flows

In radial stagnation flows, instead of a stagnation point, we have a stagnation circle and the stagnation plane is replaced by a stagnation cylinder. The radial stagnation flow is described using the cylindrical coordinate system with velocity components as follows [3] [4] [5]

where is the location of the stagnation cylinder.

Hiemenz flow

Two-dimensional stagnation point flow Stagnation2D.pdf
Two-dimensional stagnation point flow

The flow due to the presence of a solid surface at in planar stagnation-point flow was described first by Karl Hiemenz in 1911, [6] whose numerical computations for the solutions were improved later by Leslie Howarth. [7] A familiar example where Hiemenz flow is applicable is the forward stagnation line that occurs in the flow over a circular cylinder. [8] [9]

The solid surface lies on the . According to potential flow theory, the fluid motion described in terms of the stream function and the velocity components are given by

The stagnation line for this flow is . The velocity component is non-zero on the solid surface indicating that the above velocity field do not satisfy no-slip boundary condition on the wall. To find the velocity components that satisfy the no-slip boundary condition, one assumes the following form

where is the Kinematic viscosity and is the characteristic thickness where viscous effects are significant. The existence of constant value for the viscous effects thickness is due to the competing balance between the fluid convection that is directed towards the solid surface and viscous diffusion that is directed away from the surface. Thus the vorticity produced at the solid surface is able to diffuse only to distances of order ; analogous situations that resembles this behavior occurs in asymptotic suction profile and von Kármán swirling flow. The velocity components, pressure and Navier–Stokes equations then become

The requirements that at and that as translate to

The condition for as cannot be prescribed and is obtained as a part of the solution. The problem formulated here is a special case of Falkner-Skan boundary layer. The solution can be obtained from numerical integrations and is shown in the figure. The asymptotic behaviors for large are

where is the displacement thickness.

Stagnation point flow with a translating wall

Hiemenz flow when the solid wall translates with a constant velocity along the was solved by Rott (1956). [10] This problem describes the flow in the neighbourhood of the forward stagnation line occurring in a flow over a rotating cylinder. [11] The required stream function is

where the function satisfies

The solution to the above equation is given by

Oblique stagnation point flow

If the incoming stream is perpendicular to the stagnation line, but approaches obliquely, the outer flow is not potential, but has a constant vorticity . The appropriate stream function for oblique stagnation point flow is given by

Viscous effects due to the presence of a solid wall was studied by Stuart (1959), [12] Tamada (1979) [13] and Dorrepaal (1986). [14] In their approach, the streamfunction takes the form

where the function

.

Homann flow

Homann flow with injection Stagnationaxi.pdf
Homann flow with injection
Homann flow with suction Stagnationaxi2.pdf
Homann flow with suction

The solution for axisymmetric stagnation point flow in the presence of a solid wall was first obtained by Homann (1936). [15] A typical example of this flow is the forward stagnation point appearing in a flow past a sphere. Paul A. Libby (1974) [16] (1976) [17] extended Homann's work by allowing the solid wall to translate along its own plane with a constant speed and allowing constant suction or injection at the solid surface.

The solution for this problem is obtained in the cylindrical coordinate system by introducing

where is the translational speed of the wall and is the injection (or, suction) velocity at the wall. The problem is axisymmetric only when . The pressure is given by

The Navier–Stokes equations then reduce to

along with boundary conditions,

When , the classical Homann problem is recovered.

Plane counterflows

Jets emerging from a slot-jets creates stagnation point in between according to potential theory. The flow near the stagnation point can by studied using self-similar solution. This setup is widely used in combustion experiments. The initial study of impinging stagnation flows are due to C.Y. Wang. [18] [19] Let two fluids with constant properties denoted with suffix flowing from opposite direction impinge, and assume the two fluids are immiscible and the interface (located at ) is planar. The velocity is given by

where are strain rates of the fluids. At the interface, velocities, tangential stress and pressure must be continuous. Introducing the self-similar transformation,

results equations,

The no-penetration condition at the interface and free stream condition far away from the stagnation plane become

But the equations require two more boundary conditions. At , the tangential velocities , the tangential stress and the pressure are continuous. Therefore,

where (from outer inviscid problem) is used. Both are not known apriori, but derived from matching conditions. The third equation is determine variation of outer pressure due to the effect of viscosity. So there are only two parameters, which governs the flow, which are

then the boundary conditions become

.

Related Research Articles

<span class="mw-page-title-main">Lorentz transformation</span> Family of linear transformations

In physics, the Lorentz transformations are a six-parameter family of linear transformations from a coordinate frame in spacetime to another frame that moves at a constant velocity relative to the former. The respective inverse transformation is then parameterized by the negative of this velocity. The transformations are named after the Dutch physicist Hendrik Lorentz.

<span class="mw-page-title-main">Navier–Stokes equations</span> Equations describing the motion of viscous fluid substances

The Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances. They were named after French engineer and physicist Claude-Louis Navier and the Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842–1850 (Stokes).

In special relativity, four-momentum (also called momentum–energy or momenergy) is the generalization of the classical three-dimensional momentum to four-dimensional spacetime. Momentum is a vector in three dimensions; similarly four-momentum is a four-vector in spacetime. The contravariant four-momentum of a particle with relativistic energy E and three-momentum p = (px, py, pz) = γmv, where v is the particle's three-velocity and γ the Lorentz factor, is

<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.

<span class="mw-page-title-main">Ekman spiral</span> Velocity profile of wind driven current with depth

The Ekman spiral is an arrangement of ocean currents: the directions of horizontal current appear to twist as the depth changes. 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">Lamb–Oseen vortex</span> Line vortex

In fluid dynamics, the Lamb–Oseen vortex models a line vortex that decays due to viscosity. This vortex is named after Horace Lamb and Carl Wilhelm Oseen.

In physics and fluid mechanics, a Blasius boundary layer 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, i.e. flows in which the plate is not parallel to the flow.

<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 (de) Saint-Venant equations, after Adhémar Jean Claude Barré de Saint-Venant.

In mathematics and mathematical physics, raising and lowering indices are operations on tensors which change their type. Raising and lowering indices are a form of index manipulation in tensor expressions.

Shear velocity, also called friction velocity, is a form by which a shear stress may be re-written in units of velocity. It is useful as a method in fluid mechanics to compare true velocities, such as the velocity of a flow in a stream, to a velocity that relates shear between layers of flow.

In fluid dynamics, the Oseen equations describe the flow of a viscous and incompressible fluid at small Reynolds numbers, as formulated by Carl Wilhelm Oseen in 1910. Oseen flow is an improved description of these flows, as compared to Stokes flow, with the (partial) inclusion of convective acceleration.

<span class="mw-page-title-main">Radiation stress</span> Term in physical oceanography

In fluid dynamics, the radiation stress is the depth-integrated – and thereafter phase-averaged – excess momentum flux caused by the presence of the surface gravity waves, which is exerted on the mean flow. The radiation stresses behave as a second-order tensor.

<span class="mw-page-title-main">Falkner–Skan boundary layer</span> Boundary layer that forms on a wedge

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.

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, 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 problems that have an exact solution for the Navier-Stokes equations. The impulse movement of semi-infinite plate was studied by Keith Stewartson.

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.

The shear viscosity of a fluid is a material property that describes the friction between internal neighboring fluid surfaces flowing with different fluid velocities. This friction is the effect of (linear) momentum exchange caused by molecules with sufficient energy to move between these fluid sheets due to fluctuations in their motion. The viscosity is not a material constant, but a material property that depends on temperature, pressure, fluid mixture composition, local velocity variations. This functional relationship is described by a mathematical viscosity model called a constitutive equation which is usually far more complex than the defining equation of shear viscosity. One such complicating feature is the relation between the viscosity model for a pure fluid and the model for a fluid mixture which is called mixing rules. When scientists and engineers use new arguments or theories to develop a new viscosity model, instead of improving the reigning model, it may lead to the first model in a new class of models. This article will display one or two representative models for different classes of viscosity models, and these classes are:

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.

<span class="mw-page-title-main">Sullivan vortex</span> Solution to the Navier–Stokes equations

In fluid dynamics, the Sullivan vortex is an exact solution of the Navier–Stokes equations describing a two-celled vortex in an axially strained flow, that was discovered by Roger D. Sullivan in 1959. At large radial distances, the Sullivan vortex resembles a Burgers vortex, however, it exhibits a two-cell structure near the center, creating a downdraft at the axis and an updraft at a finite radial location. Specifically, in the outer cell, the fluid spirals inward and upward and in the inner cell, the fluid spirals down at the axis and spirals upwards at the boundary with the outer cell. Due to its multi-celled structure, the vortex is used to model tornadoes and large-scale complex vortex structures in turbulent flows.

References

  1. Moffatt, H. K., Kida, S., & Ohkitani, K. (1994). Stretched vortices–the sinews of turbulence; large-Reynolds-number asymptotics. Journal of Fluid Mechanics, 259, 241-264.
  2. Taylor, G. I. (1934). The formation of emulsions in definable fields of flow. Proceedings of the Royal Society of London. Series A, containing papers of a mathematical and physical character, 146(858), 501-523.
  3. Wang, C. Y. (1974). Axisymmetric stagnation flow on a cylinder. Quarterly of Applied Mathematics, 32(2), 207-213.
  4. Craik, A. D. (2009). Exact vortex solutions of the Navier–Stokes equations with axisymmetric strain and suction or injection. Journal of fluid mechanics, 626, 291-306.
  5. Rajamanickam, P., & Weiss, A. D. (2021). Steady axisymmetric vortices in radial stagnation flows. The Quarterly Journal of Mechanics and Applied Mathematics, 74(3), 367-378.
  6. Hiemenz, Karl (1911) "Die Grenzschicht an einem in den gleichförmigen Flüssigkeitsstrom eingetauchten geraden Kreiszylinder"
  7. Howarth, Leslie (1934) On the calculation of steady flow in the boundary layer near the surface of a cylinder in a stream. No. ARC-R/M-1632. AERONAUTICAL RESEARCH COUNCIL LONDON (UNITED KINGDOM)
  8. Rosenhead, Louis, editor (1963) Laminar boundary layers, Clarendon Press
  9. Batchelor, George Keith (2000) An introduction to fluid dynamics, Cambridge University Press
  10. Rott, Nicholas. "Unsteady viscous flow in the vicinity of a stagnation point." Quarterly of Applied Mathematics 13.4 (1956): 444–451.
  11. Drazin, Philip G., and Norman Riley (2006) The Navier–Stokes equations: a classification of flows and exact solutions. No. 334. Cambridge University Press
  12. J. T. Stuart (2012) "The viscous flow near a stagnation point when the external flow has uniform vorticity." Journal of the Aerospace Sciences
  13. Tamada, Ko. "Two-dimensional stagnation-point flow impinging obliquely on a plane wall." Journal of the Physical Society of Japan 46 (1979): 310.
  14. Dorrepaal, J. M. "An exact solution of the Navier–Stokes equation which describes non-orthogonal stagnation-point flow in two dimensions." Journal of Fluid Mechanics 163 (1986): 141–147.
  15. Homann, Fritz. "Der Einfluss grosser Zähigkeit bei der Strömung um den Zylinder und um die Kugel." ZAMM‐Journal of Applied Mathematics and Mechanics/Zeitschrift für Angewandte Mathematik und Mechanik 16.3 (1936): 153–164.
  16. Libby, Paul A. "Wall shear at a three-dimensional stagnation point with a moving wall." AIAA Journal 12.3 (1974): 408–409.
  17. Libby, Paul A. "Laminar flow at a three-dimensional stagnation point with large rates of injection." AIAA Journal 14.9 (1976): 1273–1279.
  18. Wang, C. Y. "Stagnation flow on the surface of a quiescent fluid—an exact solution of the Navier–Stokes equations." Quarterly of applied mathematics 43.2 (1985): 215–223.
  19. Wang, C. Y. "Impinging stagnation flows." The Physics of fluids 30.3 (1987): 915–917.