Schneider flow

Last updated

Schneider flow describes the axisymmetric outer flow induced by a laminar or turbulent jet having a large jet Reynolds number or by a laminar plume with a large Grashof number, in the case where the fluid domain is bounded by a wall. When the jet Reynolds number or the plume Grashof number is large, the full flow field constitutes two regions of different extent: a thin boundary-layer flow that may identified as the jet or as the plume and a slowly moving fluid in the large outer region encompassing the jet or the plume. The Schneider flow describing the latter motion is an exact solution of the Navier-Stokes equations, discovered by Wilhelm Schneider in 1981. [1] The solution was discovered also by A. A. Golubinskii and V. V. Sychev in 1979, [2] [3] however, was never applied to flows entrained by jets. The solution is an extension of Taylor's potential flow solution [4] to arbitrary Reynolds number.

Contents

Mathematical description

Schneider flow: Numerical solution of the function
f
(
x
)
{\displaystyle f(\xi )}
(black lines) and
f
'
(
x
)
{\displaystyle f'(\xi )}
(blue lines) for various values of
K
{\displaystyle K}
and
x
w
=
0
{\displaystyle \xi _{w}=0}
(
a
=
p
/
2
{\displaystyle \alpha =\pi /2}
). Dashed lines correspond to the Taylor's solution
f
(
x
)
=
x
{\displaystyle f(\xi )=\xi }
. Schneider flow.jpg
Schneider flow: Numerical solution of the function (black lines) and (blue lines) for various values of and (). Dashed lines correspond to the Taylor's solution .

For laminar or turbulent jets and for laminar plumes, the volumetric entertainment rate per unit axial length is constant as can be seen from the solution of Schlichting jet and Yih plume. Thus, the jet or plume can be considered as a line sink that drives the motion in the outer region, as was first done by G. I. Taylor. Prior to Schneider, it was assumed that this outer fluid motion is also a large Reynolds number flow, hence the outer fluid motion is assumed to be a potential flow solution, which was solved by G. I. Taylor in 1958. For turbulent plume, the entrainment is not constant, nevertheless, the outer fluid is still governed by Taylors solution.

Though Taylor's solution is still true for turbulent jet, for laminar jet or laminar plume, the effective Reynolds number for outer fluid is found to be of order unity since the entertainment by the sink in these cases is such that the flow is not inviscid. In this case, full Navier-Stokes equations has to be solved for the outer fluid motion and at the same time, since the fluid is bounded from the bottom by a solid wall, the solution has to satisfy the non-slip condition. Schneider obtained a self-similar solution for this outer fluid motion, which naturally reduced to Taylor's potential flow solution as the entrainment rate by the line sink is increased.

Suppose a conical wall of semi-angle with polar axis along the cone-axis and assume the vertex of the solid cone sits at the origin of the spherical coordinates extending along the negative axis. Now, put the line sink along the positive side of the polar axis. Set this way, represents the common case of flat wall with jet or plume emerging from the origin. The case corresponds to jet/plume issuing from a thin injector. The flow is axisymmetric with zero azimuthal motion, i.e., the velocity components are . The usual technique to study the flow is to introduce the Stokes stream function such that

Introducing as the replacement for and introducing the self-similar form into the axisymmetric Navier-Stokes equations, we obtain [5]

where the constant is such that the volumetric entrainment rate per unit axial length is equal to . For laminar jet, and for laminar plume, it depends on the Prandtl number , for example with , we have and with , we have . For turbulent jet, this constant is the order of the jet Reynolds number, which is a large number.

The above equation can easily be reduced to a Riccati equation by integrating thrice, a procedure that is same as in the Landau–Squire jet (main difference between Landau-Squire jet and the current problem are the boundary conditions). The boundary conditions on the conical wall become

and along the line sink , we have

The problem has been solved numerically from here. The numerical solution also provides the values (the radial velocity at the axis), which must be accounted in the first-order boundary analysis of the inner jet problem at the axis.

Taylor's potential flow

For turbulent jet, , the linear terms in the equation can be neglected everywhere except near a small boundary layer along the wall. Then neglecting the non-slip conditions () at the wall, the solution, which was provided by G. I. Taylor in 1958, is given by [4]

In the case of axisymmetric turbulent plumes where the entrainment rate per unit axial length of the plume increases like , [6] Taylor's solution is given by where is a constant, is the specific buoyancy flux and [5]

in which denotes the associated Legendre function of the first kind with degree and order .

Composite solution for laminar jets and plumes

Equi-spaced contours of
ps
/
n
{\displaystyle \psi /\nu }
of the composite expansion, projected onto the
y
=
0
{\displaystyle y=0}
-plane, for the laminar jet with
R
e
=
50
{\displaystyle Re=50}
and
a
=
p
/
2
{\displaystyle \alpha =\pi /2}
. Schneider stream function.jpg
Equi-spaced contours of of the composite expansion, projected onto the -plane, for the laminar jet with and .
Equi-spaced contours of
ps
/
n
{\displaystyle \psi /\nu }
of the composite expansion, projected onto the
y
=
0
{\displaystyle y=0}
-plane, for the laminar jet with
R
e
=
50
{\displaystyle Re=50}
and
a
-
p
{\displaystyle \alpha \to \pi }
. Schneider stream2.jpg
Equi-spaced contours of of the composite expansion, projected onto the -plane, for the laminar jet with and .

The Schneider flow describes the outer motion driven by the jets or plumes and it becomes invalid in a thin region encompassing the axis where the jet or plume resides. For laminrar jets, the inner solution is described by the Schlichting jet and for laminar plumes, the inner solution is prescribed by Yih plume. A composite solution by stitching the inner thin Schlichting solution and the outer Schneider solution can be constructed by the method of matched asymptotic expansions. For the laminar jet, the composite solution is given by [5]

in which the first term respresents the Schlichting jet (with a characteristic jet thickness ), the second term represents the Schneider flow and the third term is the subtraction of the matching conditions. Here is the Reynolds number of the jet and is the kinematic momentum flux of the jet.

A similar composite solution can be constructed for the laminar plumes.

Other considerations

The exact solution of the Navier-Stokes solutions was verified experimentally by Zauner in 1985. [7] Further analysis [8] [9] showed that the axial momentum flux decays slowly along the axis unlike the Schlichting jet solution and it is found that the Schneider flow becomes invalid when distance from the origin increases to a distance of the order exponential of square of the jet Reynolds number, thus the domain of validity of Schneider solution increases with increasing jet Reynolds number.

Presence of swirl

The presence of swirling motion, i.e., is shown not to influence the axial motion given by provided . If is very large, the presence of swirl completely alters the motion on the axial plane. For , the azimuthal solution can be solved in terms of the circulation , where . The solution can be described in terms of the self-similar solution of the second kind, , where is an unknown constant and is an eigenvalue. The function satisfies [5]

subjected to the boundary conditions and as .

See also

Related Research Articles

<span class="mw-page-title-main">Laplace's equation</span> Second-order partial differential equation

In mathematics and physics, Laplace's equation is a second-order partial differential equation named after Pierre-Simon Laplace, who first studied its properties. This is often written as or where is the Laplace operator, is the divergence operator, is the gradient operator, and is a twice-differentiable real-valued function. The Laplace operator therefore maps a scalar function to another scalar function.

<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).

<span class="mw-page-title-main">Gamma distribution</span> Probability distribution

In probability theory and statistics, the gamma distribution is a versatile two-parameter family of continuous probability distributions. The exponential distribution, Erlang distribution, and chi-squared distribution are special cases of the gamma distribution. There are two equivalent parameterizations in common use:

  1. With a shape parameter k and a scale parameter θ
  2. With a shape parameter and an inverse scale parameter , called a rate parameter.

A directional derivative is a concept in multivariable calculus that measures the rate at which a function changes in a particular direction at a given point.

In theoretical physics, a supermultiplet is a representation of a supersymmetry algebra, possibly with extended supersymmetry.

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.

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">Gravitational lensing formalism</span>

In general relativity, a point mass deflects a light ray with impact parameter by an angle approximately equal to

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">Cnoidal wave</span> Nonlinear and exact periodic wave solution of the Korteweg–de Vries equation

In fluid dynamics, a cnoidal wave is a nonlinear and exact periodic wave solution of the Korteweg–de Vries equation. These solutions are in terms of the Jacobi elliptic function cn, which is why they are coined cnoidal waves. They are used to describe surface gravity waves of fairly long wavelength, as compared to the water depth.

In optics, the Fraunhofer diffraction equation is used to model the diffraction of waves when the diffraction pattern is viewed at a long distance from the diffracting object, and also when it is viewed at the focal plane of an imaging lens.

In the science of fluid flow, Stokes' paradox is the phenomenon that there can be no creeping flow of a fluid around a disk in two dimensions; or, equivalently, the fact there is no non-trivial steady-state solution for the Stokes equations around an infinitely long cylinder. This is opposed to the 3-dimensional case, where Stokes' method provides a solution to the problem of flow around a sphere.

Triple-deck theory is a theory that describes a three-layered boundary-layer structure when sufficiently large disturbances are present in the boundary layer. This theory is able to successfully explain the phenomenon of boundary layer separation, but it has found applications in many other flow setups as well, including the scaling of the lower-branch instability (T-S) of the Blasius flow, etc. James Lighthill, Lev Landau and others were the first to realize that to explain boundary layer separation, different scales other than the classical boundary-layer scales need to be introduced. These scales were first introduced independently by James Lighthill and E. A. Müller in 1953. The triple-layer structure itself was independently discovered by Keith Stewartson (1969) and V. Y. Neiland (1969) and by A. F. Messiter (1970). Stewartson and Messiter considered the separated flow near the trailing edge of a flat plate, whereas Neiland studied the case of a shock impinging on a boundary layer.

In image analysis, the generalized structure tensor (GST) is an extension of the Cartesian structure tensor to curvilinear coordinates. It is mainly used to detect and to represent the "direction" parameters of curves, just as the Cartesian structure tensor detects and represents the direction in Cartesian coordinates. Curve families generated by pairs of locally orthogonal functions have been the best studied.

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.

<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, Frank-Kamenetskii theory explains the thermal explosion of a homogeneous mixture of reactants, kept inside a closed vessel with constant temperature walls. It is named after a Russian scientist David A. Frank-Kamenetskii, who along with Nikolay Semenov developed the theory in the 1930s.

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 fluid dynamics, Hicks equation, sometimes also referred as Bragg–Hawthorne equation or Squire–Long equation, is a partial differential equation that describes the distribution of stream function for axisymmetric inviscid fluid, named after William Mitchinson Hicks, who derived it first in 1898. The equation was also re-derived by Stephen Bragg and William Hawthorne in 1950 and by Robert R. Long in 1953 and by Herbert Squire in 1956. The Hicks equation without swirl was first introduced by George Gabriel Stokes in 1842. The Grad–Shafranov equation appearing in plasma physics also takes the same form as the Hicks equation.

Moffatt eddies are sequences of eddies that develop in corners bounded by plane walls due to an arbitrary disturbance acting at asymptotically large distances from the corner. Although the source of motion is the arbitrary disturbance at large distances, the eddies develop quite independently and thus solution of these eddies emerges from an eigenvalue problem, a self-similar solution of the second kind.

References

  1. Schneider, W. (1981). Flow induced by jets and plumes. Journal of Fluid Mechanics, 108, 55–65.
  2. A. A. Golubinskii and V. V. Sychev, A similar solution of the Navier–Stokes equations, Uch. Zap. TsAGI 7 (1976) 11–17.
  3. Rajamanickam, P., & Weiss, A. D. (2020). A note on viscous flow induced by half-line sources bounded by conical surfaces. The Quarterly Journal of Mechanics and Applied Mathematics, 73(1), 24-35.
  4. 1 2 Taylor, G. (1958). Flow induced by jets. Journal of the Aerospace Sciences, 25(7), 464–465.
  5. 1 2 3 4 Coenen, W., Rajamanickam, P., Weiss, A. D., Sánchez, A. L., & Williams, F. A. (2019). Swirling flow induced by jets and plumes. Acta Mechanica, 230(6), 2221-2231.
  6. Batchelor, G. K. (1954). Heat convection and buoyancy effects in fluids. Quarterly journal of the royal meteorological society, 80(345), 339-358.
  7. Zauner, E. (1985). Visualization of the viscous flow induced by a round jet. Journal of Fluid Mechanics, 154, 111–119
  8. Mitsotakis, K., Schneider, W., & Zauner, E. (1984). Second-order boundary-layer theory of laminar jet flows. Acta mechanica, 53(1-2), 115–123.
  9. Schneider, W. (1985). Decay of momentum flux in submerged jets. Journal of Fluid Mechanics, 154, 91–110.