Law of the wall

Last updated
law of the wall, horizontal velocity near the wall with mixing length model Law of the wall (English).svg
law of the wall, horizontal velocity near the wall with mixing length model

In fluid dynamics, the law of the wall (also known as the logarithmic law of the wall) states that the average velocity of a turbulent flow at a certain point is proportional to the logarithm of the distance from that point to the "wall", or the boundary of the fluid region. This law of the wall was first published in 1930 by Hungarian-American mathematician, aerospace engineer, and physicist Theodore von Kármán. [1] It is only technically applicable to parts of the flow that are close to the wall (<20% of the height of the flow), though it is a good approximation for the entire velocity profile of natural streams. [2]

Contents

General logarithmic formulation

The logarithmic law of the wall is a self similar solution for the mean velocity parallel to the wall, and is valid for flows at high Reynolds numbers — in an overlap region with approximately constant shear stress and far enough from the wall for (direct) viscous effects to be negligible: [3]

  with     and  

where

is the wall coordinate: the distance y to the wall, made dimensionless with the friction velocity uτ and kinematic viscosity ν,
is the dimensionless velocity: the velocity u parallel to the wall as a function of y (distance from the wall), divided by the friction velocity uτ,
is the wall shear stress,
is the fluid density,
is called the friction velocity or shear velocity,
is the Von Kármán constant,
is a constant, and
is the natural logarithm.

From experiments, the von Kármán constant is found to be and for a smooth wall. [3]

With dimensions, the logarithmic law of the wall can be written as: [4]

where y0 is the distance from the boundary at which the idealized velocity given by the law of the wall goes to zero. This is necessarily nonzero because the turbulent velocity profile defined by the law of the wall does not apply to the laminar sublayer. The distance from the wall at which it reaches zero is determined by comparing the thickness of the laminar sublayer with the roughness of the surface over which it is flowing. For a near-wall laminar sublayer of thickness and a characteristic roughness length-scale , [2]

 : hydraulically smooth flow,
 : transitional flow,
 : hydraulically rough flow.

Intuitively, this means that if the roughness elements are hidden within the laminar sublayer, they have a much different effect on the turbulent law of the wall velocity profile than if they are sticking out into the main part of the flow.

This is also often more formally formulated in terms of a boundary Reynolds number, , where

The flow is hydraulically smooth for , hydraulically rough for , and transitional for intermediate values. [2]

Values for are given by: [2] [5]

 for hydraulically smooth flow
for hydraulically rough flow.

Intermediate values are generally given by the empirically derived Nikuradse diagram, [2] though analytical methods for solving for this range have also been proposed. [6]

For channels with a granular boundary, such as natural river systems,

where is the average diameter of the 84th largest percentile of the grains of the bed material. [7]

Power law solutions

Works by Barenblatt and others have shown that besides the logarithmic law of the wall — the limit for infinite Reynolds numbers — there exist power-law solutions, which are dependent on the Reynolds number. [8] [9] In 1996, Cipra submitted experimental evidence in support of these power-law descriptions. [10] This evidence itself has not been fully accepted by other experts. [11] In 2001, Oberlack claimed to have derived both the logarithmic law of the wall, as well as power laws, directly from the Reynolds-averaged Navier–Stokes equations, exploiting the symmetries in a Lie group approach. [3] [12] However, in 2014, Frewer et al. [13] refuted these results.

For scalars

For scalars (most notably temperature), the self-similar logarithmic law of the wall has been theorized (first formulated by B. A. Kader [14] ) and observed in experimental and computational studies. [15] [16] [17] [18] In many cases, extensions to the original law of the wall formulation (usually through integral transformations) are generally needed to account for compressibility, variable-property and real fluid effects.

Near the wall

Below the region where the law of the wall is applicable, there are other estimations for friction velocity. [19]

Viscous sublayer

In the region known as the viscous sublayer, below 5 wall units, the variation of to is approximately 1:1, such that:

For  

where,

is the wall coordinate: the distance y to the wall, made dimensionless with the friction velocity and kinematic viscosity ,
is the dimensionless velocity: the velocity u parallel to the wall as a function of y (distance from the wall), divided by the friction velocity ,

This approximation can be used farther than 5 wall units, but by the error is more than 25%.

Buffer layer

In the buffer layer, between 5 wall units and 30 wall units, neither law holds, such that:

For  

with the largest variation from either law occurring approximately where the two equations intercept, at . That is, before 11 wall units the linear approximation is more accurate and after 11 wall units the logarithmic approximation should be used, though neither are relatively accurate at 11 wall units.

The mean streamwise velocity profile is improved for with an eddy viscosity formulation based on a near-wall turbulent kinetic energy function and the van Driest mixing length equation. Comparisons with DNS data of fully developed turbulent channel flows for showed good agreement. [20]

Notes

  1. von Kármán, Th. (1930), "Mechanische Ähnlichkeit und Turbulenz", Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen, Fachgruppe 1 (Mathematik), 5: 58–76 (also as: “Mechanical Similitude and Turbulence”, Tech. Mem. NACA, no. 611, 1931).
  2. 1 2 3 4 5 Mohrig, David (2004). "Conservation of Mass and Momentum" (PDF). 12.110: Sedimentary Geology, Fall 2004. MIT OCW. Retrieved 2009-03-27.
  3. 1 2 3 Schlichting & Gersten (2000) pp. 522–524.
  4. Schlichting & Gersten (2000) p. 530.
  5. Whipple, Kelin (2004). "Hydraulic Roughness" (PDF). 12.163: Surface processes and landscape evolution. MIT OCW. Retrieved 2009-03-27.
  6. Le Roux, J.P. (2004), "An integrated law of the wall for hydrodynamically transitional flow over plane beds", Sedimentary Geology, 163 (3–4): 311–321, Bibcode:2004SedG..163..311L, doi:10.1016/j.sedgeo.2003.07.005
  7. Haws, Benjamin. "Equivalent sand roughness of Nikuradse (ks)" . Retrieved 2009-03-27.[ dead link ]
  8. Lynn Yarris. "A flaw in the law". Berkeley Lab: Highlights 97–98. Lawrence Berkeley National Laboratory, U.S. Department of Energy.
  9. Barenblatt, G.I. (1993), "Scaling laws for fully developed turbulent shear flows. Part 1. Basic hypotheses and analysis", Journal of Fluid Mechanics, 248: 513–520, Bibcode:1993JFM...248..513B, doi:10.1017/S0022112093000874, S2CID   123639410
    Barenblatt, G.I.; Prostokishin, V.M. (1993), "Scaling laws for fully developed turbulent shear flows. Part 2. Processing of experimental data", Journal of Fluid Mechanics, 248: 521–529, Bibcode:1993JFM...248..521B, doi:10.1017/S0022112093000886, S2CID   121328837
    Barenblatt, G.I.; Goldenfeld, N. (1995), "Does fully developed turbulence exist? Reynolds number independence versus asymptotic covariance", Physics of Fluids, 7 (12): 3078–3084, arXiv: cond-mat/9507132 , Bibcode:1995PhFl....7.3078B, doi:10.1063/1.868685, S2CID   15138376
    Barenblatt, G. I.; Chorin, A. J. (1998), "Scaling laws and vanishing-viscosity limits for wall-bounded shear flows and for local structure in developed turbulence", Communications on Pure and Applied Mathematics, 50 (4): 381–398, doi:10.1002/(SICI)1097-0312(199704)50:4<381::AID-CPA5>3.0.CO;2-6
  10. Cipra, Barry Arthur (May 1996), "A New Theory of Turbulence Causes a Stir Among Experts", Science, 272 (5264): 951, Bibcode:1996Sci...272..951C, doi:10.1126/science.272.5264.951, S2CID   117371905
  11. Zagarola, M.V.; Perry, A.E.; Smits, A.J. (1997), "Log Laws or Power Laws: The Scaling in the Overlap Region", Physics of Fluids, 9 (7): 2094–2100, Bibcode:1997PhFl....9.2094Z, CiteSeerX   10.1.1.503.989 , doi:10.1063/1.869328
  12. Oberlack, Martin (2001), "A unified approach for symmetries in plane parallel turbulent shear flows", Journal of Fluid Mechanics, 427: 299–328, Bibcode:2001JFM...427..299O, doi:10.1017/S0022112000002408, S2CID   122979735
  13. Frewer, Michael; Khujadze, George; Foysi, Holger (2014), Is the log-law a first principle result from Lie-group invariance analysis?, pp. 1–32, arXiv: 1412.3069 , Bibcode:2014arXiv1412.3069F
  14. Kader, B. A. (1981-09-01). "Temperature and concentration profiles in fully turbulent boundary layers". International Journal of Heat and Mass Transfer. 24 (9): 1541–1544. doi:10.1016/0017-9310(81)90220-9. ISSN   0017-9310.
  15. Simonich, J. C.; Bradshaw, P. (1978-11-01). "Effect of Free-Stream Turbulence on Heat Transfer through a Turbulent Boundary Layer". Journal of Heat Transfer. 100 (4): 671–677. doi:10.1115/1.3450875. ISSN   0022-1481.
  16. Patel, Ashish; Boersma, Bendiks J.; Pecnik, Rene (2017-08-21). "Scalar statistics in variable property turbulent channel flows". Physical Review Fluids. 2 (8): 084604. doi:10.1103/PhysRevFluids.2.084604.
  17. Toki, Takahiko; Teramoto, Susumu; Okamoto, Koji (2020-01-01). "Velocity and Temperature Profiles in Turbulent Channel Flow at Supercritical Pressure". Journal of Propulsion and Power. 36 (1): 3–13. doi:10.2514/1.B37381. S2CID   209963353.
  18. Guo, J.; Yang, X. I. A.; Ihme, M. (March 2022). "Structure of the thermal boundary layer in turbulent channel flows at transcritical conditions". Journal of Fluid Mechanics. 934. doi: 10.1017/jfm.2021.1157 . ISSN   0022-1120.
  19. Turbulent Flows (2000) pp. 273–274.Pope, Stephen (2000), Turbulent Flows (1st revised ed.), Cambridge University Press, ISBN   0-521-59125-2
  20. Absi, Rafik (2009), "A simple eddy viscosity formulation for turbulent boundary layers near smooth walls", Comptes Rendus Mécanique, 337 (3): 158–165, arXiv: 1106.0985 , Bibcode:2009CRMec.337..158A, doi:10.1016/j.crme.2009.03.010, S2CID   40907005

Related Research Articles

In fluid dynamics, turbulence or turbulent flow is fluid motion characterized by chaotic changes in pressure and flow velocity. It is in contrast to a laminar flow, which occurs when a fluid flows in parallel layers, with no disruption between those layers.

In fluid dynamics, the Darcy–Weisbach equation is an empirical equation that relates the head loss, or pressure loss, due to friction along a given length of pipe to the average velocity of the fluid flow for an incompressible fluid. The equation is named after Henry Darcy and Julius Weisbach. Currently, there is no formula more accurate or universally applicable than the Darcy-Weisbach supplemented by the Moody diagram or Colebrook equation.

<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">Parasitic drag</span> Aerodynamic resistance against the motion of an object

Parasitic drag, also known as profile drag, is a type of aerodynamic drag that acts on any object when the object is moving through a fluid. Parasitic drag is a combination of form drag and skin friction drag. It affects all objects regardless of whether they are capable of generating lift.

<span class="mw-page-title-main">Shear stress</span> Component of stress coplanar with a material cross section

Shear stress, often denoted by τ, is the component of stress coplanar with a material cross section. It arises from the shear force, the component of force vector parallel to the material cross section. Normal stress, on the other hand, arises from the force vector component perpendicular to the material cross section on which it acts.

<span class="mw-page-title-main">Surface layer</span> Layer of a turbulent fluid affected by interaction with a surface

The surface layer is the layer of a turbulent fluid most affected by interaction with a solid surface or the surface separating a gas and a liquid where the characteristics of the turbulence depend on distance from the interface. Surface layers are characterized by large normal gradients of tangential velocity and large concentration gradients of any substances transported to or from the interface.

<span class="mw-page-title-main">Large eddy simulation</span>

Large eddy simulation (LES) is a mathematical model for turbulence used in computational fluid dynamics. It was initially proposed in 1963 by Joseph Smagorinsky to simulate atmospheric air currents, and first explored by Deardorff (1970). LES is currently applied in a wide variety of engineering applications, including combustion, acoustics, and simulations of the atmospheric boundary layer.

In the field of fluid dynamics the point at which the boundary layer changes from laminar to turbulent is called the transition point. Where and how this transition occurs depends on the Reynolds number, the pressure gradient, pressure fluctuations due to sound, surface vibration, the initial turbulence level of the flow, boundary layer suction, surface heat flows, and surface roughness. The effects of a boundary layer turned turbulent are an increase in drag due to skin friction. As speed increases, the upper surface transition point tends to move forward. As the angle of attack increases, the upper surface transition point also tends to move forward.

<span class="mw-page-title-main">Plug flow</span>

In fluid mechanics, plug flow is a simple model of the velocity profile of a fluid flowing in a pipe. In plug flow, the velocity of the fluid is assumed to be constant across any cross-section of the pipe perpendicular to the axis of the pipe. The plug flow model assumes there is no boundary layer adjacent to the inner wall of the pipe.

In fluid dynamics, Couette flow is the flow of a viscous fluid in the space between two surfaces, one of which is moving tangentially relative to the other. The relative motion of the surfaces imposes a shear stress on the fluid and induces flow. Depending on the definition of the term, there may also be an applied pressure gradient in the flow direction.

The Fanning friction factor, named after John Thomas Fanning, is a dimensionless number used as a local parameter in continuum mechanics calculations. It is defined as the ratio between the local shear stress and the local flow kinetic energy density:

<span class="mw-page-title-main">Turbulence modeling</span>

Turbulence modeling is the construction and use of a mathematical model to predict the effects of turbulence. Turbulent flows are commonplace in most real life scenarios, including the flow of blood through the cardiovascular system, the airflow over an aircraft wing, the re-entry of space vehicles, besides others. In spite of decades of research, there is no analytical theory to predict the evolution of these turbulent flows. The equations governing turbulent flows can only be solved directly for simple cases of flow. For most real life turbulent flows, CFD simulations use turbulent models to predict the evolution of turbulence. These turbulence models are simplified constitutive equations that predict the statistical evolution of turbulent flows.

<span class="mw-page-title-main">Friction loss</span>

The term friction loss has a number of different meanings, depending on its context.

<span class="mw-page-title-main">Sediment transport</span> Movement of solid particles, typically by gravity and fluid entrainment

Sediment transport is the movement of solid particles (sediment), typically due to a combination of gravity acting on the sediment, and/or the movement of the fluid in which the sediment is entrained. Sediment transport occurs in natural systems where the particles are clastic rocks, mud, or clay; the fluid is air, water, or ice; and the force of gravity acts to move the particles along the sloping surface on which they are resting. Sediment transport due to fluid motion occurs in rivers, oceans, lakes, seas, and other bodies of water due to currents and tides. Transport is also caused by glaciers as they flow, and on terrestrial surfaces under the influence of wind. Sediment transport due only to gravity can occur on sloping surfaces in general, including hillslopes, scarps, cliffs, and the continental shelf—continental slope boundary.

In fluid dynamics, the von Kármán constant, named for Theodore von Kármán, is a dimensionless constant involved in the logarithmic law describing the distribution of the longitudinal velocity in the wall-normal direction of a turbulent fluid flow near a boundary with a no-slip condition. The equation for such boundary layer flow profiles is:

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.

<span class="mw-page-title-main">Reynolds number</span> Dimensionless quantity in fluid mechanics

In fluid mechanics, the Reynolds number is a dimensionless quantity that helps predict fluid flow patterns in different situations by measuring the ratio between inertial and viscous forces. At low Reynolds numbers, flows tend to be dominated by laminar (sheet-like) flow, while at high Reynolds numbers, flows tend to be turbulent. The turbulence results from differences in the fluid's speed and direction, which may sometimes intersect or even move counter to the overall direction of the flow. These eddy currents begin to churn the flow, using up energy in the process, which for liquids increases the chances of cavitation.

Particle-laden flows refers to a class of two-phase fluid flow, in which one of the phases is continuously connected and the other phase is made up of small, immiscible, and typically dilute particles. Fine aerosol particles in air is an example of a particle-laden flow; the aerosols are the dispersed phase, and the air is the carrier phase.

Reynolds stress equation model (RSM), also referred to as second moment closures are the most complete classical turbulence model. In these models, the eddy-viscosity hypothesis is avoided and the individual components of the Reynolds stress tensor are directly computed. These models use the exact Reynolds stress transport equation for their formulation. They account for the directional effects of the Reynolds stresses and the complex interactions in turbulent flows. Reynolds stress models offer significantly better accuracy than eddy-viscosity based turbulence models, while being computationally cheaper than Direct Numerical Simulations (DNS) and Large Eddy Simulations.

Skin friction drag is a type of aerodynamic or hydrodynamic drag, which is resistant force exerted on an object moving in a fluid. Skin friction drag is caused by the viscosity of fluids and is developed from laminar drag to turbulent drag as a fluid moves on the surface of an object. Skin friction drag is generally expressed in terms of the Reynolds number, which is the ratio between inertial force and viscous force.

References

Further reading