Inviscid flow

Last updated

In fluid dynamics, inviscid flow is the flow of an inviscid fluid which is a fluid with zero viscosity. [1]

Contents

The Reynolds number of inviscid flow approaches infinity as the viscosity approaches zero. When viscous forces are neglected, such as the case of inviscid flow, the Navier–Stokes equation can be simplified to a form known as the Euler equation. This simplified equation is applicable to inviscid flow as well as flow with low viscosity and a Reynolds number much greater than one. Using the Euler equation, many fluid dynamics problems involving low viscosity are easily solved, however, the assumed negligible viscosity is no longer valid in the region of fluid near a solid boundary (the boundary layer) or, more generally in regions with large velocity gradients which are evidently accompanied by viscous forces. [1] [2] [3]

The flow of a superfluid is inviscid. [4]

Inviscid flows are broadly classified into potential flows (or, irrotational flows) and rotational inviscid flows.

Prandtl hypothesis

These diagrams show the dividing streamlines associated with an airfoil in two-dimensional inviscid flow.
The upper diagram shows zero circulation and zero lift. It implies high-speed vortex flow at the trailing edge which is known to be inaccurate in a model of the steady state.
The lower diagram shows the Kutta condition which implies finite circulation, finite lift, and no vortex flow at the trailing edge. These characteristics are known to be accurate as models of the steady state in a real fluid. Airfoil Kutta condition.jpg
These diagrams show the dividing streamlines associated with an airfoil in two-dimensional inviscid flow.
The upper diagram shows zero circulation and zero lift. It implies high-speed vortex flow at the trailing edge which is known to be inaccurate in a model of the steady state.
The lower diagram shows the Kutta condition which implies finite circulation, finite lift, and no vortex flow at the trailing edge. These characteristics are known to be accurate as models of the steady state in a real fluid.

Ludwig Prandtl developed the modern concept of the boundary layer. His hypothesis establishes that for fluids of low viscosity, shear forces due to viscosity are evident only in thin regions at the boundary of the fluid, adjacent to solid surfaces. Outside these regions, and in regions of favorable pressure gradient, viscous shear forces are absent so the fluid flow field can be assumed to be the same as the flow of an inviscid fluid. By employing the Prandtl hypothesis it is possible to estimate the flow of a real fluid in regions of favorable pressure gradient by assuming inviscid flow and investigating the irrotational flow pattern around the solid body. [5]

Real fluids experience separation of the boundary layer and resulting turbulent wakes but these phenomena cannot be modelled using inviscid flow. Separation of the boundary layer usually occurs where the pressure gradient reverses from favorable to adverse so it is inaccurate to use inviscid flow to estimate the flow of a real fluid in regions of unfavorable pressure gradient. [5]

Reynolds number

The Reynolds number (Re) is a dimensionless quantity that is commonly used in fluid dynamics and engineering. [6] [7] Originally described by George Gabriel Stokes in 1850, it became popularized by Osborne Reynolds after whom the concept was named by Arnold Sommerfeld in 1908. [7] [8] [9] The Reynolds number is calculated as:

SymbolDescriptionUnits
characteristic length m
fluid velocitym/s
fluid densitykg/m3
fluid viscosityPa*s

The value represents the ratio of inertial forces to viscous forces in a fluid, and is useful in determining the relative importance of viscosity. [6] In inviscid flow, since the viscous forces are zero, the Reynolds number approaches infinity. [1] When viscous forces are negligible, the Reynolds number is much greater than one. [1] In such cases (Re>>1), assuming inviscid flow can be useful in simplifying many fluid dynamics problems.

Euler equations

Inviscid flow around a wing, assuming circulation that achieves the Kutta condition Flow around a wing.gif
Inviscid flow around a wing, assuming circulation that achieves the Kutta condition

In a 1757 publication, Leonhard Euler described a set of equations governing inviscid flow: [10]

SymbolDescriptionUnits
material derivative
del operator
pressurePa
acceleration vector due to gravitym/s2

Assuming inviscid flow allows the Euler equation to be applied to flows in which viscous forces are insignificant. [1] Some examples include flow around an airplane wing, upstream flow around bridge supports in a river, and ocean currents. [1]

In 1845, George Gabriel Stokes published another important set of equations, today known as the Navier-Stokes equations. [1] [11] Claude-Louis Navier developed the equations first using molecular theory, which was further confirmed by Stokes using continuum theory. [1] The Navier-Stokes equations describe the motion of fluids: [1]

When the fluid is inviscid, or the viscosity can be assumed to be negligible, the Navier-Stokes equation simplifies to the Euler equation: [1] This simplification is much easier to solve, and can apply to many types of flow in which viscosity is negligible. [1] Some examples include flow around an airplane wing, upstream flow around bridge supports in a river, and ocean currents. [1]

The Navier-Stokes equation reduces to the Euler equation when . Another condition that leads to the elimination of viscous force is , and this results in an "inviscid flow arrangement". [12] Such flows are found to be vortex-like.

Flow developing over a solid surface Boundary Layer en.png
Flow developing over a solid surface

Solid boundaries

It is important to note, that negligible viscosity can no longer be assumed near solid boundaries, such as the case of the airplane wing. [1] In turbulent flow regimes (Re >> 1), viscosity can typically be neglected, however this is only valid at distances far from solid interfaces. [1] When considering flow in the vicinity of a solid surface, such as flow through a pipe or around a wing, it is convenient to categorize four distinct regions of flow near the surface: [1]

Although these distinctions can be a useful tool in illustrating the significance of viscous forces near solid interfaces, it is important to note that these regions are fairly arbitrary. [1] Assuming inviscid flow can be a useful tool in solving many fluid dynamics problems, however, this assumption requires careful consideration of the fluid sub layers when solid boundaries are involved.

Superfluids

Superfluid helium Liquid helium Rollin film.jpg
Superfluid helium

Superfluid is the state of matter that exhibits frictionless flow, zero viscosity, also known as inviscid flow. [4]

To date, helium is the only fluid to exhibit superfluidity that has been discovered. Helium-4 becomes a superfluid once it is cooled to below 2.2K, a point known as the lambda point. [13] At temperatures above the lambda point, helium exists as a liquid exhibiting normal fluid dynamic behavior. Once it is cooled to below 2.2K it begins to exhibit quantum behavior. For example, at the lambda point there is a sharp increase in heat capacity, as it is continued to be cooled, the heat capacity begins to decrease with temperature. [14] In addition, the thermal conductivity is very large, contributing to the excellent coolant properties of superfluid helium. [15] Similarly, Helium-3 is found become a superfluid at 2.491mK.

Applications

Spectrometers are kept at a very low temperature using helium as the coolant. This allows for minimal background flux in far-infrared readings. Some of the designs for the spectrometers may be simple, but even the frame is at its warmest less than 20 Kelvin. These devices are not commonly used as it is very expensive to use superfluid helium over other coolants. [16]

Large Hadron Collider Large Hadron Collider dipole magnets IMG 0955.jpg
Large Hadron Collider

Superfluid helium has a very high thermal conductivity, which makes it very useful for cooling superconductors. Superconductors such as the ones used at the LHC (Large Hadron Collider) are cooled to temperatures of approximately 1.9 Kelvin. This temperature allows the niobium-titanium magnets to reach a superconductor state. Without the use of the superfluid helium, this temperature would not be possible. Using helium to cool to these temperatures is very expensive and cooling systems that use alternative fluids are more numerous. [17]

Another application of the superfluid helium is its uses in understanding quantum mechanics. Using lasers to look at small droplets allows scientists to view behaviors that may not normally be viewable. This is due to all the helium in each droplet being at the same quantum state. This application does not have any practical uses by itself, but it helps us better understand quantum mechanics which has its own applications.

See also

Related Research Articles

<span class="mw-page-title-main">Aerodynamics</span> Branch of dynamics concerned with studying the motion of air

Aerodynamics is the study of the motion of air, particularly when affected by a solid object, such as an airplane wing. It involves topics covered in the field of fluid dynamics and its subfield of gas dynamics, and is an important domain of study in aeronautics. The term aerodynamics is often used synonymously with gas dynamics, the difference being that "gas dynamics" applies to the study of the motion of all gases, and is not limited to air. The formal study of aerodynamics began in the modern sense in the eighteenth century, although observations of fundamental concepts such as aerodynamic drag were recorded much earlier. Most of the early efforts in aerodynamics were directed toward achieving heavier-than-air flight, which was first demonstrated by Otto Lilienthal in 1891. Since then, the use of aerodynamics through mathematical analysis, empirical approximations, wind tunnel experimentation, and computer simulations has formed a rational basis for the development of heavier-than-air flight and a number of other technologies. Recent work in aerodynamics has focused on issues related to compressible flow, turbulence, and boundary layers and has become increasingly computational in nature.

<span class="mw-page-title-main">Fluid dynamics</span> Aspects of fluid mechanics involving flow

In physics, physical chemistry and engineering, fluid dynamics is a subdiscipline of fluid mechanics that describes the flow of fluids—liquids and gases. It has several subdisciplines, including aerodynamics and hydrodynamics. Fluid dynamics has a wide range of applications, including calculating forces and moments on aircraft, determining the mass flow rate of petroleum through pipelines, predicting weather patterns, understanding nebulae in interstellar space and modelling fission weapon detonation.

<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).

The vorticity equation of fluid dynamics describes the evolution of the vorticity ω of a particle of a fluid as it moves with its flow; that is, the local rotation of the fluid. The governing equation 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.

Quantum turbulence is the name given to the turbulent flow – the chaotic motion of a fluid at high flow rates – of quantum fluids, such as superfluids. The idea that a form of turbulence might be possible in a superfluid via the quantized vortex lines was first suggested by Richard Feynman. The dynamics of quantum fluids are governed by quantum mechanics, rather than classical physics which govern classical (ordinary) fluids. Some examples of quantum fluids include superfluid helium, Bose–Einstein condensates (BECs), polariton condensates, and nuclear pasta theorized to exist inside neutron stars. Quantum fluids exist at temperatures below the critical temperature at which Bose-Einstein condensation takes place.

In fluid dynamics, the Boussinesq approximation is used in the field of buoyancy-driven flow. It ignores density differences except where they appear in terms multiplied by g, the acceleration due to gravity. The essence of the Boussinesq approximation is that the difference in inertia is negligible but gravity is sufficiently strong to make the specific weight appreciably different between the two fluids. Sound waves are impossible/neglected when the Boussinesq approximation is used since sound waves move via density variations.

In continuum mechanics, the Froude number is a dimensionless number defined as the ratio of the flow inertia to the external field. The Froude number is based on the speed–length ratio which he defined as:

<span class="mw-page-title-main">D'Alembert's paradox</span>

In fluid dynamics, d'Alembert's paradox is a contradiction reached in 1752 by French mathematician Jean le Rond d'Alembert. D'Alembert proved that – for incompressible and inviscid potential flow – the drag force is zero on a body moving with constant velocity relative to the fluid. Zero drag is in direct contradiction to the observation of substantial drag on bodies moving relative to fluids, such as air and water; especially at high velocities corresponding with high Reynolds numbers. It is a particular example of the reversibility paradox.

<span class="mw-page-title-main">Stokes flow</span> Type of fluid flow

Stokes flow, also named creeping flow or creeping motion, is a type of fluid flow where advective inertial forces are small compared with viscous forces. The Reynolds number is low, i.e. . This is a typical situation in flows where the fluid velocities are very slow, the viscosities are very large, or the length-scales of the flow are very small. Creeping flow was first studied to understand lubrication. In nature, this type of flow occurs in the swimming of microorganisms and sperm. In technology, it occurs in paint, MEMS devices, and in the flow of viscous polymers generally.

In fluid dynamics, drag is a force acting opposite to the relative motion of any object moving with respect to a surrounding fluid. This can exist between two fluid layers or between a fluid and a solid surface.

Aeroacoustics is a branch of acoustics that studies noise generation via either turbulent fluid motion or aerodynamic forces interacting with surfaces. Noise generation can also be associated with periodically varying flows. A notable example of this phenomenon is the Aeolian tones produced by wind blowing over fixed objects.

Fluid mechanics is the branch of physics concerned with the mechanics of fluids and the forces on them. It has applications in a wide range of disciplines, including mechanical, aerospace, civil, chemical, and biomedical engineering, as well as geophysics, oceanography, meteorology, astrophysics, and biology.

The intent of this article is to highlight the important points of the derivation of the Navier–Stokes equations as well as its application and formulation for different families of fluids.

Pressure-correction method is a class of methods used in computational fluid dynamics for numerically solving the Navier-Stokes equations normally for incompressible flows.

Volume viscosity is a material property relevant for characterizing fluid flow. Common symbols are or . It has dimensions, and the corresponding SI unit is the pascal-second (Pa·s).

In a body submerged in a fluid, unsteady forces due to acceleration of that body with respect to the fluid, can be divided into two parts: the virtual mass effect and the Basset force.

<span class="mw-page-title-main">Hydrodynamic stability</span> Subfield of fluid dynamics

In fluid dynamics, hydrodynamic stability is the field which analyses the stability and the onset of instability of fluid flows. The study of hydrodynamic stability aims to find out if a given flow is stable or unstable, and if so, how these instabilities will cause the development of turbulence. The foundations of hydrodynamic stability, both theoretical and experimental, were laid most notably by Helmholtz, Kelvin, Rayleigh and Reynolds during the nineteenth century. These foundations have given many useful tools to study hydrodynamic stability. These include Reynolds number, the Euler equations, and the Navier–Stokes equations. When studying flow stability it is useful to understand more simplistic systems, e.g. incompressible and inviscid fluids which can then be developed further onto more complex flows. Since the 1980s, more computational methods are being used to model and analyse the more complex flows.

<span class="mw-page-title-main">Reynolds number</span> Ratio of inertial to viscous forces acting on a liquid

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.

In fluid mechanics, non-dimensionalization of the Navier–Stokes equations is the conversion of the Navier–Stokes equation to a nondimensional form. This technique can ease the analysis of the problem at hand, and reduce the number of free parameters. Small or large sizes of certain dimensionless parameters indicate the importance of certain terms in the equations for the studied flow. This may provide possibilities to neglect terms in the considered flow. Further, non-dimensionalized Navier–Stokes equations can be beneficial if one is posed with similar physical situations – that is problems where the only changes are those of the basic dimensions of the system.

References

  1. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 E., Stewart, Warren; N., Lightfoot, Edwin (2007-01-01). Transport phenomena. Wiley. ISBN   9780470115398. OCLC   762715172.{{cite book}}: CS1 maint: multiple names: authors list (link)
  2. Clancy, L.J., Aerodynamics, p.xviii
  3. Kundu, P.K., Cohen, I.M., & Hu, H.H., Fluid Mechanics, Chapter 10, sub-chapter 1
  4. 1 2 S., Stringari (2016). Bose-Einstein condensation and superfluidity. Oxford University Press. ISBN   9780198758884. OCLC   936040211.
  5. 1 2 Streeter, Victor L. (1966) Fluid Mechanics, sections 5.6 and 7.1, 4th edition, McGraw-Hill Book Co., Library of Congress Catalog Card Number 66-15605
  6. 1 2 L., Bergman, Theodore; S., Lavine, Adrienne; P., Incropera, Frank; P., Dewitt, David (2011-01-01). Fundamentals of heat and mass transfer. Wiley. ISBN   9780470501979. OCLC   875769912.{{cite book}}: CS1 maint: multiple names: authors list (link)
  7. 1 2 Rott, N (2003-11-28). "Note on the History of the Reynolds Number". Annual Review of Fluid Mechanics. 22 (1): 1–12. Bibcode:1990AnRFM..22....1R. doi:10.1146/annurev.fl.22.010190.000245.
  8. Reynolds, Osborne (1883-01-01). "An Experimental Investigation of the Circumstances Which Determine Whether the Motion of Water Shall Be Direct or Sinuous, and of the Law of Resistance in Parallel Channels". Philosophical Transactions of the Royal Society of London. 174: 935–982. Bibcode:1883RSPT..174..935R. doi: 10.1098/rstl.1883.0029 . ISSN   0261-0523.
  9. Stokes, G. G. (1851-01-01). "On the Effect of the Internal Friction of Fluids on the Motion of Pendulums". Transactions of the Cambridge Philosophical Society. 9: 8. Bibcode:1851TCaPS...9....8S.
  10. Euler, Leonhard (1757). ""Principes généraux de l'état d'équilibre d'un fluide" [General principles of the state of equilibrium]". Mémoires de l'académie des sciences de Berlin. 11: 217–273.
  11. Stokes, G. G. (1845). "On the Theories of the Internal Friction of Fluids in Motion and of the Equilibrium and Motion of Elastic Solids". Proc. Camb. Phil. Soc. 8: 287–319.
  12. Runstedtler, Allan (2013). "Inviscid Flow Arrangements in Fluid Dynamics". International Journal of Fluid Mechanics Research. 40 (2): 148–158. doi:10.1615/interjfluidmechres.v40.i2.50. ISSN   1064-2277.
  13. "This Month in Physics History". www.aps.org. Retrieved 2017-03-07.
  14. Landau, L. (1941). "Theory of the Superfluidity of Helium II". Physical Review. 60 (4): 356–358. Bibcode:1941PhRv...60..356L. doi:10.1103/physrev.60.356.
  15. "nature physics portal - looking back - Going with the flow -- superfluidity observed". www.nature.com. Retrieved 2017-03-07.
  16. HOUCK, J. R.; WARD, DENNIS (1979-01-01). "A liquid-helium-cooled grating spectrometer for far infrared astronomical observations". Publications of the Astronomical Society of the Pacific. 91 (539): 140–142. Bibcode:1979PASP...91..140H. doi: 10.1086/130456 . JSTOR   40677459. S2CID   120273071.
  17. "Cryogenics: Low temperatures, high performance | CERN". home.cern. Retrieved 2017-02-14.