Capillary length

Last updated
The capillary length will vary for different liquids and different conditions. Here is a picture of a water droplet on a lotus leaf. If the temperature is 20 then
l
c
{\displaystyle \lambda _{c}}
= 2.71mm Water droplets in lotus2.JPG
The capillary length will vary for different liquids and different conditions. Here is a picture of a water droplet on a lotus leaf. If the temperature is 20 then = 2.71mm

The capillary length or capillary constant is a length scaling factor that relates gravity and surface tension. It is a fundamental physical property that governs the behavior of menisci, and is found when body forces (gravity) and surface forces (Laplace pressure) are in equilibrium.

Contents

The pressure of a static fluid does not depend on the shape, total mass or surface area of the fluid. It is directly proportional to the fluid's specific weight – the force exerted by gravity over a specific volume, and its vertical height. However, a fluid also experiences pressure that is induced by surface tension, commonly referred to as the Young–Laplace pressure. [1] Surface tension originates from cohesive forces between molecules, and in the bulk of the fluid, molecules experience attractive forces from all directions. The surface of a fluid is curved because exposed molecules on the surface have fewer neighboring interactions, resulting in a net force that contracts the surface. There exists a pressure difference either side of this curvature, and when this balances out the pressure due to gravity, one can rearrange to find the capillary length. [2]

In the case of a fluid–fluid interface, for example a drop of water immersed in another liquid, the capillary length denoted or is most commonly given by the formula,

,

where is the surface tension of the fluid interface, is the gravitational acceleration and is the mass density difference of the fluids. The capillary length is sometimes denoted in relation to the mathematical notation for curvature. The term capillary constant is somewhat misleading, because it is important to recognize that is a composition of variable quantities, for example the value of surface tension will vary with temperature and the density difference will change depending on the fluids involved at an interface interaction. However if these conditions are known, the capillary length can be considered a constant for any given liquid, and be used in numerous fluid mechanical problems to scale the derived equations such that they are valid for any fluid. [3] For molecular fluids, the interfacial tensions and density differences are typically of the order of mN m−1 and g mL−1 respectively resulting in a capillary length of mm for water and air at room temperature on earth. [4] On the other hand, the capillary length would be mm for water-air on the moon. For a soap bubble, the surface tension must be divided by the mean thickness, resulting in a capillary length of about meters in air! [5] The equation for can also be found with an extra term, most often used when normalising the capillary height. [6]

Origin

Theoretical

One way to theoretically derive the capillary length, is to imagine a liquid droplet at the point where surface tension balances gravity.

Let there be a spherical droplet with radius ,

The characteristic Laplace pressure , due to surface tension, is equal to

,

where is the surface tension. The pressure due to gravity (hydrostatic pressure) of a column of liquid is given by

,

where is the droplet density, the gravitational acceleration, and is the height of the droplet.

At the point where the Laplace pressure balances out the pressure due to gravity ,

.

Relationship with the Eötvös number

The above derivation can be used when dealing with the Eötvös number, a dimensionless quantity that represents the ratio between the buoyancy forces and surface tension of the liquid. Despite being introduced by Loránd Eötvös in 1886, he has since become fairly dissociated with it, being replaced with Wilfrid Noel Bond such that it is now referred to as the Bond number in recent literature.

The Bond number can be written such that it includes a characteristic length- normally the radius of curvature of a liquid, and the capillary length [7]

,

with parameters defined above, and the radius of curvature.

Therefore the bond number can be written as

,

with the capillary length.

If the bond number is set to 1, then the characteristic length is the capillary length.

Experimental

The capillary length can also be found through the manipulation of many different physical phenomenon. One method is to focus on capillary action, which is the attraction of a liquids surface to a surrounding solid. [8]

Association with Jurin's law

Jurin's law is a quantitative law that shows that the maximum height that can be achieved by a liquid in a capillary tube is inversely proportional to the diameter of the tube. The law can be illustrated mathematically during capillary uplift, which is a traditional experiment measuring the height of a liquid in a capillary tube. When a capillary tube is inserted into a liquid, the liquid will rise or fall in the tube, due to an imbalance in pressure. The characteristic height is the distance from the bottom of the meniscus to the base, and exists when the Laplace pressure and the pressure due to gravity are balanced. One can reorganize to show the capillary length as a function of surface tension and gravity.

,

with the height of the liquid, the radius of the capillary tube, and the contact angle.

The contact angle is defined as the angle formed by the intersection of the liquid-solid interface and the liquid–vapour interface. [2] The size of the angle quantifies the wettability of liquid, i.e., the interaction between the liquid and solid surface. A contact angle of can be considered, perfect wetting.

.

Thus the forms a cyclical 3 factor equation with .

This property is usually used by physicists to estimate the height a liquid will rise in a particular capillary tube, radius known, without the need for an experiment. When the characteristic height of the liquid is sufficiently less than the capillary length, then the effect of hydrostatic pressure due to gravity can be neglected. [9]

Using the same premises of capillary rise, one can find the capillary length as a function of the volume increase, and wetting perimeter of the capillary walls. [10]

Association with a sessile droplet

Another way to find the capillary length is using different pressure points inside a sessile droplet, with each point having a radius of curvature, and equate them to the Laplace pressure equation. This time the equation is solved for the height of the meniscus level which again can be used to give the capillary length.

The shape of a sessile droplet is directly proportional to whether the radius is greater than or less than the capillary length. Microdrops are droplets with radius smaller than the capillary length, and their shape is governed solely by surface tension, forming a spherical cap shape. If a droplet has a radius larger than the capillary length, they are known as macrodrops and the gravitational forces will dominate. Macrodrops will be 'flattened' by gravity and the height of the droplet will be reduced. [11]

The capillary length against radii of a droplet Sessile drop Capillary Length.jpg
The capillary length against radii of a droplet

History

The investigations in capillarity stem back as far as Leonardo da Vinci, however the idea of capillary length was not developed until much later. Fundamentally the capillary length is a product of the work of Thomas Young and Pierre Laplace. They both appreciated that surface tension arose from cohesive forces between particles and that the shape of a liquid's surface reflected the short range of these forces. At the turn of the 19th century they independently derived pressure equations, but due to notation and presentation, Laplace often gets the credit. The equation showed that the pressure within a curved surface between two static fluids is always greater than that outside of a curved surface, but the pressure will decrease to zero as the radius approached infinity. Since the force is perpendicular to the surface and acts towards the centre of the curvature, a liquid will rise when the surface is concave and depress when convex. [12] This was a mathematical explanation of the work published by James Jurin in 1719, [13] where he quantified a relationship between the maximum height taken by a liquid in a capillary tube and its diameter – Jurin's law. [10] The capillary length evolved from the use of the Laplace pressure equation at the point it balanced the pressure due to gravity, and is sometimes called the Laplace capillary constant, after being introduced by Laplace in 1806. [14]

In nature

Bubbles

The size of soap bubbles are limited by the capillary length. Soap bubble 10 (cropped).JPG
The size of soap bubbles are limited by the capillary length.

Like a droplet, bubbles are round because cohesive forces pull its molecules into the tightest possible grouping, a sphere. Due to the trapped air inside the bubble, it is impossible for the surface area to shrink to zero, hence the pressure inside the bubble is greater than outside, because if the pressures were equal, then the bubble would simply collapse. [15] This pressure difference can be calculated from Laplace's pressure equation,

.

For a soap bubble, there exists two boundary surfaces, internal and external, and therefore two contributions to the excess pressure and Laplace's formula doubles to

. [16]

The capillary length can then be worked out the same way except that the thickness of the film, must be taken into account as the bubble has a hollow center, unlike the droplet which is a solid. Instead of thinking of a droplet where each side is as in the above derivation, for a bubble is now

,

with and the radius and thickness of the bubble respectively.

As above, the Laplace and hydrostatic pressure are equated resulting in

.

Thus the capillary length contributes to a physiochemical limit that dictates the maximum size a soap bubble can take. [5]

See also

Related Research Articles

<span class="mw-page-title-main">Surface tension</span> Tendency of a liquid surface to shrink to reduce surface area

Surface tension is the tendency of liquid surfaces at rest to shrink into the minimum surface area possible. Surface tension is what allows objects with a higher density than water such as razor blades and insects to float on a water surface without becoming even partly submerged.

<span class="mw-page-title-main">Foam</span> Form of matter

Foams are materials formed by trapping pockets of gas in a liquid or solid.

<span class="mw-page-title-main">Capillary wave</span> Wave on the surface of a fluid, dominated by surface tension

A capillary wave is a wave traveling along the phase boundary of a fluid, whose dynamics and phase velocity are dominated by the effects of surface tension.

In physics, Washburn's equation describes capillary flow in a bundle of parallel cylindrical tubes; it is extended with some issues also to imbibition into porous materials. The equation is named after Edward Wight Washburn; also known as Lucas–Washburn equation, considering that Richard Lucas wrote a similar paper three years earlier, or the Bell-Cameron-Lucas-Washburn equation, considering J.M. Bell and F.K. Cameron's discovery of the form of the equation in 1906.

<span class="mw-page-title-main">Wetting</span> Ability of a liquid to maintain contact with a solid surface

Wetting is the ability of a liquid to maintain contact with a solid surface, resulting from intermolecular interactions when the two are brought together. This happens in presence of a gaseous phase or another liquid phase not miscible with the first one. The degree of wetting (wettability) is determined by a force balance between adhesive and cohesive forces. There are two types of wetting: non-reactive wetting and reactive wetting.

<span class="mw-page-title-main">Marangoni effect</span> Physical phenomenon between two fluids

The Marangoni effect is the mass transfer along an interface between two phases due to a gradient of the surface tension. In the case of temperature dependence, this phenomenon may be called thermo-capillary convection.

The Ostwald–Freundlich equation governs boundaries between two phases; specifically, it relates the surface tension of the boundary to its curvature, the ambient temperature, and the vapor pressure or chemical potential in the two phases.

The Tolman length measures the extent by which the surface tension of a small liquid drop deviates from its planar value. It is conveniently defined in terms of an expansion in , with the equimolar radius of the liquid drop, of the pressure difference across the droplet's surface:

In fluid statics, capillary pressure is the pressure between two immiscible fluids in a thin tube, resulting from the interactions of forces between the fluids and solid walls of the tube. Capillary pressure can serve as both an opposing or driving force for fluid transport and is a significant property for research and industrial purposes. It is also observed in natural phenomena.

In fluid dynamics the Eötvös number (Eo), also called the Bond number (Bo), is a dimensionless number measuring the importance of gravitational forces compared to surface tension forces for the movement of liquid front. Alongside the Capillary number, commonly denoted , which represents the contribution of viscous drag, is useful for studying the movement of fluid in porous or granular media, such as soil. The Bond number is also used to characterize the shape of bubbles or drops moving in a surrounding fluid. The two names used for this dimensionless term commemorate the Hungarian physicist Loránd Eötvös (1848–1919) and the English physicist Wilfrid Noel Bond (1897–1937), respectively. The term Eötvös number is more frequently used in Europe, while Bond number is commonly used in other parts of the world.

The Kelvin equation describes the change in vapour pressure due to a curved liquid–vapor interface, such as the surface of a droplet. The vapor pressure at a convex curved surface is higher than that at a flat surface. The Kelvin equation is dependent upon thermodynamic principles and does not allude to special properties of materials. It is also used for determination of pore size distribution of a porous medium using adsorption porosimetry. The equation is named in honor of William Thomson, also known as Lord Kelvin.

In physics, the Young–Laplace equation is an algebraic equation that describes the capillary pressure difference sustained across the interface between two static fluids, such as water and air, due to the phenomenon of surface tension or wall tension, although use of the latter is only applicable if assuming that the wall is very thin. The Young–Laplace equation relates the pressure difference to the shape of the surface or wall and it is fundamentally important in the study of static capillary surfaces. It is a statement of normal stress balance for static fluids meeting at an interface, where the interface is treated as a surface :

In fluid mechanics and mathematics, a capillary surface is a surface that represents the interface between two different fluids. As a consequence of being a surface, a capillary surface has no thickness in slight contrast with most real fluid interfaces.

In fluid dynamics, Airy wave theory gives a linearised description of the propagation of gravity waves on the surface of a homogeneous fluid layer. The theory assumes that the fluid layer has a uniform mean depth, and that the fluid flow is inviscid, incompressible and irrotational. This theory was first published, in correct form, by George Biddell Airy in the 19th century.

<span class="mw-page-title-main">Laplace pressure</span> Pressure difference between the inside and the outside of a curved surface

The Laplace pressure is the pressure difference between the inside and the outside of a curved surface that forms the boundary between two fluid regions. The pressure difference is caused by the surface tension of the interface between liquid and gas, or between two immiscible liquids.

The spinning drop method or rotating drop method is one of the methods used to measure interfacial tension. Measurements are carried out in a rotating horizontal tube which contains a dense fluid. A drop of a less dense liquid or a gas bubble is placed inside the fluid. Since the rotation of the horizontal tube creates a centrifugal force towards the tube walls, the liquid drop will start to deform into an elongated shape; this elongation stops when the interfacial tension and centrifugal forces are balanced. The surface tension between the two liquids can then be derived from the shape of the drop at this equilibrium point. A device used for such measurements is called a “spinning drop tensiometer”.

<span class="mw-page-title-main">Jurin's law</span>

Jurin's law, or capillary rise, is the simplest analysis of capillary action—the induced motion of liquids in small channels—and states that the maximum height of a liquid in a capillary tube is inversely proportional to the tube's diameter. Capillary action is one of the most common fluid mechanical effects explored in the field of microfluidics. Jurin's law is named after James Jurin, who discovered it between 1718 and 1719. His quantitative law suggests that the maximum height of liquid in a capillary tube is inversely proportional to the tube's diameter. The difference in height between the surroundings of the tube and the inside, as well as the shape of the meniscus, are caused by capillary action. The mathematical expression of this law can be derived directly from hydrostatic principles and the Young–Laplace equation. Jurin's law allows the measurement of the surface tension of a liquid and can be used to derive the capillary length.

<span class="mw-page-title-main">Visco-elastic jets</span>

Visco-elastic jets are the jets of viscoelastic fluids, i.e. fluids that disobey Newton's law of Viscocity. A Viscoelastic fluid that returns to its original shape after the applied stress is released.

<span class="mw-page-title-main">Capillary bridges</span> Minimised surface of liquid commecting two wetted objects

A capillary bridge is a minimized surface of liquid or membrane created between two rigid bodies of arbitrary shape. Capillary bridges also may form between two liquids. Plateau defined a sequence of capillary shapes known as (1) nodoid with 'neck', (2) catenoid, (3) unduloid with 'neck', (4) cylinder, (5) unduloid with 'haunch' (6) sphere and (7) nodoid with 'haunch'. The presence of capillary bridge, depending on their shapes, can lead to attraction or repulsion between the solid bodies. The simplest cases of them are the axisymmetric ones. We distinguished three important classes of bridging, depending on connected bodies surface shapes:

In the theory of capillarity, Bosanquet equation is an improved modification of the simpler Lucas–Washburn theory for the motion of a liquid in a thin capillary tube or a porous material that can be approximated as a large collection of capillaries. In the Lucas–Washburn model, the inertia of the fluid is ignored, leading to the assumption that flow is continuous under constant viscous laminar Poiseuille flow conditions without considering the effects of mass transport undergoing acceleration occurring at the start of flow and at points of changing internal capillary geometry. The Bosanquet equation is a differential equation that is second-order in the time derivative, similar to Newton's Second Law, and therefore takes into account the fluid inertia. Equations of motion, like the Washburn's equation, that attempt to explain a velocity as proportional to a driving force are often described with the term Aristotelian mechanics.

References

  1. Nguyen, Anh V.; Schulze, Hans Joachim (2004). Colloidal science of flotation. New York: Marcel Dekker. ISBN   978-0824747824. OCLC   53390392.
  2. 1 2 Yuan, Yuehua; Lee, T. Randall (2013), Bracco, Gianangelo; Holst, Bodil (eds.), "Contact Angle and Wetting Properties", Surface Science Techniques, Springer Series in Surface Sciences, vol. 51, Springer Berlin Heidelberg, pp. 3–34, doi:10.1007/978-3-642-34243-1_1, ISBN   9783642342424, S2CID   133761573
  3. Rapp, Bastian E. (2016-12-13). Microfluidics : modeling, mechanics and mathematics. Kidlington, Oxford, United Kingdom. ISBN   9781455731510. OCLC   966685733.{{cite book}}: CS1 maint: location missing publisher (link)
  4. Aarts, D. G. A. L. (2005). "Capillary Length in a Fluid−Fluid Demixed Colloid−Polymer Mixture". The Journal of Physical Chemistry B. 109 (15): 7407–7411. doi:10.1021/jp044312q. hdl: 1874/14751 . ISSN   1520-6106. PMID   16851848. S2CID   32362123.
  5. 1 2 Clanet, Christophe; Quéré, David; Snoeijer, Jacco H.; Reyssat, Etienne; Texier, Baptiste Darbois; Cohen, Caroline (2017-03-07). "On the shape of giant soap bubbles". Proceedings of the National Academy of Sciences. 114 (10): 2515–2519. Bibcode:2017PNAS..114.2515C. doi: 10.1073/pnas.1616904114 . ISSN   0027-8424. PMC   5347548 . PMID   28223485.
  6. Boucher, E A (1980-04-01). "Capillary phenomena: Properties of systems with fluid/fluid interfaces". Reports on Progress in Physics. 43 (4): 497–546. doi:10.1088/0034-4885/43/4/003. ISSN   0034-4885. S2CID   250817869.
  7. Liu, Tingyi “Leo”; Kim, Chang-Jin “CJ” (2017). "Contact Angle Measurement of Small Capillary Length Liquid in Super-repelled State". Scientific Reports. 7 (1): 740. Bibcode:2017NatSR...7..740L. doi:10.1038/s41598-017-00607-9. ISSN   2045-2322. PMC   5428877 . PMID   28389672.
  8. Cleveland, Cutler J.; Morris, Christopher G. (2014-10-20). Dictionary of energy. Cleveland, Cutler J.,, Morris, Christopher G. (Second ed.). Amsterdam, Netherlands. ISBN   9780080968124. OCLC   896841847.{{cite book}}: CS1 maint: location missing publisher (link)
  9. Noam, Eliaz; Gileadi, Eliezer (2018-09-13). Physical electrochemistry: fundamentals, techniques and applications (Second ed.). Weinheim. ISBN   9783527341405. OCLC   1080923071.{{cite book}}: CS1 maint: location missing publisher (link)
  10. 1 2 Kashin, V. V.; Shakirov, K. M.; Poshevneva, A. I. (2011). "The capillary constant in calculating the surface tension of liquids". Steel in Translation. 41 (10): 795–798. doi:10.3103/S0967091211100093. ISSN   0967-0912. S2CID   137015683.
  11. Berthier, Jean; Silberzan, Pacal (2010). Microfluidics for biotechnology (2nd ed.). Boston: Artech House. ISBN   9781596934443. OCLC   642685865.
  12. West, John B. (1996). Respiratory Physiology: People and Ideas. New York, NY: Springer New York. ISBN   9781461475200. OCLC   852791684.
  13. "Jurin". Philosophical Transactions of the Royal Society of London. 30 (355): 739–747. 1719. doi:10.1098/rstl.1717.0026. S2CID   186211806.
  14. L. Landau and B. Levich, “Dragging of a liquid by a moving plate,” Acta Physicochimica U.R.S.S., Vol. 17, No. 1-2, 1942, pp. 42-54
  15. Agarwal, P.K. IIT Physics-I. Krishna Prakashan Media.
  16. Darvell, B.W. (2009-04-29). Materials science for dentistry (Ninth ed.). Cambridge, England. ISBN   9781845696672. OCLC   874155175.{{cite book}}: CS1 maint: location missing publisher (link)