Capillary bridges

Last updated

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. [1] Plateau defined a sequence of capillary shapes [2] 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:

Contents

fig.1 Concave capillary bridge between two planes(schematic representation) Concave capullary bridge.jpg
fig.1 Concave capillary bridge between two planes(schematic representation)
fig.2 Concave capillary bridge between particles and flat surface(schematic representation) Capillary bridge flat spherical surface.png
fig.2 Concave capillary bridge between particles and flat surface(schematic representation)
fig.3 Concave capillary bridge between two particles(schematic representation) Grains cap bridge.png
fig.3 Concave capillary bridge between two particles(schematic representation)

Capillary bridges and their properties may also be influenced by Earth gravity and by properties of the bridged surfaces. The bridging substance may be a liquid or a gas. The enclosing boundary is called the interface (capillary surface). The interface is characterized by a particular surface tension.

History

Capillary bridges have been studied for over 200 years. The question was raised for the first time by Josef Louis Lagrange in 1760, and interest was further spread by the French astronomer and mathematician C. Delaunay. [3] Delaunay found an entirely new class of axially symmetrical surfaces of constant mean curvature. The formulation and the proof of his theorem had a long story. It began with Euler's [4] proposition of new figure, called catenoid. (Much later, Kenmotsu [5] solved the complex nonlinear equations, describing this class of surfaces. However, his solution is of little practical importance because it has no geometrical interpretation.) J. Plateau showed the existence of such shapes with given boundaries. The problem was named after him Plateau's problem. [6]
Many scientists contributed to the solution of the problem. One of them is Thomas Young. [7] Pierre Simon Laplace contributed the notion of capillary tension. Laplace even formulated the widely known nowadays condition for mechanical equilibrium between two fluids, divided by a capillary surface PγP i.e. capillary pressure between two phases is balanced by their adjacent pressure difference.
A general survey on capillary bridge behavior in gravity field is completed by Myshkis and Babskii. [8]
In the last century a lot of efforts were put of study of surface forces that drive capillary effects of bridging. There was established that these forces result from intermolecular forces and become significant in thin fluid gaps (<10 nm) between two surfaces. [9] [10]
The instability of capillary bridges was discussed in first time by Rayleigh. [11] He demonstrated that a liquid jet or capillary cylindrical surface became unstable when the ratio between its length, H to the radius R, becomes bigger than 2π. In these conditions of small sinusoidal perturbations with wavelength bigger than its perimeter, the cylinder surface area becomes larger than the one of unperturbed cylinder with the same volume and thus it becomes unstable. Later, Hove [12] formulated the variational requirements for the stability of axisymmetric capillary surfaces (unbounded) in absence of gravity and with disturbances constrained to constant volume. He first solved Young-Laplace equation for equilibrium shapes and showed that the Legendre condition for the second variation is always satisfied. Therefore, the stability is determined by the absence of negative eigenvalue of the linearized Young-Laplace equation. This approach of determining stability from second variation is used now widely. [8] Perturbation methods became very successful despite that nonlinear nature of capillary interaction can limit their application. Other methods now include direct simulation. [13] [14] To that moment most methods for stability determination required calculation of equilibrium as a basis for perturbations. There appeared a new idea that stability may be deduced from equilibrium states. [15] [16] The proposition was further proven by Pitts [17] for axisymmetric constant volume. In the following years Vogel [18] [19] extended the theory. He examined the case of axisymmetric capillary bridges with constant volumes and the stability changes correspond to turning points. The recent development of bifurcation theory proved that exchange of stability between turning points and branch points is a general phenomenon. [20] [21]

Applications and occurrences

Recent studies indicated that ancient Egyptians used the properties of sand to create capillary bridges by using water on it. [22] In this way, they reduced surface friction and were capable to move statues and heavy pyramid stones. Some contemporary arts, like sand art, are also close related to capability of water to bridge particles. In atomic force microscopy, when one works in higher humidity environment, their studies might be affected by the appearance of nano sized capillary bridges. [23] These bridges appear when the working tip approaches the studied sample. Capillary bridges also play important role in soldering process. [24]


Capillary bridges also widely spread in living nature. Bugs, flies, grasshoppers and tree frogs are capable to adhere to vertical rough surfaces because of their ability to inject wetting liquid into the pad-substrate contact area. This way is created long range attractive interaction due to the formation of capillary bridges. [25] Many medical problems involving respiratory diseases, and the health of the body joints depend on tiny capillary bridges. [26] Liquid bridges are now commonly used in growth of cell cultures because of the need to mimic work of living tissues in scientific research. [27] [28]

General equations

General solution for the profile of capillary is known from consideration of unduloid or nodoid curvature. [29]
Let's assume the following cylindrical coordinate system: z shows axis of revolution; r represents radial coordinate and φ is the angle between the normal and the positive z axis. The nodoid has vertical tangents at r = r1 and r = r2 and horizontal tangent at r = r3. When φ is the angle between the normal to the interface and positive z axis then φ is equal to 90°, 0°, -90° for nodoid.

The Young-Laplace equation may be written in a form convenient for integration for axial symmetry :

where R1, R2 are the radii of curvature and γ is interfacial surface tension.
The integration of the equation is called the first integral and it yields:

Since:

One finds:

After the integration, the obtained equation is called the second integral:

where: F and E are elliptic integrals of first and second kind, and φ is related to r according to

.

The unduloid has only vertical tangents at r=r1 and r=r2, where φ = + 90. In a completely analogous way:

The second integral for unduloid is obtained:

where the relation between parameters k and φ are defined the same way as above. In the limiting case r1=0, both nodoid and unduloid consist of a series of spheres. When r1=r2. The last and the very interesting limiting case is catenoid. The Laplace equation is reduced to:

It integration can be represented in very convenient form, in cylindrical coordinate system, called catenary equation: [29]

Fig. 4. Catenoid existence domain (1) height is scaled by radius R, (2) height is scaled by cubic root of volume (only valid for capillary bridges of C=0) Catenoid existence domain.png
Fig. 4. Catenoid existence domain (1) height is scaled by radius R, (2) height is scaled by cubic root of volume (only valid for capillary bridges of C=0)

Equation (9) is important because it shows in some simplification all issues, related to the capillary bridges, transparent. Drawing in dimensionless coordinates exhibit a maximum, that distinguishes two branches. One of them is energetically favorable and come up to existence in statics while the other (in dashed line) is not energetically favorable. Maximum is important because when stretching quasi-equilibrium way capillary bridge, if maximum is reached, it breakage takes place. Catenoids with energetically unfavorable dimensions may form during process of dynamical stretching/pressing. [30] Zero capillary pressure C=0 is natural for classical catenoid (capillary soap surface stretched between two coaxial rings). When typical capillary bridge comes to catenoidal state of C = 0, despite that it surface properties are the same as classical catenoid, it is more appropriate to be presented as scaled by cube root of its volume rather than the radius, R.

The solution of the second integral is different in cases of oblate capillary bridges (nodoid and unduloid):

where: F and E are again elliptic integrals of first and second kind, and φ is related to r according to: .
It is important to note that all described curves are found by rolling a conic section without slip along z axis. The unduloid is described by the focus of rolling ellipse, which can degenerate into a line, a sphere or a parabola, leading to the corresponding limiting cases. Similarly, a nodoid is described by the focus of a rolling hyperbola.

Well systematized summary of capillary bridges shapes is given in table 11.1 of Kralchevsky and Nagayama's book. [2]

Statics between two flat surfaces

The mechanical equilibrium comprises the pressure balance on liquid/gas interface and the external force on plates, ΔP, balancing the capillary attraction or repulsion, , i.e. . Upon neglecting gravity effects and other external fields, the pressure balance is ΔP=Pi - Pe (The indexes "i" and "e" denote correspondingly internal and external pressures). In case of axial symmetry, the equation for capillary pressure takes the form:

where γ is interfacial liquid/gas tension; r is radial coordinate and φ is the angle between the axis symmetry and normal to interface generatrix.
The first integral is easily obtained regarding dimensionless capillary pressure at the contact with surface:

where , dimensionless radius at the contact is and θ is the contact angle. The relation shows that capillary pressure can be positive or negative. The shape of capillary bridges is governed by the equation: [2]

where the equation is obtained after substitution is made in Eq. ( 11 ) and scaling is introduced.

Thin liquid bridge

In contrast to cases with increasing height of capillary bridges, that poses variety of profile shapes, the flattening (thinning) toward zero thickness has much more universal character. The universality appears when H<<R (fig. 1). Equation (11) may be written: [31]

The generatrix converges to equation:

Upon integration, the equation yields:

fig.5. Thin liquid bridge Thin layer (liquid capillary bridge).png
fig.5. Thin liquid bridge

The dimensionless circular radii 1/2C coincides with capillary bridge radii of curvature. The positive sign '+' represents generatrix profile of concave bridge and negative '-', oblate. For the convex capillary bridges, the circular generatrix is retained until the boundary of definition domain is reached while stretching. Near the beginning of self-initiated breakage kinetics, the bridge profile evolves consequently to an ellipse, parabola and possibly to hyperbola. [32]

Definition domain

The observations, presented in fig. 5 indicate that a domain of capillary bridges existence can be defined. Therefore, if stretching of a liquid bridge it might discontinue its existence not only because of raising instabilities but also because of reaching of some points that the shape can not exist anymore. The estimation of definition domain requires manipulation of integrated equations for capillary bridge height and its volume. Both they are integrable but the integrals are improper. The applied method includes splitting of the integrals on two parts: singular but integrable analytically and regular but integrable only numerical way.
After the integration, for the capillary bridge height is obtained [31]

Similar way for contact radius R, is obtained the integrated equation [31]

where and

Fig. 6. Isogones demonstrating capillary bridges static domain, red curve shows catenoidal state of C=0 Relation of liquid static capillary bridge dimensionless height and dimensionless radius in cylindrical coordinates.jpg
Fig. 6. Isogones demonstrating capillary bridges static domain, red curve shows catenoidal state of C=0

In fig. 6 are shown number of stable static states of liquid capillary bridge, represented by two characteristic parameters: (i) dimensionless height that is obtained by scaling of capillary bridge height by cubic root of its volume Eq. ( 16 ) and (ii) its radius, also scaled by cubic root of volume, Eq. ( 17 ). The partially analytical solutions, obtained for these two parameters, are presented above. The solutions somehow differs from widely accepted Plateau's approach [by elliptical functions, Eq. ( 7 )], because they offer convenient numerical approach for integration of regular integrals, while irregular part of the equation was integrated analytically. These solutions became further a basis for prediction of capillary bridges quasi-equilibrium stretching and breakage for contact angles below 45°. The practical implementation allows to be identified not only the end of definition domain but also the exact behavior during the capillary bridge stretching, [32] because in coordinates stretching forms an inclined line, where the inclination angle is proportional to the contact angle.

Concave capillary bridge

The case of concave capillary bridge is presented by isogones for contact angles below in fig. 6, . The isogones show well defined maximum . This maximum is noted by dot for each isogone. It again, similarly to a simple catenoid, separates two branches. The left branch is energetically favorable while the right one is energetically unfavourable.

Cylindrical capillary bridge

This case is analyzed well by Rayleigh. Note that the definition domain in his case shows no limitations and it goes to infinity, fig. 6, . However, the breakage of cylindrical capillary bridges is usually observed. It takes place as result from well studied instability known now as Rayleigh instability. [11] The definition domain for 90° isogone in shown in fig 6 by dashed line.

Convex capillary bridge

The case of convex capillary bridges is presented in fig. 6, left from the domain of cylindrical case.

Stability between two flat surfaces

Equilibrium shapes and stability limits for capillary liquid bridges are subject to many theoretical and experimental studies. [33] Studies are mostly concentrated on investigation of bridges between equals disks under gravitational conditions. It is well known that for each value of the Bond number, defined as [34] (where: g is Earth gravitational acceleration, γ is the surface tension and R is radius of the contact) the stability diagram can be represented by a single closed piecewise curve on the slenderness/dimensionless volume plane. Slenderness is defined as , and the dimensionless volume is capillary bridge volume divided on cylinder volume with the same height, H and radius R: .

If both slenderness and liquid volume are small enough, the stability limits are governed by detachment of liquid shape from the edges of the disks (three-phase contact line), AB line in fig. 7. The line BC represents minimum in volume that corresponds to axisymmetrical breakage. It is known in literature as minimum volume stability limit. The curve CA represents another limit to stability, characterizing maximum volume. It is upper bound to the stability region. There also exists a transition region between minimum and maximum volume stability. It is not yet clearly defined and thus is noted by dashed line in fig. 7.[ where? ]

See also

Related Research Articles

<span class="mw-page-title-main">Ideal gas</span> Mathematical model which approximates the behavior of real gases

An ideal gas is a theoretical gas composed of many randomly moving point particles that are not subject to interparticle interactions. The ideal gas concept is useful because it obeys the ideal gas law, a simplified equation of state, and is amenable to analysis under statistical mechanics. The requirement of zero interaction can often be relaxed if, for example, the interaction is perfectly elastic or regarded as point-like collisions.

<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">Surface energy</span> Excess energy at the surface of a material relative to its interior

In surface science, surface energy quantifies the disruption of intermolecular bonds that occurs when a surface is created. In solid-state physics, surfaces must be intrinsically less energetically favorable than the bulk of the material, otherwise there would be a driving force for surfaces to be created, removing the bulk of the material by sublimation. The surface energy may therefore be defined as the excess energy at the surface of a material compared to the bulk, or it is the work required to build an area of a particular surface. Another way to view the surface energy is to relate it to the work required to cut a bulk sample, creating two surfaces. There is "excess energy" as a result of the now-incomplete, unrealized bonding between the two created surfaces.

In mathematics, the discrete Laplace operator is an analog of the continuous Laplace operator, defined so that it has meaning on a graph or a discrete grid. For the case of a finite-dimensional graph, the discrete Laplace operator is more commonly called the Laplacian matrix.

<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 displace gas 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">Contact angle</span> Angle between a liquid–vapor interface and a solid surface

The contact angle is the angle between a liquid surface and a solid surface where they meet. More specifically, it is the angle between the surface tangent on the liquid–vapor interface and the tangent on the solid–liquid interface at their intersection. It quantifies the wettability of a solid surface by a liquid via the Young equation.

<span class="mw-page-title-main">Dendrite (metal)</span>

A dendrite in metallurgy is a characteristic tree-like structure of crystals growing as molten metal solidifies, the shape produced by faster growth along energetically favourable crystallographic directions. This dendritic growth has large consequences in regard to material properties.

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 Green's function for the Laplacian in three variables is used to describe the response of a particular type of physical system to a point source. In particular, this Green's function arises in systems that can be described by Poisson's equation, a partial differential equation (PDE) of the form where is the Laplace operator in , is the source term of the system, and is the solution to the equation. Because is a linear differential operator, the solution to a general system of this type can be written as an integral over a distribution of source given by : where the Green's function for Laplacian in three variables describes the response of the system at the point to a point source located at : and the point source is given by , the Dirac delta function.

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 : where is the Laplace pressure, the pressure difference across the fluid interface, is the surface tension, is the unit normal pointing out of the surface, is the mean curvature, and and are the principal radii of curvature. Note that only normal stress is considered, because a static interface is possible only in the absence of tangential stress.

<span class="mw-page-title-main">Capillary length</span>

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 are in equilibrium.

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.

<span class="mw-page-title-main">Capillary condensation</span> Ability of porous media to condense liquid from an adsorbed vapor

In materials science and biology, capillary condensation is the "process by which multilayer adsorption from the vapor [phase] into a porous medium proceeds to the point at which pore spaces become filled with condensed liquid from the vapor [phase]." The unique aspect of capillary condensation is that vapor condensation occurs below the saturation vapor pressure, Psat, of the pure liquid. This result is due to an increased number of van der Waals interactions between vapor phase molecules inside the confined space of a capillary. Once condensation has occurred, a meniscus immediately forms at the liquid-vapor interface which allows for equilibrium below the saturation vapor pressure. Meniscus formation is dependent on the surface tension of the liquid and the shape of the capillary, as shown by the Young-Laplace equation. As with any liquid-vapor interface involving a meniscus, the Kelvin equation provides a relation for the difference between the equilibrium vapor pressure and the saturation vapor pressure. A capillary does not necessarily have to be a tubular, closed shape, but can be any confined space with respect to its surroundings.

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.

The Gibbs–Thomson effect, in common physics usage, refers to variations in vapor pressure or chemical potential across a curved surface or interface. The existence of a positive interfacial energy will increase the energy required to form small particles with high curvature, and these particles will exhibit an increased vapor pressure. See Ostwald–Freundlich equation. More specifically, the Gibbs–Thomson effect refers to the observation that small crystals are in equilibrium with their liquid melt at a lower temperature than large crystals. In cases of confined geometry, such as liquids contained within porous media, this leads to a depression in the freezing point / melting point that is inversely proportional to the pore size, as given by the Gibbs–Thomson equation.

<span class="mw-page-title-main">Electrodynamic droplet deformation</span> Liquid droplets suspended in a liquid exposed to an oscillating electric field

Electrohydrodynamic droplet deformation is a phenomenon that occurs when liquid droplets suspended in a second immiscible liquid are exposed to an oscillating electric field. Under these conditions, the droplet will periodically deform between prolate and oblate ellipsoidal shapes. The characteristic frequency and magnitude of the deformation is determined by a balance of electrodynamic, hydrodynamic, and capillary stresses acting on the droplet interface. This phenomenon has been studied extensively both mathematically and experimentally because of the complex fluid dynamics that occur. Characterization and modulation of electrodynamic droplet deformation is of particular interest for engineering applications because of the growing need to improve the performance of complex industrial processes(e.g. two-phase cooling, crude oil demulsification). The primary advantage of using oscillatory droplet deformation to improve these engineering processes is that the phenomenon does not require sophisticated machinery or the introduction of heat sources. This effectively means that improving performance via oscillatory droplet deformation is simple and in no way diminishes the effectiveness of the existing engineering system.

References

  1. Ross, Sydney (1950). "The Inhibition of Foaming. II. A Mechanism for the Rupture of Liquid Films by Anti-foaming Agents". J. Phys. Chem. 54 (3): 429–436. doi:10.1021/j150477a018.
  2. 1 2 3 Kralchevsky, P.; Nagayama, K. (2001). Particles at Fluid Interfaces and Membranes. Amsterdam: Elsevier. pp. 469–502.
  3. Delaunay (1841). "Sur la Surface de Revolution dont la Courbure Moyenne est Constante". J. Math. Pures Appl. 6: 309–314.
  4. L. Euler, Methodus inveniendi lineas curvas maximi minimive proprietate gaudentes, Opera omnia, I, 24, (1744)
  5. Kenmotsu, K., Surfaces of revolution with prescribed mean curvature, Tohoku Math. J. 32(1980), 147–153.
  6. Plateau, Josef (1873). Statique expérimentale et théorique des liquides soumis aux seulesforces moléculaires. Gauthier-Villars.
  7. Young, T (1805). "An Essay on the Cohesion of Fluids". Philos. Trans. R. Soc. Lond. 95: 65–87. doi: 10.1098/rstl.1805.0005 .
  8. 1 2 A.D. Myshkis and V.G. Babskii, Low-Gravity Fluid Mechanics: Mathematical Theory of Capillary Phenomena, Springer-Verlag 1987
  9. Nikolai V. Churaev, B.V. Derjaguin, V.M. Muller, Surface Forces, Springer Scoence and Business Media, 1987
  10. J. Israelashvilly, Intermolecular and Surface Forces, Third Edition: Revised, Elsevier, 2011
  11. 1 2 Lord Rayleigh (1878). "On the instability of jets". Proceedings of the London Mathematical Society. 10: 4–13.
  12. Hove, W., Ph.D. Dissertation, Friendlich-Wilhelms, Universitat zu Berlin (1887)
  13. Meseguer, J.; Sanz, A. (1985). "Numerical and experimental study of the dynamics of axisymmetric liquid bridges". J. Fluid Mech. 153: 83. doi:10.1017/s002211208500115x.
  14. Martinez; Perales, J. M. (1986). "Liquid bridge stability data". J. Cryst. Growth. 78 (2): 369. Bibcode:1986JCrGr..78..369M. doi:10.1016/0022-0248(86)90073-4.
  15. J. F. Padday, A. R. Pitt, The Stability of Axisymmetric Menisci, Philosophical Transactions of the Royal Society A, (1973)
  16. Boucher, E. A.; Evans, M. J. B. (1975). "Pendent Drop Profiles and Related Capillary Phenomena". Proceedings of the Royal Society of London A: Mathematical and Physical Sciences. 346 (1646): 349–374. Bibcode:1975RSPSA.346..349B. doi:10.1098/rspa.1975.0180.
  17. Pitts, R., The Stability of a Drop Hanging from a Tube, IMA J Appl Math (1976) 17 (3): 387-397
  18. Vogel, Thomas I., Stability of a liquid drop trapped between two parallel planes, SIAM J. Appl. Math. 47 (1987), 516–525
  19. Vogel, Thomas I., Stability of a liquid drop trapped between two parallel planes II, SIAM J. Appl. Math. 49 (1989), 1009–1028
  20. Michael, D. H., Annual Review of Fluid Mechanics Vol. 13: 189-216 (Volume publication date January 1981)
  21. Brian James, Lowry; Steen, Paul H. (1995). "Capillary Surfaces: Stability from Families of Equilibria with Application to the Liquid Bridge". Proceedings of the Royal Society A. 449 (1937): 411–439. Bibcode:1995RSPSA.449..411L. doi:10.1098/rspa.1995.0051.
  22. A. Fall, B. Weber, M. Pakpour, N. Lenoir, N. Shahidzadeh, J. Fiscina, C. Wagner, and D. Bonn, Sliding Friction on Wet and Dry Sand Phys. Rev. Lett. 112, 175502, Published 29 April 2014
  23. Yumei Men, Xianren Zhang and Wenchuan Wang, Capillary liquid bridges in atomic force microscopy(AFM): Formation, rupture, and hysteresis, J. Chem. Phys. 131, 184702 (2009)
  24. Ft. B. EDWARDS, Joint Tolerances in Capillary Copper Piping Joints, Welding Journal, v06, pp 321-(1972)
  25. Persson, B N J (2007). "Wet adhesion with application to tree frog adhesive toe pads and tires". J. Phys.: Condens. Matter. 19 (37): 376110. Bibcode:2007JPCM...19K6110P. doi:10.1088/0953-8984/19/37/376110. S2CID   73607791.
  26. Alencar, A.M.; Majumdar, A.; Hantos, Z.; Buldyrev, S.V.; Stanley, H.E.; Suki, B. (2005). "Crackles and instabilities during lung inflation". Physica A. 357: 18–26. Bibcode:2005PhyA..357...18A. doi:10.1016/j.physa.2005.05.047.
  27. Pampaloni, F.; Reynaud, E.G.; Stelzer, E.H.K. (2007). "The third dimension bridges the gap between cell culture and live tissue". Nature Reviews Molecular Cell Biology. 8 (10): 839–845. doi:10.1038/nrm2236. PMID   17684528.
  28. G. DiLisi, R. Dempsey, R. Rarick, and C. Rosenblatt, Using Parabolic Flights to Examine Quantitatively the Stability of Liquid Bridges under Varying Total Body Force, Sci. and Tech. (submitted)
  29. 1 2 Princen N.M., "The Equilibrium Shape of Interfaces, Drops and Bubbles Rigid and Deformable Particles at Interfaces" in Surface and Colloid Science, ed. E. Matijievich, (1969), v.3, New York
  30. Masato Ito, Taku Sato, In-situ observation of a soap film catenoid - a simple educational physics experiment, Eur. J. Phys. 31 (2010) 357-365
  31. 1 2 3 Petkov, P.V.; Radoev, B. R. (2014). "Statics and dynamics of capillary bridges". Colloids and Surfaces A. 460: 18–27. doi:10.1016/j.colsurfa.2014.03.038.
  32. 1 2 Petkov, P.V.; Radoev, B. R. (2019). "Investigation of Single and Binary of "Sandwich" Type Convex Liquid Capillary Bridges, Stretched between Two Flat Surfaces (Experimental Approach)". Colloids and Interfaces. 3 (68): 68. doi: 10.3390/colloids3040068 .
  33. Bezdenejnykh, N. A.; Meseguer, J.; Perales, J. M. (1992). "Experimental analysis of stability limits of capillary liquid bridges". Phys. Fluids A. 4 (4): 677. Bibcode:1992PhFlA...4..677B. doi:10.1063/1.858286.
  34. Shi, Z.; et al. (2018). "Dynamic contact angle hysteresis in liquid bridges". Colloids and Surfaces A: Physicochemical and Engineering Aspects. 555: 365–371. arXiv: 1712.04703 . doi:10.1016/j.colsurfa.2018.07.004.