fbpx
Wikipedia

Contact angle

The contact angle (symbol θC) 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.

Schematic of a liquid drop showing the quantities in the Young equation.

A given system of solid, liquid, and vapor at a given temperature and pressure has a unique equilibrium contact angle. However, in practice a dynamic phenomenon of contact angle hysteresis is often observed, ranging from the advancing (maximal) contact angle to the receding (minimal) contact angle.[1] The equilibrium contact is within those values, and can be calculated from them. The equilibrium contact angle reflects the relative strength of the liquid, solid, and vapour molecular interaction.

The contact angle depends upon the medium above the free surface of the liquid, and the nature of the liquid and solid in contact. It is independent of the inclination of solid to the liquid surface. It changes with surface tension and hence with the temperature and purity of the liquid.

Thermodynamics edit

 
Cloth, treated to be hydrophobic, shows a high contact angle.

The theoretical description of contact angle arises from the consideration of a thermodynamic equilibrium between the three phases: the liquid phase (L), the solid phase (S), and the gas or vapor phase (G) (which could be a mixture of ambient atmosphere and an equilibrium concentration of the liquid vapor). (The "gaseous" phase could be replaced by another immiscible liquid phase.) If the solid–vapor interfacial energy is denoted by γSG, the solid–liquid interfacial energy by γSL, and the liquid–vapor interfacial energy (i.e. the surface tension) by γLG, then the equilibrium contact angle θC is determined from these quantities by the Young equation:

 

The contact angle can also be related to the work of adhesion via the Young–Dupré equation:

 

where   is the solid – liquid adhesion energy per unit area when in the medium G.

Modified Young’s equation edit

The earliest study on the relationship between contact angle and surface tensions for sessile droplets on flat surfaces was reported by Thomas Young in 1805.[2] A century later Gibbs[3] proposed a modification to Young's equation to account for the volumetric dependence of the contact angle. Gibbs postulated the existence of a line tension, which acts at the three-phase boundary and accounts for the excess energy at the confluence of the solid-liquid-gas phase interface, and is given as:

 

where κ is the line tension in Newtons and a is the droplet radius in meters. Although experimental data validates an affine relationship between the cosine of the contact angle and the inverse line radius, it does not account for the correct sign of κ and overestimates its value by several orders of magnitude.

Contact angle prediction while accounting for line tension and Laplace pressure edit

 
Schematic Diagrams for droplets on flat (a) concave (b) and convex (c) surfaces[4]

With improvements in measuring techniques such as atomic force microscopy, confocal microscopy, and scanning electron microscope, researchers were able to produce and image droplets at ever smaller scales. With the reduction in droplet size came new experimental observations of wetting. These observations confirmed that the modified Young's equation does not hold at the micro-nano scales. Jasper[5][4] proposed that including a V dP term in the variation of the free energy may be the key to solving the contact angle problem at such small scales. Given that the variation in free energy is zero at equilibrium:

 

The variation in the pressure at the free liquid-vapor boundary is due to Laplace pressure, which is proportional to the mean curvature. Solving the above equation for both convex and concave surfaces yields:[4]

 

where

 

This equation relates the contact angle, a geometric property of a sessile droplet to the bulk thermodynamics, the energy at the three phase contact boundary, and the mean curvature of the droplet. For the special case of a sessile droplet on a flat surface (α = 0):

 

In the above equation, the first two terms are the modified Young's equation, while the third term is due to the Laplace pressure. This nonlinear equation correctly predicts the sign and magnitude of κ, the flattening of the contact angle at very small scales, and contact angle hysteresis.

Contact angle hysteresis edit

A given substrate-liquid-vapor combination yields a continuous range of contact angle values in practice. The maximum contact angle is referred to as the advancing contact angle and the minimum contact angle is referred to as the receding contact angle. The advancing and receding contact angles are measured from dynamic experiments where droplets or liquid bridges are in movement.[1] In contrast, the equilibrium contact angle described by the Young-Laplace equation is measured from a static state. Static measurements yield values in-between the advancing and receding contact angle depending on deposition parameters (e.g. velocity, angle, and drop size) and drop history (e.g. evaporation from time of deposition). Contact angle hysteresis is defined as θAθR although the term is also used to describe the expression cos θR – cos θA. The static, advancing, or receding contact angle can be used in place of the equilibrium contact angle depending on the application. The overall effect can be seen as closely analogous to static friction, i.e., a minimal amount of work per unit distance is required to move the contact line.[6]

The advancing contact angle can be described as a measure of the liquid-solid cohesion while the receding contact angle is a measure of liquid-solid adhesion. The advancing and receding contact angles can be measured directly using different methods and can also be calculated from other wetting measurements such as force tensiometry (aka Wilhemy-Plate method).

Advancing and receding contact angles can be measured directly from the same measurement if drops are moved linearly on a surface. For example, a drop of liquid will adopt a given contact angle when static, but when the surface is tilted the drop will initially deform so that the contact area between the drop and surface remains constant. The "downhill" side of the drop will adopt a higher contact angle while the "uphill" side of the drop will adopt a lower contact angle. As the tilt angle increases the contact angles will continue to change but the contact area between the drop and surface will remain constant. At a given surface tilt angle, the advancing and receding contact angles will be met and the drop will move on the surface. In practice, the measurement can be influenced by shear forces and momentum if the tilt velocity is high. The measurement method can also be challenging in practice for systems with high (>30 degrees) or low (<10 degrees) contact angle hysteresis.

Advancing and receding contact angle measurements can be carried out by adding and removing liquid from a drop deposited on a surface. If a sufficiently small volume of liquid is added to a drop, the contact line will still be pinned, and the contact angle will increase. Similarly, if a small amount of liquid is removed from a drop, the contact angle will decrease.

The Young's equation assumes a homogeneous surface and does not account for surface texture or outside forces such as gravity. Real surfaces are not atomically smooth or chemically homogeneous so a drop will assume contact angle hysteresis. The equilibrium contact angle (θC) can be calculated from θA and θR as was shown theoretically by Tadmor[7] and confirmed experimentally by Chibowski[8] as,

 

where

 

On a surface that is rough or contaminated, there will also be contact angle hysteresis, but now the local equilibrium contact angle (the Young equation is now only locally valid) may vary from place to place on the surface.[9] According to the Young–Dupré equation, this means that the adhesion energy varies locally – thus, the liquid has to overcome local energy barriers in order to wet the surface. One consequence of these barriers is contact angle hysteresis: the extent of wetting, and therefore the observed contact angle (averaged along the contact line), depends on whether the liquid is advancing or receding on the surface.

Because liquid advances over previously dry surface but recedes from previously wet surface, contact angle hysteresis can also arise if the solid has been altered due to its previous contact with the liquid (e.g., by a chemical reaction, or absorption). Such alterations, if slow, can also produce measurably time-dependent contact angles.

Effect of roughness to contact angles edit

Surface roughness has a strong effect on the contact angle and wettability of a surface. The effect of roughness depends on if the droplet will wet the surface grooves or if air pockets will be left between the droplet and the surface.[10]

If the surface is wetted homogeneously, the droplet is in Wenzel state.[11] In Wenzel state, adding surface roughness will enhance the wettability caused by the chemistry of the surface. The Wenzel correlation can be written as

 
where θm is the measured contact angle, θY is the Young contact angle and r is the roughness ratio. The roughness ratio is defined as the ratio between the actual and projected solid surface area.

If the surface is wetted heterogeneously, the droplet is in Cassie-Baxter state.[12] The most stable contact angle can be connected to the Young contact angle. The contact angles calculated from the Wenzel and Cassie-Baxter equations have been found to be good approximations of the most stable contact angles with real surfaces.[13]

Dynamic contact angles edit

For liquid moving quickly over a surface, the contact angle can be altered from its value at rest. The advancing contact angle will increase with speed, and the receding contact angle will decrease. The discrepancies between static and dynamic contact angles are closely proportional to the capillary number, noted  .[1]

Contact angle curvature edit

On the basis of interfacial energies, the profile of a surface droplet or a liquid bridge between two surfaces can be described by the Young–Laplace equation.[1] This equation is applicable for three-dimensional axisymmetric conditions and is highly non-linear. This is due to the mean curvature term which includes products of first- and second-order derivatives of the drop shape function  :

 

Solving this elliptic partial differential equation that governs the shape of a three-dimensional drop, in conjunction with appropriate boundary conditions, is complicated, and an alternate energy minimization approach to this is generally adopted. The shapes of three-dimensional sessile and pendant drops have been successfully predicted using this energy minimisation method.[14]

Typical contact angles edit

 
Image from a video contact angle device. Water drop on glass, with reflection below.
 
A water drop on a lotus leaf surface showing contact angles of approximately 147°.

Contact angles are extremely sensitive to contamination; values reproducible to better than a few degrees are generally only obtained under laboratory conditions with purified liquids and very clean solid surfaces. If the liquid molecules are strongly attracted to the solid molecules then the liquid drop will completely spread out on the solid surface, corresponding to a contact angle of 0°. This is often the case for water on bare metallic or ceramic surfaces,[15] although the presence of an oxide layer or contaminants on the solid surface can significantly increase the contact angle. Generally, if the water contact angle is smaller than 90°, the solid surface is considered hydrophilic[16] and if the water contact angle is larger than 90°, the solid surface is considered hydrophobic. Many polymers exhibit hydrophobic surfaces. Highly hydrophobic surfaces made of low surface energy (e.g. fluorinated) materials may have water contact angles as high as ≈ 120°.[15] Some materials with highly rough surfaces may have a water contact angle even greater than 150°, due to the presence of air pockets under the liquid drop. These are called superhydrophobic surfaces.

If the contact angle is measured through the gas instead of through the liquid, then it should be replaced by 180° minus their given value. Contact angles are equally applicable to the interface of two liquids, though they are more commonly measured in solid products such as non-stick pans and waterproof fabrics.

Control of contact angles edit

Control of the wetting contact angle can often be achieved through the deposition or incorporation of various organic and inorganic molecules onto the surface. This is often achieved through the use of specialty silane chemicals which can form a SAM (self-assembled monolayers) layer. With the proper selection of the organic molecules with varying molecular structures and amounts of hydrocarbon and/or perfluorinated terminations, the contact angle of the surface can tune. The deposition of these specialty silanes[17] can be achieved in the gas phase through the use of a specialized vacuum ovens or liquid-phase process. Molecules that can bind more perfluorinated terminations to the surface can results in lowering the surface energy (high water contact angle).

Effect of surface fluorine on contact angle Water contact angle
Precursor on polished silicon (deg.)
Henicosyl-1,1,2,2-tetrahydrododecyldimethyltris(dimethylaminosilane) 118.0
Heptadecafluoro-1,1,2,2-tetrahydrodecyltrichlorosilane – (FDTS) 110.0
Nonafluoro-1,1,2,2-tetrahydrohexyltris(dimethylamino)silane 110.0
3,3,3,4,4,5,5,6,6-Nonafluorohexyltrichlorosilane 108.0
Tridecafluoro-1,1,2,2-tetrahydrooctyltrichlorosilane – (FOTS) 108.0
BIS(Tridecafluoro-1,1,2,2-tetrahydrooctyl)dimethylsiloxymethylchlorosilane 107.0
Dodecyltrichlorosilane – (DDTS) 105.0
Dimethyldichlorosilane – (DDMS) 103.0
10-Undecenyltrichlorosilane – (V11) 100.0
Pentafluorophenylpropyltrichlorosilane 90.0

Measuring methods edit

 
A contact angle goniometer is used to measure the contact angle.
 
Dynamic sessile drop method

The static sessile drop method edit

The sessile drop contact angle is measured by a contact angle goniometer using an optical subsystem to capture the profile of a pure liquid on a solid substrate. The angle formed between the liquid–solid interface and the liquid–vapor interface is the contact angle. Older systems used a microscope optical system with a back light. Current-generation systems employ high resolution cameras and software to capture and analyze the contact angle. Angles measured in such a way are often quite close to advancing contact angles. Equilibrium contact angles can be obtained through the application of well defined vibrations.[18][19]

The pendant drop method edit

Measuring contact angles for pendant drops is much more complicated than for sessile drops due to the inherent unstable nature of inverted drops. This complexity is further amplified when one attempts to incline the surface. Experimental apparatus to measure pendant drop contact angles on inclined substrates has been developed recently.[20] This method allows for the deposition of multiple microdrops on the underside of a textured substrate, which can be imaged using a high resolution CCD camera. An automated system allows for tilting the substrate and analysing the images for the calculation of advancing and receding contact angles.

The dynamic sessile drop method edit

The dynamic sessile drop is similar to the static sessile drop but requires the drop to be modified. A common type of dynamic sessile drop study determines the largest contact angle possible without increasing its solid–liquid interfacial area by adding volume dynamically. This maximum angle is the advancing angle. Volume is removed to produce the smallest possible angle, the receding angle. The difference between the advancing and receding angle is the contact angle hysteresis.[19]

Dynamic Wilhelmy method edit

 
Measuring dynamic contact angle of a rod/fiber with a force tensiometer.

The dynamic Wilhelmy method is a method for calculating average advancing and receding contact angles on solids of uniform geometry. Both sides of the solid must have the same properties. Wetting force on the solid is measured as the solid is immersed in or withdrawn from a liquid of known surface tension. Also in that case it is possible to measure the equilibrium contact angle by applying a very controlled vibration. That methodology, called VIECA, can be implemented in a quite simple way on every Wilhelmy balance.[21]

Single-fiber Wilhelmy method edit

Dynamic Wilhelmy method applied to single fibers to measure advancing and receding contact angles.

 
Single-fiber meniscus contact angle measurement.

Single-fiber meniscus method edit

An optical variation of the single-fiber Wilhelmy method. Instead of measuring with a balance, the shape of the meniscus on the fiber is directly imaged using a high resolution camera. Automated meniscus shape fitting can then directly measure the static, advancing or receding contact angle on the fiber.

Washburn's equation capillary rise method edit

In case of a porous materials many issues have been raised both about the physical meaning of the calculated pore diameter and the real possibility to use this equation for the calculation of the contact angle of the solid, even if this method is often offered by much software as consolidated.[22][clarification needed] Change of weight as a function of time is measured.[23]

See also edit

References edit

  1. ^ a b c d 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. S2CID 51916594.
  2. ^ "III. An essay on the cohesion of fluids". Philosophical Transactions of the Royal Society of London. 95: 65–87. January 1805. doi:10.1098/rstl.1805.0005. ISSN 0261-0523. S2CID 116124581.
  3. ^ Gibbs, J. Willard (Josiah Willard) (1961). Scientific papers. Dover Publications. ISBN 978-0486607214. OCLC 964884.
  4. ^ a b c Jasper, Warren J.; Anand, Nadish (May 2019). "A generalized variational approach for predicting contact angles of sessile nano-droplets on both flat and curved surfaces". Journal of Molecular Liquids. 281: 196–203. doi:10.1016/j.molliq.2019.02.039. ISSN 0167-7322. S2CID 104412970.
  5. ^ Jasper, Warren J.; Rasipuram, Srinivasan (December 2017). "Relationship between contact angle and contact line radius for micro to atto [10−6 to 10−18] liter size oil droplets". Journal of Molecular Liquids. 248: 920–926. doi:10.1016/j.molliq.2017.10.134. ISSN 0167-7322.
  6. ^ Hattori, Tsuyoshi; Koshizuka, Seiichi (2019). "Numerical simulation of droplet behavior on an inclined plate using the Moving Particle Semi-implicit method". Mechanical Engineering Journal. 6 (5): 19-00204–19-00204. doi:10.1299/mej.19-00204. ISSN 2187-9745.
  7. ^ Tadmor, Rafael (2004). "Line energy and the relation between advancing, receding, and Young contact angles". Langmuir. 20 (18): 7659–64. doi:10.1021/la049410h. PMID 15323516.
  8. ^ Chibowski, Emil (2008). "Surface free energy of sulfur—Revisited I. Yellow and orange samples solidified against glass surface". Journal of Colloid and Interface Science. 319 (2): 505–13. Bibcode:2008JCIS..319..505C. doi:10.1016/j.jcis.2007.10.059. PMID 18177886.
  9. ^ de Gennes, P.G. (1985). "Wetting: statics and dynamics". Reviews of Modern Physics. 57 (3): 827–863. Bibcode:1985RvMP...57..827D. doi:10.1103/RevModPhys.57.827.
  10. ^ "Influence of surface roughness on contact angle and wettability" (PDF).
  11. ^ Wenzel, Robert N. (1936-08-01). "Resistance of Solid Surfaces to Wetting by Water". Industrial & Engineering Chemistry. 28 (8): 988–994. doi:10.1021/ie50320a024. ISSN 0019-7866.
  12. ^ Cassie, A. B. D.; Baxter, S. (1944-01-01). "Wettability of porous surfaces". Transactions of the Faraday Society. 40: 546. doi:10.1039/tf9444000546. ISSN 0014-7672.
  13. ^ Marmur, Abraham (2009-07-06). "Solid-Surface Characterization by Wetting". Annual Review of Materials Research. 39 (1): 473–489. Bibcode:2009AnRMS..39..473M. doi:10.1146/annurev.matsci.38.060407.132425. ISSN 1531-7331.
  14. ^ Chen Y, He B, Lee J, Patankar NA (2005). (PDF). Journal of Colloid and Interface Science. 281 (2): 458–464. Bibcode:2005JCIS..281..458C. doi:10.1016/j.jcis.2004.07.038. PMID 15571703. Archived from the original (PDF) on 2017-08-10. Retrieved 2017-03-31.
  15. ^ a b Zisman, W.A. (1964). F. Fowkes (ed.). Contact Angle, Wettability, and Adhesion. ACS. pp. 1–51.
  16. ^ Renate Förch; Holger Schönherr; A. Tobias A. Jenkins (2009). Surface design: applications in bioscience and nanotechnology. Wiley-VCH. p. 471. ISBN 978-3-527-40789-7.
  17. ^ Kobrin, B.; Zhang, T.; Chinn, J. "Choice of precursors in Vapor-phase Surface Modification". 209th Electrochemical Society meeting, May 7–12, 2006, Denver, CO.
  18. ^ Volpe, C. D.; Brugnara, M.; Maniglio, D.; Siboni, S.; Wangdu, T. (2006). "About the possibility of experimentally measuring an equilibrium contact angle and its theoretical and practical consequences". Contact Angle, Wettability and Adhesion. 4: 79–100.
  19. ^ a b Huhtamäki, Tommi; Tian, Xuelin; Korhonen, Juuso T.; Ras, Robin H. A. (2018). "Surface-wetting characterization using contact-angle measurements". Nature Protocols. 13 (7): 1521–1538. doi:10.1038/s41596-018-0003-z. ISSN 1754-2189. PMID 29988109. S2CID 51605807.
  20. ^ Bhutani, Gaurav; Muralidhar, K.; Khandekar, Sameer (2013). "Determination of apparent contact angle and shape of a static pendant drop on a physically textured inclined surface". Interfacial Phenomena and Heat Transfer. 1: 29–49. doi:10.1615/InterfacPhenomHeatTransfer.2013007038.
  21. ^ Volpe, C. D.; Maniglio, D.; Siboni, S.; Morra, M. (2001). "An experimental procedure to obtain the equilibrium contact angle from the Wilhelmy method" (PDF). Oil & Gas Science and Technology. 56: 9–22. doi:10.2516/ogst:2001002.
  22. ^ Marco, Brugnara; Claudio, Della Volpe; Stefano, Siboni (2006). "Wettability of porous materials. II. Can we obtain the contact angle from the Washburn equation?". In Mittal, K. L. (ed.). Contact Angle, Wettability and Adhesion. Mass. VSP.
  23. ^ Washburn, Edward W. (1921). "The Dynamics of Capillary Flow". Physical Review. 17 (3): 273. Bibcode:1921PhRv...17..273W. doi:10.1103/PhysRev.17.273.

Further reading edit

  • Pierre-Gilles de Gennes, Françoise Brochard-Wyart, David Quéré, Capillarity and Wetting Phenomena: Drops, Bubbles, Pearls, Waves, Springer (2004)
  • Jacob Israelachvili, Intermolecular and Surface Forces, Academic Press (1985–2004)
  • D.W. Van Krevelen, Properties of Polymers, 2nd revised edition, Elsevier Scientific Publishing Company, Amsterdam-Oxford-New York (1976)
  • Yuan, Yuehua; Lee, T. Randall (2013). "Contact Angle and Wetting Properties". Surface Science Techniques. Springer Series in Surface Sciences. Vol. 51. doi:10.1007/978-3-642-34243-1. ISBN 978-3-642-34242-4. ISSN 0931-5195. S2CID 137147527.
  • Clegg, Carl Contact Angle Made Easy, ramé-hart (2013), ISBN 978-1-300-66298-3

contact, angle, contact, angle, symbol, angle, between, liquid, surface, solid, surface, where, they, meet, more, specifically, angle, between, surface, tangent, liquid, vapor, interface, tangent, solid, liquid, interface, their, intersection, quantifies, wett. The contact angle symbol 8C 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 Schematic of a liquid drop showing the quantities in the Young equation A given system of solid liquid and vapor at a given temperature and pressure has a unique equilibrium contact angle However in practice a dynamic phenomenon of contact angle hysteresis is often observed ranging from the advancing maximal contact angle to the receding minimal contact angle 1 The equilibrium contact is within those values and can be calculated from them The equilibrium contact angle reflects the relative strength of the liquid solid and vapour molecular interaction The contact angle depends upon the medium above the free surface of the liquid and the nature of the liquid and solid in contact It is independent of the inclination of solid to the liquid surface It changes with surface tension and hence with the temperature and purity of the liquid Contents 1 Thermodynamics 1 1 Modified Young s equation 1 2 Contact angle prediction while accounting for line tension and Laplace pressure 1 3 Contact angle hysteresis 1 4 Effect of roughness to contact angles 1 5 Dynamic contact angles 2 Contact angle curvature 3 Typical contact angles 4 Control of contact angles 5 Measuring methods 5 1 The static sessile drop method 5 2 The pendant drop method 5 3 The dynamic sessile drop method 5 4 Dynamic Wilhelmy method 5 5 Single fiber Wilhelmy method 5 6 Single fiber meniscus method 5 7 Washburn s equation capillary rise method 6 See also 7 References 8 Further readingThermodynamics edit nbsp Cloth treated to be hydrophobic shows a high contact angle The theoretical description of contact angle arises from the consideration of a thermodynamic equilibrium between the three phases the liquid phase L the solid phase S and the gas or vapor phase G which could be a mixture of ambient atmosphere and an equilibrium concentration of the liquid vapor The gaseous phase could be replaced by another immiscible liquid phase If the solid vapor interfacial energy is denoted by gSG the solid liquid interfacial energy by gSL and the liquid vapor interfacial energy i e the surface tension by gLG then the equilibrium contact angle 8C is determined from these quantities by the Young equation g S G g S L g L G cos 8 C 0 displaystyle gamma rm SG gamma rm SL gamma rm LG cos theta rm C 0 nbsp The contact angle can also be related to the work of adhesion via the Young Dupre equation g L G 1 cos 8 C D W S L G displaystyle gamma rm LG 1 cos theta rm C Delta W rm SLG nbsp where D W S L G displaystyle Delta W rm SLG nbsp is the solid liquid adhesion energy per unit area when in the medium G Modified Young s equation edit The earliest study on the relationship between contact angle and surface tensions for sessile droplets on flat surfaces was reported by Thomas Young in 1805 2 A century later Gibbs 3 proposed a modification to Young s equation to account for the volumetric dependence of the contact angle Gibbs postulated the existence of a line tension which acts at the three phase boundary and accounts for the excess energy at the confluence of the solid liquid gas phase interface and is given as cos 8 g S G g S L g L G k g L G 1 a displaystyle cos theta frac gamma rm SG gamma rm SL gamma rm LG frac kappa gamma rm LG frac 1 a nbsp where k is the line tension in Newtons and a is the droplet radius in meters Although experimental data validates an affine relationship between the cosine of the contact angle and the inverse line radius it does not account for the correct sign of k and overestimates its value by several orders of magnitude Contact angle prediction while accounting for line tension and Laplace pressure edit nbsp Schematic Diagrams for droplets on flat a concave b and convex c surfaces 4 With improvements in measuring techniques such as atomic force microscopy confocal microscopy and scanning electron microscope researchers were able to produce and image droplets at ever smaller scales With the reduction in droplet size came new experimental observations of wetting These observations confirmed that the modified Young s equation does not hold at the micro nano scales Jasper 5 4 proposed that including a V dP term in the variation of the free energy may be the key to solving the contact angle problem at such small scales Given that the variation in free energy is zero at equilibrium 0 d A L G d A S L g S L g S G g L G k g L G d L d A S L V g L G d P d A S L displaystyle 0 frac dA rm LG dA rm SL frac gamma rm SL gamma rm SG gamma rm LG frac kappa gamma rm LG frac dL dA rm SL frac V gamma rm LG frac dP dA rm SL nbsp The variation in the pressure at the free liquid vapor boundary is due to Laplace pressure which is proportional to the mean curvature Solving the above equation for both convex and concave surfaces yields 4 cos 8 a A B cos a a C sin 8 a 1 cos 8 2 sin a 2 cos a 1 cos a 2 sin 8 2 cos 8 1 cos 8 2 displaystyle cos theta mp alpha A B frac cos alpha a pm C sin theta mp alpha 1 cos theta 2 biggl frac sin alpha 2 cos alpha 1 cos alpha 2 mp frac sin theta 2 cos theta 1 cos theta 2 biggr nbsp whereA g S G g S L g L G B k g L G C g 3 g L G displaystyle A frac gamma rm SG gamma rm SL gamma rm LG quad B frac kappa gamma rm LG quad C frac gamma 3 gamma rm LG nbsp This equation relates the contact angle a geometric property of a sessile droplet to the bulk thermodynamics the energy at the three phase contact boundary and the mean curvature of the droplet For the special case of a sessile droplet on a flat surface a 0 cos 8 g S G g S L g L G k g L G 1 a g 3 g L G 2 cos 8 2 cos 2 8 cos 3 8 displaystyle cos theta frac gamma rm SG gamma rm SL gamma rm LG frac kappa gamma rm LG frac 1 a frac gamma 3 gamma rm LG 2 cos theta 2 cos 2 theta cos 3 theta nbsp In the above equation the first two terms are the modified Young s equation while the third term is due to the Laplace pressure This nonlinear equation correctly predicts the sign and magnitude of k the flattening of the contact angle at very small scales and contact angle hysteresis Contact angle hysteresis edit A given substrate liquid vapor combination yields a continuous range of contact angle values in practice The maximum contact angle is referred to as the advancing contact angle and the minimum contact angle is referred to as the receding contact angle The advancing and receding contact angles are measured from dynamic experiments where droplets or liquid bridges are in movement 1 In contrast the equilibrium contact angle described by the Young Laplace equation is measured from a static state Static measurements yield values in between the advancing and receding contact angle depending on deposition parameters e g velocity angle and drop size and drop history e g evaporation from time of deposition Contact angle hysteresis is defined as 8A 8R although the term is also used to describe the expression cos 8R cos 8A The static advancing or receding contact angle can be used in place of the equilibrium contact angle depending on the application The overall effect can be seen as closely analogous to static friction i e a minimal amount of work per unit distance is required to move the contact line 6 The advancing contact angle can be described as a measure of the liquid solid cohesion while the receding contact angle is a measure of liquid solid adhesion The advancing and receding contact angles can be measured directly using different methods and can also be calculated from other wetting measurements such as force tensiometry aka Wilhemy Plate method Advancing and receding contact angles can be measured directly from the same measurement if drops are moved linearly on a surface For example a drop of liquid will adopt a given contact angle when static but when the surface is tilted the drop will initially deform so that the contact area between the drop and surface remains constant The downhill side of the drop will adopt a higher contact angle while the uphill side of the drop will adopt a lower contact angle As the tilt angle increases the contact angles will continue to change but the contact area between the drop and surface will remain constant At a given surface tilt angle the advancing and receding contact angles will be met and the drop will move on the surface In practice the measurement can be influenced by shear forces and momentum if the tilt velocity is high The measurement method can also be challenging in practice for systems with high gt 30 degrees or low lt 10 degrees contact angle hysteresis Advancing and receding contact angle measurements can be carried out by adding and removing liquid from a drop deposited on a surface If a sufficiently small volume of liquid is added to a drop the contact line will still be pinned and the contact angle will increase Similarly if a small amount of liquid is removed from a drop the contact angle will decrease The Young s equation assumes a homogeneous surface and does not account for surface texture or outside forces such as gravity Real surfaces are not atomically smooth or chemically homogeneous so a drop will assume contact angle hysteresis The equilibrium contact angle 8C can be calculated from 8A and 8R as was shown theoretically by Tadmor 7 and confirmed experimentally by Chibowski 8 as 8 c arccos r A cos 8 A r R cos 8 R r A r R displaystyle theta rm c arccos left frac r rm A cos theta rm A r rm R cos theta rm R r rm A r rm R right nbsp wherer A sin 3 8 A 2 3 cos 8 A cos 3 8 A 3 r R sin 3 8 R 2 3 cos 8 R cos 3 8 R 3 displaystyle begin aligned r rm A amp sqrt 3 frac sin 3 theta rm A 2 3 cos theta rm A cos 3 theta rm A 4pt r rm R amp sqrt 3 frac sin 3 theta rm R 2 3 cos theta rm R cos 3 theta rm R end aligned nbsp On a surface that is rough or contaminated there will also be contact angle hysteresis but now the local equilibrium contact angle the Young equation is now only locally valid may vary from place to place on the surface 9 According to the Young Dupre equation this means that the adhesion energy varies locally thus the liquid has to overcome local energy barriers in order to wet the surface One consequence of these barriers is contact angle hysteresis the extent of wetting and therefore the observed contact angle averaged along the contact line depends on whether the liquid is advancing or receding on the surface Because liquid advances over previously dry surface but recedes from previously wet surface contact angle hysteresis can also arise if the solid has been altered due to its previous contact with the liquid e g by a chemical reaction or absorption Such alterations if slow can also produce measurably time dependent contact angles Effect of roughness to contact angles edit Surface roughness has a strong effect on the contact angle and wettability of a surface The effect of roughness depends on if the droplet will wet the surface grooves or if air pockets will be left between the droplet and the surface 10 If the surface is wetted homogeneously the droplet is in Wenzel state 11 In Wenzel state adding surface roughness will enhance the wettability caused by the chemistry of the surface The Wenzel correlation can be written ascos 8 m r cos 8 Y displaystyle cos theta m r cos theta Y nbsp where 8m is the measured contact angle 8Y is the Young contact angle and r is the roughness ratio The roughness ratio is defined as the ratio between the actual and projected solid surface area If the surface is wetted heterogeneously the droplet is in Cassie Baxter state 12 The most stable contact angle can be connected to the Young contact angle The contact angles calculated from the Wenzel and Cassie Baxter equations have been found to be good approximations of the most stable contact angles with real surfaces 13 Dynamic contact angles edit For liquid moving quickly over a surface the contact angle can be altered from its value at rest The advancing contact angle will increase with speed and the receding contact angle will decrease The discrepancies between static and dynamic contact angles are closely proportional to the capillary number noted C a displaystyle Ca nbsp 1 Contact angle curvature editOn the basis of interfacial energies the profile of a surface droplet or a liquid bridge between two surfaces can be described by the Young Laplace equation 1 This equation is applicable for three dimensional axisymmetric conditions and is highly non linear This is due to the mean curvature term which includes products of first and second order derivatives of the drop shape function f x y displaystyle f x y nbsp k m 1 2 1 f x 2 f y y 2 f x f y f x y 1 f y 2 f x x 1 f x 2 f y 2 3 2 displaystyle kappa m frac 1 2 frac 1 f x 2 f yy 2f x f y f xy 1 f y 2 f xx 1 f x 2 f y 2 3 2 nbsp Solving this elliptic partial differential equation that governs the shape of a three dimensional drop in conjunction with appropriate boundary conditions is complicated and an alternate energy minimization approach to this is generally adopted The shapes of three dimensional sessile and pendant drops have been successfully predicted using this energy minimisation method 14 Typical contact angles edit nbsp Image from a video contact angle device Water drop on glass with reflection below nbsp A water drop on a lotus leaf surface showing contact angles of approximately 147 Contact angles are extremely sensitive to contamination values reproducible to better than a few degrees are generally only obtained under laboratory conditions with purified liquids and very clean solid surfaces If the liquid molecules are strongly attracted to the solid molecules then the liquid drop will completely spread out on the solid surface corresponding to a contact angle of 0 This is often the case for water on bare metallic or ceramic surfaces 15 although the presence of an oxide layer or contaminants on the solid surface can significantly increase the contact angle Generally if the water contact angle is smaller than 90 the solid surface is considered hydrophilic 16 and if the water contact angle is larger than 90 the solid surface is considered hydrophobic Many polymers exhibit hydrophobic surfaces Highly hydrophobic surfaces made of low surface energy e g fluorinated materials may have water contact angles as high as 120 15 Some materials with highly rough surfaces may have a water contact angle even greater than 150 due to the presence of air pockets under the liquid drop These are called superhydrophobic surfaces If the contact angle is measured through the gas instead of through the liquid then it should be replaced by 180 minus their given value Contact angles are equally applicable to the interface of two liquids though they are more commonly measured in solid products such as non stick pans and waterproof fabrics Control of contact angles editControl of the wetting contact angle can often be achieved through the deposition or incorporation of various organic and inorganic molecules onto the surface This is often achieved through the use of specialty silane chemicals which can form a SAM self assembled monolayers layer With the proper selection of the organic molecules with varying molecular structures and amounts of hydrocarbon and or perfluorinated terminations the contact angle of the surface can tune The deposition of these specialty silanes 17 can be achieved in the gas phase through the use of a specialized vacuum ovens or liquid phase process Molecules that can bind more perfluorinated terminations to the surface can results in lowering the surface energy high water contact angle Effect of surface fluorine on contact angle Water contact anglePrecursor on polished silicon deg Henicosyl 1 1 2 2 tetrahydrododecyldimethyltris dimethylaminosilane 118 0Heptadecafluoro 1 1 2 2 tetrahydrodecyltrichlorosilane FDTS 110 0Nonafluoro 1 1 2 2 tetrahydrohexyltris dimethylamino silane 110 03 3 3 4 4 5 5 6 6 Nonafluorohexyltrichlorosilane 108 0Tridecafluoro 1 1 2 2 tetrahydrooctyltrichlorosilane FOTS 108 0BIS Tridecafluoro 1 1 2 2 tetrahydrooctyl dimethylsiloxymethylchlorosilane 107 0Dodecyltrichlorosilane DDTS 105 0Dimethyldichlorosilane DDMS 103 010 Undecenyltrichlorosilane V11 100 0Pentafluorophenylpropyltrichlorosilane 90 0Measuring methods edit nbsp A contact angle goniometer is used to measure the contact angle nbsp Dynamic sessile drop methodThe static sessile drop method edit Main article Sessile drop technique The sessile drop contact angle is measured by a contact angle goniometer using an optical subsystem to capture the profile of a pure liquid on a solid substrate The angle formed between the liquid solid interface and the liquid vapor interface is the contact angle Older systems used a microscope optical system with a back light Current generation systems employ high resolution cameras and software to capture and analyze the contact angle Angles measured in such a way are often quite close to advancing contact angles Equilibrium contact angles can be obtained through the application of well defined vibrations 18 19 The pendant drop method edit Measuring contact angles for pendant drops is much more complicated than for sessile drops due to the inherent unstable nature of inverted drops This complexity is further amplified when one attempts to incline the surface Experimental apparatus to measure pendant drop contact angles on inclined substrates has been developed recently 20 This method allows for the deposition of multiple microdrops on the underside of a textured substrate which can be imaged using a high resolution CCD camera An automated system allows for tilting the substrate and analysing the images for the calculation of advancing and receding contact angles The dynamic sessile drop method edit Main article Sessile drop technique The dynamic sessile drop is similar to the static sessile drop but requires the drop to be modified A common type of dynamic sessile drop study determines the largest contact angle possible without increasing its solid liquid interfacial area by adding volume dynamically This maximum angle is the advancing angle Volume is removed to produce the smallest possible angle the receding angle The difference between the advancing and receding angle is the contact angle hysteresis 19 Dynamic Wilhelmy method edit nbsp Measuring dynamic contact angle of a rod fiber with a force tensiometer The dynamic Wilhelmy method is a method for calculating average advancing and receding contact angles on solids of uniform geometry Both sides of the solid must have the same properties Wetting force on the solid is measured as the solid is immersed in or withdrawn from a liquid of known surface tension Also in that case it is possible to measure the equilibrium contact angle by applying a very controlled vibration That methodology called VIECA can be implemented in a quite simple way on every Wilhelmy balance 21 Single fiber Wilhelmy method edit Dynamic Wilhelmy method applied to single fibers to measure advancing and receding contact angles nbsp Single fiber meniscus contact angle measurement Single fiber meniscus method edit An optical variation of the single fiber Wilhelmy method Instead of measuring with a balance the shape of the meniscus on the fiber is directly imaged using a high resolution camera Automated meniscus shape fitting can then directly measure the static advancing or receding contact angle on the fiber Washburn s equation capillary rise method edit Main article Washburn s equation In case of a porous materials many issues have been raised both about the physical meaning of the calculated pore diameter and the real possibility to use this equation for the calculation of the contact angle of the solid even if this method is often offered by much software as consolidated 22 clarification needed Change of weight as a function of time is measured 23 See also editGoniometer Meniscus liquid Porosimetry Sessile drop technique Surface tension WettingReferences edit a b c d 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 S2CID 51916594 III An essay on the cohesion of fluids Philosophical Transactions of the Royal Society of London 95 65 87 January 1805 doi 10 1098 rstl 1805 0005 ISSN 0261 0523 S2CID 116124581 Gibbs J Willard Josiah Willard 1961 Scientific papers Dover Publications ISBN 978 0486607214 OCLC 964884 a b c Jasper Warren J Anand Nadish May 2019 A generalized variational approach for predicting contact angles of sessile nano droplets on both flat and curved surfaces Journal of Molecular Liquids 281 196 203 doi 10 1016 j molliq 2019 02 039 ISSN 0167 7322 S2CID 104412970 Jasper Warren J Rasipuram Srinivasan December 2017 Relationship between contact angle and contact line radius for micro to atto 10 6 to 10 18 liter size oil droplets Journal of Molecular Liquids 248 920 926 doi 10 1016 j molliq 2017 10 134 ISSN 0167 7322 Hattori Tsuyoshi Koshizuka Seiichi 2019 Numerical simulation of droplet behavior on an inclined plate using the Moving Particle Semi implicit method Mechanical Engineering Journal 6 5 19 00204 19 00204 doi 10 1299 mej 19 00204 ISSN 2187 9745 Tadmor Rafael 2004 Line energy and the relation between advancing receding and Young contact angles Langmuir 20 18 7659 64 doi 10 1021 la049410h PMID 15323516 Chibowski Emil 2008 Surface free energy of sulfur Revisited I Yellow and orange samples solidified against glass surface Journal of Colloid and Interface Science 319 2 505 13 Bibcode 2008JCIS 319 505C doi 10 1016 j jcis 2007 10 059 PMID 18177886 de Gennes P G 1985 Wetting statics and dynamics Reviews of Modern Physics 57 3 827 863 Bibcode 1985RvMP 57 827D doi 10 1103 RevModPhys 57 827 Influence of surface roughness on contact angle and wettability PDF Wenzel Robert N 1936 08 01 Resistance of Solid Surfaces to Wetting by Water Industrial amp Engineering Chemistry 28 8 988 994 doi 10 1021 ie50320a024 ISSN 0019 7866 Cassie A B D Baxter S 1944 01 01 Wettability of porous surfaces Transactions of the Faraday Society 40 546 doi 10 1039 tf9444000546 ISSN 0014 7672 Marmur Abraham 2009 07 06 Solid Surface Characterization by Wetting Annual Review of Materials Research 39 1 473 489 Bibcode 2009AnRMS 39 473M doi 10 1146 annurev matsci 38 060407 132425 ISSN 1531 7331 Chen Y He B Lee J Patankar NA 2005 Anisotropy in the wetting of rough surfaces PDF Journal of Colloid and Interface Science 281 2 458 464 Bibcode 2005JCIS 281 458C doi 10 1016 j jcis 2004 07 038 PMID 15571703 Archived from the original PDF on 2017 08 10 Retrieved 2017 03 31 a b Zisman W A 1964 F Fowkes ed Contact Angle Wettability and Adhesion ACS pp 1 51 Renate Forch Holger Schonherr A Tobias A Jenkins 2009 Surface design applications in bioscience and nanotechnology Wiley VCH p 471 ISBN 978 3 527 40789 7 Kobrin B Zhang T Chinn J Choice of precursors in Vapor phase Surface Modification 209th Electrochemical Society meeting May 7 12 2006 Denver CO Volpe C D Brugnara M Maniglio D Siboni S Wangdu T 2006 About the possibility of experimentally measuring an equilibrium contact angle and its theoretical and practical consequences Contact Angle Wettability and Adhesion 4 79 100 a b Huhtamaki Tommi Tian Xuelin Korhonen Juuso T Ras Robin H A 2018 Surface wetting characterization using contact angle measurements Nature Protocols 13 7 1521 1538 doi 10 1038 s41596 018 0003 z ISSN 1754 2189 PMID 29988109 S2CID 51605807 Bhutani Gaurav Muralidhar K Khandekar Sameer 2013 Determination of apparent contact angle and shape of a static pendant drop on a physically textured inclined surface Interfacial Phenomena and Heat Transfer 1 29 49 doi 10 1615 InterfacPhenomHeatTransfer 2013007038 Volpe C D Maniglio D Siboni S Morra M 2001 An experimental procedure to obtain the equilibrium contact angle from the Wilhelmy method PDF Oil amp Gas Science and Technology 56 9 22 doi 10 2516 ogst 2001002 Marco Brugnara Claudio Della Volpe Stefano Siboni 2006 Wettability of porous materials II Can we obtain the contact angle from the Washburn equation In Mittal K L ed Contact Angle Wettability and Adhesion Mass VSP Washburn Edward W 1921 The Dynamics of Capillary Flow Physical Review 17 3 273 Bibcode 1921PhRv 17 273W doi 10 1103 PhysRev 17 273 Further reading editPierre Gilles de Gennes Francoise Brochard Wyart David Quere Capillarity and Wetting Phenomena Drops Bubbles Pearls Waves Springer 2004 Jacob Israelachvili Intermolecular and Surface Forces Academic Press 1985 2004 D W Van Krevelen Properties of Polymers 2nd revised edition Elsevier Scientific Publishing Company Amsterdam Oxford New York 1976 Yuan Yuehua Lee T Randall 2013 Contact Angle and Wetting Properties Surface Science Techniques Springer Series in Surface Sciences Vol 51 doi 10 1007 978 3 642 34243 1 ISBN 978 3 642 34242 4 ISSN 0931 5195 S2CID 137147527 Clegg Carl Contact Angle Made Easy rame hart 2013 ISBN 978 1 300 66298 3 Retrieved from https en wikipedia org w index php title Contact angle amp oldid 1195280627, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.