Premelting

Last updated

Premelting (also surface melting) refers to a quasi-liquid film that can occur on the surface of a solid even below melting point (). The thickness of the film is temperature () dependent. This effect is common for all crystalline materials. Premelting shows its effects in frost heave, and, taking grain boundary interfaces into account, maybe even in the movement of glaciers.

Contents

Considering a solid-vapour interface, complete and incomplete premelting can be distinguished. During a temperature rise from below to above , in the case of complete premelting, the solid melts homogeneously from the outside to the inside; in the case of incomplete premelting, the liquid film stays very thin during the beginning of the melting process, but droplets start to form on the interface. In either case, the solid always melts from the outside inwards, never from the inside.

History

The first to mention premelting might have been Michael Faraday in 1842 for ice surfaces. [1] He compared the effect which holds a snowball together to that which makes buildings from moistured sand stable. Another interesting thing he mentioned is that two blocks of ice can freeze together. Later Tammann (1910) and Stranski (1942) suggested that all crystals might, due to the reduction of surface energy, start melting at their surfaces. [2] [3] Frenkel strengthened this by noting that, in contrast to liquids, no overheating can be found for solids. [4] After extensive studies on many materials, it can be concluded that it is a common attribute of the solid state that the melting process begins at the surface. [5]

Theoretical explanations

There are several ways to approach the topic of premelting, the most figurative way might be thermodynamically. A more detailed or abstract view on what physics is important for premelting is given by the Lifshitz and the Landau theories. One always starts with looking at a crystalline solid phase (fig. 1: (1) solid) and another phase. This second phase (fig. 1: (2)) can either be vapour, liquid or solid. Further it can consist of the same chemical material or another. In the case of the second phase being a solid of the same chemical material one speaks of grain boundaries. This case is very important when looking at polycrystalline materials.

Thermodynamical picture for solid gas interface

Boundary between phase (1) and phase (2) without and with intermediate phase (3) Premelting1.JPG
Boundary between phase (1) and phase (2) without and with intermediate phase (3)

In the following thermodynamical equilibrium is assumed, as well as for simplicity (2) should be a vaporous phase.

The first (1) and the second (2) phase are always divided by some form of interface, what results in an interfacial energy . One can now ask whether this energy can be lowered by inserting a third phase (l) in between (1) and (2). Written in interfacial energies this would mean:

If this is the case then it is more efficient for the system to form a separating phase (3). The only possibility for the system to form such a layer is to take material of the solid and "melt" it to a quasi-liquid. In further notation there will be no distinction between quasi-liquid and liquid but one should always keep in mind that there is a difference. This difference to a real liquid becomes clear when looking at a very thin layer (l). As, due to the long range forces of the molecules of the solid material the liquid very near the solid still "feels" the order of crystalline solid and hence itself is in a state providing a not liquid like amount of order. As considering a very thin layer at the moment it is clear that the whole separating layer (l) is too well ordered for a liquid. Further comments on ordering can be found in the paragraph on Landau theory.

Now, looking closer at the thermodynamics of the newly introduced phase (l), its Gibbs energy can be written as:

,

where is the temperature, the pressure, the thickness of (l) corresponding to the number or particles in this case. and are the atomic density and the chemical potential in (l) and . Note that one has to consider that the interfacial energies can just be added to the Gibbs energy in this case. As noted before corresponds so the derivation to results in:

Where . Hence and differ and can be defined. Assuming that a Taylor expansion around the melting point is possible and using the Clausius–Clapeyron equation one can get the following results:

Where is in the order of molecular dimensions the specific melting heat and

These formulas also show that the more the temperature increases, the more increases the thickness of the premelt as this is energetically advantageous. This is the explanation why no overheating exists for this type of phase transition. [5]

Lifshitz theory: Complete and incomplete premelting

With the help of the Lifshitz Theory on Casimir, respectively van der Waals, interactions of macroscopic bodies premelting can be viewed from an electrodynamical perspective. A good example for determining the difference between complete and incomplete premelting is ice. From vacuum ultraviolet (VUV) frequencies upwards the polarizability of ice is greater than that of water, at lower frequencies this is reversed. Assuming there is already a film of thickness on the solid it is easy for any components of electromagnetic waves to travel through the film in the direction perpendicular to the solid surface as long is small. Hence as long as the film is thin compared to the frequency interaction from the solid to the whole film is possible. But when gets large compared to typical VUV frequencies the electronic structure of the film will be too slow to pass the high frequencies to the other end of the liquid phase. Thus this end of the liquid phase feels only a retarded van der Waals interaction with the solid phase. Hence the attraction between the liquid molecules themselves will predominate and they will start forming droplets instead of thickening the film further. So the speed of light limits complete premelting. This makes it a question of solid and surface free energies whether complete premelting occurs. Complete surface melting will occur when is monotonically decreasing. If instead shows a global minimum at finite than the premelting will be incomplete. This implies: When the long range interactions in the system are attractive than there will be incomplete premelting — assuming the film thickness is larger than any repulsive interactions. Is the film thickness small compared to the range of the repulsive interactions present and the repulsive interactions are stronger than the attractive ones than complete premelting can occur. For van der Waals interactions Lifshitz theory can now calculate which type of premelting should occur for a special system. In fact small differences in systems can affect the type of premelting. For example, ice in an atmosphere of water vapour shows incomplete premelting, whereas the premelting of ice in air is complete.

For solid–solid interfaces it cannot be predicted in general whether the premelting is complete or incomplete when only considering van der Waals interactions. Here other types of interactions become very important. This also accounts for grain boundaries. [5]

Landau theory

A qualitative picture of the order parameter of a premelting solid for temperatures below the melting point. One can see that there is still a high amount of order in the liquid, which decreases with rising Temperature Premelting2.JPG
A qualitative picture of the order parameter of a premelting solid for temperatures below the melting point. One can see that there is still a high amount of order in the liquid, which decreases with rising Temperature

Most insight in the problem probably emerges when approaching the effect form Landau Theory. Which is a little bit problematic as the melting of a bulk in general has to be considered as a first order phase transition, meaning the order parameter jumps at . The derivation of Lipowski (basic geometry shown in fig.2) leads to the following results when :

Where is the order parameter at the border between (2) and (l), the so-called extrapolation length and a constant that enters the model and has to be determined using experiment and other models. Hence one can see that the order parameter in the liquid film can undergo a continuous phase transition for large enough extrapolation length. A further result is that what corresponds to the result of the thermodynamical model in the case of short range interactions. Landau Theory does not consider fluctuations like capillary waves, this could change the results qualitatively. [6]

Experimental proof for premelting

Shadowing and blocking diffraction experiment to show the occurrence of premelting. The incident beam follows a crystal direction, so does the angle under which the detector is. The disorder of the quasi liquid premelt changes the scattering spectrum. Premelting3.JPG
Shadowing and blocking diffraction experiment to show the occurrence of premelting. The incident beam follows a crystal direction, so does the angle under which the detector is. The disorder of the quasi liquid premelt changes the scattering spectrum.

There are several techniques to prove the existence of a liquid layer on a well-ordered surface. Basically it is all about showing that there is a phase on top of the solid which has hardly any order (quasi-liquid, see fig. order parameter). One possibility was done by Frenken and van der Veen using proton scattering on a lead (Pb) single crystal (110) surface. First the surface was atomically cleaned in [UHV], because one obviously has to have a very well ordered surface for such experiments. Then they did proton shadowing and blocking measurements. An ideal shadowing and blocking measurements results in an energy spectrum of the scattered protons that shows only a peak for the first surface layer and nothing else. Due to the non ideality of the experiment the spectrum also shows effects of the underlying layers. That means the spectrum is not one well defined peak but has a tail to lower energies due to protons scattered on deeper layers which results in losing energies because of stopping. This is different for a liquid film on the surface: This film does hardly (to the meaning of hardly see Landau theory) have any order. So the effects of shadowing and blocking vanish what means all the liquid film contributes the same amount of scattered electrons to the signal. Therefore, the peak does not only have a tail, but also becomes broadened. During their measurements Frenken and van der Veen raised the temperature to the melting point and hence could show that with increasing temperature a disordered film formed on the surface in equilibrium with a still well ordered Pb crystal. [7]

Curvature, disorder and impurities

Up to here, an ideal surface was considered, but going beyond the idealized case there are several effects which influence premelting:

Ice skating

The friction coefficient for ice, without a liquid film on the surface, is measured to be . [8] A comparable friction coefficient is that of rubber or bitumen (roughly 0.8), which would be very difficult to ice skate on. The friction coefficients needs to be around or below 0.005 for ice skating to be possible. [9] The reason ice skating is possible is because there is a thin film of water present between the blade of the ice skate and the ice. The origin of this water film has been a long-standing debate. There are three proposed mechanisms that could account for a film of liquid water on the ice surface: [10]

While contributions from all three of these factors are usually in effect when ice skating, the scientific community has long debated over which is the dominating mechanism. For several decades it was common to explain the low friction of the skates on ice by pressure melting, but there are several recent arguments that contradict this hypothesis. [10] The strongest argument against pressure melting is that ice skating is still possible below temperatures under -20 °C (253K). At this temperature, a great deal of pressure (>100MPa) is required to induce melting. Just below -23 °C (250K), increasing the pressure can only form a different solid structure of ice (Ice III) since the isotherm no longer passes through the liquid phase on the phase diagram. While impurities in the ice will suppress the melting temperature, many materials scientists agree that pressure melting is not the dominant mechanism. [11] The thickness of the water film due to premelting is also limited at low temperatures. While the water film can reach thicknesses on the order of μm, at temperatures around -10 °C the thickness is on the order of nm. Although, De Koning et al. found in their measurements that the adding of impurities to the ice can lower the friction coefficient up to 15%. The friction coefficient increases with skating speed, which could yield different results depending on the skating technique and speeds. [9] While the pressure melting hypothesis may have been put to rest, the debate between premelting and friction as the dominant mechanism still rages on.

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">Sintering</span> Process of forming and bonding material by heat or pressure

Sintering or frittage is the process of compacting and forming a solid mass of material by pressure or heat without melting it to the point of liquefaction. Sintering happens as part of a manufacturing process used with metals, ceramics, plastics, and other materials. The nanoparticles in the sintered material diffuse across the boundaries of the particles, fusing the particles together and creating a solid piece.

In thermodynamics, the chemical potential of a species is the energy that can be absorbed or released due to a change of the particle number of the given species, e.g. in a chemical reaction or phase transition. The chemical potential of a species in a mixture is defined as the rate of change of free energy of a thermodynamic system with respect to the change in the number of atoms or molecules of the species that are added to the system. Thus, it is the partial derivative of the free energy with respect to the amount of the species, all other species' concentrations in the mixture remaining constant. When both temperature and pressure are held constant, and the number of particles is expressed in moles, the chemical potential is the partial molar Gibbs free energy. At chemical equilibrium or in phase equilibrium, the total sum of the product of chemical potentials and stoichiometric coefficients is zero, as the free energy is at a minimum. In a system in diffusion equilibrium, the chemical potential of any chemical species is uniformly the same everywhere throughout the system.

<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.

<span class="mw-page-title-main">Granular material</span> Conglomeration of discrete solid, macroscopic particles

A granular material is a conglomeration of discrete solid, macroscopic particles characterized by a loss of energy whenever the particles interact. The constituents that compose granular material are large enough such that they are not subject to thermal motion fluctuations. Thus, the lower size limit for grains in granular material is about 1 μm. On the upper size limit, the physics of granular materials may be applied to ice floes where the individual grains are icebergs and to asteroid belts of the Solar System with individual grains being asteroids.

<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">Dewetting</span> Retraction of a fluid from a surface it was forced to cover

In fluid mechanics, dewetting is one of the processes that can occur at a solid–liquid, solid–solid or liquid–liquid interface. Generally, dewetting describes the process of retraction of a fluid from a non-wettable surface it was forced to cover. The opposite process—spreading of a liquid on a substrate—is called wetting. The factor determining the spontaneous spreading and dewetting for a drop of liquid placed on a solid substrate with ambient gas, is the so-called spreading coefficient S:

<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.

<span class="mw-page-title-main">Contact angle</span> The 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">Langmuir–Blodgett film</span> Thin film obtained by depositing multiple monolayers onto a surface

A Langmuir–Blodgett (LB) film is a nanostructured system formed when Langmuir films—or Langmuir monolayers (LM)—are transferred from the liquid-gas interface to solid supports during the vertical passage of the support through the monolayers. LB films can contain one or more monolayers of an organic material, deposited from the surface of a liquid onto a solid by immersing the solid substrate into the liquid. A monolayer is adsorbed homogeneously with each immersion or emersion step, thus films with very accurate thickness can be formed. This thickness is accurate because the thickness of each monolayer is known and can therefore be added to find the total thickness of a Langmuir–Blodgett film.

The Gibbs adsorption isotherm for multicomponent systems is an equation used to relate the changes in concentration of a component in contact with a surface with changes in the surface tension, which results in a corresponding change in surface energy. For a binary system, the Gibbs adsorption equation in terms of surface excess is:

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.

The Marangoni number (Ma) is, as usually defined, the dimensionless number that compares the rate of transport due to Marangoni flows, with the rate of transport of diffusion. The Marangoni effect is flow of a liquid due to gradients in the surface tension of the liquid. Diffusion is of whatever is creating the gradient in the surface tension. Thus as the Marangoni number compares flow and diffusion timescales it is a type of Péclet number.

Melting-point depression is the phenomenon of reduction of the melting point of a material with a reduction of its size. This phenomenon is very prominent in nanoscale materials, which melt at temperatures hundreds of degrees lower than bulk materials.

<span class="mw-page-title-main">Viscosity</span> Resistance of a fluid to shear deformation

The viscosity of a fluid is a measure of its resistance to deformation at a given rate. For liquids, it corresponds to the informal concept of "thickness": for example, syrup has a higher viscosity than water. Viscosity is defined scientifically as a force multiplied by a time divided by an area. Thus its SI units are newton-seconds per square meter, or pascal-seconds.

In surface chemistry, disjoining pressure according to an IUPAC definition arises from an attractive interaction between two surfaces. For two flat and parallel surfaces, the value of the disjoining pressure can be calculated as the derivative of the Gibbs energy of interaction per unit area in respect to distance. There is also a related concept of disjoining force, which can be viewed as disjoining pressure times the surface area of the interacting surfaces.

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">Chemistry of photolithography</span> Overview article

Photolithography is a process in removing select portions of thin films used in microfabrication. Microfabrication is the production of parts on the micro- and nano- scale, typically on the surface of silicon wafers, for the production of integrated circuits, microelectromechanical systems (MEMS), solar cells, and other devices. Photolithography makes this process possible through the combined use of hexamethyldisilazane (HMDS), photoresist, spin coating, photomask, an exposure system and other various chemicals. By carefully manipulating these factors it is possible to create nearly any geometry microstructure on the surface of a silicon wafer. The chemical interaction between all the different components and the surface of the silicon wafer makes photolithography an interesting chemistry problem. Current engineering has been able to create features on the surface of silicon wafers between 1 and 100 μm.

In fluid mechanics, the thin-film equation is a partial differential equation that approximately predicts the time evolution of the thickness h of a liquid film that lies on a surface. The equation is derived via lubrication theory which is based on the assumption that the length-scales in the surface directions are significantly larger than in the direction normal to the surface. In the non-dimensional form of the Navier-Stokes equation the requirement is that terms of order ε2 and ε2Re are negligible, where ε ≪ 1 is the aspect ratio and Re is the Reynolds number. This significantly simplifies the governing equations. However, lubrication theory, as the name suggests, is typically derived for flow between two solid surfaces, hence the liquid forms a lubricating layer. The thin-film equation holds when there is a single free surface. With two free surfaces, the flow must be treated as a viscous sheet.

Liquid phase sintering is a sintering technique that uses a liquid phase to accelerate the interparticle bonding of the solid phase. In addition to rapid initial particle rearrangement due to capillary forces, mass transport through liquid is generally orders of magnitude faster than through solid, enhancing the diffusional mechanisms that drive densification. The liquid phase can be obtained either through mixing different powders—melting one component or forming a eutectic—or by sintering at a temperature between the liquidus and solidus. Additionally, since the softer phase is generally the first to melt, the resulting microstructure typically consists of hard particles in a ductile matrix, increasing the toughness of an otherwise brittle component. However, liquid phase sintering is inherently less predictable than solid phase sintering due to the complexity added by the presence of additional phases and rapid solidification rates. Activated sintering is the solid-state analog to the process of liquid phase sintering.

References

  1. Faraday, Michael (1933). Faraday's Diary. Vol. IV. London, England: Bell and Sons. p. 79 (entry for September 8, 1842).
  2. Tammann, G. (1910). "Zur Überhitzung von Kristallen" [On the superheating of crystals]. Zeitschrift für physikalische Chemie, Stöchiometrie und Verwandtschaftslehre (in German). 68 (3): 257–269.
  3. Stranski, I. N. (July 1942). "Über den Schmelzvorgang bei nichtpolaren Kristallen" [On the melting process in nonpolar crystals]. Naturwissenschaften (in German). 30: 425–433.
  4. Frenkel, J. (1946). Kinetic Theory of Liquids. Oxford, England, UK: Oxford University Press. pp. 136–156.
  5. 1 2 3 4 Dash, J.G.; Rempel, A.; Wettlaufer, J. (2006). "The physics of premelted ice and its geophysical consequences". Rev Mod Phys. 78 (3): 695. Bibcode:2006RvMP...78..695D. CiteSeerX   10.1.1.462.1061 . doi:10.1103/RevModPhys.78.695.
  6. Lipowski, R. (1982). "Critical Surface Phenomena at First-Order Bulk Transitions". Phys. Rev. Lett. 49 (21): 1575–1578. Bibcode:1982PhRvL..49.1575L. doi:10.1103/PhysRevLett.49.1575.
  7. Frenken, J.W.M. c; Van Der Veen, JF (1985). "Observation of Surface Melting" (PDF). Phys. Rev. Lett. 54 (2): 134–137. Bibcode:1985PhRvL..54..134F. doi:10.1103/PhysRevLett.54.134. hdl: 1887/71364 . PMID   10031263.
  8. Bluhm, H.; T. Inoue; M. Salmeron (2000). "Friction of ice measured using lateral force microscopy". Phys. Rev. B. 61 (11): 7760. Bibcode:2000PhRvB..61.7760B. doi:10.1103/PhysRevB.61.7760.
  9. 1 2 De Koning, J.J.; G. De Groot; G.J. Van Ingen Schenau (1992). "Ice friction during speed skating". J Biomech. 25 (6): 565–71. doi:10.1016/0021-9290(92)90099-M. PMID   1517252.
  10. 1 2 Why is Ice slippery? Rosenberg. pdf
  11. Colbeck, S.C. (1995). "Pressure melting and ice skating". Am J Phys. 63 (10): 888–890. Bibcode:1995AmJPh..63..888C. doi:10.1119/1.18028.