Stefan problem

Last updated

In mathematics and its applications, particularly to phase transitions in matter, a Stefan problem is a particular kind of boundary value problem for a system of partial differential equations (PDE), in which the boundary between the phases can move with time. The classical Stefan problem aims to describe the evolution of the boundary between two phases of a material undergoing a phase change, for example the melting of a solid, such as ice to water. This is accomplished by solving heat equations in both regions, subject to given boundary and initial conditions. At the interface between the phases (in the classical problem) the temperature is set to the phase change temperature. To close the mathematical system a further equation, the Stefan condition, is required. This is an energy balance which defines the position of the moving interface. Note that this evolving boundary is an unknown (hyper-)surface; hence, Stefan problems are examples of free boundary problems.

Contents

Analogous problems occur, for example, in the study of porous media flow, mathematical finance and crystal growth from monomer solutions. [1]

Historical note

The problem is named after Josef Stefan (Jožef Stefan), the Slovenian physicist who introduced the general class of such problems around 1890 in a series of four papers concerning the freezing of the ground and the formation of sea ice. [2] However, some 60 years earlier, in 1831, an equivalent problem, concerning the formation of the Earth's crust, had been studied by Lamé and Clapeyron. Stefan's problem admits a similarity solution, this is often termed the Neumann solution, which was allegedly presented in a series of lectures in the early 1860s.

A comprehensive description of the history of Stefan problems may be found in Rubinstein. [3]

Premises to the mathematical description

From a mathematical point of view, the phases are merely regions in which the solutions of the underlying PDE are continuous and differentiable up to the order of the PDE. In physical problems such solutions represent properties of the medium for each phase. The moving boundaries (or interfaces) are infinitesimally thin surfaces that separate adjacent phases; therefore, the solutions of the underlying PDE and its derivatives may suffer discontinuities across interfaces.

The underlying PDEs are not valid at the phase change interfaces; therefore, an additional condition—the Stefan condition—is needed to obtain closure. The Stefan condition expresses the local velocity of a moving boundary, as a function of quantities evaluated at either side of the phase boundary, and is usually derived from a physical constraint. In problems of heat transfer with phase change, for instance, conservation of energy dictates that the discontinuity of heat flux at the boundary must be accounted for by the rate of latent heat release (which is proportional to the local velocity of the interface).

The regularity of the equation has been studied mainly by Luis Caffarelli [4] [5] and further refined by work of Alessio Figalli, Xavier Ros-Oton and Joaquim Serra [6] [7]

Mathematical formulation

The one-dimensional one-phase Stefan problem

The one-phase Stefan problem is based on an assumption that one of the material phases may be neglected. Typically this is achieved by assuming that a phase is at the phase change temperature and hence any variation from this leads to a change of phase. This is a mathematically convenient approximation, which simplifies analysis whilst still demonstrating the essential ideas behind the process. A further standard simplification is to work in non-dimensional format, such that the temperature at the interface may be set to zero and far-field values to or .

Consider a semi-infinite one-dimensional block of ice initially at melting temperature for . The most well-known form of Stefan problem involves melting via an imposed constant temperature at the left hand boundary, leaving a region occupied by water. The melted depth, denoted by , is an unknown function of time. The Stefan problem is defined by

  • The heat equation:
  • A fixed temperature, above the melt temperature, on the left boundary:
  • The interface at the melting temperature is set to
  • The Stefan condition: where is the Stefan number, the ratio of latent to specific sensible heat (where specific indicates it is divided by the mass). Note this definition follows naturally from the nondimensionalisation and is used in many texts [8] [9] however it may also be defined as the inverse of this.
  • The initial temperature distribution:
  • The initial depth of the melted ice block:
The Neumann solution, obtained by using self-similar variables, indicates that the position of the boundary is given by where satisfies the transcendental equation
The temperature in the liquid is then given by

Applications

Apart from modelling melting of solids, Stefan problem is also used as a model for the asymptotic behaviour (in time) of more complex problems. For example, Pego [10] uses matched asymptotic expansions to prove that Cahn-Hilliard solutions for phase separation problems behave as solutions to a non-linear Stefan problem at an intermediate time scale. Additionally, the solution of the Cahn–Hilliard equation for a binary mixture is reasonably comparable with the solution of a Stefan problem. [11] In this comparison, the Stefan problem was solved using a front-tracking, moving-mesh method with homogeneous Neumann boundary conditions at the outer boundary. Also, Stefan problems can be applied to describe phase transformations other than solid-fluid or fluid-fluid. [12]

Application of Stefan problem to metal crystallization in electrochemical deposition of metal powders was envisaged by Călușaru [13]

The Stefan problem also has a rich inverse theory; in such problems, the meting depth (or curve or hyper-surface) s is the known datum and the problem is to find u or f. [14]

Advanced forms of Stefan problem

The classical Stefan problem deals with stationary materials with constant thermophysical properties (usually irrespective of phase), a constant phase change temperature and, in the example above, an instantaneous switch from the initial temperature to a distinct value at the boundary. In practice thermal properties may vary and specifically always do when the phase changes. The jump in density at phase change induces a fluid motion: the resultant kinetic energy does not figure in the standard energy balance. With an instantaneous temperature switch the initial fluid velocity is infinite, resulting in an initial infinite kinetic energy. In fact the liquid layer is often in motion, thus requiring advection or convection terms in the heat equation. The melt temperature may vary with size, curvature or speed of the interface. It is impossible to instantaneously switch temperatures and then difficult to maintain an exact fixed boundary temperature. Further, at the nanoscale the temperature may not even follow Fourier's law.

A number of these issues have been tackled in recent years for a variety of physical applications. In the solidification of supercooled melts an analysis where the phase change temperature depends on the interface velocity may be found in Font et al. [15] Nanoscale solidification, with variable phase change temperature and energy/density effects are modelled in. [16] [17] Solidification with flow in a channel has been studied, in the context of lava [18] and microchannels, [19] or with a free surface in the context of water freezing over an ice layer. [20] [21] A general model including different properties in each phase, variable phase change temperature and heat equations based on either Fourier's law or the Guyer-Krumhansl equation is analysed in. [22]

See also

Notes

  1. Applied partial differential equations. Ockendon, J. R. (Rev. ed.). Oxford: Oxford University Press. 2003. ISBN   0-19-852770-5. OCLC   52486357.{{cite book}}: CS1 maint: others (link)
  2. ( Vuik 1993 , p. 157).
  3. RUBINSTEIN, L. I. (2016). STEFAN PROBLEM. [Place of publication not identified]: American Mathematical Society. ISBN   978-1-4704-2850-1. OCLC   973324855.
  4. Caffarelli, Luis A. (1977). "The regularity of free boundaries in higher dimensions". Acta Mathematica. 139 (none): 155–184. doi: 10.1007/BF02392236 . ISSN   0001-5962. S2CID   123660704.
  5. CAFFARELLI, LUIS A. (1978). "Some Aspects of the One-Phase Stefan Problem". Indiana University Mathematics Journal. 27 (1): 73–77. doi: 10.1512/iumj.1978.27.27006 . ISSN   0022-2518. JSTOR   24891579.
  6. Figalli, Alessio; Ros-Oton, Xavier; Serra, Joaquim (2021-03-24). "The singular set in the Stefan problem". arXiv: 2103.13379v1 .{{cite journal}}: Cite journal requires |journal= (help)
  7. Rorvig, Mordechai (2021-10-06). "Mathematicians Prove Melting Ice Stays Smooth". Quanta Magazine. Retrieved 2021-10-08.
  8. Davis, Stephen H., 1939-. Theory of solidification. Cambridge. ISBN   978-0-511-01924-1. OCLC   232161077.{{cite book}}: CS1 maint: multiple names: authors list (link) CS1 maint: numeric names: authors list (link)
  9. Fowler, A. C. (Andrew Cadle), 1953- (1997). Mathematical models in the applied sciences. Cambridge: Cambridge University Press. ISBN   0-521-46140-5. OCLC   36621805.{{cite book}}: CS1 maint: multiple names: authors list (link) CS1 maint: numeric names: authors list (link)
  10. R. L. Pego. (1989). Front Migration in the Nonlinear Cahn-Hilliard Equation. Proc. R. Soc. Lond. A.,422:261278.
  11. Vermolen, F. J.; Gharasoo, M. G.; Zitha, P. L. J.; Bruining, J. (2009). "Numerical Solutions of Some Diffuse Interface Problems: The Cahn–Hilliard Equation and the Model of Thomas and Windle". International Journal for Multiscale Computational Engineering. 7 (6): 523–543. doi:10.1615/IntJMultCompEng.v7.i6.40.
  12. Alvarenga HD, Van de Putter T, Van Steenberge N, Sietsma J, Terryn H (Apr 2009). "Influence of Carbide Morphology and Microstructure on the Kinetics of Superficial Decarburization of C-Mn Steels". Metallurgical and Materials Transactions A. 46 (1): 123–133. Bibcode:2015MMTA...46..123A. doi:10.1007/s11661-014-2600-y. S2CID   136871961.
  13. A. Călușaru, Electrochemical Deposition of Metal Powders, Elsevier, Amsterdam,1978.
  14. ( Kirsch 1996 ).
  15. Font, F.; Mitchell, S. L.; Myers, T. G. (2013-07-01). "One-dimensional solidification of supercooled melts". International Journal of Heat and Mass Transfer. 62: 411–421. doi:10.1016/j.ijheatmasstransfer.2013.02.070. hdl: 2072/205484 . ISSN   0017-9310.
  16. Myers, T. G. (2016-08-01). "Mathematical modelling of phase change at the nanoscale". International Communications in Heat and Mass Transfer. 76: 59–62. doi:10.1016/j.icheatmasstransfer.2016.05.005. ISSN   0735-1933.
  17. Font, F.; Myers, T. G.; Mitchell, S. L. (February 2015). "A mathematical model for nanoparticle melting with density change". Microfluidics and Nanofluidics. 18 (2): 233–243. doi:10.1007/s10404-014-1423-x. ISSN   1613-4982. S2CID   54087370.
  18. Lister, J.R. (1994). "The solidification of buoyancy-driven flow in a flexible-walled channel. Part 1. Constant-volume release". Journal of Fluid Mechanics. 272: 21–44. Bibcode:1994JFM...272...21L. doi:10.1017/S0022112094004362. S2CID   124068245.
  19. Myers, T. G.; Low, J. (October 2011). "An approximate mathematical model for solidification of a flowing liquid in a microchannel". Microfluidics and Nanofluidics. 11 (4): 417–428. doi:10.1007/s10404-011-0807-4. hdl: 2072/169268 . ISSN   1613-4982. S2CID   97060677.
  20. Myers, T. G.; Charpin, J. P. F.; Chapman, S. J. (August 2002). "The flow and solidification of a thin fluid film on an arbitrary three-dimensional surface". Physics of Fluids. 14 (8): 2788–2803. Bibcode:2002PhFl...14.2788M. doi:10.1063/1.1488599. hdl: 2117/102903 . ISSN   1070-6631.
  21. Myers, T.G.; Charpin, J.P.F. (December 2004). "A mathematical model for atmospheric ice accretion and water flow on a cold surface". International Journal of Heat and Mass Transfer. 47 (25): 5483–5500. doi:10.1016/j.ijheatmasstransfer.2004.06.037.
  22. Myers, T. G.; Hennessy, M. G.; Calvo-Schwarzwälder, M. (2020-03-01). "The Stefan problem with variable thermophysical properties and phase change temperature". International Journal of Heat and Mass Transfer. 149: 118975. arXiv: 1904.05698 . doi:10.1016/j.ijheatmasstransfer.2019.118975. hdl:2072/445741. ISSN   0017-9310. S2CID   115147121.

Related Research Articles

<span class="mw-page-title-main">Partial differential equation</span> Type of differential equation

In mathematics, a partial differential equation (PDE) is an equation which computes a function between various partial derivatives of a multivariable function.

Conduction is the process by which heat is transferred from the hotter end to the colder end of an object. The ability of the object to conduct heat is known as its thermal conductivity, and is denoted k.

<span class="mw-page-title-main">Heat equation</span> Partial differential equation describing the evolution of temperature in a region

In mathematics and physics, the heat equation is a certain partial differential equation. Solutions of the heat equation are sometimes known as caloric functions. The theory of the heat equation was first developed by Joseph Fourier in 1822 for the purpose of modeling how a quantity such as heat diffuses through a given region.

<span class="mw-page-title-main">Boundary layer</span> Layer of fluid in the immediate vicinity of a bounding surface

In physics and fluid mechanics, a boundary layer is the thin layer of fluid in the immediate vicinity of a bounding surface formed by the fluid flowing along the surface. The fluid's interaction with the wall induces a no-slip boundary condition. The flow velocity then monotonically increases above the surface until it returns to the bulk flow velocity. The thin layer consisting of fluid whose velocity has not yet returned to the bulk flow velocity is called the velocity boundary layer.

<span class="mw-page-title-main">Boltzmann equation</span> Equation of statistical mechanics

The Boltzmann equation or Boltzmann transport equation (BTE) describes the statistical behaviour of a thermodynamic system not in a state of equilibrium; it was devised by Ludwig Boltzmann in 1872. The classic example of such a system is a fluid with temperature gradients in space causing heat to flow from hotter regions to colder ones, by the random but biased transport of the particles making up that fluid. In the modern literature the term Boltzmann equation is often used in a more general sense, referring to any kinetic equation that describes the change of a macroscopic quantity in a thermodynamic system, such as energy, charge or particle number.

<span class="mw-page-title-main">Differential equation</span> Type of functional equation (mathematics)

In mathematics, a differential equation is an equation that relates one or more unknown functions and their derivatives. In applications, the functions generally represent physical quantities, the derivatives represent their rates of change, and the differential equation defines a relationship between the two. Such relations are common; therefore, differential equations play a prominent role in many disciplines including engineering, physics, economics, and biology.

In mathematics, a hyperbolic partial differential equation of order is a partial differential equation (PDE) that, roughly speaking, has a well-posed initial value problem for the first derivatives. More precisely, the Cauchy problem can be locally solved for arbitrary initial data along any non-characteristic hypersurface. Many of the equations of mechanics are hyperbolic, and so the study of hyperbolic equations is of substantial contemporary interest. The model hyperbolic equation is the wave equation. In one spatial dimension, this is

In mathematics, the convergence condition by Courant–Friedrichs–Lewy is a necessary condition for convergence while solving certain partial differential equations numerically. It arises in the numerical analysis of explicit time integration schemes, when these are used for the numerical solution. As a consequence, the time step must be less than a certain upper bound, given a fixed spatial increment, in many explicit time-marching computer simulations; otherwise, the simulation produces incorrect or unstable results. The condition is named after Richard Courant, Kurt Friedrichs, and Hans Lewy who described it in their 1928 paper.

In mathematics, the viscosity solution concept was introduced in the early 1980s by Pierre-Louis Lions and Michael G. Crandall as a generalization of the classical concept of what is meant by a 'solution' to a partial differential equation (PDE). It has been found that the viscosity solution is the natural solution concept to use in many applications of PDE's, including for example first order equations arising in dynamic programming, differential games or front evolution problems, as well as second-order equations such as the ones arising in stochastic optimal control or stochastic differential games.

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

A parabolic partial differential equation is a type of partial differential equation (PDE). Parabolic PDEs are used to describe a wide variety of time-dependent phenomena, including heat conduction, particle diffusion, and pricing of derivative investment instruments.

<span class="mw-page-title-main">Thermal effusivity</span> Ability of a material to exchange thermal energy with surroundings

In thermodynamics, a material's thermal effusivity, also known as thermal responsivity, is a measure of its ability to exchange thermal energy with its surroundings. It is defined as the square root of the product of the material's thermal conductivity and its volumetric heat capacity or as the ratio of thermal conductivity to the square root of thermal diffusivity.

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.

A phase-field model is a mathematical model for solving interfacial problems. It has mainly been applied to solidification dynamics, but it has also been applied to other situations such as viscous fingering, fracture mechanics, hydrogen embrittlement, and vesicle dynamics.

In engineering, physics, and chemistry, the study of transport phenomena concerns the exchange of mass, energy, charge, momentum and angular momentum between observed and studied systems. While it draws from fields as diverse as continuum mechanics and thermodynamics, it places a heavy emphasis on the commonalities between the topics covered. Mass, momentum, and heat transport all share a very similar mathematical framework, and the parallels between them are exploited in the study of transport phenomena to draw deep mathematical connections that often provide very useful tools in the analysis of one field that are directly derived from the others.

In mathematics, a free boundary problem is a partial differential equation to be solved for both an unknown function and an unknown domain . The segment of the boundary of which is not known at the outset of the problem is the free boundary.

Riemann invariants are mathematical transformations made on a system of conservation equations to make them more easily solvable. Riemann invariants are constant along the characteristic curves of the partial differential equations where they obtain the name invariant. They were first obtained by Bernhard Riemann in his work on plane waves in gas dynamics.

Computational Fluid Dynamics (CFD) modeling and simulation for phase change materials (PCMs) is a technique to analyze the performance and behavior of PCMs. The CFD models have been successful in studying and analyzing the air quality, natural ventilation and stratified ventilation, air flow initiated by buoyancy forces and temperature space for the systems integrated with PCMs. Simple shapes like flat plates, cylinders or annular tubes, fins, macro- and micro-encapsulations with containers of different shape are often modeled in CFD software's to study.

In the study of partial differential equations, particularly in fluid dynamics, a self-similar solution is a form of solution which is similar to itself if the independent and dependent variables are appropriately scaled. Self-similar solutions appear whenever the problem lacks a characteristic length or time scale. These include, for example, the Blasius boundary layer or the Sedov–Taylor shell.

<span class="mw-page-title-main">Physics-informed neural networks</span> Technique to solve partial differential equations

Physics-informed neural networks (PINNs) are a type of universal function approximators that can embed the knowledge of any physical laws that govern a given data-set in the learning process, and can be described by partial differential equations (PDEs). They overcome the low data availability of some biological and engineering systems that makes most state-of-the-art machine learning techniques lack robustness, rendering them ineffective in these scenarios. The prior knowledge of general physical laws acts in the training of neural networks (NNs) as a regularization agent that limits the space of admissible solutions, increasing the correctness of the function approximation. This way, embedding this prior information into a neural network results in enhancing the information content of the available data, facilitating the learning algorithm to capture the right solution and to generalize well even with a low amount of training examples.

References

Historical references

Scientific and general references