fbpx
Wikipedia

Fick's laws of diffusion

Fick's laws of diffusion describe diffusion and were derived by Adolf Fick in 1855.[1] They can be used to solve for the diffusion coefficient, D. Fick's first law can be used to derive his second law which in turn is identical to the diffusion equation.

Molecular diffusion from a microscopic and macroscopic point of view. Initially, there are solute molecules on the left side of a barrier (purple line) and none on the right. The barrier is removed, and the solute diffuses to fill the whole container. Top: A single molecule moves around randomly. Middle: With more molecules, there is a clear trend where the solute fills the container more and more uniformly. Bottom: With an enormous number of solute molecules, randomness becomes undetectable: The solute appears to move smoothly and systematically from high-concentration areas to low-concentration areas. This smooth flow is described by Fick's laws.

A diffusion process that obeys Fick's laws is called normal or Fickian diffusion; otherwise, it is called anomalous diffusion or non-Fickian diffusion.

History

In 1855, physiologist Adolf Fick first reported[1] his now well-known laws governing the transport of mass through diffusive means. Fick's work was inspired by the earlier experiments of Thomas Graham, which fell short of proposing the fundamental laws for which Fick would become famous. Fick's law is analogous to the relationships discovered at the same epoch by other eminent scientists: Darcy's law (hydraulic flow), Ohm's law (charge transport), and Fourier's Law (heat transport).

Fick's experiments (modeled on Graham's) dealt with measuring the concentrations and fluxes of salt, diffusing between two reservoirs through tubes of water. It is notable that Fick's work primarily concerned diffusion in fluids, because at the time, diffusion in solids was not considered generally possible.[2] Today, Fick's Laws form the core of our understanding of diffusion in solids, liquids, and gases (in the absence of bulk fluid motion in the latter two cases). When a diffusion process does not follow Fick's laws (which happens in cases of diffusion through porous media and diffusion of swelling penetrants, among others),[3][4] it is referred to as non-Fickian.

Fick's first law

Fick's first law relates the diffusive flux to the gradient of the concentration. It postulates that the flux goes from regions of high concentration to regions of low concentration, with a magnitude that is proportional to the concentration gradient (spatial derivative), or in simplistic terms the concept that a solute will move from a region of high concentration to a region of low concentration across a concentration gradient. In one (spatial) dimension, the law can be written in various forms, where the most common form (see[5][6]) is in a molar basis:

 

where

  • J is the diffusion flux, of which the dimension is the amount of substance per unit area per unit time. J measures the amount of substance that will flow through a unit area during a unit time interval.
  • D is the diffusion coefficient or diffusivity. Its dimension is area per unit time.
  • φ (for ideal mixtures) is the concentration, of which the dimension is the amount of substance per unit volume.
  • x is position, the dimension of which is length.

D is proportional to the squared velocity of the diffusing particles, which depends on the temperature, viscosity of the fluid and the size of the particles according to the Stokes–Einstein relation. In dilute aqueous solutions the diffusion coefficients of most ions are similar and have values that at room temperature are in the range of (0.6–2)×10−9 m2/s. For biological molecules the diffusion coefficients normally range from 10−10 to 10−11 m2/s.

In two or more dimensions we must use , the del or gradient operator, which generalises the first derivative, obtaining

 

where J denotes the diffusion flux vector.

The driving force for the one-dimensional diffusion is the quantity φ/x, which for ideal mixtures is the concentration gradient.

Alternative formulations of the first law

Another form for the first law is to write it with the primary variable as mass fraction (yi, given for example in kg/kg), then the equation changes to:

 

where

  • the index i denotes the ith species,
  • Ji is the diffusion flux vector of the ith species (for example in mol/m2-s),
  • Mi is the molar mass of the ith species, and
  • ρ is the mixture density (for example in kg/m3).

Note that the   is outside the gradient operator. This is because:

 

where ρsi is the partial density of the ith species.

Beyond this, in chemical systems other than ideal solutions or mixtures, the driving force for diffusion of each species is the gradient of chemical potential of this species. Then Fick's first law (one-dimensional case) can be written

 

where

  • the index i denotes the ith species.
  • c is the concentration (mol/m3).
  • R is the universal gas constant (J/K/mol).
  • T is the absolute temperature (K).
  • μ is the chemical potential (J/mol).

The driving force of Fick's law can be expressed as a fugacity difference:

 

Fugacity   has Pa units.   is a partial pressure of component i in a vapor   or liquid   phase. At vapor liquid equilibrium the evaporation flux is zero because  .

Derivation of Fick's first law for gases

Four versions of Fick's law for binary gas mixtures are given below. These assume: thermal diffusion is negligible; the body force per unit mass is the same on both species; and either pressure is constant or both species have the same molar mass. Under these conditions, Ref. [7] shows in detail how the diffusion equation from the kinetic theory of gases reduces to this version of Fick's law:

 

where Vi is the diffusion velocity of species i. In terms of species flux this is

 

If, additionally,  , this reduces to the most common form of Fick's law,

 

If (instead of or in addition to  ) both species have the same molar mass, Fick's law becomes

 

where   is the mole fraction of species i.

Fick's second law

Fick's second law predicts how diffusion causes the concentration to change with respect to time. It is a partial differential equation which in one dimension reads:

 

where

  • φ is the concentration in dimensions of [(amount of substance) length−3], example mol/m3; φ = φ(x,t) is a function that depends on location x and time t
  • t is time, example s
  • D is the diffusion coefficient in dimensions of [length2 time−1], example m2/s
  • x is the position [length], example m

In two or more dimensions we must use the Laplacian Δ = ∇2, which generalises the second derivative, obtaining the equation

 

Fick's second law has the same mathematical form as the Heat equation and its fundamental solution is the same as the Heat kernel, except switching thermal conductivity   with diffusion coefficient  :

 

Derivation of Fick's second law

Fick's second law can be derived from Fick's first law and the mass conservation in absence of any chemical reactions:

 

Assuming the diffusion coefficient D to be a constant, one can exchange the orders of the differentiation and multiply by the constant:

 

and, thus, receive the form of the Fick's equations as was stated above.

For the case of diffusion in two or more dimensions Fick's second law becomes

 

which is analogous to the heat equation.

If the diffusion coefficient is not a constant, but depends upon the coordinate or concentration, Fick's second law yields

 

An important example is the case where φ is at a steady state, i.e. the concentration does not change by time, so that the left part of the above equation is identically zero. In one dimension with constant D, the solution for the concentration will be a linear change of concentrations along x. In two or more dimensions we obtain

 

which is Laplace's equation, the solutions to which are referred to by mathematicians as harmonic functions.

Example solutions and generalization

Fick's second law is a special case of the convection–diffusion equation in which there is no advective flux and no net volumetric source. It can be derived from the continuity equation:

 

where j is the total flux and R is a net volumetric source for φ. The only source of flux in this situation is assumed to be diffusive flux:

 

Plugging the definition of diffusive flux to the continuity equation and assuming there is no source (R = 0), we arrive at Fick's second law:

 

If flux were the result of both diffusive flux and advective flux, the convection–diffusion equation is the result.

Example solution 1: constant concentration source and diffusion length

A simple case of diffusion with time t in one dimension (taken as the x-axis) from a boundary located at position x = 0, where the concentration is maintained at a value n0 is

 

where erfc is the complementary error function. This is the case when corrosive gases diffuse through the oxidative layer towards the metal surface (if we assume that concentration of gases in the environment is constant and the diffusion space – that is, the corrosion product layer – is semi-infinite, starting at 0 at the surface and spreading infinitely deep in the material). If, in its turn, the diffusion space is infinite (lasting both through the layer with n(x, 0) = 0, x > 0 and that with n(x, 0) = n0, x ≤ 0), then the solution is amended only with coefficient 1/2 in front of n0 (as the diffusion now occurs in both directions). This case is valid when some solution with concentration n0 is put in contact with a layer of pure solvent. (Bokstein, 2005) The length 2Dt is called the diffusion length and provides a measure of how far the concentration has propagated in the x-direction by diffusion in time t (Bird, 1976).

As a quick approximation of the error function, the first two terms of the Taylor series can be used:

 

If D is time-dependent, the diffusion length becomes

 

This idea is useful for estimating a diffusion length over a heating and cooling cycle, where D varies with temperature.

Example solution 2: Brownian particle and Mean squared displacement

Another simple case of diffusion is the Brownian motion of one particle. The particle's Mean squared displacement from its original position is:

 

where   is the dimension of the particle's Brownian motion. For example, the diffusion of a molecule across a cell membrane 8 nm thick is 1-D diffusion because of the spherical symmetry; However, the diffusion of a molecule from the membrane to the center of a eukaryotic cell is a 3-D diffusion. For a cylindrical cactus, the diffusion from photosynthetic cells on its surface to its center (the axis of its cylindrical symmetry) is a 2-D diffusion.

The square root of MSD,  , is often used as a characterization of how far has the particle moved after time   has elapsed. The MSD is symmetrically distributed over the 1D, 2D, and 3D space. Thus, the probability distribution of the magnitude of MSD in 1D is Gaussian and 3D is a Maxwell-Boltzmann distribution.

Generalizations

  • In non-homogeneous media, the diffusion coefficient varies in space, D = D(x). This dependence does not affect Fick's first law but the second law changes:
     
  • In anisotropic media, the diffusion coefficient depends on the direction. It is a symmetric tensor Dji = Dij. Fick's first law changes to
     
    it is the product of a tensor and a vector:
     
    For the diffusion equation this formula gives
     
    The symmetric matrix of diffusion coefficients Dij should be positive definite. It is needed to make the right hand side operator elliptic.
  • For inhomogeneous anisotropic media these two forms of the diffusion equation should be combined in
     
  • The approach based on Einstein's mobility and Teorell formula gives the following generalization of Fick's equation for the multicomponent diffusion of the perfect components:
     
    where φi are concentrations of the components and Dij is the matrix of coefficients. Here, indices i and j are related to the various components and not to the space coordinates.

The Chapman–Enskog formulae for diffusion in gases include exactly the same terms. These physical models of diffusion are different from the test models tφi = Σj Dij Δφj which are valid for very small deviations from the uniform equilibrium. Earlier, such terms were introduced in the Maxwell–Stefan diffusion equation.

For anisotropic multicomponent diffusion coefficients one needs a rank-four tensor, for example Dij,αβ, where i, j refer to the components and α, β = 1, 2, 3 correspond to the space coordinates.

Applications

Equations based on Fick's law have been commonly used to model transport processes in foods, neurons, biopolymers, pharmaceuticals, porous soils, population dynamics, nuclear materials, plasma physics, and semiconductor doping processes. The theory of voltammetric methods is based on solutions of Fick's equation. On the other hand, in some cases a "Fickian (another common approximation of the transport equation is that of the diffusion theory)[8]" description is inadequate. For example, in polymer science and food science a more general approach is required to describe transport of components in materials undergoing a glass transition. One more general framework is the Maxwell–Stefan diffusion equations[9] of multi-component mass transfer, from which Fick's law can be obtained as a limiting case, when the mixture is extremely dilute and every chemical species is interacting only with the bulk mixture and not with other species. To account for the presence of multiple species in a non-dilute mixture, several variations of the Maxwell–Stefan equations are used. See also non-diagonal coupled transport processes (Onsager relationship).

Fick's flow in liquids

When two miscible liquids are brought into contact, and diffusion takes place, the macroscopic (or average) concentration evolves following Fick's law. On a mesoscopic scale, that is, between the macroscopic scale described by Fick's law and molecular scale, where molecular random walks take place, fluctuations cannot be neglected. Such situations can be successfully modeled with Landau-Lifshitz fluctuating hydrodynamics. In this theoretical framework, diffusion is due to fluctuations whose dimensions range from the molecular scale to the macroscopic scale.[10]

In particular, fluctuating hydrodynamic equations include a Fick's flow term, with a given diffusion coefficient, along with hydrodynamics equations and stochastic terms describing fluctuations. When calculating the fluctuations with a perturbative approach, the zero order approximation is Fick's law. The first order gives the fluctuations, and it comes out that fluctuations contribute to diffusion. This represents somehow a tautology, since the phenomena described by a lower order approximation is the result of a higher approximation: this problem is solved only by renormalizing the fluctuating hydrodynamics equations.

Sorption rate and collision frequency of diluted solute

 
Scheme of molecular diffusion in the solution. Orange dots are solute molecules, solvent molecules are not drawn, black arrow is an example random walk trajectory, and the red curve is the diffusive Gaussian broadening probability function from the Fick's law of diffusion.[11]:Fig. 9

The adsorption or absorption rate of a dilute solute to a surface or interface in a (gas or liquid) solution can be calculated using Fick's laws of diffusion. The accumulated number of molecules adsorbed on the surface is expressed by the Langmuir-Schaefer equation at the short-time limit by integrating the diffusion flux equation over time:[12]

 
  •   is number of molecules in unit # molecules adsorbed during the time  .
  • A is the surface area in unit  .
  • C is the number concentration of the adsorber molecules in the bulk solution in unit # molecules/ .
  • D is diffusion coefficient of the adsorber in unit  .
  • t is elapsed time in unit  .

The equation is named after American chemists Irving Langmuir and Vincent Schaefer.

The Langmuir-Schaefer equation can be extended to the Ward-Tordai Equation to account for the "back-diffusion" of rejected molecules from the surface:[13]

 

where   is the bulk concentration,   is the sub-surface concentration (which is a function of time depending on the reaction model of the adsorption), and   is a dummy variable.

Monte Carlo simulations show that these two equations work to predict the adsorption rate of systems that form predictable concentration gradients near the surface but have troubles for systems without or with unpredictable concentration gradients, such as typical biosensing systems or when flow and convection are significant.[14]

 
A brief history of the theories on diffusive adsorption.[14]

A brief history of diffusive adsorption is shown in the right figure.[14] A noticeable challenge of understanding the diffusive adsorption at the single-molecule level is the fractal nature of diffusion. Most computer simulations pick a time step for diffusion which ignores the fact that there are self-similar finer diffusion events (fractal) within each step. Simulating the fractal diffusion shows that a factor of two corrections should be introduced for the result of a fixed time-step adsorption simulation, bringing it to be consistent with the above two equations.[14]

In the ultrashort time limit, in the order of the diffusion time a2/D, where a is the particle radius, the diffusion is described by the Langevin equation. At a longer time, the Langevin equation merges into the Stokes–Einstein equation. The latter is appropriate for the condition of the diluted solution, where long-range diffusion is considered. According to the fluctuation-dissipation theorem based on the Langevin equation in the long-time limit and when the particle is significantly denser than the surrounding fluid, the time-dependent diffusion constant is:[15]

 

where (all in SI units)

For a single molecule such as organic molecules or biomolecules (e.g. proteins) in water, the exponential term is negligible due to the small product of in the picosecond region.

When the area of interest is the size of a molecule (specifically, a long cylindrical molecule such as DNA), the adsorption rate equation represents the collision frequency of two molecules in a diluted solution, with one molecule a specific side and the other no steric dependence, i.e., a molecule (random orientation) hit one side of the other. The diffusion constant need to be updated to the relative diffusion constant between two diffusing molecules. This estimation is especially useful in studying the interaction between a small molecule and a larger molecule such as a protein. The effective diffusion constant is dominated by the smaller one whose diffusion constant can be used instead.

The above hitting rate equation is also useful to predict the kinetics of molecular self-assembly on a surface. Molecules are randomly oriented in the bulk solution. Assuming 1/6 of the molecules has the right orientation to the surface binding sites, i.e. 1/2 of the z-direction in x, y, z three dimensions, thus the concentration of interest is just 1/6 of the bulk concentration. Put this value into the equation one should be able to calculate the theoretical adsorption kinetic curve using the Langmuir adsorption model. In a more rigid picture, 1/6 can be replaced by the steric factor of the binding geometry.

Biological perspective

The first law gives rise to the following formula:[16]

 

in which

  • P is the permeability, an experimentally determined membrane "conductance" for a given gas at a given temperature.
  • c2c1 is the difference in concentration of the gas across the membrane for the direction of flow (from c1 to c2).

Fick's first law is also important in radiation transfer equations. However, in this context, it becomes inaccurate when the diffusion constant is low and the radiation becomes limited by the speed of light rather than by the resistance of the material the radiation is flowing through. In this situation, one can use a flux limiter.

The exchange rate of a gas across a fluid membrane can be determined by using this law together with Graham's law.

Under the condition of a diluted solution when diffusion takes control, the membrane permeability mentioned in the above section can be theoretically calculated for the solute using the equation mentioned in the last section (use with particular care because the equation is derived for dense solutes, while biological molecules are not denser than water):[11]

 

where

  •   is the total area of the pores on the membrane (unit m2).
  •   transmembrane efficiency (unitless), which can be calculated from the stochastic theory of chromatography.
  • D is the diffusion constant of the solute unit m2s−1.
  • t is time unit s.
  • c2, c1 concentration should use unit mol m−3, so flux unit becomes mol s−1.

The flux is decay over the square root of time because a concentration gradient builds up near the membrane over time under ideal conditions. When there is flow and convection, the flux can be significantly different than the equation predicts and show an effective time t with a fixed value,[14] which makes the flux stable instead of decay over time. This strategy is adopted in biology such as blood circulation.

Semiconductor fabrication applications

The semiconductor is a collective term for a series of devices. It mainly includes three categories:two-terminal devices, three-terminal devices, and four-terminal devices. The combination of the semiconductors is called an integrated circuit.

The relationship between Fick's law and semiconductors: the principle of the semiconductor is transferring chemicals or dopants from a layer to a layer. Fick's law can be used to control and predict the diffusion by knowing how much the concentration of the dopants or chemicals move per meter and second through mathematics.

Therefore, different types and levels of semiconductors can be fabricated.

Integrated circuit fabrication technologies, model processes like CVD, thermal oxidation, wet oxidation, doping, etc. use diffusion equations obtained from Fick's law.

CVD method of fabricate semiconductor

The wafer is a kind of semiconductor whose silicon substrate is coated with a layer of CVD-created polymer chain and films. This film contains n-type and p-type dopants and takes responsibility for dopant conductions. The principle of CVD relies on the gas phase and gas-solid chemical reaction to create thin films.

The viscous flow regime of CVD is driven by a pressure gradient. CVD also includes a diffusion component distinct from the surface diffusion of adatoms. In CVD, reactants and products must also diffuse through a boundary layer of stagnant gas that exists next to the substrate. The total number of steps required for CVD film growth are gas phase diffusion of reactants through the boundary layer, adsorption and surface diffusion of adatoms, reactions on the substrate, and gas phase diffusion of products away through the boundary layer.

The velocity profile for gas flow is:

 
where
  •   is the thickness
  •   is the Reynolds number
  • x is the length of the subtrate.
  • v = 0 at any surface
  •   is viscosity
  •   is density.

Integrated the x from 0 to L, it gives the average thickness:

 

To keep the reaction balanced, reactants must diffuse through the stagnant boundary layer to reach the substrate. So a thin boundary layer is desirable. According to the equations, increasing vo would result in more wasted reactants. The reactants will not reach the substrate uniformly if the flow becomes turbulent. Another option is to switch to a new carrier gas with lower viscosity or density.

The Fick's first law describes diffusion through the boundary layer. As a function of pressure (P) and temperature (T) in a gas, diffusion is determined.

 
where
  •   is the standard pressure.
  •   is the standard temperature.
  •   is the standard diffusitivity.

The equation tells that increasing the temperature or decreasing the pressure can increase the diffusivity.

Fick's first law predicts the flux of the reactants to the substrate and product away from the substrate:

 
where
  •   is the thickness  
  •   is the first reactant's concentration.

In ideal gas law  , the concentration of the gas is expressed by partial pressure.

 
where
  •   is the gas constant.
  •   is the partial pressure gradient.

As a result, Fick's first law tells us we can use a partial pressure gradient to control the diffusivity and control the growth of thin films of semiconductors.

In many realistic situations, the simple Fick's law is not an adequate formulation for the semiconductor problem. It only applies to certain conditions, for example, given the semiconductor boundary conditions: constant source concentration diffusion, limited source concentration, or moving boundary diffusion (where junction depth keeps moving into the substrate).

Food production and cooking

The formulation of Fick's first law can explain a variety of complex phenomena in the context of food and cooking: Diffusion of molecules such as ethylene promotes plant growth and ripening, salt and sugar molecules promotes meat brining and marinating, and water molecules promote dehydration. Fick's first law can also be used to predict the changing moisture profiles across a spaghetti noodle as it hydrates during cooking. These phenomena are all about the spontaneous movement of particles of solutes driven by the concentration gradient. In different situations, there is different diffusivity which is a constant.[17]

By controlling the concentration gradient, the cooking time, shape of the food, and salting can be controlled.

See also

Citations

  1. ^ a b * Fick, A. (1855). "Ueber Diffusion". Annalen der Physik (in German). 94 (1): 59–86. Bibcode:1855AnP...170...59F. doi:10.1002/andp.18551700105.
    • Fick, A. (1855). "V. On liquid diffusion". Phil. Mag. 10 (63): 30–39. doi:10.1080/14786445508641925.
  2. ^ Philibert, Jean (2005). (PDF). Diffusion Fundamentals. 2: 1.1–1.10. Archived from the original (PDF) on 5 February 2009.
  3. ^ Vázquez, J. L. (2006). "The Porous Medium Equation". Mathematical Theory. Oxford Univ. Press.
  4. ^ Gorban, A. N.; Sargsyan, H. P.; Wahab, H. A. (2011). "Quasichemical Models of Multicomponent Nonlinear Diffusion". Mathematical Modelling of Natural Phenomena. 6 (5): 184–262. arXiv:1012.2908. doi:10.1051/mmnp/20116509. S2CID 18961678.
  5. ^ Atkins, Peter; de Paula, Julio (2006). Physical Chemistry for the Life Science.
  6. ^ Conlisk, A. Terrence (2013). Essentials of Micro- and Nanofluidics: With Applications to the Biological and Chemical Sciences. Cambridge University Press. p. 43. ISBN 9780521881685.
  7. ^ Williams, F.A. (1985). "Appendix E". Combustion Theory. Benjamin/Cummings.
  8. ^ "Fickian Diffusion - an overview | ScienceDirect Topics". www.sciencedirect.com. Retrieved 11 May 2022.
  9. ^ Taylor, Ross; Krishna, R. (1993). "Multicomponent mass transfer". Wiley. {{cite journal}}: Cite journal requires |journal= (help)
  10. ^ Brogioli, D.; Vailati, A. (2001). "Diffusive mass transfer by nonequilibrium fluctuations: Fick's law revisited". Phys. Rev. E. 63 (1–4): 012105. arXiv:cond-mat/0006163. Bibcode:2000PhRvE..63a2105B. doi:10.1103/PhysRevE.63.012105. PMID 11304296. S2CID 1302913.
  11. ^ a b Pyle, Joseph R.; Chen, Jixin (2 November 2017). "Photobleaching of YOYO-1 in super-resolution single DNA fluorescence imaging". Beilstein Journal of Nanotechnology. 8: 2292–2306. doi:10.3762/bjnano.8.229. PMC 5687005. PMID 29181286.
  12. ^ Langmuir, I.; Schaefer, V.J. (1937). "The Effect of Dissolved Salts on Insoluble Monolayers". Journal of the American Chemical Society. 29 (11): 2400–2414. doi:10.1021/ja01290a091.
  13. ^ Ward, A.F.H.; Tordai, L. (1946). "Time-dependence of Boundary Tensions of Solutions I. The Role of Diffusion in Time-effects". Journal of Chemical Physics. 14 (7): 453–461. Bibcode:1946JChPh..14..453W. doi:10.1063/1.1724167.
  14. ^ a b c d e Chen, J. (2022). "Simulating stochastic adsorption of diluted solute molecules at interfaces". AIP Advances. 12 (1): 015318. Bibcode:2022AIPA...12a5318C. doi:10.1063/5.0064140. PMC 8758205. PMID 35070490.
  15. ^ Bian, Xin; Kim, Changho; Karniadakis, George Em (14 August 2016). "111 years of Brownian motion". Soft Matter. 12 (30): 6331–6346. Bibcode:2016SMat...12.6331B. doi:10.1039/c6sm01153e. PMC 5476231. PMID 27396746.
  16. ^ Nosek, Thomas M. . Essentials of Human Physiology. Archived from the original on 24 March 2016.
  17. ^ Zhou, Larissa; Nyberg, Kendra; Rowat, Amy C. (1 September 2015). "Understanding diffusion theory and Fick's law through food and cooking". Advances in Physiology Education. 39 (3): 192–197. doi:10.1152/advan.00133.2014. ISSN 1043-4046. PMID 26330037. S2CID 3921833.

General and cited references

  • Smith, W. F. (2004). Foundations of Materials Science and Engineering (3rd ed.). McGraw-Hill.
  • Berg, H. C. (1977). Random Walks in Biology. Princeton.
  • Bird, R. B.; Stewart, W. E.; Lightfoot, E. N. (1976). Transport Phenomena. John Wiley & Sons.
  • Crank, J. (1980). The Mathematics of Diffusion. Oxford University Press.
  • Bokshtein, B. S.; Mendelev, M. I.; Srolovitz, D. J., eds. (2005). Thermodynamics and Kinetics in Materials Science: A Short Course. Oxford: Oxford University Press. pp. 167–171.
  • Fick, A. (1855). "On liquid diffusion". Annalen der Physik und Chemie. 94: 59. – reprinted in Fick, Adolph (1995). "On liquid diffusion". Journal of Membrane Science. 100: 33–38. doi:10.1016/0376-7388(94)00230-v.

External links

  • Fick's equations, Boltzmann's transformation, etc. (with figures and animations)
  • Fick's Second Law on OpenStax

fick, laws, diffusion, technique, measuring, cardiac, output, fick, principle, describe, diffusion, were, derived, adolf, fick, 1855, they, used, solve, diffusion, coefficient, fick, first, used, derive, second, which, turn, identical, diffusion, equation, mol. For the technique of measuring cardiac output see Fick principle Fick s laws of diffusion describe diffusion and were derived by Adolf Fick in 1855 1 They can be used to solve for the diffusion coefficient D Fick s first law can be used to derive his second law which in turn is identical to the diffusion equation Molecular diffusion from a microscopic and macroscopic point of view Initially there are solute molecules on the left side of a barrier purple line and none on the right The barrier is removed and the solute diffuses to fill the whole container Top A single molecule moves around randomly Middle With more molecules there is a clear trend where the solute fills the container more and more uniformly Bottom With an enormous number of solute molecules randomness becomes undetectable The solute appears to move smoothly and systematically from high concentration areas to low concentration areas This smooth flow is described by Fick s laws A diffusion process that obeys Fick s laws is called normal or Fickian diffusion otherwise it is called anomalous diffusion or non Fickian diffusion Contents 1 History 2 Fick s first law 2 1 Alternative formulations of the first law 2 2 Derivation of Fick s first law for gases 3 Fick s second law 3 1 Derivation of Fick s second law 4 Example solutions and generalization 4 1 Example solution 1 constant concentration source and diffusion length 4 2 Example solution 2 Brownian particle and Mean squared displacement 4 3 Generalizations 5 Applications 5 1 Fick s flow in liquids 5 2 Sorption rate and collision frequency of diluted solute 5 3 Biological perspective 5 4 Semiconductor fabrication applications 5 4 1 CVD method of fabricate semiconductor 5 5 Food production and cooking 6 See also 7 Citations 8 General and cited references 9 External linksHistory EditIn 1855 physiologist Adolf Fick first reported 1 his now well known laws governing the transport of mass through diffusive means Fick s work was inspired by the earlier experiments of Thomas Graham which fell short of proposing the fundamental laws for which Fick would become famous Fick s law is analogous to the relationships discovered at the same epoch by other eminent scientists Darcy s law hydraulic flow Ohm s law charge transport and Fourier s Law heat transport Fick s experiments modeled on Graham s dealt with measuring the concentrations and fluxes of salt diffusing between two reservoirs through tubes of water It is notable that Fick s work primarily concerned diffusion in fluids because at the time diffusion in solids was not considered generally possible 2 Today Fick s Laws form the core of our understanding of diffusion in solids liquids and gases in the absence of bulk fluid motion in the latter two cases When a diffusion process does not follow Fick s laws which happens in cases of diffusion through porous media and diffusion of swelling penetrants among others 3 4 it is referred to as non Fickian Fick s first law EditFick s first law relates the diffusive flux to the gradient of the concentration It postulates that the flux goes from regions of high concentration to regions of low concentration with a magnitude that is proportional to the concentration gradient spatial derivative or in simplistic terms the concept that a solute will move from a region of high concentration to a region of low concentration across a concentration gradient In one spatial dimension the law can be written in various forms where the most common form see 5 6 is in a molar basis J D d f d x displaystyle J D frac d varphi dx where J is the diffusion flux of which the dimension is the amount of substance per unit area per unit time J measures the amount of substance that will flow through a unit area during a unit time interval D is the diffusion coefficient or diffusivity Its dimension is area per unit time f for ideal mixtures is the concentration of which the dimension is the amount of substance per unit volume x is position the dimension of which is length D is proportional to the squared velocity of the diffusing particles which depends on the temperature viscosity of the fluid and the size of the particles according to the Stokes Einstein relation In dilute aqueous solutions the diffusion coefficients of most ions are similar and have values that at room temperature are in the range of 0 6 2 10 9 m2 s For biological molecules the diffusion coefficients normally range from 10 10 to 10 11 m2 s In two or more dimensions we must use the del or gradient operator which generalises the first derivative obtaining J D f displaystyle mathbf J D nabla varphi where J denotes the diffusion flux vector The driving force for the one dimensional diffusion is the quantity f x which for ideal mixtures is the concentration gradient Alternative formulations of the first law Edit Another form for the first law is to write it with the primary variable as mass fraction yi given for example in kg kg then the equation changes to J i r D M i y i displaystyle mathbf J i frac rho D M i nabla y i where the index i denotes the i th species Ji is the diffusion flux vector of the i th species for example in mol m2 s Mi is the molar mass of the i th species and r is the mixture density for example in kg m3 Note that the r displaystyle rho is outside the gradient operator This is because y i r s i r displaystyle y i frac rho si rho where rsi is the partial density of the i th species Beyond this in chemical systems other than ideal solutions or mixtures the driving force for diffusion of each species is the gradient of chemical potential of this species Then Fick s first law one dimensional case can be written J i D c i R T m i x displaystyle J i frac Dc i RT frac partial mu i partial x where the index i denotes the i th species c is the concentration mol m3 R is the universal gas constant J K mol T is the absolute temperature K m is the chemical potential J mol The driving force of Fick s law can be expressed as a fugacity difference J i D R T f i x displaystyle J i frac D RT frac partial f i partial x Fugacity f i displaystyle f i has Pa units f i displaystyle f i is a partial pressure of component i in a vapor f i G displaystyle f i G or liquid f i L displaystyle f i L phase At vapor liquid equilibrium the evaporation flux is zero because f i G f i L displaystyle f i G f i L Derivation of Fick s first law for gases Edit Four versions of Fick s law for binary gas mixtures are given below These assume thermal diffusion is negligible the body force per unit mass is the same on both species and either pressure is constant or both species have the same molar mass Under these conditions Ref 7 shows in detail how the diffusion equation from the kinetic theory of gases reduces to this version of Fick s law V i D ln y i displaystyle mathbf V i D nabla ln y i where Vi is the diffusion velocity of species i In terms of species flux this isJ i r D M i y i displaystyle mathbf J i frac rho D M i nabla y i If additionally r 0 displaystyle nabla rho 0 this reduces to the most common form of Fick s law J i D f displaystyle mathbf J i D nabla varphi If instead of or in addition to r 0 displaystyle nabla rho 0 both species have the same molar mass Fick s law becomesJ i r D M i x i displaystyle mathbf J i frac rho D M i nabla x i where x i displaystyle x i is the mole fraction of species i Fick s second law EditFick s second law predicts how diffusion causes the concentration to change with respect to time It is a partial differential equation which in one dimension reads f t D 2 f x 2 displaystyle frac partial varphi partial t D frac partial 2 varphi partial x 2 where f is the concentration in dimensions of amount of substance length 3 example mol m3 f f x t is a function that depends on location x and time t t is time example s D is the diffusion coefficient in dimensions of length2 time 1 example m2 s x is the position length example mIn two or more dimensions we must use the Laplacian D 2 which generalises the second derivative obtaining the equation f t D D f displaystyle frac partial varphi partial t D Delta varphi Fick s second law has the same mathematical form as the Heat equation and its fundamental solution is the same as the Heat kernel except switching thermal conductivity k displaystyle k with diffusion coefficient D displaystyle D f x t 1 4 p D t exp x 2 4 D t displaystyle varphi x t frac 1 sqrt 4 pi Dt exp left frac x 2 4Dt right Derivation of Fick s second law Edit Fick s second law can be derived from Fick s first law and the mass conservation in absence of any chemical reactions f t x J 0 f t x D x f 0 displaystyle frac partial varphi partial t frac partial partial x J 0 Rightarrow frac partial varphi partial t frac partial partial x left D frac partial partial x varphi right 0 Assuming the diffusion coefficient D to be a constant one can exchange the orders of the differentiation and multiply by the constant x D x f D x x f D 2 f x 2 displaystyle frac partial partial x left D frac partial partial x varphi right D frac partial partial x frac partial partial x varphi D frac partial 2 varphi partial x 2 and thus receive the form of the Fick s equations as was stated above For the case of diffusion in two or more dimensions Fick s second law becomes f t D 2 f displaystyle frac partial varphi partial t D nabla 2 varphi which is analogous to the heat equation If the diffusion coefficient is not a constant but depends upon the coordinate or concentration Fick s second law yields f t D f displaystyle frac partial varphi partial t nabla cdot D nabla varphi An important example is the case where f is at a steady state i e the concentration does not change by time so that the left part of the above equation is identically zero In one dimension with constant D the solution for the concentration will be a linear change of concentrations along x In two or more dimensions we obtain 2 f 0 displaystyle nabla 2 varphi 0 which is Laplace s equation the solutions to which are referred to by mathematicians as harmonic functions Example solutions and generalization EditFick s second law is a special case of the convection diffusion equation in which there is no advective flux and no net volumetric source It can be derived from the continuity equation f t j R displaystyle frac partial varphi partial t nabla cdot mathbf j R where j is the total flux and R is a net volumetric source for f The only source of flux in this situation is assumed to be diffusive flux j diffusion D f displaystyle mathbf j text diffusion D nabla varphi Plugging the definition of diffusive flux to the continuity equation and assuming there is no source R 0 we arrive at Fick s second law f t D 2 f x 2 displaystyle frac partial varphi partial t D frac partial 2 varphi partial x 2 If flux were the result of both diffusive flux and advective flux the convection diffusion equation is the result Example solution 1 constant concentration source and diffusion length Edit A simple case of diffusion with time t in one dimension taken as the x axis from a boundary located at position x 0 where the concentration is maintained at a value n0 is n x t n 0 erfc x 2 D t displaystyle n left x t right n 0 operatorname erfc left frac x 2 sqrt Dt right where erfc is the complementary error function This is the case when corrosive gases diffuse through the oxidative layer towards the metal surface if we assume that concentration of gases in the environment is constant and the diffusion space that is the corrosion product layer is semi infinite starting at 0 at the surface and spreading infinitely deep in the material If in its turn the diffusion space is infinite lasting both through the layer with n x 0 0 x gt 0 and that with n x 0 n0 x 0 then the solution is amended only with coefficient 1 2 in front of n0 as the diffusion now occurs in both directions This case is valid when some solution with concentration n0 is put in contact with a layer of pure solvent Bokstein 2005 The length 2 Dt is called the diffusion length and provides a measure of how far the concentration has propagated in the x direction by diffusion in time t Bird 1976 As a quick approximation of the error function the first two terms of the Taylor series can be used n x t n 0 1 2 x 2 D t p displaystyle n x t n 0 left 1 2 left frac x 2 sqrt Dt pi right right If D is time dependent the diffusion length becomes 2 0 t D t d t displaystyle 2 sqrt int 0 t D tau d tau This idea is useful for estimating a diffusion length over a heating and cooling cycle where D varies with temperature Example solution 2 Brownian particle and Mean squared displacement Edit Another simple case of diffusion is the Brownian motion of one particle The particle s Mean squared displacement from its original position is MSD x x 0 2 2 n D t displaystyle text MSD equiv langle mathbf x mathbf x 0 2 rangle 2nDt where n displaystyle n is the dimension of the particle s Brownian motion For example the diffusion of a molecule across a cell membrane 8 nm thick is 1 D diffusion because of the spherical symmetry However the diffusion of a molecule from the membrane to the center of a eukaryotic cell is a 3 D diffusion For a cylindrical cactus the diffusion from photosynthetic cells on its surface to its center the axis of its cylindrical symmetry is a 2 D diffusion The square root of MSD 2 n D t displaystyle sqrt 2nDt is often used as a characterization of how far has the particle moved after time t displaystyle t has elapsed The MSD is symmetrically distributed over the 1D 2D and 3D space Thus the probability distribution of the magnitude of MSD in 1D is Gaussian and 3D is a Maxwell Boltzmann distribution Generalizations Edit In non homogeneous media the diffusion coefficient varies in space D D x This dependence does not affect Fick s first law but the second law changes f x t t D x f x t D x D f x t i 1 3 D x x i f x t x i displaystyle frac partial varphi x t partial t nabla cdot bigl D x nabla varphi x t bigr D x Delta varphi x t sum i 1 3 frac partial D x partial x i frac partial varphi x t partial x i In anisotropic media the diffusion coefficient depends on the direction It is a symmetric tensor Dji Dij Fick s first law changes to J D f displaystyle J D nabla varphi it is the product of a tensor and a vector J i j 1 3 D i j f x j displaystyle J i sum j 1 3 D ij frac partial varphi partial x j For the diffusion equation this formula gives f x t t D f x t i 1 3 j 1 3 D i j 2 f x t x i x j displaystyle frac partial varphi x t partial t nabla cdot bigl D nabla varphi x t bigr sum i 1 3 sum j 1 3 D ij frac partial 2 varphi x t partial x i partial x j The symmetric matrix of diffusion coefficients Dij should be positive definite It is needed to make the right hand side operator elliptic For inhomogeneous anisotropic media these two forms of the diffusion equation should be combined in f x t t D x f x t i j 1 3 D i j x 2 f x t x i x j D i j x x i f x t x i displaystyle frac partial varphi x t partial t nabla cdot bigl D x nabla varphi x t bigr sum i j 1 3 left D ij x frac partial 2 varphi x t partial x i partial x j frac partial D ij x partial x i frac partial varphi x t partial x i right The approach based on Einstein s mobility and Teorell formula gives the following generalization of Fick s equation for the multicomponent diffusion of the perfect components f i t j D i j f i f j f j displaystyle frac partial varphi i partial t sum j nabla cdot left D ij frac varphi i varphi j nabla varphi j right where fi are concentrations of the components and Dij is the matrix of coefficients Here indices i and j are related to the various components and not to the space coordinates The Chapman Enskog formulae for diffusion in gases include exactly the same terms These physical models of diffusion are different from the test models tfi Sj Dij Dfj which are valid for very small deviations from the uniform equilibrium Earlier such terms were introduced in the Maxwell Stefan diffusion equation For anisotropic multicomponent diffusion coefficients one needs a rank four tensor for example Dij ab where i j refer to the components and a b 1 2 3 correspond to the space coordinates Applications EditEquations based on Fick s law have been commonly used to model transport processes in foods neurons biopolymers pharmaceuticals porous soils population dynamics nuclear materials plasma physics and semiconductor doping processes The theory of voltammetric methods is based on solutions of Fick s equation On the other hand in some cases a Fickian another common approximation of the transport equation is that of the diffusion theory 8 description is inadequate For example in polymer science and food science a more general approach is required to describe transport of components in materials undergoing a glass transition One more general framework is the Maxwell Stefan diffusion equations 9 of multi component mass transfer from which Fick s law can be obtained as a limiting case when the mixture is extremely dilute and every chemical species is interacting only with the bulk mixture and not with other species To account for the presence of multiple species in a non dilute mixture several variations of the Maxwell Stefan equations are used See also non diagonal coupled transport processes Onsager relationship Fick s flow in liquids Edit When two miscible liquids are brought into contact and diffusion takes place the macroscopic or average concentration evolves following Fick s law On a mesoscopic scale that is between the macroscopic scale described by Fick s law and molecular scale where molecular random walks take place fluctuations cannot be neglected Such situations can be successfully modeled with Landau Lifshitz fluctuating hydrodynamics In this theoretical framework diffusion is due to fluctuations whose dimensions range from the molecular scale to the macroscopic scale 10 In particular fluctuating hydrodynamic equations include a Fick s flow term with a given diffusion coefficient along with hydrodynamics equations and stochastic terms describing fluctuations When calculating the fluctuations with a perturbative approach the zero order approximation is Fick s law The first order gives the fluctuations and it comes out that fluctuations contribute to diffusion This represents somehow a tautology since the phenomena described by a lower order approximation is the result of a higher approximation this problem is solved only by renormalizing the fluctuating hydrodynamics equations Sorption rate and collision frequency of diluted solute Edit Scheme of molecular diffusion in the solution Orange dots are solute molecules solvent molecules are not drawn black arrow is an example random walk trajectory and the red curve is the diffusive Gaussian broadening probability function from the Fick s law of diffusion 11 Fig 9 The adsorption or absorption rate of a dilute solute to a surface or interface in a gas or liquid solution can be calculated using Fick s laws of diffusion The accumulated number of molecules adsorbed on the surface is expressed by the Langmuir Schaefer equation at the short time limit by integrating the diffusion flux equation over time 12 G 2 A C D t p displaystyle Gamma 2AC sqrt frac Dt pi G displaystyle Gamma is number of molecules in unit molecules adsorbed during the time t displaystyle t A is the surface area in unit m 2 displaystyle m 2 C is the number concentration of the adsorber molecules in the bulk solution in unit molecules m 3 displaystyle m 3 D is diffusion coefficient of the adsorber in unit m 2 s displaystyle m 2 s t is elapsed time in unit s displaystyle s The equation is named after American chemists Irving Langmuir and Vincent Schaefer The Langmuir Schaefer equation can be extended to the Ward Tordai Equation to account for the back diffusion of rejected molecules from the surface 13 G 2 A C D t p A D p 0 t C b t t t d t displaystyle Gamma 2AC sqrt frac Dt pi A sqrt frac D pi int 0 sqrt t frac C b tau sqrt t tau d tau where C displaystyle C is the bulk concentration C b displaystyle C b is the sub surface concentration which is a function of time depending on the reaction model of the adsorption and t displaystyle tau is a dummy variable Monte Carlo simulations show that these two equations work to predict the adsorption rate of systems that form predictable concentration gradients near the surface but have troubles for systems without or with unpredictable concentration gradients such as typical biosensing systems or when flow and convection are significant 14 A brief history of the theories on diffusive adsorption 14 A brief history of diffusive adsorption is shown in the right figure 14 A noticeable challenge of understanding the diffusive adsorption at the single molecule level is the fractal nature of diffusion Most computer simulations pick a time step for diffusion which ignores the fact that there are self similar finer diffusion events fractal within each step Simulating the fractal diffusion shows that a factor of two corrections should be introduced for the result of a fixed time step adsorption simulation bringing it to be consistent with the above two equations 14 In the ultrashort time limit in the order of the diffusion time a2 D where a is the particle radius the diffusion is described by the Langevin equation At a longer time the Langevin equation merges into the Stokes Einstein equation The latter is appropriate for the condition of the diluted solution where long range diffusion is considered According to the fluctuation dissipation theorem based on the Langevin equation in the long time limit and when the particle is significantly denser than the surrounding fluid the time dependent diffusion constant is 15 D t m k B T 1 e t m m displaystyle D t mu k rm B T left 1 e t m mu right where all in SI units kB is Boltzmann s constant T is the absolute temperature m is the mobility of the particle in the fluid or gas which can be calculated using the Einstein relation kinetic theory m is the mass of the particle t is time For a single molecule such as organic molecules or biomolecules e g proteins in water the exponential term is negligible due to the small product of mm in the picosecond region When the area of interest is the size of a molecule specifically a long cylindrical molecule such as DNA the adsorption rate equation represents the collision frequency of two molecules in a diluted solution with one molecule a specific side and the other no steric dependence i e a molecule random orientation hit one side of the other The diffusion constant need to be updated to the relative diffusion constant between two diffusing molecules This estimation is especially useful in studying the interaction between a small molecule and a larger molecule such as a protein The effective diffusion constant is dominated by the smaller one whose diffusion constant can be used instead The above hitting rate equation is also useful to predict the kinetics of molecular self assembly on a surface Molecules are randomly oriented in the bulk solution Assuming 1 6 of the molecules has the right orientation to the surface binding sites i e 1 2 of the z direction in x y z three dimensions thus the concentration of interest is just 1 6 of the bulk concentration Put this value into the equation one should be able to calculate the theoretical adsorption kinetic curve using the Langmuir adsorption model In a more rigid picture 1 6 can be replaced by the steric factor of the binding geometry Biological perspective Edit The first law gives rise to the following formula 16 flux P c 2 c 1 displaystyle text flux P left c 2 c 1 right in which P is the permeability an experimentally determined membrane conductance for a given gas at a given temperature c2 c1 is the difference in concentration of the gas across the membrane for the direction of flow from c1 to c2 Fick s first law is also important in radiation transfer equations However in this context it becomes inaccurate when the diffusion constant is low and the radiation becomes limited by the speed of light rather than by the resistance of the material the radiation is flowing through In this situation one can use a flux limiter The exchange rate of a gas across a fluid membrane can be determined by using this law together with Graham s law Under the condition of a diluted solution when diffusion takes control the membrane permeability mentioned in the above section can be theoretically calculated for the solute using the equation mentioned in the last section use with particular care because the equation is derived for dense solutes while biological molecules are not denser than water 11 P 2 A p h t m D p t displaystyle P 2A p eta tm sqrt D pi t where A P displaystyle A P is the total area of the pores on the membrane unit m2 h t m displaystyle eta tm transmembrane efficiency unitless which can be calculated from the stochastic theory of chromatography D is the diffusion constant of the solute unit m2s 1 t is time unit s c2 c1 concentration should use unit mol m 3 so flux unit becomes mol s 1 The flux is decay over the square root of time because a concentration gradient builds up near the membrane over time under ideal conditions When there is flow and convection the flux can be significantly different than the equation predicts and show an effective time t with a fixed value 14 which makes the flux stable instead of decay over time This strategy is adopted in biology such as blood circulation Semiconductor fabrication applications Edit The semiconductor is a collective term for a series of devices It mainly includes three categories two terminal devices three terminal devices and four terminal devices The combination of the semiconductors is called an integrated circuit The relationship between Fick s law and semiconductors the principle of the semiconductor is transferring chemicals or dopants from a layer to a layer Fick s law can be used to control and predict the diffusion by knowing how much the concentration of the dopants or chemicals move per meter and second through mathematics Therefore different types and levels of semiconductors can be fabricated Integrated circuit fabrication technologies model processes like CVD thermal oxidation wet oxidation doping etc use diffusion equations obtained from Fick s law CVD method of fabricate semiconductor Edit The wafer is a kind of semiconductor whose silicon substrate is coated with a layer of CVD created polymer chain and films This film contains n type and p type dopants and takes responsibility for dopant conductions The principle of CVD relies on the gas phase and gas solid chemical reaction to create thin films The viscous flow regime of CVD is driven by a pressure gradient CVD also includes a diffusion component distinct from the surface diffusion of adatoms In CVD reactants and products must also diffuse through a boundary layer of stagnant gas that exists next to the substrate The total number of steps required for CVD film growth are gas phase diffusion of reactants through the boundary layer adsorption and surface diffusion of adatoms reactions on the substrate and gas phase diffusion of products away through the boundary layer The velocity profile for gas flow is d x 5 x R e 1 2 R e v r L h displaystyle delta x left frac 5x mathrm Re 1 2 right mathrm Re frac v rho L eta where d displaystyle delta is the thickness R e displaystyle mathrm Re is the Reynolds number x is the length of the subtrate v 0 at any surface h displaystyle eta is viscosity r displaystyle rho is density Integrated the x from 0 to L it gives the average thickness d 10 L 3 R e 1 2 displaystyle delta frac 10L 3 mathrm Re 1 2 To keep the reaction balanced reactants must diffuse through the stagnant boundary layer to reach the substrate So a thin boundary layer is desirable According to the equations increasing vo would result in more wasted reactants The reactants will not reach the substrate uniformly if the flow becomes turbulent Another option is to switch to a new carrier gas with lower viscosity or density The Fick s first law describes diffusion through the boundary layer As a function of pressure P and temperature T in a gas diffusion is determined D D 0 P 0 P T T 0 3 2 displaystyle D D 0 left frac P 0 P right left frac T T 0 right 3 2 where P 0 displaystyle P 0 is the standard pressure T 0 displaystyle T 0 is the standard temperature D 0 displaystyle D 0 is the standard diffusitivity The equation tells that increasing the temperature or decreasing the pressure can increase the diffusivity Fick s first law predicts the flux of the reactants to the substrate and product away from the substrate J D i d c i d x displaystyle J D i left frac dc i dx right where x displaystyle x is the thickness d displaystyle delta d c i displaystyle dc i is the first reactant s concentration In ideal gas law P V n R T displaystyle PV nRT the concentration of the gas is expressed by partial pressure J D i P i P 0 d R T displaystyle J D i left frac P i P 0 delta RT right where R displaystyle R is the gas constant P i P 0 d displaystyle frac P i P 0 delta is the partial pressure gradient As a result Fick s first law tells us we can use a partial pressure gradient to control the diffusivity and control the growth of thin films of semiconductors In many realistic situations the simple Fick s law is not an adequate formulation for the semiconductor problem It only applies to certain conditions for example given the semiconductor boundary conditions constant source concentration diffusion limited source concentration or moving boundary diffusion where junction depth keeps moving into the substrate Food production and cooking Edit The formulation of Fick s first law can explain a variety of complex phenomena in the context of food and cooking Diffusion of molecules such as ethylene promotes plant growth and ripening salt and sugar molecules promotes meat brining and marinating and water molecules promote dehydration Fick s first law can also be used to predict the changing moisture profiles across a spaghetti noodle as it hydrates during cooking These phenomena are all about the spontaneous movement of particles of solutes driven by the concentration gradient In different situations there is different diffusivity which is a constant 17 By controlling the concentration gradient the cooking time shape of the food and salting can be controlled See also EditAdvection Churchill Bernstein equation Diffusion False diffusion Gas exchange Mass flux Maxwell Stefan diffusion Nernst Planck equation OsmosisCitations Edit a b Fick A 1855 Ueber Diffusion Annalen der Physik in German 94 1 59 86 Bibcode 1855AnP 170 59F doi 10 1002 andp 18551700105 Fick A 1855 V On liquid diffusion Phil Mag 10 63 30 39 doi 10 1080 14786445508641925 Philibert Jean 2005 One and a Half Centuries of Diffusion Fick Einstein before and beyond PDF Diffusion Fundamentals 2 1 1 1 10 Archived from the original PDF on 5 February 2009 Vazquez J L 2006 The Porous Medium Equation Mathematical Theory Oxford Univ Press Gorban A N Sargsyan H P Wahab H A 2011 Quasichemical Models of Multicomponent Nonlinear Diffusion Mathematical Modelling of Natural Phenomena 6 5 184 262 arXiv 1012 2908 doi 10 1051 mmnp 20116509 S2CID 18961678 Atkins Peter de Paula Julio 2006 Physical Chemistry for the Life Science Conlisk A Terrence 2013 Essentials of Micro and Nanofluidics With Applications to the Biological and Chemical Sciences Cambridge University Press p 43 ISBN 9780521881685 Williams F A 1985 Appendix E Combustion Theory Benjamin Cummings Fickian Diffusion an overview ScienceDirect Topics www sciencedirect com Retrieved 11 May 2022 Taylor Ross Krishna R 1993 Multicomponent mass transfer Wiley a href Template Cite journal html title Template Cite journal cite journal a Cite journal requires journal help Brogioli D Vailati A 2001 Diffusive mass transfer by nonequilibrium fluctuations Fick s law revisited Phys Rev E 63 1 4 012105 arXiv cond mat 0006163 Bibcode 2000PhRvE 63a2105B doi 10 1103 PhysRevE 63 012105 PMID 11304296 S2CID 1302913 a b Pyle Joseph R Chen Jixin 2 November 2017 Photobleaching of YOYO 1 in super resolution single DNA fluorescence imaging Beilstein Journal of Nanotechnology 8 2292 2306 doi 10 3762 bjnano 8 229 PMC 5687005 PMID 29181286 Langmuir I Schaefer V J 1937 The Effect of Dissolved Salts on Insoluble Monolayers Journal of the American Chemical Society 29 11 2400 2414 doi 10 1021 ja01290a091 Ward A F H Tordai L 1946 Time dependence of Boundary Tensions of Solutions I The Role of Diffusion in Time effects Journal of Chemical Physics 14 7 453 461 Bibcode 1946JChPh 14 453W doi 10 1063 1 1724167 a b c d e Chen J 2022 Simulating stochastic adsorption of diluted solute molecules at interfaces AIP Advances 12 1 015318 Bibcode 2022AIPA 12a5318C doi 10 1063 5 0064140 PMC 8758205 PMID 35070490 Bian Xin Kim Changho Karniadakis George Em 14 August 2016 111 years of Brownian motion Soft Matter 12 30 6331 6346 Bibcode 2016SMat 12 6331B doi 10 1039 c6sm01153e PMC 5476231 PMID 27396746 Nosek Thomas M Section 3 3ch9 s3ch9 2 Essentials of Human Physiology Archived from the original on 24 March 2016 Zhou Larissa Nyberg Kendra Rowat Amy C 1 September 2015 Understanding diffusion theory and Fick s law through food and cooking Advances in Physiology Education 39 3 192 197 doi 10 1152 advan 00133 2014 ISSN 1043 4046 PMID 26330037 S2CID 3921833 General and cited references EditSmith W F 2004 Foundations of Materials Science and Engineering 3rd ed McGraw Hill Berg H C 1977 Random Walks in Biology Princeton Bird R B Stewart W E Lightfoot E N 1976 Transport Phenomena John Wiley amp Sons Crank J 1980 The Mathematics of Diffusion Oxford University Press Bokshtein B S Mendelev M I Srolovitz D J eds 2005 Thermodynamics and Kinetics in Materials Science A Short Course Oxford Oxford University Press pp 167 171 Fick A 1855 On liquid diffusion Annalen der Physik und Chemie 94 59 reprinted in Fick Adolph 1995 On liquid diffusion Journal of Membrane Science 100 33 38 doi 10 1016 0376 7388 94 00230 v External links EditFick s equations Boltzmann s transformation etc with figures and animations Fick s Second Law on OpenStax Retrieved from https en wikipedia org w index php title Fick 27s laws of diffusion amp oldid 1133949658, 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.