fbpx
Wikipedia

Classical nucleation theory

Classical nucleation theory (CNT) is the most common theoretical model used to quantitatively study the kinetics of nucleation.[1][2][3][4]

Nucleation is the first step in the spontaneous formation of a new thermodynamic phase or a new structure, starting from a state of metastability. The kinetics of formation of the new phase is frequently dominated by nucleation, such that the time to nucleate determines how long it will take for the new phase to appear. The time to nucleate can vary by orders of magnitude, from negligible to exceedingly large, far beyond reach of experimental timescales. One of the key achievements of classical nucleation theory is to explain and quantify this immense variation.[5]

Description Edit

The central result of classical nucleation theory is a prediction for the rate of nucleation  , in units of (number of events)/(volume·time). For instance, a rate   in a supersaturated vapor would correspond to an average of 1000 droplets nucleating in a volume of 1 cubic meter in 1 second.

The CNT prediction for   is[3]

 

where

  •   is the free energy cost of the nucleus at the top of the nucleation barrier, and   is the average thermal energy with   the absolute temperature and   the Boltzmann constant.
  •   is the number of nucleation sites.
  •   is the rate at which molecules attach to the nucleus.
  •   is the Zeldovich factor, (named after Yakov Zeldovich[6]) which gives the probability that a nucleus at the top of the barrier will go on to form the new phase, rather than dissolve.

This expression for the rate can be thought of as a product of two factors: the first,  , is the number of nucleation sites multiplied by the probability that a nucleus of critical size has grown around it. It can be interpreted as the average, instantaneous number of nuclei at the top of the nucleation barrier. Free energies and probabilities are closely related by definition.[7] The probability of a nucleus forming at a site is proportional to  . So if   is large and positive the probability of forming a nucleus is very low and nucleation will be slow. Then the average number will be much less than one, i.e., it is likely that at any given time none of the sites has a nucleus.

The second factor in the expression for the rate is the dynamic part,  . Here,   expresses the rate of incoming matter and   is the probability that a nucleus of critical size (at the maximum of the energy barrier) will continue to grow and not dissolve. The Zeldovich factor is derived by assuming that the nuclei near the top of the barrier are effectively diffusing along the radial axis. By statistical fluctuations, a nucleus at the top of the barrier can grow diffusively into a larger nucleus that will grow into a new phase, or it can lose molecules and shrink back to nothing. The probability that a given nucleus goes forward is  .

Taking into consideration kinetic theory and assuming that there is the same transition probability in each direction, it is known that  . As   determines the hopping rate, the previous formula can be rewritten in terms of the mean free path and the mean free time  . Consequently, a relation of   in terms of the diffusion coefficient is obtained

 .

Further considerations can be made in order to study temperature dependence. Therefore, Einstein-Stokes relation is introduced under the consideration of a spherical shape

 , where   is the material's viscosity.

Considering the last two expressions, it is seen that    . If  , being   the melting temperature, the ensemble gains high velocity and makes   and   to increase and hence,   decreases. If  , the ensemble has a low mobility, which makes   to decrease as well.

To see how this works in practice we can look at an example. Sanz and coworkers[8] have used computer simulation to estimate all the quantities in the above equation, for the nucleation of ice in liquid water. They did this for a simple but approximate model of water called TIP4P/2005. At a supercooling of 19.5 °C, i.e., 19.5 °C below the freezing point of water in their model, they estimate a free energy barrier to nucleation of ice of  . They also estimate a rate of addition of water molecules to an ice nucleus near the top of the barrier of   and a Zeldovich factor  . The number of water molecules in 1 m3 of water is approximately 1028. These leads to the prediction  , which means that on average one would have to wait 1083s (1076 years) to see a single ice nucleus forming in 1 m3 of water at -20 °C!

This is a rate of homogeneous nucleation estimated for a model of water, not real water — in experiments one cannot grow nuclei of water and so cannot directly determine the values of the barrier  , or the dynamic parameters such as  , for real water. However, it may be that indeed the homogeneous nucleation of ice at temperatures near -20 °C and above is extremely slow and so that whenever water freezes at temperatures of -20 °C and above this is due to heterogeneous nucleation, i.e., the ice nucleates in contact with a surface.

Homogeneous nucleation Edit

Homogeneous nucleation is much rarer than heterogeneous nucleation.[1][9] However, homogeneous nucleation is simpler and easier to understand than heterogeneous nucleation, so the easiest way to understand heterogeneous nucleation is to start with homogeneous nucleation. So we will outline the CNT calculation for the homogeneous nucleation barrier  .

 
The green curve is the total (Gibbs if this is at constant pressure) free energy as a function of radius. Shown is the free energy barrier,  , and radius at the top of the barrier,  . This total free energy is a sum of two terms. The first is a bulk term, which is plotted in red. This scales with volume and is always negative. The second term is an interfacial term, which is plotted in black. This is the origin of the barrier. It is always positive and scales with surface area.

To understand if nucleation is fast or slow,  , the Gibbs free energy change as a function of the size of the nucleus, needs to be calculated. The classical theory[10] assumes that even for a microscopic nucleus of the new phase, we can write the free energy of a droplet   as the sum of a bulk term that is proportional to the volume of the nucleus, and a surface term, that is proportional to its surface area

 

The first term is the volume term, and as we are assuming that the nucleus is spherical, this is the volume of a sphere of radius  .   is the difference in free energy per unit of volume between the phase that nucleates and the thermodynamic phase nucleation is occurring in. For example, if water is nucleating in supersaturated air, then   is the free energy per unit of volume of water minus that of supersaturated air at the same pressure. As nucleation only occurs when the air is supersaturated,   is always negative. The second term comes from the interface at surface of the nucleus, which is why it is proportional to the surface area of a sphere.   is the surface tension of the interface between the nucleus and its surroundings, which is always positive. In case of nucleation in a solid matrix, there is a third energy component in addition to the two mentioned above. This third energy appears as the strain that is caused by the density difference between the product and the parent phase, and is positive (unfavorable for nucleation). [11]

For small   the second surface term dominates and  . The free energy is the sum of an   and   terms. Now the   terms varies more rapidly with   than the   term, so as small   the   term dominates and the free energy is positive while for large  , the   term dominates and the free energy is negative. This shown in the figure to the right. Thus at some intermediate value of  , the free energy goes through a maximum, and so the probability of formation of a nucleus goes through a minimum. There is a least-probable nucleus size, i.e., the one with the highest value of  

where

 

Addition of new molecules to nuclei larger than this critical radius,  , decreases the free energy, so these nuclei are more probable. The rate at which nucleation occurs is then limited by, i.e., determined by the probability, of forming the critical nucleus. This is just the exponential of minus the free energy of the critical nucleus  , which is

 

This is the free energy barrier needed in the CNT expression for   above.

In the discussion above, we assumed the growing nucleus to be three-dimensional and spherical. Similar equations can be set up for other dimensions and/or other shapes, using the appropriate expressions for the analogues of volume and surface area of the nucleus. One will then find out that any non-spherical nucleus has a higher barrier height   than the corresponding spherical nucleus. This can be understood from the fact that a sphere has the lowest possible surface area to volume ratio, thereby minimizing the (unfavourable) surface contribution with respect to the (favourable) bulk volume contribution to the free energy. Assuming equal kinetic prefactors, the fact that   is higher for non-spherical nuclei implies that their formation rate is lower. This explains why in homogeneous nucleation usually only spherical nuclei are taken into account.

From an experimental standpoint, this theory grants tuning of the critical radius through the dependence of   on temperature. The variable  , described above, can be expressed as

 

where   is the melting point and   is the enthalpy of formation for the material. Furthermore, the critical radius can be expressed as

 

revealing a dependence of reaction temperature. Thus as you increase the temperature near  , the critical radius will increase. Same happens when you move away from the melting point, the critical radius and the free energy decrease.

Heterogeneous nucleation Edit

 
Three droplets on a surface, illustrating decreasing contact angles. The contact angle the droplet surface makes with the solid horizontal surface decreases from left to right.
 
A diagram featuring all of the factors that affect heterogeneous nucleation

Unlike homogeneous nucleation, heterogeneous nucleation occurs on a surface or impurity. It is much more common than homogeneous nucleation. This is because the nucleation barrier for heterogeneous nucleation is much lower than for homogeneous nucleation. To see this, note that the nucleation barrier is determined by the positive term in the free energy  , which is proportional to the total exposed surface area of a nucleus. For homogeneous nucleation the surface area is simply that of a sphere. For heterogeneous nucleation, however, the surface area is smaller since part of the nucleus boundary is accommodated by the surface or impurity onto which it is nucleating.[12]

There are several factors which determine the precise reduction in the exposed surface area.[12] As shown in a diagram on the left, these factors include the size of the droplet, the contact angle,  , between the droplet and surface, and the interactions at the three phase interfaces: liquid-solid, solid-vapor, and liquid-vapor.

The free energy needed for heterogeneous nucleation,  , is equal to the product of homogeneous nucleation,  , and a function of the contact angle,  :

 .

The schematic to the right illustrates the decrease in the exposed surface area of the droplet as the contact angle decreases. Deviations from a flat interface decrease the exposed surface even further: there exist expressions for this reduction for simple surface geometries.[13] In practice, this means that nucleation will tend to occur on surface imperfections.

 
Difference in energy barriers

Statistical mechanical treatment Edit

The classical nucleation theory hypothesis for the form of   can be examined more rigorously using the tools of statistical mechanics.[14] Specifically, the system is modeled as a gas of non-interacting clusters in the grand canonical ensemble. A state of metastable equilibrium is assumed, such that the methods of statistical mechanics hold at least approximately.[15] The grand partition function is[16]

 

Here the inner summation is over all microstates   which contain exactly   particles. It can be decomposed into contributions from each possible combination of clusters which results in   total particles.[17] For instance,

 

where   is the configuration integral of a cluster with   particles and potential energy  :

 

The quantity   is the thermal de Broglie wavelength of the particle, which enters due to the integration over the   momentum degrees of freedom. The inverse factorials are included to compensate for overcounting, since particles and clusters alike are assumed indistinguishable.

More compactly,

 .

Then, by expanding   in powers of  , the probability   of finding exactly   clusters which each has   particles is

 

The number density   of  -clusters can therefore be calculated as

 

This is also called the cluster size distribution.

The grand potential   is equal to  , which, using the thermodynamic relationship  , leads to the following expansion for the pressure:

 

If one defines the right hand side of the above equation as the function  , then various other thermodynamic quantities can be calculated in terms of derivatives of   with respect to  .[18]

The connection with the simple version of the theory is made by assuming perfectly spherical clusters, in which case   depends only on  , in the form

 

where   is the binding energy of a single particle in the interior of a cluster, and   is the excess energy per unit area of the cluster surface. Then,  , and the cluster size distribution is

 

which implies an effective free energy landscape  , in agreement with the form proposed by the simple theory.

On the other hand, this derivation reveals the significant approximation in assuming spherical clusters with  . In reality, the configuration integral   contains contributions from the full set of particle coordinates  , thus including deviations from spherical shape as well as cluster degrees of freedom such as translation, vibration, and rotation. Various attempts have been made to include these effects in the calculation of  , although the interpretation and application of these extended theories has been debated.[5][19][20] A common feature is the addition of a logarithmic correction   to  , which plays an important role near the critical point of the fluid.[21]

Limitations Edit

Classical nucleation theory makes a number of assumptions which limit its applicability. Most fundamentally, in the so-called capillarity approximation it treats the nucleus interior as a bulk, incompressible fluid and ascribes to the nucleus surface the macroscopic interfacial tension  , even though it is not obvious that such macroscopic equilibrium properties apply to a typical nucleus of, say, 50 molecules across.[22][23] In fact, it has been shown that the effective surface tension of small droplets is smaller than that of the bulk liquid.[24]

In addition, the classical theory places restrictions on the kinetic pathways by which nucleation occurs, assuming clusters grow or shrink only by single particle adsorption/emission. In reality, merging and fragmentation of entire clusters cannot be excluded as important kinetic pathways in some systems. Particularly in dense systems or near the critical point – where clusters acquire an extended and ramified structure – such kinetic pathways are expected to contribute significantly.[24] The behavior near the critical point also suggests the inadequacy, at least in some cases, of treating clusters as purely spherical.[25]

Various attempts have been made to remedy these limitations and others by explicitly accounting for the microscopic properties of clusters. However, the validity of such extended models is debated. One difficulty is the exquisite sensitivity of the nucleation rate   to the free energy  : even small discrepancies in the microscopic parameters can lead to enormous changes in the predicted nucleation rate. This fact makes first-principles predictions nearly impossible. Instead, models must be fit directly to experimental data, which limits the ability to test their fundamental validity.[26]

Comparison with simulation and experiment Edit

For simple model systems, modern computers are powerful enough to calculate exact nucleation rates numerically. An example is the nucleation of the crystal phase in a system of hard spheres, which is a simple model of colloids consisting of perfectly hard spheres in thermal motion. The agreement of CNT with the simulated rates for this system confirms that the classical theory is a reasonable approximation.[27] For simple models CNT works quite well; however it is unclear if it describes complex (e.g., molecular) systems equally well. For example, in the context of vapor to liquid nucleation, the CNT predictions for the nucleation rate are incorrect by several orders of magnitude on an absolute scale — that is, without renormalizing with respect to experimental data. Nevertheless, certain variations on the classical theory have been claimed to represent the temperature dependence adequately, even if the absolute magnitude is inaccurate.[28] Jones et al. computationally explored the nucleation of small water clusters using a classical water model. It was found that CNT could describe the nucleation of clusters of 8-50 water molecules well, but failed to describe smaller clusters.[29] Corrections to CNT, obtained from higher accuracy methods such as quantum chemical calculations, may improve the agreement with experiment.[30]

References Edit

  1. ^ a b H. R. Pruppacher and J. D. Klett, Microphysics of Clouds and Precipitation, Kluwer (1997)
  2. ^ P.G. Debenedetti, Metastable Liquids: Concepts and Principles, Princeton University Press (1997)
  3. ^ a b Sear, R. P. (2007). "Nucleation: theory and applications to protein solutions and colloidal suspensions". J. Phys.: Condens. Matter. 19 (3): 033101. Bibcode:2007JPCM...19c3101S. CiteSeerX 10.1.1.605.2550. doi:10.1088/0953-8984/19/3/033101. S2CID 4992555.
  4. ^ Kreer, Markus (1993). "Classical Becker‐Döring cluster equations: Rigorous results on metastability and long‐time behaviour". Annalen der Physik. 505 (4): 398–417. Bibcode:1993AnP...505..398K. doi:10.1002/andp.19935050408.
  5. ^ a b Oxtoby, David W. (1992), "Homogeneous nucleation: theory and experiment", Journal of Physics: Condensed Matter, 4 (38): 7627–7650, Bibcode:1992JPCM....4.7627O, doi:10.1088/0953-8984/4/38/001, S2CID 250827558
  6. ^ Zeldovich, Y. B. (1943). On the theory of new phase formation: cavitation. Acta Physicochem., USSR, 18, 1.
  7. ^ Frenkel, Daan; Smit, Berent (2001). Understanding Molecular Simulation, Second Edition: From Algorithms to Applications. p. Academic Press. ISBN 978-0122673511.
  8. ^ Sanz, Eduardo; Vega, Carlos; Espinosa, J. R.; Cabellero-Bernal, R.; Abascal, J. L. F.; Valeriani, Chantal (2013). "Homogeneous Ice Nucleation at Moderate Supercooling from Molecular Simulation". Journal of the American Chemical Society. 135 (40): 15008–15017. arXiv:1312.0822. Bibcode:2013arXiv1312.0822S. doi:10.1021/ja4028814. PMID 24010583. S2CID 2304292.
  9. ^ Sear, Richard P. (2014). "Quantitative Studies of Crystal Nucleation at Constant Supersaturation: Experimental Data and Models". CrystEngComm. 16 (29): 6506–6522. doi:10.1039/C4CE00344F.
  10. ^ F. F. Abraham (1974) Homogeneous nucleation theory (Academic Press, NY).
  11. ^ Shirzad K.; Viney C. (2023). "A critical review on applications of the Avrami equation beyond materials science". Journal of the Royal Society Interface. 20 (203). doi:10.1098/rsif.2023.0242. PMC 10282574. PMID 37340781.
  12. ^ a b Liu, X. Y. (31 May 2000). "Heterogeneous nucleation or homogeneous nucleation?". The Journal of Chemical Physics. 112 (22): 9949–9955. Bibcode:2000JChPh.112.9949L. doi:10.1063/1.481644. ISSN 0021-9606.
  13. ^ Sholl, C. A.; N. H. Fletcher (1970). "Decoration criteria for surface steps". Acta Metall. 18 (10): 1083–1086. doi:10.1016/0001-6160(70)90006-4. hdl:1885/213299.
  14. ^ The discussion which follows draws from Kalikmanov (2001), unless noted otherwise.
  15. ^ Kalikmanov, V.I. (2013), Nucleation Theory, Lecture Notes in Physics, vol. 860, Springer Netherlands, pp. 17–19, Bibcode:2013LNP...860.....K, doi:10.1007/978-90-481-3643-8, ISBN 978-90-481-3643-8, ISSN 0075-8450
  16. ^ Kardar, Mehran (2007), Statistical Physics of Particles, Cambridge University Press, p. 118, ISBN 978-0-521-87342-0
  17. ^ Kalikmanov, V.I. (2001), Statistical Physics of Fluids: Basic Concepts and Applications, Springer-Verlag, pp. 170–172, ISBN 978-3-540-417-47-7, ISSN 0172-5998
  18. ^ Kalikmanov (2001) pp. 172-173
  19. ^ Kiang, C. S. and Stauffer, D. and Walker, G. H. and Puri, O. P. and Wise, J. D. and Patterson, E. M., C.S.; Stauffer, D.; Walker, G. H.; Puri, O.P.; Wise, J.D.; Patterson, E.M. (1971), "A Reexamination of Homogeneous Nucleation Theory", Journal of the Atmospheric Sciences, 28 (7): 1222–1232, Bibcode:1971JAtS...28.1222K, doi:10.1175/1520-0469(1971)028<1222:AROHNT>2.0.CO;2{{citation}}: CS1 maint: multiple names: authors list (link)
  20. ^ Reguera, D.; Rubí, J.M. (2001), "Nonequilibrium translational-rotational effects in nucleation", The Journal of Chemical Physics, 115 (15): 7100–7106, arXiv:cond-mat/0109270, Bibcode:2001JChPh.115.7100R, doi:10.1063/1.1405122, S2CID 95887792
  21. ^ Sator, N. (2003), "Clusters in simple fluids", Physics Reports, 376 (1): 1–39, arXiv:cond-mat/0210566, Bibcode:2003PhR...376....1S, doi:10.1016/S0370-1573(02)00583-5, ISSN 0370-1573, S2CID 119492597
  22. ^ Kalikmanov (2013), p. 21
  23. ^ Oxtoby (1992), p. 7631
  24. ^ a b Kiang, et al (1971)
  25. ^ Sator (2003)
  26. ^ Oxtoby (1992), p. 7638–7640
  27. ^ Auer, S.; D. Frenkel (2004). "Numerical prediction of absolute crystallization rates in hard-sphere colloids" (PDF). J. Chem. Phys. 120 (6): 3015–29. Bibcode:2004JChPh.120.3015A. doi:10.1063/1.1638740. hdl:1874/12074. PMID 15268449. S2CID 15747794.
  28. ^ Fladerer, Alexander; Strey, Reinhard (2006-04-28). "Homogeneous nucleation and droplet growth in supersaturated argon vapor: The cryogenic nucleation pulse chamber". The Journal of Chemical Physics. AIP Publishing. 124 (16): 164710. Bibcode:2006JChPh.124p4710F. doi:10.1063/1.2186327. ISSN 0021-9606. PMID 16674160.
  29. ^ Merikanto, Joonas; Zapadinsky, Evgeni; Lauri, Antti; Vehkamäki, Hanna (4 April 2007). "Origin of the Failure of Classical Nucleation Theory: Incorrect Description of the Smallest Clusters". Physical Review Letters. 98 (14): 145702. Bibcode:2007PhRvL..98n5702M. doi:10.1103/PhysRevLett.98.145702. PMID 17501289.
  30. ^ Temelso, Berhane; Morrell, Thomas E.; Shields, Robert M.; Allodi, Marco A.; Wood, Elena K.; Kirschner, Karl N.; Castonguay, Thomas C.; Archer, Kaye A.; Shields, George C. (22 February 2012). "Quantum Mechanical Study of Sulfuric Acid Hydration: Atmospheric Implications". The Journal of Physical Chemistry A. 116 (9): 2209–2224. Bibcode:2012JPCA..116.2209T. doi:10.1021/jp2119026. PMID 22296037.

classical, nucleation, theory, most, common, theoretical, model, used, quantitatively, study, kinetics, nucleation, nucleation, first, step, spontaneous, formation, thermodynamic, phase, structure, starting, from, state, metastability, kinetics, formation, pha. Classical nucleation theory CNT is the most common theoretical model used to quantitatively study the kinetics of nucleation 1 2 3 4 Nucleation is the first step in the spontaneous formation of a new thermodynamic phase or a new structure starting from a state of metastability The kinetics of formation of the new phase is frequently dominated by nucleation such that the time to nucleate determines how long it will take for the new phase to appear The time to nucleate can vary by orders of magnitude from negligible to exceedingly large far beyond reach of experimental timescales One of the key achievements of classical nucleation theory is to explain and quantify this immense variation 5 Contents 1 Description 1 1 Homogeneous nucleation 1 2 Heterogeneous nucleation 2 Statistical mechanical treatment 3 Limitations 4 Comparison with simulation and experiment 5 ReferencesDescription EditThe central result of classical nucleation theory is a prediction for the rate of nucleation R displaystyle R nbsp in units of number of events volume time For instance a rate R 1000 m 3 s 1 displaystyle R 1000 text m 3 text s 1 nbsp in a supersaturated vapor would correspond to an average of 1000 droplets nucleating in a volume of 1 cubic meter in 1 second The CNT prediction for R displaystyle R nbsp is 3 R N S Z j exp D G k B T displaystyle R N S Zj exp left frac Delta G k B T right nbsp dd where D G displaystyle Delta G nbsp is the free energy cost of the nucleus at the top of the nucleation barrier and k B T displaystyle k B T nbsp is the average thermal energy with T displaystyle T nbsp the absolute temperature and k B displaystyle k B nbsp the Boltzmann constant N S displaystyle N S nbsp is the number of nucleation sites j displaystyle j nbsp is the rate at which molecules attach to the nucleus Z displaystyle Z nbsp is the Zeldovich factor named after Yakov Zeldovich 6 which gives the probability that a nucleus at the top of the barrier will go on to form the new phase rather than dissolve This expression for the rate can be thought of as a product of two factors the first N S exp D G k B T displaystyle N S exp left Delta G k B T right nbsp is the number of nucleation sites multiplied by the probability that a nucleus of critical size has grown around it It can be interpreted as the average instantaneous number of nuclei at the top of the nucleation barrier Free energies and probabilities are closely related by definition 7 The probability of a nucleus forming at a site is proportional to exp D G k T displaystyle exp Delta G kT nbsp So if D G displaystyle Delta G nbsp is large and positive the probability of forming a nucleus is very low and nucleation will be slow Then the average number will be much less than one i e it is likely that at any given time none of the sites has a nucleus The second factor in the expression for the rate is the dynamic part Z j displaystyle Zj nbsp Here j displaystyle j nbsp expresses the rate of incoming matter and Z displaystyle Z nbsp is the probability that a nucleus of critical size at the maximum of the energy barrier will continue to grow and not dissolve The Zeldovich factor is derived by assuming that the nuclei near the top of the barrier are effectively diffusing along the radial axis By statistical fluctuations a nucleus at the top of the barrier can grow diffusively into a larger nucleus that will grow into a new phase or it can lose molecules and shrink back to nothing The probability that a given nucleus goes forward is Z displaystyle Z nbsp Taking into consideration kinetic theory and assuming that there is the same transition probability in each direction it is known that x 2 2 D t displaystyle x 2 2Dt nbsp As Z j displaystyle Zj nbsp determines the hopping rate the previous formula can be rewritten in terms of the mean free path and the mean free time l 2 2 D t displaystyle lambda 2 2D tau nbsp Consequently a relation of Z j displaystyle Zj nbsp in terms of the diffusion coefficient is obtainedZ j 1 t 2 D l 2 displaystyle Zj frac 1 tau frac 2D lambda 2 nbsp Further considerations can be made in order to study temperature dependence Therefore Einstein Stokes relation is introduced under the consideration of a spherical shapeD k B T 6 p h l displaystyle D frac k B T 6 pi eta lambda nbsp where h displaystyle eta nbsp is the material s viscosity Considering the last two expressions it is seen that Z j displaystyle Zj nbsp T displaystyle propto T nbsp If T T m displaystyle T approx T m nbsp being T m displaystyle T m nbsp the melting temperature the ensemble gains high velocity and makes Z j displaystyle Zj nbsp and D G displaystyle Delta G nbsp to increase and hence R displaystyle R nbsp decreases If T T m displaystyle T ll T m nbsp the ensemble has a low mobility which makes R displaystyle R nbsp to decrease as well To see how this works in practice we can look at an example Sanz and coworkers 8 have used computer simulation to estimate all the quantities in the above equation for the nucleation of ice in liquid water They did this for a simple but approximate model of water called TIP4P 2005 At a supercooling of 19 5 C i e 19 5 C below the freezing point of water in their model they estimate a free energy barrier to nucleation of ice of D G 275 k B T displaystyle Delta G 275k B T nbsp They also estimate a rate of addition of water molecules to an ice nucleus near the top of the barrier of j 10 11 s 1 displaystyle j 10 11 text s 1 nbsp and a Zeldovich factor Z 10 3 displaystyle Z 10 3 nbsp The number of water molecules in 1 m3 of water is approximately 1028 These leads to the prediction R 10 83 s 1 displaystyle R 10 83 text s 1 nbsp which means that on average one would have to wait 1083s 1076 years to see a single ice nucleus forming in 1 m3 of water at 20 C This is a rate of homogeneous nucleation estimated for a model of water not real water in experiments one cannot grow nuclei of water and so cannot directly determine the values of the barrier D G displaystyle Delta G nbsp or the dynamic parameters such as j displaystyle j nbsp for real water However it may be that indeed the homogeneous nucleation of ice at temperatures near 20 C and above is extremely slow and so that whenever water freezes at temperatures of 20 C and above this is due to heterogeneous nucleation i e the ice nucleates in contact with a surface Homogeneous nucleation Edit Homogeneous nucleation is much rarer than heterogeneous nucleation 1 9 However homogeneous nucleation is simpler and easier to understand than heterogeneous nucleation so the easiest way to understand heterogeneous nucleation is to start with homogeneous nucleation So we will outline the CNT calculation for the homogeneous nucleation barrier D G displaystyle Delta G nbsp nbsp The green curve is the total Gibbs if this is at constant pressure free energy as a function of radius Shown is the free energy barrier D G displaystyle Delta G nbsp and radius at the top of the barrier r displaystyle r nbsp This total free energy is a sum of two terms The first is a bulk term which is plotted in red This scales with volume and is always negative The second term is an interfacial term which is plotted in black This is the origin of the barrier It is always positive and scales with surface area To understand if nucleation is fast or slow D G r displaystyle Delta G r nbsp the Gibbs free energy change as a function of the size of the nucleus needs to be calculated The classical theory 10 assumes that even for a microscopic nucleus of the new phase we can write the free energy of a droplet D G displaystyle Delta G nbsp as the sum of a bulk term that is proportional to the volume of the nucleus and a surface term that is proportional to its surface area D G 4 3 p r 3 D g v 4 p r 2 s displaystyle Delta G frac 4 3 pi r 3 Delta g v 4 pi r 2 sigma nbsp The first term is the volume term and as we are assuming that the nucleus is spherical this is the volume of a sphere of radius r displaystyle r nbsp D g v displaystyle Delta g v nbsp is the difference in free energy per unit of volume between the phase that nucleates and the thermodynamic phase nucleation is occurring in For example if water is nucleating in supersaturated air then D g v displaystyle Delta g v nbsp is the free energy per unit of volume of water minus that of supersaturated air at the same pressure As nucleation only occurs when the air is supersaturated D g v displaystyle Delta g v nbsp is always negative The second term comes from the interface at surface of the nucleus which is why it is proportional to the surface area of a sphere s displaystyle sigma nbsp is the surface tension of the interface between the nucleus and its surroundings which is always positive In case of nucleation in a solid matrix there is a third energy component in addition to the two mentioned above This third energy appears as the strain that is caused by the density difference between the product and the parent phase and is positive unfavorable for nucleation 11 For small r displaystyle r nbsp the second surface term dominates and D G r gt 0 displaystyle Delta G r gt 0 nbsp The free energy is the sum of an r 2 displaystyle r 2 nbsp and r 3 displaystyle r 3 nbsp terms Now the r 3 displaystyle r 3 nbsp terms varies more rapidly with r displaystyle r nbsp than the r 2 displaystyle r 2 nbsp term so as small r displaystyle r nbsp the r 2 displaystyle r 2 nbsp term dominates and the free energy is positive while for large r displaystyle r nbsp the r 3 displaystyle r 3 nbsp term dominates and the free energy is negative This shown in the figure to the right Thus at some intermediate value of r displaystyle r nbsp the free energy goes through a maximum and so the probability of formation of a nucleus goes through a minimum There is a least probable nucleus size i e the one with the highest value of D G displaystyle Delta G nbsp where d G d r r r c 0 r c 2 s D g v displaystyle left frac dG dr right r r c 0 implies r c frac 2 sigma Delta g v nbsp Addition of new molecules to nuclei larger than this critical radius r c displaystyle r c nbsp decreases the free energy so these nuclei are more probable The rate at which nucleation occurs is then limited by i e determined by the probability of forming the critical nucleus This is just the exponential of minus the free energy of the critical nucleus D G displaystyle Delta G nbsp which isD G 16 p s 3 3 D g v 2 displaystyle Delta G frac 16 pi sigma 3 3 Delta g v 2 nbsp dd This is the free energy barrier needed in the CNT expression for R displaystyle R nbsp above In the discussion above we assumed the growing nucleus to be three dimensional and spherical Similar equations can be set up for other dimensions and or other shapes using the appropriate expressions for the analogues of volume and surface area of the nucleus One will then find out that any non spherical nucleus has a higher barrier height D G displaystyle Delta G nbsp than the corresponding spherical nucleus This can be understood from the fact that a sphere has the lowest possible surface area to volume ratio thereby minimizing the unfavourable surface contribution with respect to the favourable bulk volume contribution to the free energy Assuming equal kinetic prefactors the fact that D G displaystyle Delta G nbsp is higher for non spherical nuclei implies that their formation rate is lower This explains why in homogeneous nucleation usually only spherical nuclei are taken into account From an experimental standpoint this theory grants tuning of the critical radius through the dependence of D G displaystyle Delta G nbsp on temperature The variable D g v displaystyle Delta g v nbsp described above can be expressed as D G D H f T m T T m D g v D H f T m T V a t T m displaystyle Delta G frac Delta H f T m T T m implies Delta g v frac Delta H f T m T V at T m nbsp where T m displaystyle T m nbsp is the melting point and H f displaystyle H f nbsp is the enthalpy of formation for the material Furthermore the critical radius can be expressed as r c 2 s D H f V a t T m T m T D G 16 p s 3 3 D H f 2 V a t T m T m T 2 displaystyle r c frac 2 sigma Delta H f frac V at T m T m T implies Delta G frac 16 pi sigma 3 3 Delta H f 2 left frac V at T m T m T right 2 nbsp revealing a dependence of reaction temperature Thus as you increase the temperature near T m displaystyle T m nbsp the critical radius will increase Same happens when you move away from the melting point the critical radius and the free energy decrease Heterogeneous nucleation Edit nbsp Three droplets on a surface illustrating decreasing contact angles The contact angle the droplet surface makes with the solid horizontal surface decreases from left to right nbsp A diagram featuring all of the factors that affect heterogeneous nucleationUnlike homogeneous nucleation heterogeneous nucleation occurs on a surface or impurity It is much more common than homogeneous nucleation This is because the nucleation barrier for heterogeneous nucleation is much lower than for homogeneous nucleation To see this note that the nucleation barrier is determined by the positive term in the free energy D G displaystyle Delta G nbsp which is proportional to the total exposed surface area of a nucleus For homogeneous nucleation the surface area is simply that of a sphere For heterogeneous nucleation however the surface area is smaller since part of the nucleus boundary is accommodated by the surface or impurity onto which it is nucleating 12 There are several factors which determine the precise reduction in the exposed surface area 12 As shown in a diagram on the left these factors include the size of the droplet the contact angle 8 displaystyle theta nbsp between the droplet and surface and the interactions at the three phase interfaces liquid solid solid vapor and liquid vapor The free energy needed for heterogeneous nucleation D G h e t displaystyle Delta G het nbsp is equal to the product of homogeneous nucleation D G h o m displaystyle Delta G hom nbsp and a function of the contact angle f 8 displaystyle f theta nbsp D G h e t f 8 D G h o m f 8 2 3 cos 8 cos 3 8 4 displaystyle Delta G het f theta Delta G hom qquad f theta frac 2 3 cos theta cos 3 theta 4 nbsp The schematic to the right illustrates the decrease in the exposed surface area of the droplet as the contact angle decreases Deviations from a flat interface decrease the exposed surface even further there exist expressions for this reduction for simple surface geometries 13 In practice this means that nucleation will tend to occur on surface imperfections nbsp Difference in energy barriersStatistical mechanical treatment EditThe classical nucleation theory hypothesis for the form of D G displaystyle Delta G nbsp can be examined more rigorously using the tools of statistical mechanics 14 Specifically the system is modeled as a gas of non interacting clusters in the grand canonical ensemble A state of metastable equilibrium is assumed such that the methods of statistical mechanics hold at least approximately 15 The grand partition function is 16 Q N 0 z N m S N e b H N m S z e b m displaystyle Q sum N 0 infty z N sum mu S N e beta mathcal H N mu S qquad z equiv e beta mu nbsp Here the inner summation is over all microstates m S displaystyle mu S nbsp which contain exactly N displaystyle N nbsp particles It can be decomposed into contributions from each possible combination of clusters which results in N displaystyle N nbsp total particles 17 For instance m S N 3 e b H 3 m S q 3 q 2 q 1 1 3 q 1 3 displaystyle sum mu S N 3 e beta mathcal H 3 mu S q 3 q 2 q 1 frac 1 3 q 1 3 nbsp where q n displaystyle q n nbsp is the configuration integral of a cluster with n displaystyle n nbsp particles and potential energy U n r n displaystyle U n left mathbf r n right nbsp q n 1 n 1 L 3 n d r n e b U n r n displaystyle begin aligned q n frac 1 n frac 1 Lambda 3n int d mathbf r n e beta U n left mathbf r n right end aligned nbsp The quantity L displaystyle Lambda nbsp is the thermal de Broglie wavelength of the particle which enters due to the integration over the 3 n displaystyle 3n nbsp momentum degrees of freedom The inverse factorials are included to compensate for overcounting since particles and clusters alike are assumed indistinguishable More compactly Q exp n 1 q n z n displaystyle Q exp left sum n 1 infty q n z n right nbsp Then by expanding Q displaystyle Q nbsp in powers of q m z m displaystyle q m z m nbsp the probability P m ℓ displaystyle P m ell nbsp of finding exactly ℓ displaystyle ell nbsp clusters which each has m displaystyle m nbsp particles is P m ℓ z ℓ m q m ℓ ℓ exp n m q n z n exp n 1 q n z n z ℓ m q m ℓ ℓ exp q m z m displaystyle begin aligned P m ell amp frac z ell m q m ell ell cdot frac exp left sum n neq m infty q n z n right exp left sum n 1 infty q n z n right amp frac z ell m q m ell ell exp left q m z m right end aligned nbsp The number density r m displaystyle rho m nbsp of m displaystyle m nbsp clusters can therefore be calculated as r m 1 V ℓ 0 ℓ z ℓ m q m ℓ ℓ exp q m z m z m q m V displaystyle begin aligned rho m amp frac 1 V sum ell 0 infty ell frac z ell m q m ell ell exp left q m z m right amp frac z m q m V end aligned nbsp This is also called the cluster size distribution The grand potential W displaystyle Omega nbsp is equal to k B T ln Q displaystyle k B T ln Q nbsp which using the thermodynamic relationship W p V displaystyle Omega pV nbsp leads to the following expansion for the pressure p k B T n 1 q n V z n displaystyle frac p k B T sum n 1 infty frac q n V z n nbsp If one defines the right hand side of the above equation as the function P T z displaystyle Pi T z nbsp then various other thermodynamic quantities can be calculated in terms of derivatives of P displaystyle Pi nbsp with respect to z displaystyle z nbsp 18 The connection with the simple version of the theory is made by assuming perfectly spherical clusters in which case U n r n displaystyle U n left mathbf r n right nbsp depends only on n displaystyle n nbsp in the form U n E 0 n w n 1 2 displaystyle U n E 0 n wn 1 2 nbsp where E 0 displaystyle E 0 nbsp is the binding energy of a single particle in the interior of a cluster and w gt 0 displaystyle w gt 0 nbsp is the excess energy per unit area of the cluster surface Then q n exp b E 0 n w n 1 2 displaystyle q n propto exp beta cdot E 0 n wn 1 2 nbsp and the cluster size distribution is r n e b E 0 m n w n 1 2 displaystyle rho n propto e beta E 0 mu n wn 1 2 nbsp which implies an effective free energy landscape D G n E 0 m n w n 1 2 displaystyle Delta G n E 0 mu n wn 1 2 nbsp in agreement with the form proposed by the simple theory On the other hand this derivation reveals the significant approximation in assuming spherical clusters with U n r n E 0 n w n 1 2 displaystyle U n left mathbf r n right E 0 n wn 1 2 nbsp In reality the configuration integral q n displaystyle q n nbsp contains contributions from the full set of particle coordinates r n displaystyle mathbf r n nbsp thus including deviations from spherical shape as well as cluster degrees of freedom such as translation vibration and rotation Various attempts have been made to include these effects in the calculation of q n displaystyle q n nbsp although the interpretation and application of these extended theories has been debated 5 19 20 A common feature is the addition of a logarithmic correction ln n displaystyle sim ln n nbsp to D G n displaystyle Delta G n nbsp which plays an important role near the critical point of the fluid 21 Limitations EditClassical nucleation theory makes a number of assumptions which limit its applicability Most fundamentally in the so called capillarity approximation it treats the nucleus interior as a bulk incompressible fluid and ascribes to the nucleus surface the macroscopic interfacial tension s displaystyle sigma nbsp even though it is not obvious that such macroscopic equilibrium properties apply to a typical nucleus of say 50 molecules across 22 23 In fact it has been shown that the effective surface tension of small droplets is smaller than that of the bulk liquid 24 In addition the classical theory places restrictions on the kinetic pathways by which nucleation occurs assuming clusters grow or shrink only by single particle adsorption emission In reality merging and fragmentation of entire clusters cannot be excluded as important kinetic pathways in some systems Particularly in dense systems or near the critical point where clusters acquire an extended and ramified structure such kinetic pathways are expected to contribute significantly 24 The behavior near the critical point also suggests the inadequacy at least in some cases of treating clusters as purely spherical 25 Various attempts have been made to remedy these limitations and others by explicitly accounting for the microscopic properties of clusters However the validity of such extended models is debated One difficulty is the exquisite sensitivity of the nucleation rate R displaystyle R nbsp to the free energy D G displaystyle Delta G nbsp even small discrepancies in the microscopic parameters can lead to enormous changes in the predicted nucleation rate This fact makes first principles predictions nearly impossible Instead models must be fit directly to experimental data which limits the ability to test their fundamental validity 26 Comparison with simulation and experiment EditFor simple model systems modern computers are powerful enough to calculate exact nucleation rates numerically An example is the nucleation of the crystal phase in a system of hard spheres which is a simple model of colloids consisting of perfectly hard spheres in thermal motion The agreement of CNT with the simulated rates for this system confirms that the classical theory is a reasonable approximation 27 For simple models CNT works quite well however it is unclear if it describes complex e g molecular systems equally well For example in the context of vapor to liquid nucleation the CNT predictions for the nucleation rate are incorrect by several orders of magnitude on an absolute scale that is without renormalizing with respect to experimental data Nevertheless certain variations on the classical theory have been claimed to represent the temperature dependence adequately even if the absolute magnitude is inaccurate 28 Jones et al computationally explored the nucleation of small water clusters using a classical water model It was found that CNT could describe the nucleation of clusters of 8 50 water molecules well but failed to describe smaller clusters 29 Corrections to CNT obtained from higher accuracy methods such as quantum chemical calculations may improve the agreement with experiment 30 References Edit a b H R Pruppacher and J D Klett Microphysics of Clouds and Precipitation Kluwer 1997 P G Debenedetti Metastable Liquids Concepts and Principles Princeton University Press 1997 a b Sear R P 2007 Nucleation theory and applications to protein solutions and colloidal suspensions J Phys Condens Matter 19 3 033101 Bibcode 2007JPCM 19c3101S CiteSeerX 10 1 1 605 2550 doi 10 1088 0953 8984 19 3 033101 S2CID 4992555 Kreer Markus 1993 Classical Becker Doring cluster equations Rigorous results on metastability and long time behaviour Annalen der Physik 505 4 398 417 Bibcode 1993AnP 505 398K doi 10 1002 andp 19935050408 a b Oxtoby David W 1992 Homogeneous nucleation theory and experiment Journal of Physics Condensed Matter 4 38 7627 7650 Bibcode 1992JPCM 4 7627O doi 10 1088 0953 8984 4 38 001 S2CID 250827558 Zeldovich Y B 1943 On the theory of new phase formation cavitation Acta Physicochem USSR 18 1 Frenkel Daan Smit Berent 2001 Understanding Molecular Simulation Second Edition From Algorithms to Applications p Academic Press ISBN 978 0122673511 Sanz Eduardo Vega Carlos Espinosa J R Cabellero Bernal R Abascal J L F Valeriani Chantal 2013 Homogeneous Ice Nucleation at Moderate Supercooling from Molecular Simulation Journal of the American Chemical Society 135 40 15008 15017 arXiv 1312 0822 Bibcode 2013arXiv1312 0822S doi 10 1021 ja4028814 PMID 24010583 S2CID 2304292 Sear Richard P 2014 Quantitative Studies of Crystal Nucleation at Constant Supersaturation Experimental Data and Models CrystEngComm 16 29 6506 6522 doi 10 1039 C4CE00344F F F Abraham 1974 Homogeneous nucleation theory Academic Press NY Shirzad K Viney C 2023 A critical review on applications of the Avrami equation beyond materials science Journal of the Royal Society Interface 20 203 doi 10 1098 rsif 2023 0242 PMC 10282574 PMID 37340781 a b Liu X Y 31 May 2000 Heterogeneous nucleation or homogeneous nucleation The Journal of Chemical Physics 112 22 9949 9955 Bibcode 2000JChPh 112 9949L doi 10 1063 1 481644 ISSN 0021 9606 Sholl C A N H Fletcher 1970 Decoration criteria for surface steps Acta Metall 18 10 1083 1086 doi 10 1016 0001 6160 70 90006 4 hdl 1885 213299 The discussion which follows draws from Kalikmanov 2001 unless noted otherwise Kalikmanov V I 2013 Nucleation Theory Lecture Notes in Physics vol 860 Springer Netherlands pp 17 19 Bibcode 2013LNP 860 K doi 10 1007 978 90 481 3643 8 ISBN 978 90 481 3643 8 ISSN 0075 8450 Kardar Mehran 2007 Statistical Physics of Particles Cambridge University Press p 118 ISBN 978 0 521 87342 0 Kalikmanov V I 2001 Statistical Physics of Fluids Basic Concepts and Applications Springer Verlag pp 170 172 ISBN 978 3 540 417 47 7 ISSN 0172 5998 Kalikmanov 2001 pp 172 173 Kiang C S and Stauffer D and Walker G H and Puri O P and Wise J D and Patterson E M C S Stauffer D Walker G H Puri O P Wise J D Patterson E M 1971 A Reexamination of Homogeneous Nucleation Theory Journal of the Atmospheric Sciences 28 7 1222 1232 Bibcode 1971JAtS 28 1222K doi 10 1175 1520 0469 1971 028 lt 1222 AROHNT gt 2 0 CO 2 a href Template Citation html title Template Citation citation a CS1 maint multiple names authors list link Reguera D Rubi J M 2001 Nonequilibrium translational rotational effects in nucleation The Journal of Chemical Physics 115 15 7100 7106 arXiv cond mat 0109270 Bibcode 2001JChPh 115 7100R doi 10 1063 1 1405122 S2CID 95887792 Sator N 2003 Clusters in simple fluids Physics Reports 376 1 1 39 arXiv cond mat 0210566 Bibcode 2003PhR 376 1S doi 10 1016 S0370 1573 02 00583 5 ISSN 0370 1573 S2CID 119492597 Kalikmanov 2013 p 21 Oxtoby 1992 p 7631 a b Kiang et al 1971 Sator 2003 Oxtoby 1992 p 7638 7640 Auer S D Frenkel 2004 Numerical prediction of absolute crystallization rates in hard sphere colloids PDF J Chem Phys 120 6 3015 29 Bibcode 2004JChPh 120 3015A doi 10 1063 1 1638740 hdl 1874 12074 PMID 15268449 S2CID 15747794 Fladerer Alexander Strey Reinhard 2006 04 28 Homogeneous nucleation and droplet growth in supersaturated argon vapor The cryogenic nucleation pulse chamber The Journal of Chemical Physics AIP Publishing 124 16 164710 Bibcode 2006JChPh 124p4710F doi 10 1063 1 2186327 ISSN 0021 9606 PMID 16674160 Merikanto Joonas Zapadinsky Evgeni Lauri Antti Vehkamaki Hanna 4 April 2007 Origin of the Failure of Classical Nucleation Theory Incorrect Description of the Smallest Clusters Physical Review Letters 98 14 145702 Bibcode 2007PhRvL 98n5702M doi 10 1103 PhysRevLett 98 145702 PMID 17501289 Temelso Berhane Morrell Thomas E Shields Robert M Allodi Marco A Wood Elena K Kirschner Karl N Castonguay Thomas C Archer Kaye A Shields George C 22 February 2012 Quantum Mechanical Study of Sulfuric Acid Hydration Atmospheric Implications The Journal of Physical Chemistry A 116 9 2209 2224 Bibcode 2012JPCA 116 2209T doi 10 1021 jp2119026 PMID 22296037 Retrieved from https en wikipedia org w index php title Classical nucleation theory amp oldid 1171688753, 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.