Next Article in Journal
Investigation of Effect of Milling Atmosphere and Starting Composition on Mg2FeH6 Formation
Next Article in Special Issue
Production of Liquid Metal Spheres by Molding
Previous Article in Journal
On the Physics of Machining Titanium Alloys: Interactions between Cutting Parameters, Microstructure and Tool Wear
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Crystallization of Supercooled Liquid Elements Induced by Superclusters Containing Magic Atom Numbers

by
Robert F. Tournier
CRETA/CNRS, Université Joseph Fourier, B.P. 166, 38042 Grenoble cedex 09, France
Metals 2014, 4(3), 359-387; https://doi.org/10.3390/met4030359
Submission received: 30 April 2014 / Revised: 10 July 2014 / Accepted: 16 July 2014 / Published: 6 August 2014
(This article belongs to the Special Issue Liquid Metals)

Abstract

:
A few experiments have detected icosahedral superclusters in undercooled liquids. These superclusters survive above the crystal melting temperature Tm because all their surface atoms have the same fusion heat as their core atoms, and are melted by liquid homogeneous and heterogeneous nucleation in their core, depending on superheating time and temperature. They act as heterogeneous growth nuclei of crystallized phase at a temperature Tc of the undercooled melt. They contribute to the critical barrier reduction, which becomes smaller than that of crystals containing the same atom number n. After strong superheating, the undercooling rate is still limited because the nucleation of 13-atom superclusters always reduces this barrier, and increases Tc above a homogeneous nucleation temperature equal to Tm/3 in liquid elements. After weak superheating, the most stable superclusters containing n = 13, 55, 147, 309 and 561 atoms survive or melt and determine Tc during undercooling, depending on n and sample volume. The experimental nucleation temperatures Tc of 32 liquid elements and the supercluster melting temperatures are predicted with sample volumes varying by 18 orders of magnitude. The classical Gibbs free energy change is used, adding an enthalpy saving related to the Laplace pressure change associated with supercluster formation, which is quantified for n = 13 and 55.

Graphical Abstract

1. Introduction

An undercooled liquid develops special clusters that minimize the energy locally which are incompatible with space filling [1,2,3]. Such entities are homogeneously formed in glass-forming melts, and act as growth nuclei of crystals above the glass transition [4]. The formation of icosahedral nanoclusters has often been studied by molecular dynamics simulations into or out of liquids [5,6,7,8]. Silver superclusters containing the magic atom numbers n = 13, 55, 147, 309, 561 are more stable. Their formation temperature out of melt and their radius have been determined [5]. Icosahedral gold nanoclusters do not premelt below their bulk melting temperature [6]. Nanoclusters have been prepared out of liquids [9,10,11,12,13,14]. The density of states of conduction electrons at the Fermi energy being strongly reduced for particle diameters smaller than one nanometer leads to a gap opening [9,10]. The growth nuclei are expected to have analogous electronic properties.
Superclusters containing magic atom numbers are tentatively viewed for the first time as being the main growth nuclei of crystallized phases in all liquid elements. I already considered that an energy saving resulting from the equalization of Fermi energies of nuclei and melts cannot be neglected in the classical crystal nucleation model [15,16]. An enthalpy saving εv per volume unit of critical radius clusters equal to εls × ΔHm/Vm was introduced in the Gibbs free energy change ΔG2ls which gives rise to spherical clusters that transform the critical energy barrier into a less effective energy barrier, thereby inducing crystal growth around them at a temperature Tc much higher than the theoretical homogeneous nucleation temperature equal to Tm/3. This enthalpy depends on ΔHm the melting heat per mole at the melting temperature Tm, Vm the molar volume and εls a numerical coefficient. The experimental growth temperature Tc is often interpreted in the literature as a homogeneous nucleation temperature. This view is not correct, because the Tc of liquid elements is highly dependent on the sample volume v [17]. The crystallization temperatures are known to be driven by an effective critical energy barrier that is strongly weakened by the Gibbs free energy change associated with impurity clusters in the liquid [18,19]. The presence of εv has for consequence to prevent the melting above Tm of the smallest clusters acting as intrinsic growth nuclei reducing the critical energy barrier in undercooled liquids. The critical energy saving coefficient εls was shown for the first time as depending on θ2 = [(TTm)/Tm]2 in liquid elements with a maximum at Tm equal to 0.217 [15,16].
In this article, each cluster having a radius smaller than the critical radius has its own energy saving coefficient εnm depending on θ2, n and its radius Rnm. In this case too, the cluster surface energy is a linear function of εnm instead of a function of θ or T [20,21,22,23,24,25]. The Gibbs free energy change derivative [d(ΔG2ls)/dT]p = −ΔSm at Tm continues to be equal to the entropy change whatever the particle radius is because (dεls/dT)T=Tm is equal to zero. All the surface atoms of growth nuclei have the same fusion heat as their core atoms [21]. They survive for a limited time above the melting temperature because they are not submitted to surface melting. A melt bath needs time to attain the thermodynamic equilibrium above the melting temperature Tm. This finding is the basic property permitting to assume for the first time that the growth nuclei in all liquid elements are superclusters instead of crystals. These superclusters are melted by homogeneous nucleation of liquid in their core instead of surface melting. A prediction of superheating effects is also presented for the first time for 38 liquid elements together with the predictions of undercooling rates depending on sample volumes and supercluster magic atom numbers n. The undercooling temperatures of gold and titanium have already been predicted using a continuous variation of growth nucleus radii and quantified values of εv [22,23].
The equalization of Fermi energies of liquid and superclusters is not realized by a transfer of conduction electrons from nuclei to melts as I assumed in the past [15,16,24]. I recently suggested that a Laplace pressure change Δp applied to conducting and nonconducting superclusters accompanied by an enthalpy saving per mole equal to Vm × Δp = εls × ΔHm is acting [25]. This quantity is proportional to 1/Rnm down to values of the radius Rnm, for which the potential energy is still equal to the quantified energy. Superclusters containing 13 and 55 atoms have an energy saving coefficient εnm0 which is quantified. This coefficient εnm0 associated with an n-atom supercluster strongly depends on n up to the critical number nc of atoms, giving rise to crystal spontaneous growth when εnm0 is equal to 0.217 in liquid elements [15].
The quantified values of εv are known solutions of the Schrödinger equation which are obtained assuming that the same complementary Laplace pressure Δp could be created by a virtual s-electron transfer from the crystal to the melt or from the melt to the crystal, creating a virtual surface charge screening associated with a spherical attractive potential [24]. All values of εv for radii smaller than the critical values lead to a progressive reduction of electron s-state density as a function of n [23]. Reduced s-state density of superclusters depending on their radius and electronic specific heat of Cu, Ag and Au n-atom superclusters are studied, imposing a relative variation of Fermi energies during their formation in noble metal liquid state equal to −2/3 of the relative volume change. The radii of Ag superclusters calculated by molecular dynamics simulations in [5] are comparable with the critical radius values R*2ls(T) deduced from this constraint.
This article follows the plan below:
2.
The supercluster formation equations leading to crystallization.
2.1.
Gibbs free energy change associated with growth nucleus formation.
2.2.
Thermal dependence of the energy saving coefficient εnm of an n-atom condensed supercluster.
2.3.
Crystal homogeneous nucleation temperature and effective nucleation temperature.
3.
The model of quantification of the energy saving of superclusters.
4.
Prediction of crystallization temperatures of 38 supercooled liquid elements at constant molar volume.
5.
Homogeneous nucleation of 13-atom superclusters and undercooling rate predictions.
6.
Maximum overheating temperature applied to melt superclusters at constant molar volume.
7.
Electronic properties of Cu, Ag and Au superclusters.
8.
Silver supercluster formation into and out of undercooled liquid.
9.
Melting of Cu, Ag and Au superclusters, varying the overheating temperatures and times.
9.1.
Superheating of Cu, Ag and Au superclusters.
9.2.
Analysis of the influence of Cu superheating time on the undercooling rate.
10.
Conclusions.

2. Supercluster Formation Equations Leading to Crystallization

2.1. Gibbs Free Energy Change Associated with Growth Nucleus Formation

The classical Gibbs free energy change for a growth nucleus formation in a melt is given in Equation (1):
Δ G 1ls = θ Δ H m V m 4 π R 3 3 + 4 π R 2 σ 1ls
where R is the nucleus radius, σ1ls the surface energy, ΔHm the melting heat, Vm the molar volume and θ = (TTm)/Tm the reduced temperature. Turnbull has defined a surface energy coefficient α1ls in Equation (2) which is equal to Equation (3) [19,26]:
σ 1ls ( V m N A ) 1 / 3 = α 1ls Δ H m V m
α 1ls = [ N A k B ln ( K ls ) 36 πΔ S m ] 1 / 3
where NA is the Avogadro number, kB the Boltzmann constant, ΔSm the melting entropy and ln(Kls) = 90 ± 2.
An energy saving per volume unit εls × ΔHm/Vm is introduced in Equation (1); the new Gibbs free energy change is given by Equation (4), where σ2ls is the new surface energy [15,27]:
Δ G 2ls = ( θ ε ls ) Δ H m V m 4 π R 3 3 + 4 π R 2 σ 2ls
The new surface energy coefficient α2ls is given by Equation (5):
σ 2ls ( V m N A ) 1 / 3 = α 2ls Δ H m V m
The critical radius R*2ls in Equation (6) and the critical thermally-activated energy barrier ΔG*2ls/kBT in Equation (7) are calculated assuming (dεls/dR)R=R*2ls = 0:
R 2ls * = 2 α 2ls θ ε ls ( V m N A ) 1 / 3
Δ G 2ls * k B T = 16 πΔ S m α 2ls 3 3 N A k B ( θ ε ls ) 2 ( 1 + θ )
They are not infinite at the melting temperature Tm because εls is no longer equal to zero [15,16]. The homogeneous nucleation temperature T2 (or θ2) occurs when the nucleation rate J in Equation (8) is equal to 1, lnKls = 90 ± 2 in Equations (9) and (10) respected with ΔG*2ls/kBT = 90 neglecting the lnKls thermal variation [28]:
J = K ls exp ( Δ G 2ls * k B T )
Δ G 2ls * k B T = ln ( K ls )
The unknown surface energy coefficient α2ls in Equation (10) is deduced from Equations (7) and (9):
α 2ls 3 = 3 N A k B ( θ 2 ε ls ) 2 ( 1 + θ 2 ) ln ( K ls ) 16 πΔ S m
The surface energy σ2ls in Equation (5) has to be minimized to obtain the homogeneous nucleation temperature T2 (or θ2) for a fixed value of εls. The derivative dα2ls/dθ is equal to zero at the temperature T2 (or θ2) given by Equation (11), assuming that ln(Kls) does not depend on the temperature:
θ 2 = T 2 T m T m = ε ls 2 3
The homogeneous nucleation temperature T2 is equal to Tm/3 (or θ2 = −2/3) in liquid elements and εls (θ) is equal to zero at this temperature [15,24].
The surface energy coefficient α2ls is now given by Equation (12), replacing θ by Equation (11) in Equation (10) for each value of εls:
α 2ls = ( 1 + ε ls ) [ N A k B ln ( K ls ) 36 πΔ S m ] 1 / 3 = ( 1 + ε ls ) α 1ls
The classical crystal nucleation Equation (4) is transformed into Equation (13) with the introduction of the energy saving coefficient εls:
Δ G 2ls ( R , θ ) = Δ H m V m ( θ ε ls ) 4 π R 3 3 + 4 π R 2 Δ H m V m ( 1 + ε ls ) ( 12 k B V m ln K ls 432 π × Δ S m ) 1 / 3
The Laplace pressure p and the complementary Laplace pressure Δp applied on the critical nucleus are calculated from the surface energy σ2ls with the Equations (13) and (6) and Δp is given by Equation (14) [21,25]:
Δ p = Δ H m V m × ε ls ( θ ) = 2 ( δ σ 2ls ) R
where δσ2ls is the complement proportional to εls in the surface energy in Equation (13). The complement Δp is equal to the energy saving εls(θ) × ΔHm/Vm. The Gibbs free energy change ΔG2ls in Equation (13) directly depends on the cluster atom number n and the energy saving coefficient εnm of the cluster instead of depending on its molar volume Vm and its radius R as shown in Equation (15):
Δ G nm ( n , θ , ε nm ) = Δ H m n N A ( θ ε nm ) + ( 4 π ) 1 / 3 N A Δ H m α 2ls ( 3 n ) 2 / 3
The formation of superclusters having a weaker effective energy barrier than that of crystals precedes the formation of crystallized nuclei in an undercooled melt [5,29]. A spherical surface containing n atoms being minimized, a supercluster having a radius smaller than the critical radius cannot be easily transformed into a non-spherical crystal of n atoms because the surface energy would increase. The critical radius of superclusters could be larger than that of crystals because the supercluster density could be smaller, as already predicted for Ag [5] and confirmed in part 7. In these conditions, the transformation of a supercluster into a crystal is expected to occur above the critical radius for crystal growth when the Gibbs free energy change begins to decrease with the radius, while that of a supercluster increases up to its critical radius. It is shown in parts 3 and 4 that the supercluster energy saving εnm × ΔHm is quantified, depends on cluster radius R and atom number n, and is larger than the critical energy saving 0.217 × ΔHm. The cluster’s previous formation during undercooling determines the spontaneous growth temperature Tc reducing the effective critical energy barrier. The smallest homogeneously-condensed cluster controls the heterogeneous growth of crystals at temperatures higher than the homogeneous nucleation temperature Tm/3 (θ2 = −2/3) even in liquids which are at thermodynamic equilibrium at Tm before cooling.

2.2. Thermal Dependence of the Energy Saving Coefficient εnm of an n-Atom Condensed Cluster

All growth nuclei that are formed in an undercooled melt are submitted to a complementary Laplace pressure. The energy saving coefficient εnm of an n-atom cluster given in Equation (16), being a function of θ2 as already shown [15], is maximum at Tm, with (dεnm/dT)T=Tm equal to zero:
ε v = ε nm 0 Δ H m V m = ε nm 0 ( 1 θ 2 θ 0m 2 ) Δ H m V m
where εnm0 is the quantified energy saving coefficient of an n-atom cluster at Tm depending on the spherical nucleus radius R [24].
This thermal variation has for consequence that the fusion entropy per mole of a cluster of radius R is equal to the fusion entropy ΔSm of the bulk solid [15,24]
3 4 π R 3 [ d ( Δ G nm ) d T ] T = T m = Δ S m V m
In these conditions, cluster surface atoms having the same fusion heat as core atoms, the cluster melts above Tm by liquid droplet homogeneous nucleation above Tm rather than by surface melting as expected for superclusters [6]. This θ2 thermal variation has already been used to predict the undercooling rate of some liquid elements [22,23].
The critical parameters for spontaneous supercluster growth are determined by an energy saving coefficient called εls in Equation (18):
ε v = ε ls Δ H m V m = ε ls0 ( 1 θ 2 θ 0m 2 ) Δ H m V m
where εls0 = 0.217 is the critical value at Tm and θ−20m = 2.25 in liquid elements [15,24]. A critical supercluster contains a critical number nc of atoms given by:
n c = 32 π α 2ls 3 3 ( θ ε ls ) 3

2.3. Crystal Homogeneous Nucleation Temperature and Effective Nucleation Temperature

The thermally-activated critical energy barrier is now given by Equation (20):
Δ G 2ls * k B T = 12 ( 1 + ε ls ) 3 ln ( K ls ) 81 ( θ ε ls ) 2 ( 1 + θ )
where εls is given by Equation (18). The coefficient of ln(Kls) in Equation (20) becomes equal to 1 at the homogeneous nucleation temperature Tm/3 and the Equations (9) and (11) are respected.
Homogeneously-condensed superclusters of n-atoms act as growth nuclei at a temperature generally higher than the homogeneous nucleation temperatures Tm/3 of liquid elements because they reduce the critical energy barrier as shown in Equation (21) [18]:
ln ( J × v × t sn ) = ln ( K ls × v × t sn ) ( Δ G 2ls * k B T Δ G nm k B T ) = 0
where v is the sample volume, J the nucleation rate, tsn the steady-state nucleation time, lnKls = 90 ± 2, ΔG*2ls/kBT defined in Equation (20) and ΔGnm in Equation (15). The Equation (21) is applied, assuming that n-atom superclusters preexist in melts when they have not been melted by superheating above Tm. It can also be applied when the homogeneous condensation time of an n-atom supercluster is evolved and its own critical energy barrier crossed. The cluster thermally-activated critical energy barrier ΔG*nm/kBT and the effective thermally-activated critical energy barrier ΔGneff/kBT of an n-atom supercluster are given by Equations (22) and (23):
Δ G nm * k B T = 12 ( 1 + ε nm ) 3 ln ( K ls ) 81 ( θ ε nm ) 2 ( 1 + θ )
Δ G neff k B T = Δ G nm * k B T Δ G nm k B T = 12 ( 1 + ε nm ) 3 ln ( K ls ) 81 ( θ ε nm ) 2 ( 1 + θ ) Δ G nm k B T
where ΔGnm is given by Equation (15), εnm by Equation (16) and lnKls = 90 ± 2. The quantified value εnm0 × ΔHm of the cluster energy saving at Tm is defined in the next part. The transient nucleation time being neglected, the growth around these nuclei is only possible when the steady-state nucleation time tsn is evolved and the relation Equation (24) is respected:
ln ( J n × v × t sn ) = ln ( K ls × v × t sn ) Δ G neff k B T
where v is the sample volume and tsn the steady-state nucleation time. The crystallization follows this cluster formation time when, in addition, Equation (21) is respected. The effective nucleation temperature deduced from Equation (21) does not result from a homogeneous nucleation because it strongly depends on the sample volume v. This phenomenon explains why the effective nucleation temperature in liquid elements is observed around θ = −0.2 in sample volumes of a few mm3 instead of θ varying from − 0.58 to −0.3 in much smaller samples [17,30].

3. Quantification of Energy Saving Associated with Supercluster Formation

The potential energy saving per nucleus volume unit εls × ΔHm/Vm is equal to the Laplace pressure change Δp = 2 × δσls/R accompanying the transformation of a liquid droplet into a nucleus. The quantified energy is smaller than 2 × δσls/R at low radius R for n = 13 and 55. The calculation is made by creating a Laplace pressure on the surface of a spherical nucleus containing n atoms, which would result from a virtual transfer of n × Δz electrons in s-states from the nucleus to the melt, Δz being the fraction of transferred electrons per supercluster atom [31]. The potential energy U0 would be equal to Equation (25) and to zero beyond the nucleus radius R:
U 0 = n × Δ z × e 2 8 π ε 0 R
where e is the electron charge, and ε0 the vacuum permittivity ([32], p. 135). The quantified energy Eq at Tm is given by Equation (26):
E q = n × ε nm0 × Δ H m N A
where NA is the Avogadro number. The quantified energy saving is given by Equation (16) as a function of θ with (θ0m)−2 = 2.25 when εnm0 is known.
The Schrödinger equation depends only on the distance R of an s-state electron from the spherical potential center. The quantified solutions Eq leading to the values of εnm0 are given by the two equations in Equation (27), depending on U0 which is equal to the complementary Laplace pressure Δp acting on an n-atom cluster having a volume equal to 4πR3/3 through an intermediate parameter called k:
k = 1 × [ 2 m ( | U 0 | | E q | ) ] 1 / 2 , sin ( k × R ) k × R = ( 2 × m 0 × ε 0 × R 2 ) 1 / 2
where m0 is the electron rest mass and ħ Planck’s constant divided by 2π ([32], p. 135).
The critical radii of liquid elements are sufficiently large at Tm to assume that U0 is equal to Eq and to deduce the values of Δz from the relation Equation (25) = (26) with R = R*2ls (θ = 0) and εnm0 = εls0 = 0.217. The potential energy U0 given by Equation (25) is also equal to −4πR3/3 × Δp. Consequently, the Δz in Equation (25) does not depend on R at Tm. The value of U0 is deduced from the atom number n which depends on molar volumes Vm of solid elements extrapolated at Tm from published tables of thermal expansion [16,33]. The values of εnm0 are calculated as a function of R using Equation (28) instead of Equation (27) for n ≥ 147 because U0 is assumed equal to Eq:
   ε nm0 = ε ls0 × R 2ls * R  
The condensed-cluster energy savings εnm0 × ΔHm of 13 and 55 atoms are quantified and calculated from Equation (26). The thermal variation of εnm is given in Equation (16) using these quantified values of εnm0.

4. Prediction of Crystallization Temperatures Tc of 38 Undercooled Liquid Samples of Various Diameters

The quantified and the potential energy saving coefficients εnm0 of silver clusters have been calculated using Equations (27) and (28) and are represented in Figure 1 as a function of supercluster radius Rnm which is assumed to continuously vary. These coefficients are equal for n ≥ 147. This last approximation is used in all liquid elements.
Figure 1. The energy saving coefficient εnm0 versus the supercluster radius Rnm. The quantified (square points) and non-quantified (diamond points) energy saving coefficients εnm0 are plotted versus the silver cluster radius. This coefficient is strongly weakened when R < 0.5 nm. Quantification is necessary for an atom number n < 147.
Figure 1. The energy saving coefficient εnm0 versus the supercluster radius Rnm. The quantified (square points) and non-quantified (diamond points) energy saving coefficients εnm0 are plotted versus the silver cluster radius. This coefficient is strongly weakened when R < 0.5 nm. Quantification is necessary for an atom number n < 147.
Metals 04 00359 g001
Properties of 38 elements are classified in Table 1:
  • Column 1, the liquid elements are classified as a function of their molar fusion entropy ΔSm;
  • Column 2, the molar volume of solid elements at Tm in m3;
  • Column 3, their fusion entropy ΔSm in J/K/mole, n;
  • Column 4, their melting temperature Tm in Kelvin;
  • Column 5, the atom magic number n of the supercluster inducing crystallization of the supercooled liquid at the closest temperature to the experimental crystallization temperature;
  • Column 6, the supercluster radius Rnm in nanometers deduced from the molar volume Vm using the relation Equation (29):
    n = 4 π R nm 3 3 N A V m
  • Column 7, the energy saving coefficient εnm0 associated with the n-atom supercluster calculated using Equation (29) for n ≥ 147 and Equation (27) for n = 13 and 55, with Δz given in Table 2 column 3;
  • Column 8, the experimental reduced crystallization temperature θc exp = (TcTm)/Tm of a liquid droplet having a diameter Dexp;
  • Column 9, the reduced crystallization temperature θc calc calculated using Equation (21);
  • Column 10, the thermally-activated effective energy barrier ΔGeff/kBT given in Equations (21) and (20) leading to the crystallization of the corresponding liquid element;
  • Column 11, the calculated diameter Dcalc in mm of the liquid droplet of volume v submitted to crystallization at θc calc using Equation (21) and v × tsnv = π/6×D3 assuming that tsn = 1 s;
  • Column 12, the experimental diameter Dexp in millimeters of the liquid droplet crystallizing at θc exp
  • Column 13, references.
Table 1. Reduced crystallization temperatures of 38 supercooled liquid elements induced by condensed superclusters containing n = 13, 55, 147, 309 or 561 atoms.
Table 1. Reduced crystallization temperatures of 38 supercooled liquid elements induced by condensed superclusters containing n = 13, 55, 147, 309 or 561 atoms.
12345678910111213
Vm × 106ΔSmTmnRnmεnm0θcθc Δ G eff k B T DcalcDexpReferences
m3J/KK nm Exp.Calc. mmmm
Fe7.537.631809550.550.859−0.304−0.29861.70.10000.1000[34,35]
In15.907.694291470.980.707−0.260−0.26651.00.00310.0030[36,37]
Ti11.107.9319433091.110.546−0.180−0.19170.61.93001.8000[38]
Zr14.607.9521255611.480.447−0.167−0.17773.55.07005.0000[39,40]
Mn8.887.9815173091.030.545−0.206−0.21759.80.05000.0500[30,41]
Pb18.808.006001471.030.698−0.260−0.24957.00.02000.0200[17]
Co7.119.161768550.540.815−0.270−0.28063.70.19000.2000[36,42]
Ag11.009.1612343091.100.521−0.332−0.36041.70.00010.0001[43]
Au10.809.4313363091.100.516−0.160−0.17476.614.210015.000[30,44]
Tc8.609.472430550.570.843−0.240−0.25279.35.6400 [28]
Cr7.549.6021763090.970.512−0.130−0.18073.55.0200 [28]
Re9.509.623453550.590.862−0.241−0.25571.93.01002.9000[45]
Ir9.209.6227163091.040.512−0.190−0.18372.13.16003.3000[28,46]
Mo10.009.6328903091.070.512−0.180−0.18073.44.97004.9000[38,47]
Os8.859.6433001470.830.656−0.200−0.20871.72.8200 [28,48]
Pd9.919.6418253091.030.512−0.182−0.20962.00.11000.1000[30,49]
Pt9.669.6520423091.060.512−0.185−0.18471.62.69002.6000[38]
Cu7.579.661356550.550.781−0.259−0.25271.75.70005.7000[50,51,52]
Rh8.899.6922391470.800.654−0.204−0.20971.12.30002.3000[38]
Ta12.409.7432881470.880.653−0.200−0.20672.53.69003.7000[38,53]
Nb10.809.8227403091.100.509−0.176−0.17973.75.42005.0000[38]
Hg14.209.91232130.420.000−0.380−0.54951.60.00340.0035[17,43]
V8.9310.0721753091.030.504−0.150−0.20663.00.15000.1400[28]
Ni7.0410.141726550.540.791−0.278−0.27662.70.14000.1400[34,36]
Ru8.7510.1925231470.800.644−0.200−0.20273.75.38005.0000[28]
Hf14.9010.2025003091.220.502−0.180−0.17973.84.80004.6000[38]
Gaβ13.4010.31256130.390.000−0.500−0.52858.60.03500.0360[17,54]
Cd9.5110.445943091.180.498−0.190−0.22857.10.02100.0200[17]
Zn10.6010.536933091.050.497−0.190−0.1968.60.9810 [28]
Al10.2011.489323091.090.483−0.190−0.23657.00.02100.0200[17,55]
W16.5012.6936803091.080.467−0.155−0.17773.14.45004.2000[26,56]
Sn11.1913.46520130.440.371−0.370−0.4850.30.00220.0020[17,37]
Bi21.7020.77544130.480.551−0.410−0.40549.90.00190.0020[17,57]
Sb18.6022.15903550.740.639−0.230−0.24457.10.02100.0200[17]
Te21.0024.76723130.480.653−0.320−0.34857.10.0220.0200[17]
Se19.5027.13494130.470.507−0.305−0.29472.83.97003.8000[58]
Si12.2029.791685130.400.765−0.253−0.27175.18.69008.4000[59,60,61,62]
Ge13.9030.501210130.420.695−0.390−0.38741.00.00010.0001[36,43,52,63]
The experimental reduced crystallization temperatures θc exp are plotted in Figure 2 versus the calculated θc calc using the supercluster atom-number n leading to about the same droplet diameter Dcalc as the experimental one Dexp. A good agreement is obtained between these values in 32 liquid elements in Figure 2 and Table 1. There is no good agreement for Hg, Sn, Al, Cd, V, and Cr because these elements are known to contain impurities or oxides. Their undercooling rates are too low compared to the calculated ones.
In Figure 3, the calculated droplet diameter logarithms are plotted as a function of those of experimental droplets used to study the undercooling rate. Six orders of magnitude are studied, corresponding to 18 orders of volume magnitude. Figure 3 shows that the model is able to describe the crystallization temperature dependence on the volume sample.
Figure 2. Experimental undercooling temperatures versus calculated undercooling temperatures. The experimental reduced crystallization temperatures θc exp = (TcTm)/Tm are plotted versus the calculated ones θc calc of 38 liquid elements listed in Table 1. The smaller the atom number n, the smaller is the undercooling temperature, as shown in Table 1.
Figure 2. Experimental undercooling temperatures versus calculated undercooling temperatures. The experimental reduced crystallization temperatures θc exp = (TcTm)/Tm are plotted versus the calculated ones θc calc of 38 liquid elements listed in Table 1. The smaller the atom number n, the smaller is the undercooling temperature, as shown in Table 1.
Metals 04 00359 g002
Figure 3. Calculated liquid droplet diameters versus experimental liquid droplet diameters. The calculated and experimental droplet diameters being crystallized are compared in a logarithmic scale. Smaller liquid droplets lead to lower undercooling temperatures, as shown in Table 1.
Figure 3. Calculated liquid droplet diameters versus experimental liquid droplet diameters. The calculated and experimental droplet diameters being crystallized are compared in a logarithmic scale. Smaller liquid droplets lead to lower undercooling temperatures, as shown in Table 1.
Metals 04 00359 g003
The atom numbers n of growth nuclei in liquid elements are represented in Figure 4 as a function of the reduced experimental crystallization temperature θc exp of liquid droplets having various diameters. The undercooling rate change is two times greater when the diameter varies from 0.036 to 8.4 mm and from 0.0001 to 5 mm for n = 13 and n = 309, respectively.
The experimental undercooling reduced temperature θc of gallium is the lowest of all the liquid elements and is equal to −0.58 and is a little higher than −2/3, corresponding to a crystallization temperature Tc equal to 129 K [17] and to a melting temperature of the α phase equal to 303 K. The gallium β phase is crystallized after undercooling. Its melting temperature is 257 K instead of 303 K for the α phase and its fusion entropy is 10.91 J/K/mole, as shown in Table 1, instead of 18.4 J/K/mole [54]. Its crystallization temperature of 129 K occurs in fact at θc = −0.5. The calculated value is equal to the experimental one due to a previous condensation of 13-atom cluster, which weakens the critical energy barrier. The model works without any adjustable parameter, and is also able to predict the nucleation rate of 13-atom clusters and the diameter of gallium droplets obtained with the liquid dispersion technique.
Figure 4. N-atom superclusters acting as growth nuclei versus the experimental reduced undercooling temperatures. The number n of atoms of superclusters is plotted versus the experimental reduced temperature of crystallization θc exp.
Figure 4. N-atom superclusters acting as growth nuclei versus the experimental reduced undercooling temperatures. The number n of atoms of superclusters is plotted versus the experimental reduced temperature of crystallization θc exp.
Metals 04 00359 g004

5. Homogeneous Nucleation of 13-Atom Superclusters and Undercooling Rate Predictions

Equations (21)–(24) are now used to calculate the homogeneous formation reduced temperature θ13c of 13-atom clusters in a melt cooled below Tm from thermodynamic equilibrium state at Tm and the crystallization reduced temperature θc that they induce in liquid droplets of 10 micrometers in diameter. In Table 2, 38 liquid elements are considered. In 33 of them, the 13-atom cluster formation temperature is much larger than the crystallization temperature (θ13c >> θc). In contrast, in indium, mercury, gallium β, cadmium and zinc, the two reduced temperatures are equal within the uncertainty on the energy saving coefficient value ε13m0 given in Table 2. The crystallization temperatures of bismuth, selenium, tellurium, antimony, silicon and germanium with a growth around 13-atom clusters are predicted in good agreement with experimental values obtained with various sizes of droplets, as shown in Table 1.
In Figure 5, the homogeneous condensation reduced temperatures of 13-atom superclusters are compared with the reduced spontaneous growth temperatures, which induce crystallization. The growth is organized around these 13-atom clusters, which are formed at temperatures higher than that of spontaneous crystallization. These homogeneous and heterogeneous crystallization temperatures depend on the droplet diameters. Their values given in Table 2, Column 10 are the lowest undercooling temperatures, which can be obtained with 10 micrometer droplets.
Figure 5. Condensation temperatures of 13-atom superclusters and crystallization temperatures in 10 micrometer droplets versus the melting entropy in J/K/mole. The squares are the reduced formation temperatures of 13-atom superclusters, which are ready to grow. These temperatures are larger than or equal to the reduced temperatures of spontaneous crystallization around them represented by diamond points.
Figure 5. Condensation temperatures of 13-atom superclusters and crystallization temperatures in 10 micrometer droplets versus the melting entropy in J/K/mole. The squares are the reduced formation temperatures of 13-atom superclusters, which are ready to grow. These temperatures are larger than or equal to the reduced temperatures of spontaneous crystallization around them represented by diamond points.
Metals 04 00359 g005
In Table 2, the liquid elements are still classified as a function of their fusion entropy ΔSm given in Table 1 (Column 3):
  • Column 1, List of liquid elements;
  • Column 2, Critical radius for spontaneous growth;
  • Column 3, The number Δz per atom of s-electrons virtually transferred from superclusters to melt at Tm;
  • Column 4, The energy saving coefficient εnm0 of 13-atom superclusters calculated with Equations (25)–(27);
  • Column 5, The energy saving coefficient εnm0 of 55-atom superclusters calculated with Equations (24)–(26);
  • Column 6, The energy saving coefficient εnm0 of 147-atom superclusters calculated with Equation (28) for n ≥ 147;
  • Column 7, The energy saving coefficient εnm0 of 309-atom superclusters;
  • Column 8, The energy saving coefficient εnm0 of 561-atom superclusters;
  • Column 9, The condensation reduced temperature of 13-atom superclusters in 10 micrometer droplets;
  • Column 10, The spontaneous growth reduced temperature around 13-atom superclusters in 10 micrometer droplets.
Table 2. Energy saving coefficients εnm0 of n-atom superclusters, and reduced condensation temperatures θ13c of 13-atom superclusters inducing spontaneous growth at θc = (TcTm)/Tm in 10 micrometer droplets.
Table 2. Energy saving coefficients εnm0 of n-atom superclusters, and reduced condensation temperatures θ13c of 13-atom superclusters inducing spontaneous growth at θc = (TcTm)/Tm in 10 micrometer droplets.
12345678910
R*2lsΔzεnm0εnm0εnm0εnm0εnm0θ13c (10 m)θc (10 m)
nm n = 13n = 55n = 147n = 309n = 561n = 13n = 13
Fe2.480.1070.670.8590.7090.5530.454−0.406−0.520
In3.180.0330.090.7070.7070.5520.452−0.570−0.560
Ti2.790.1340.860.8810.7000.5460.448−0.277−0.498
Zr3.050.1610.990.9000.6990.5460.447−0.192−0.487
Mn2.580.0980.640.8420.6980.5450.447−0.408−0.514
Pb3.320.0500.370.7860.6980.5450.446−0.513−0.534
Co2.290.1160.660.8150.6670.5210.427−0.370−0.490
Ag2.650.0940.620.8080.6670.5210.427−0.348−0.492
Au2.610.1030.670.8080.6600.5160.423−0.337−0.485
Tc2.410.1740.890.8430.6590.5150.422−0.195−0.466
Cr2.300.1500.800.8260.6560.5120.420−0.248−0.470
Re2.480.2581.050.8620.6560.5120.420−0.160−0.450
Ir2.460.2010.950.8490.6560.5120.420−0.094−0.458
Mo2.530.2201.000.8550.6560.5120.420−0.123−0.455
Os2.420.2500.760.8190.6560.5120.420−0.269−0.450
Pd2.520.1331.040.8610.6560.5120.420−0.098−0.472
Pt2.490.1540.850.8320.6550.5120.419−0.217−0.467
Cu2.300.0940.540.7810.6550.5110.419−0.397−0.490
Rh2.420.1640.860.8340.6540.5110.419−0.204−0.465
Ta2.700.2661.090.8640.6530.5100.418−0.071−0.445
Nb2.570.2161.000.8500.6520.5090.417−0.119−0.451
Hg2.810.0200.000.5250.6500.5070.416−0.551−0.526
V2.400.1640.850.8230.6460.5040.413−0.200−0.459
Ni2.210.1210.660.7910.6450.5030.413−0.318−0.472
Ru2.370.1900.910.8290.6440.5020.412−0.160−0.453
Hf2.830.2251.040.8460.6430.5020.412−0.087−0.442
Gaβ2.710.0210.000.5090.6410.5000.410−0.546−0.520
Cd2.410.0520.300.7100.6380.4980.409−0.479−0.496
Zn2.430.0550.260.6980.6370.4970.407−0.489−0.497
Al2.320.0810.490.7330.6190.4830.396−0.379−0.467
W2.670.3381.040.7970.5980.4670.383−0.030−0.412
Sn2.560.0580.370.6750.5870.4580.375−0.406−0.452
Bi2.530.0890.550.6280.5080.3960.325−0.218−0.382
Sb2.350.1470.700.6390.4970.3880.318−0.109−0.366
Te2.360.1320.650.6130.4790.3740.306−0.118−0.356
Se2.230.0940.510.5750.4640.3620.297−0.197−0.353
Si1.850.2900.770.5970.4500.3510.288−0.018−0.331
Ge1.920.2080.6950.5840.4470.3490.286−0.056−0.332

6. Maximum Superheating Temperatures of Superclusters at Constant Molar Volume

6.1. Superheating and Melting of n-atom Superclusters by Liquid Homogeneous Nucleation

N-atom superclusters survive above the melting temperature Tm up to an superheating temperature which is time-dependent. They can be melted by liquid homogeneous nucleation in their core instead of surface melting. The Gibbs free energy change associated with their melting at a temperature T > Tm is given by Equation (30):
Δ G nm ( n , θ , ε nm ) = Δ H m n N A ( θ ε nm ) + ( 4 π ) 1 / 3 N A Δ H m α 2ls ( 3 n ) 2 / 3
where the energy saving coefficient εnm is given in Equation (16) even for θ > 0. The fusion enthalpy has changed sign as compared to Equation (15) and the equalization of Fermi energies always still leads to an energy saving. An n-atom supercluster melts when Equation (31) is respected:
ln ( K sl . v n . t sn ) = Δ G nm k B T
where vn is the n-atom supercluster volume deduced from its radius given in Table 1, ΔGnm deduced from Equation (29) and tsn is the superheating time at its own melting temperature because the supercluster radius is much smaller than its critical radius. The time tsn is chosen equal to 600 s and lnKsl to 90.

6.2. Overheating and Melting of n-Atom Superclusters by Liquid Heterogeneous Nucleation

Melting temperatures of superclusters are reduced by previous melting of a 13-atom droplet in their core. These entities melt when Equation (32) is respected for n = 13:
ln ( K sl × v n × t sn ) = Δ G nm k B T Δ G 13m k B T
where the critical energy barrier ΔG*nm/kBT no longer exists and is replaced by ΔGnm/kBT, εnm in Equation (16), εnm0 in Table 2, tsn = 600 s and lnKls = 90. The critical barrier is not involved in Equation (32) because the n-atom supercluster radius is much smaller than the critical radius for liquid growth and ΔG*nm >> ΔGnm.

6.3. Prediction of Melting Temperatures of Superclusters in 38 Liquid Elements by Melt Superheating above Tm

The reduced melting temperatures θ = (TTm)/Tm of superclusters depending on their atom number n are given in several columns of Table 3 and in Figure 6. They are calculated assuming that the molar volume is constant, tsn = 600 s. and lnKsl = 90. The liquid elements having fusion entropy ΔSm larger than 20 J/K/mole have a melting temperature which is determined by liquid homogeneous nucleation because the 13-atom clusters melt at higher temperatures while those with ΔSm < 20 J/K/mole are submitted to chain-melting.
Figure 6. The melting temperatures of superclusters containing 13, 55, 147, 309 and 561 atoms. These melting temperatures are given in columns 10, 11, 12 and 13 of Table 3 versus ΔSm.
Figure 6. The melting temperatures of superclusters containing 13, 55, 147, 309 and 561 atoms. These melting temperatures are given in columns 10, 11, 12 and 13 of Table 3 versus ΔSm.
Metals 04 00359 g006
Table 3. The melting temperatures of superclusters. The final melting temperatures are given in Columns 2, 11, 12, 13 and 14. The temperatures in Columns 3, 4, 5 and 6 are calculated assuming that the melting starts from 13-atom droplets acting as heterogeneous nuclei in the core of superclusters. Those in Columns 7, 8, 9 and 10 correspond to a liquid homogeneous nucleation.
Table 3. The melting temperatures of superclusters. The final melting temperatures are given in Columns 2, 11, 12, 13 and 14. The temperatures in Columns 3, 4, 5 and 6 are calculated assuming that the melting starts from 13-atom droplets acting as heterogeneous nuclei in the core of superclusters. Those in Columns 7, 8, 9 and 10 correspond to a liquid homogeneous nucleation.
1234567891011121314
θfθf (n-13)θf (n-13)θf (n-13)θf (n-13)θf (n) Hom.θf (n) Hom.θf (n) Hom.θf (n) Hom.θf (n)θf (n)θf (n)θf (n)
n = 13n = 55n = 147n = 309n = 561n = 55n = 147n = 309n = 561n = 55n = 147n = 309n = 561
Fe00.0770.2740.2840.2460.3950.4500.3920.3110.0770.2740.2840.246
In00.1300.3060.3020.2570.3750.4430.3870.3070.1300.3060.3020.257
Ti0.0250.0500.2560.2710.2370.3960.4460.3850.3040.0500.2560.2710.237
Zr0.0540.0290.2430.2640.2320.3960.4430.3840.3020.0540.2430.2640.232
Mn00.0840.2740.2800.2410.3980.4480.3860.3040.0840.2740.2800.241
Pb00.1060.2860.2870.2450.3840.4400.3810.3010.1060.2860.2870.245
Co0.0450.0990.2700.2670.2270.4200.4460.3700.2860.0990.2700.2670.227
Ag0.0270.0980.2680.2660.2260.4120.4420.3670.2850.0980.2680.2660.226
Au0.0490.0950.2640.2620.2220.4160.4410.3640.2810.0950.2640.2620.222
Tc0.1070.0690.2510.2540.2170.4210.4420.3640.2810.0690.1070.2540.217
Cr0.0940.0850.2580.2570.2180.4240.4440.3630.2790.0940.2580.2570.218
Re0.1450.0470.2380.2460.2120.4220.4420.3620.2780.1450.2380.2460.212
Ir0.1260.0610.2460.2500.2150.4220.4420.3620.2790.1260.2460.2500.215
Mo0.1330.0540.2420.2480.2130.4220.4400.3610.2780.1330.2420.2480.213
Os0.0850.0870.2590.2570.2190.4220.4420.3620.2780.0870.2590.2570.219
Pd0.1450.0480.2380.2460.2130.4220.4400.3610.2780.1450.2380.2460.213
Pt0.1020.0750.2530.2530.2170.4220.4400.3610.2780.1020.2530.2530.217
Cu0.0330.1180.2750.2650.2230.4220.4430.3620.2780.1180.2750.2650.223
Rh0.1080.0740.2520.2530.2160.4230.4420.3610.2780.1080.2520.2530.216
Ta0.1520.0400.2330.2430.2100.4220.4380.3590.2760.1520.2330.2430.210
Nb0.1400.0550.2410.2460.2110.4230.4390.3580.2760.1400.2410.2460.211
Hg00.1670.3020.2790.2300.4070.4360.3560.2740.1670.3020.2790.230
V0.1200.080.2510.2490.2120.4270.4390.3560.2730.1200.2510.2490.212
Ni0.0820.1090.2650.2560.2160.4300.4400.3560.2730.1090.2650.2560.216
Ru0.1370.0730.2470.2460.2100.4290.4380.3540.2700.1370.2470.2460.210
Hf0.1530.0490.2340.2390.2060.4230.4340.3520.2700.1530.2340.2390.206
Gaβ00.1750.3020.2750.2260.4150.4360.3520.2690.1750.3020.2750.226
Cd00.1480.2810.2630.2180.4230.4330.3490.2670.1480.2810.2630.218
Zn00.1580.2860.2650.2190.4270.4360.3490.2670.1580.2860.2650.219
Al0.0820.1360.2660.2470.2050.4350.4280.3370.2560.1360.2660.2470.205
W0.2300.0720.2270.2200.1860.4450.4200.3230.2430.2300.2270.220.186
Sn0.1050.1580.2590.2330.1910.4430.4110.3140.2050.1580.2590.2330.191
Bi0.2750.1510.2140.1850.1510.4550.3580.2560.1900.2750.2750.2560.190
Sb0.3150.1360.2030.1770.1450.4540.3500.2480.1840.3150.3150.2480.184
Te0.3330.1430.1960.1680.1370.4490.3340.2350.1740.3330.3340.2350.174
Se0.3390.1600.1930.1630.1330.4480.3210.2250.1660.3390.3210.2250.166
Si0.3890.1360.1800.1540.1250.4420.3100.2150.1590.3890.310.2150.159
Ge0.3850.1420.1800.1520.1240.4400.3060.2120.1560.3850.3060.2120.156
In Table 3,
  • Column 1, The liquid elements are classified as a function of the fusion entropy ΔSm;
  • Column 2, The melting temperature of 13-atom superclusters is high for large fusion entropies of Bi, Sb, Te, Se, Si and Ge. The highest value θ = 0.389 is obtained for Si; the lowest θ = 0 is obtained in 8 liquid elements. Homogeneous liquids do not contain any condensed cluster. They are crystallizing during undercooling because 13-atom clusters are condensed as shown in Figure 5. Their previous presence at Tm does not have any influence on the undercooling rate;
  • Column 3, The calculated melting temperatures of 55-atom clusters induced by previous formation in their core of a droplet of 13 atoms are often lower than those of the 13-atom clusters. Then, they melt at the same temperature as the 13-atom clusters;
  • Column 4, The calculated melting temperatures of 147-atom clusters induced by previous formation in their core of a droplet of 13 atoms are larger than those of the 13-atom clusters from Fe to Al. They are nearly equal for W and smaller from Sn to Ge;
  • Column 5, The calculated melting temperatures of 309-atom clusters induced by previous formation in their core of a droplet of 13 atoms are larger those of the 13-atom clusters from Fe to Sn. They are smaller from Bi to Ge;
  • Column 6, The calculated melting temperatures of 561-atom clusters induced by previous formation in their core of a droplet of 13 atoms are larger than those of the13-atom clusters from Fe to Sn except W. They are smaller from Bi to Ge;
  • Columns 7, 8, 9 and 10, The melting temperatures of 55-, 147-, 309- and 561-atom superclusters are obtained considering homogeneous liquid nucleation without introducing heterogeneous nucleation from 13-atom droplets;
  • Columns 11, 12, 13 and 14, The expected melting temperatures of 55-, 147-, 309- and 561-atom superclusters are selected in order to be coherent between them. The homogeneous nucleation temperature of some superclusters having a large fusion entropy are sometimes smaller than those of the 13-atom superclusters, as shown in Figure 6.
All these results have been obtained assuming that the superheating time at their own melting temperature is 600 s. The time effects on copper supercluster melting are examined in part 9 in relation with detailed experimental studies [51].

7. Electronic Properties of Cu, Ag and Au Superclusters

Electronic properties of superclusters can be calculated from the enthalpy saving associated with their formation temperature in noble metallic liquids because this energy is due to Fermi energy equalization of liquid and superclusters [23]. The Fermi energy difference ΔEF between condensed superclusters of radius Rnm containing n atoms and liquid state at Tm can be directly evaluated for noble metals using Equation (33) and assuming that Δz is small, as shown in Table 2:
Δ E F 2 m n Δ z = n ε ls Δ H m N A
where m is the ratio of electron masses m*/m0, m0 being the electron rest mass and m* the effective electron mass which is assumed to be the same in superclusters and liquid states, and Δz being calculated at variable temperature using the known quantified energy saving εls in Equations (18), (26) and (27). The Fermi energy difference ΔEF is plotted in Figure 7 as a function of 1/R*2ls, where R*2ls is given in Equation (6) for Cu, Ag and Au assuming that the molar volume does not depend on temperature and a continuous variation of R*2ls. The quantified value εls is given in Equation (18) and the U0 and Δz values are calculated with Equation (26). For R* > 1 nm, EF is proportional to the Laplace pressure, while for R << 1 nm there is a gap opening in the conduction electron band accompanying the quantification of the energy saving. This analysis is able to detect well-known properties of clusters out of the melt, which become much less conducting at very low radii [10].
Figure 7. Fermi energy difference ΔEF between liquid and superclusters. The ΔEF in eV/mole is plotted as a function of the reverse of the critical radius R*2ls in nm−1.
Figure 7. Fermi energy difference ΔEF between liquid and superclusters. The ΔEF in eV/mole is plotted as a function of the reverse of the critical radius R*2ls in nm−1.
Metals 04 00359 g007
A strong variation of ΔEF at constant molar volume Vm is observed in Figure 7. In principle, the ΔEF has to obey Equation (34) in the liquid state because the Fermi energy EF depends on (Vm)−2/3:
Δ E F E F = 2 3 Δ V m V m
where EF is the Fermi energy of the liquid, Vm the molar volume of a supercluster of infinite radius, ΔVm is the variation of the molar volume with the radius decrease. The supercluster molar volume Vm has to depend on the particle radius instead of being constant. Equation (34) is respected when the formation temperature T of superclusters corresponding to the critical atom number nc in Equation (19) and to a molar volume Vm depending on R*2ls is introduced. The formation temperatures of superclusters with magic atom numbers are indicated in Figure 8 using a special molar volume thermal variation Vm(T) given in Figure 9 for each liquid element.
Figure 8. The formation temperatures of superclusters containing n atoms. The formation temperatures of Ag critical superclusters are plotted versus the critical number nc of atoms that they contain. The superclusters with magic atom numbers are represented by squares.
Figure 8. The formation temperatures of superclusters containing n atoms. The formation temperatures of Ag critical superclusters are plotted versus the critical number nc of atoms that they contain. The superclusters with magic atom numbers are represented by squares.
Metals 04 00359 g008
Figure 9. The supercluster molar volume change of Cu, Ag and Au. The supercluster molar volume change with the critical radius R*2ls (being a hidden variable) is plotted as a function of their formation temperature in Kelvin up to Tm. Each point corresponds to a supercluster of radius R*2ls and to an n-atom number.
Figure 9. The supercluster molar volume change of Cu, Ag and Au. The supercluster molar volume change with the critical radius R*2ls (being a hidden variable) is plotted as a function of their formation temperature in Kelvin up to Tm. Each point corresponds to a supercluster of radius R*2ls and to an n-atom number.
Metals 04 00359 g009
The following laws are used in Figure 8 and Figure 9:
V m ( m 3 ) = 7.37 × 10 6 × θ + 11.8 × 10 6
for Cu,
V m ( m 3 ) = 12.35 × 10 6 × θ + 17 × 10 6
for Ag and Au, where θ is equal to (TTm)/Tm. All superclusters containing magic atom numbers have their molar volume obeying these laws at their critical formation temperature. The molar volumes Vm for θ = 0.198 are maximum and equal to 13.4 × 10−6, 19.68 × 10−6, and 19.68 × 10−6 m3/mole, for Cu, Ag and Au respectively. They correspond to an infinite radius for superclusters in the absence of crystallization [15]. The molar volume Vm of bulk superclusters would be attained when θ − εnm becomes equal to zero using the critical radius as a hidden variable becoming infinite instead of the temperature.
The Fermi energy change ΔEF depends on Δz in Equation (33); Δz is calculated with Equations (25)–(27) for each radius R = R*2ls(T) in Equation (6), determining n from Equation (29). Equation (34) is now respected for Cu, Ag and Au, as shown in Figure 10. The Fermi energy EF is defined in Equation (37), assuming that there is one conduction electron per atom in Cu, Ag and Au:
E F = 2 2 m 0 ( 3 π 2 V m ) 2 / 3
where Vm is the liquid molar volume at Tm which is equal to 7.95 × 10−6, 11.5 × 10−6, and 11.3 × 10−6 m3/mole for Cu, Ag and Au respectively [64].
Figure 10. Relative Fermi energy change between superclusters and liquid. The ΔEF/EF is plotted as a function of the relative volume change ΔVm/Vm of superclusters, where Vm is the molar volume of the supercluster having an infinite radius.
Figure 10. Relative Fermi energy change between superclusters and liquid. The ΔEF/EF is plotted as a function of the relative volume change ΔVm/Vm of superclusters, where Vm is the molar volume of the supercluster having an infinite radius.
Metals 04 00359 g010
The molar electronic specific heat Cel = γel × T of Cu, Ag and Au superclusters can be obtained from the knowledge of their electronic density of states D(EF) at the Fermi level, calculated with Equations (38) and (39) [64]:
Δ E F m m D ( E F ) = N A Δ z
γ e l = π 2 3 D ( E F ) k B 2
The Δz values have been previously determined from Equations (25)–(27). Each n-atom supercluster has its own molar volume Vm and its own Δz at Tm is determined with Equation (25) = (26) with R = R*2ls (Tm) depending on Vm and εnm0 = 0.217. The electronic specific heat coefficient γel is plotted in Figure 11 as a function of the supercluster molar volume.
The electronic specific heat coefficient γel of superclusters falls when their molar volume Vm and their radius decrease below Tm. The coefficients γel of Cu, Ag and Au crystals at 4 K are a little larger, being equal to 0.695, 0.646 and 0.729 instead of 0.48, 0.547 and 0.599 mJ/K2/mole at Tm respectively [64]. Small crystals are known to become insulating for radii smaller than 5 nm when they are studied out of their melt [10]. This electronic transformation is also present in superclusters and is very abrupt below their critical growth volume at Tm as shown in Figure 11. The γel at Tm is also calculated as a function of the supercluster radius R and represented in Figure 12. The coefficient Δz is obtained at Tm with Equation (25) = (26) and εnm = εls0 = 0.217. Then, the potential energy U0 depending on R is known for each value of R and the quantified coefficient εnm0 of an n-atom supercluster of radius R is deduced from Equations (26) and (27). In Figure 12, the highest points are calculated at Tm while the lowest are already shown in Figure 11. The smallest superclusters are still metallic at Tm, while they become insulating when the temperature is close to Tm/3. All these predictions are in good agreement with many properties of divided metals. They are only based on an enthalpy saving equal to 0.217 × (1 − 2.25 × θ2) × ΔHm for the supercluster formation in all liquid elements.
Figure 11. Electronic specific heat of superclusters. The supercluster electronic specific heat coefficient in mJ/K2/mole is plotted versus their molar volume in m3 when the critical radius is smaller and smaller below Tm.
Figure 11. Electronic specific heat of superclusters. The supercluster electronic specific heat coefficient in mJ/K2/mole is plotted versus their molar volume in m3 when the critical radius is smaller and smaller below Tm.
Metals 04 00359 g011
Figure 12. Electronic specific heat coefficient of Cu, Ag, and Au superclusters as a function of supercluster radius R at T = Tm (colored points) and for T < Tm when R is the critical radius (black points).
Figure 12. Electronic specific heat coefficient of Cu, Ag, and Au superclusters as a function of supercluster radius R at T = Tm (colored points) and for T < Tm when R is the critical radius (black points).
Metals 04 00359 g012

8. Silver Supercluster Formation into and Out of Undercooled Liquid

The formation of icosahedral silver clusters with magic numbers n of atoms equal to 13, 55, 147, 309, 561, 923, 1415 and 2057 has been already studied out of liquid by molecular dynamics in the temperature range 0–1300 K. Icosahedral clusters of 13, 55 and 147 are formed below room temperature and larger clusters with n = 309, 561, 923, 1415 are formed from 300 to 1000 K. The radii of these Ag stable superclusters have been found to be equal to 2.74, 5.51, 8.32, 11.14 and 14.94 Å for n = 13, 55, 147, 309, 561 respectively [5]. The Ag radii have also been calculated in the liquid using the molar volume shown in Figure 9 and their formation temperature as deduced from the critical radius. Their values for n = 13, 55, 147, 309, 561 are nearly equal to those predicted by molecular dynamics, as shown in Table 4 and Figure 13.
Table 4. The Ag supercluster radii with magic atom numbers. The radius R is deduced from molar volume Vm and equal to critical radius R*2ls(T) given in Equation (6). For T > Tm/3 = 411.33 K, the energy saving coefficient εls in Equation (18) is used with εls0 = 0.217 and (θ0m)−2 = 2.25. For T < 411.33 K, εls is equal to zero. The radii RMD result from molecular dynamics simulations [5].
Table 4. The Ag supercluster radii with magic atom numbers. The radius R is deduced from molar volume Vm and equal to critical radius R*2ls(T) given in Equation (6). For T > Tm/3 = 411.33 K, the energy saving coefficient εls in Equation (18) is used with εls0 = 0.217 and (θ0m)−2 = 2.25. For T < 411.33 K, εls is equal to zero. The radii RMD result from molecular dynamics simulations [5].
n135514730956192314152057
R (Å)3.3875.4858.54111.63014.67017.68020.67023.660
RMD (Å)2.745.518.3211.1414.94
T (K)0291604817952104411081234
Figure 13. The critical atom number in blue versus the critical radius and RMD the radius calculated by molecular dynamics simulations in red square [5].
Figure 13. The critical atom number in blue versus the critical radius and RMD the radius calculated by molecular dynamics simulations in red square [5].
Metals 04 00359 g013

9. Melting of Cu, Ag and Au Superclusters Varying the Superheating

9.1. Overheating of Cu, Ag and Au Superclusters

The melting temperatures are now calculated using the molar volume associated with the supercluster radius as shown in Figure 9. The superheating time continues to be equal to 600 s. The supercluster radius variation is continuous while the radius of magic number clusters is indicated in Figure 14, Figure 15 and Figure 16. In these three figures, the Cu, Ag, and Au supercluster radius is plotted versus the reduced temperature θ = (TTm)/Tm. The points labeled “homogeneous” are calculated assuming that supercluster melting is produced by liquid homogeneous nucleation using Equations (30) and (31). The triangles labeled (n-13) are calculated assuming that the supercluster melting is induced by previous formation of liquid droplets of 13 atoms into superclusters using Equation (32). The homogeneous nucleation temperatures are much too high compared to the (n-13) temperatures. The undercooling temperatures depend on the volume sample v. The square points are determined for ln(Kls·v·tsn) = 71.8 corresponding to v·tsn = 12 × 10−9 m3·s and a heterogeneous nucleation induced by superclusters of radius R when the applied superheating temperature is smaller than those indicated by triangles. Another supercooling temperature represented by triangle points is added in Figure 15 and Figure 16. In Figure 15, v·tsn is equal to 7.08 × 10−22 m3·s while, in Figure 16, v·tsn = 15 × 10−7 m3·s. These three figures show that an undercooling rate of about 20% is generally obtained when the sample volume is of the order of a few mm3 and the applied superheating rate is less than about 25%. The undercooling temperature is very stable when the superheating is less than 25%. Larger undercooling rates are obtained using much smaller volume samples [17].
Figure 14. Supercooling temperatures of liquid copper controlled by unmelted superclusters having melting temperatures depending on overheating rate applied during 600 s.
Figure 14. Supercooling temperatures of liquid copper controlled by unmelted superclusters having melting temperatures depending on overheating rate applied during 600 s.
Metals 04 00359 g014
Figure 15. Supercooling temperatures of liquid silver depending on sample volume and controlled by unmelted superclusters having melting temperatures depending on overheating rate applied during 600 s.
Figure 15. Supercooling temperatures of liquid silver depending on sample volume and controlled by unmelted superclusters having melting temperatures depending on overheating rate applied during 600 s.
Metals 04 00359 g015
Figure 16. Supercooling temperatures of liquid gold depending on sample volume and controlled by unmelted superclusters having melting temperatures depending on overheating rate applied during 600 s.
Figure 16. Supercooling temperatures of liquid gold depending on sample volume and controlled by unmelted superclusters having melting temperatures depending on overheating rate applied during 600 s.
Metals 04 00359 g016

9.2. Analysis of the Influence of Cu Superheating Time on the Undercooling Rate

The superheating time has a strong influence on supercluster melting, as shown by studies of Cu undercooling [51]. It has been found that a minimum superheating temperature of 40 K is required in order to achieve any undercooling prior to crystallization nucleation. This phenomenon is also observed in many magnetic texturing experiments [65]. In Table 3, the first Cu supercluster to be melted at θ = 0.033 in 600 s, corresponding to a superheating of 44.7 K, contains 13 atoms in perfect agreement with the observation. There is no other supercluster melting. A temperature below which no small supercluster melts is predicted in this model. The lowest value of the undercooling temperature is obtained when 6 thermal cycles are applied prior to nucleation after 6 steps of 2400 s at 1473 K. The total time evolved at 1473 K is 14,400 s. In Figure 17, the time necessary to melt all superclusters surviving in copper melt is calculated. With lnKls = 89.26 instead of 90, the time to melt all the 13-atom clusters is 141 s, while that to melt 55-atom clusters is 14,541 s, which is in very good agreement with the measurements [51]. The other superclusters are melted in very short times after the melting of the 55-atom clusters. The reduced undercooling temperature becomes equal to θ = −0.259 after these thermal treatments of a sample of 5.7 mm in diameter. A homogeneous nucleation of 13- and 55-atom clusters leads to θ = −0.252. These results show that superclusters can be chain melted with a weaker superheating if the time evolved at the overheating temperature is increased substantially beyond 600 s. Our model can be used to evaluate the approximate superheating time leading to a thermodynamic equilibrium of a melt without condensed superclusters at any temperature above Tm.
Figure 17. Chain melting of superclusters versus superheating time. The first cluster to be melted at θ = 0.086 contains 13 atoms, the next ones 55, 147, 309 and 561 because the liquid droplet grows in the core of the largest particles.
Figure 17. Chain melting of superclusters versus superheating time. The first cluster to be melted at θ = 0.086 contains 13 atoms, the next ones 55, 147, 309 and 561 because the liquid droplet grows in the core of the largest particles.
Metals 04 00359 g017

10. Conclusions

The undercooling temperatures of 32 of the 38 liquid elements are predicted for the first time in good agreement with experimental values depending on the sample volume, without using any adjustable barrier energy, and only assuming the existence of growth nuclei containing stable magic atom numbers n equal to 13, 55, 147, 309 and 561 that are generally devoted to icosahedral structures. The model is based on a volume enthalpy saving εv previously determined to be equal to 0.217 × ΔHm/Vm at Tm and added to the classical Gibbs free energy change for a critical nucleus formation in a melt. This enthalpy is due to the Laplace pressure change Δp acting on the growth nuclei and equalizing the Fermi energies of liquid and nuclei in metallic liquids. The Gibbs free energy change has to contain a contribution −Vm × Δp which has been neglected up to now because its magnitude was unknown. This missing enthalpy has serious consequences because the critical radius for crystal growth is considered, in the classical view, as being infinite at the melting temperature and all solid traces being eliminated in melts. This is in contradiction with many experiments on the superheating influence on undercooling rates and on magnetic texturing efficiency [27,51,65,66,67,68]. Nuclei having radii smaller than the critical radius at Tm are melted at higher temperatures depending on the superheating time and on their atom number.
Some growth nuclei survive above Tm because they are superclusters that are not melted by surface melting. This new property of superclusters is a consequence of the thermal variation of εv, which is a unique function of θ2 = [(TTm)/Tm]2 being maximum at Tm, and a fusion heat equal to that of bulk crystals. The surface atom fusion heat is not weakened and there is no premelting of these entities depending on their radius. This thermal variation was established, for the first time, from our study of the maximum undercooling rate of the same liquid elements. In addition, it is the only law validating the existence of non-melted intrinsic entities.
The energy saving is proportional to the supercluster reverse radius R−1 when n ≥ 147 and is quantified for n < 147. The quantified energy at Tm is calculated by creating a virtual s-electron transfer from the nucleus of radius R to the melt and an electrostatic spherical potential induced by the surface charges and also varying with R−1. The Schrödinger equation solutions are known and used to predict the condensation temperatures of 13-atom superclusters in undercooled melts, which govern the crystallization temperatures of liquids having fusion entropy larger than 20 J/K/mole.
The superclusters are melted by homogeneous or heterogeneous liquid nucleation in their core. The liquid homogeneous nucleation is effective in all superclusters when ΔSm ≥ 20 J/K/mole while a chain melting is produced, starting with a 13-atom droplet induced in the core of the supercluster and being magnified with the time increase at the superheating temperature. The model is able to predict an approximate value of the minimum time necessary to melt superclusters and to attain the true thermodynamic equilibrium of the melt at any superheating temperature.
The electronic specific heat of superclusters submitted to Laplace pressure in metals is determined for the first time from the enthalpy saving deduced from undercooling experiments. It strongly declines with radius as compared to that of a bulk metal, in agreement with the conductance properties of tiny clusters having radii smaller than 5 Å. The electronic s-state density of superclusters is greatly weakened compared to that of bulk crystals when their radius decreases. The supercluster critical radii deduced from the nucleation model are in quantitative agreement with recent molecular dynamics simulations devoted to Ag cluster radii.
The transformation of superclusters in crystals occurs for a radius between the critical radius for crystal growth and that for supercluster growth because the superclusters have a much lower density than crystals. The Gibbs free energy change from the liquid state to crystal becomes smaller than that of the supercluster just above its maximum at the crystal critical radius.

Acknowledgments

Thanks are due to Bernard Dreyfus and Jacques Friedel. The idea of this work started 50 years ago with criticisms of Bernard Dreyfus who considered that the nucleation equation would have to depend on the Fermi energy of conduction electrons. My first model developed in 2006, introduces an energy saving in the classical Gibbs free energy change for a nucleus formation that I had attributed to a transfer of conduction electrons from nucleus to liquid. Jacques Friedel noted that this transfer does not exist. I have proposed in 2011 that a Laplace pressure change acting on growth nuclei induces the enthalpy saving and the equalization of Fermi energies without electron transfer. A virtual transfer is nevertheless considered to quantify the energy saving in agreement with undercooling rates. The author thanks Andrew Mullis for the confirmation that the copper sample weights used in [51] are equal to 0.6 to 1 g.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Frank, F.C. Supercooling of liquids. Proc. R. Soc. Lond. 1952, A215, 43–46. [Google Scholar] [CrossRef]
  2. Mauro, N.A.; Bendert, J.C.; Vogt, A.J.; Gewin, J.M.; Kelton, K.F. High energy X-ray scattering studies of the local order in liquid Al. J. Chem. Phys. 2011, 135, 044502. [Google Scholar] [CrossRef] [PubMed]
  3. Tanaka, H. Relationship among glass-forming ability, fragility, and short-range bond ordering of liquids. J. Non-Cryst. Solids 2005, 351, 678–690. [Google Scholar]
  4. Shen, Y.T.; Kim, T.H.; Gangopadhyay, A.K.; Kelton, K.F. Icosahedral order, frustration, and the glass transition: Evidence from time-dependent nucleation and supercooled liquid structure studies. Phys. Rev. Lett. 2009, 102, 057801. [Google Scholar] [CrossRef] [PubMed]
  5. Kuzmin, V.I.; Tytik, D.L.; Belashchenko, D.K.; Sirenko, A.N. Structure of silver clusters with magic numbers of atoms by data of molecular dynamics. Colloid J. 2008, 70, 284–296. [Google Scholar] [CrossRef]
  6. Wang, Y.; Teitel, S.; Dellago, C. Melting of icosahedral gold nanoclusters from molecular dynamics simulations. J. Chem. Phys. 2005, 122, 21722. [Google Scholar]
  7. Gafner, Y.Y.; Gafner, S.L.; Entel, P. Formation of an icosahedral structure during crystallization of nickel nanoclusters. Phys. Sol. State 2004, 46, 1327–1330. [Google Scholar] [CrossRef]
  8. Cleveland, C.L.; Luedtke, W.D.; Landman, U. Melting of gold clusters: Icosahedral precursors. Phys. Rev. Lett. 1998, 81, 2036–2039. [Google Scholar] [CrossRef]
  9. Aiyer, H.N.; Vijayakrishnan, V.; Subbana, G.N.; Rao, C.N.R. Investigations of Pd clusters by the combined use of HREM, STM, high-energy spectroscopies and tunneling conductance measurements. Surf. Sci. 1994, 313, 392–398. [Google Scholar] [CrossRef]
  10. Vinod, C.P.; Kulkarni, G.U.; Rao, C.N.R. Size-dependent changes in the electronic structure of metal clusters as investigated by scanning tunneling spectroscopy. Chem. Phys. Lett. 1998, 289, 329–333. [Google Scholar] [CrossRef]
  11. Van Leeuwen, D.A.; van Ruitenbeck, J.M.; Schmid, G.; de Jongh, L.J. Size-dependent magnetisation of Pd clusters and colloids. Phys. Lett. A 1992, 170, 325–333. [Google Scholar] [CrossRef]
  12. Harbola, M.K. Magic numbers for metallic clusters and the principle of maximum hardness. Proc. Natl. Acad. Sci. USA 1992, 89, 1036–1039. [Google Scholar] [CrossRef] [PubMed]
  13. Zhdanov, V.P.; Schwind, M.; Zoric, I.; Kasemo, B. Overheating and undercooling during melting and crystallization of metal nanoparticles. Phys. E 2010, 42, 1990–1994. [Google Scholar] [CrossRef]
  14. Kusche, R.; Hippler, T.; Schmidt, M.; von Issendorf, B.H.; Haberland, H. Melting of free sodium clusters. Eur. Phys. J. D 1999, 9, 1–4. [Google Scholar]
  15. Tournier, R.F. Presence of intrinsic growth nuclei in overheated and undercooled liquid elements. Phys. B Condens. Matt. 2007, 392, 79–91. [Google Scholar] [CrossRef]
  16. Tournier, R.F. Tiny Crystals Surviving above the melting temperature and acting as growth nuclei of the high-Tc superconductor microstructure. Mater. Sci. Forum 2007, 546–549, 1827–1840. [Google Scholar]
  17. Perezpezko, J.H. Nucleation in undercooled liquids. Mater. Sci. Eng. 1984, 65, 125–135. [Google Scholar] [CrossRef]
  18. Gutzow, I.; Schmeltzer, J. The Vitreous State/Thermodynamics, Structure, Rheology and Crystallization; Springer-Verlag: Berlin/Heidelberg, Germany; New York, NY, USA, 1995; ISBN:3-540-59087-0. [Google Scholar]
  19. Turnbull, D. Kinetics of solidification of supercooled liquid mercury droplets. J. Chem. Phys. 1952, 20, 411–424. [Google Scholar] [CrossRef]
  20. Wu, D.T.; Granasy, L.; Spaepen, F. Nucleation and the solid-liquid interfacial free energy. MRS Bull. 2004, 29, 945–950. [Google Scholar] [CrossRef]
  21. Tournier, R.F. Thermodynamic origin of the vitreous transition. Materials 2011, 4, 869–892. [Google Scholar] [CrossRef]
  22. Tournier, R.F. Expected properties of gold melt containing intrinsic nuclei. In Proceedings of the 6th International Conference on Electromagnetic Processing of Materials (EPM), Dresden, Germany, 19–23 October 2009; Forschungszentrum Dresden-Rossendorf: Dresden, Germany, 2009; pp. 304–307. [Google Scholar]
  23. Tournier, R.F. Nucleation of crystallization in titanium and vitreous state in glass-forming melt. In Proceedings of the 12th World Conference on Titanium (Ti-2011), Beijing, China, 19–24 June 2011; Chang, H., Lu, Y., Xu, D., Zhou, L., Eds.; Science Press: Beijing, China, 2012; Volume II, pp. 1527–1531. [Google Scholar]
  24. Tournier, R.F. Crystal growth nucleation and Fermi energy equalization of intrinsic growth nuclei in glass-forming melts. Sci. Technol. Adv. Mater. 2009, 10, 014607. [Google Scholar] [CrossRef]
  25. Tournier, R.F. Thermodynamics of the vitreous transition. Rev. Metall. 2012, 109, 27–33. [Google Scholar] [CrossRef]
  26. Vinet, B.; Cortella, L.; Favier, J.J.; Desré, P.J. Highly undercooled W and Re drops in an ultra-high vacuum drop tube. Appl. Phys. Lett. 1991, 58, 97–99. [Google Scholar] [CrossRef]
  27. Tournier, R.F.; Beaugnon, E. Texturing by cooling a metallic melt in a magnetic field. Sci. Technol. Adv. Mater. 2009, 10, 014501. [Google Scholar] [CrossRef]
  28. Vinet, V.; Magnusson, L.; Frederiksson, H.; Desre, P.J. Correlations between surface and interface energies with respect to crystal nucleation. J. Coll. Interf. Sci. 2002, 255, 363–374. [Google Scholar] [CrossRef]
  29. Kelton, K.F.; Lee, G.W.; Gangopadhyay, A.K.; Hyers, R.W.; Rathz, T.J.; Rogers, J.R.; Robinson, M.B.; Robinson, D.S. First X-ray scattering studies on electrostatically levitated metallic liquids: Demonstrated influence of local icosahedral order on the nucleation barrier. Phys. Rev. Lett. 2003, 90, 195504. [Google Scholar] [CrossRef] [PubMed]
  30. Turnbull, D.; Cech, R.E. Microscopic observation of the solidification of small metal droplets. J. Appl. Phys. 1950, 21, 804–810. [Google Scholar] [CrossRef]
  31. Tournier, R.F. Crystal growth nucleation and equalization of Fermi energies of intrinsic nuclei and glass-forming melts. J. Conf. Ser. 2009, 144, 012116. [Google Scholar] [CrossRef]
  32. Landau, L.; Lifchitz, E. Physique Théorique: Mecanique Quantique; MIR: Moscow, Russia, 1966; Volume III, pp. 1–135. [Google Scholar]
  33. Touloukian, Y.S.; Kirby, R.K.; Taylor, R.E.; Desai, P.D. Thermal Expansion: Metallic Elements and Alloys: Thermophysical Properties of Matter; Plenum Press: New York, NY, USA, 1970; Volume 12. [Google Scholar]
  34. Schade, J.; MacLean, A.; Miller, W.A. Undercooled Alloy Phases; Collings, E.W., Koch, C.C., Eds.; The Metallurgical Society: Warrendale, PA, USA, 1986; pp. 1–233. [Google Scholar]
  35. Dukhin, I. Sb Problemi Metallovedenia I Fiziki Metallov; [Collection: Problems of physical metallurgy and metal physics]; Metallurgizdat: Moscow, Russia, 1959; pp. 1–9. [Google Scholar]
  36. Turnbull, D. Undercoolability and the Exposure of Metastable Structures; Collings, E.W., Koch, C.C., Eds.; The Metallurgical Society: Warrendale, PA, USA, 1986; pp. 3–22. [Google Scholar]
  37. Perepezko, J.H.; Paik, J.S. Rapidly Solidified Amorphous and Crystalline Alloys; Keer, B.H., Griessen, B.C., Cohen, M., Eds.; North Holland: Amsterdam, The Netherlands, 1982; pp. 49–63. [Google Scholar]
  38. Hofmeister, W.H.; Robinson, M.B.; Bayuzick, R.J. Undercooling of pure metals in a containerless microgravity environment. Appl. Phys. Lett. 1986, 49, 1342–1344. [Google Scholar] [CrossRef]
  39. Herlach, D.M. Containerless undercooling ans solidification of pure metals. Ann. Rev. Mater. Sci. 1991, 21, 23–44. [Google Scholar] [CrossRef]
  40. Morton, C.W.; Hofmeister, W.H.; Bayuzick, R.J.; Rulison, A.J.; Watkins, J.L. The kinetics of solid nucleation in zirconium. Acta Mater. 1998, 46, 6033–6039. [Google Scholar] [CrossRef]
  41. Vonnegut, B. Variation with temperature of the nucleation rate of supercooled liquid tin and water drops. J. Coll. Sci. 1948, 3, 563–569. [Google Scholar] [CrossRef]
  42. Perepezko, J.H.; Follstaedt, D.M.; Peercy, P.S. Nucleation of allotropic phases during pulsed laser annealing of manganese. MRS Proc. 1985, 51. [Google Scholar] [CrossRef]
  43. Skripov, V.P. Homogeneous Nucleation in Melts and Amorphous Films: Crystal growth and materials; Kaldis, E., Scheel, H.J., Eds.; North Holland: Amsterdam, The Netherlands, 1977; pp. 328–376. [Google Scholar]
  44. Wilde, G.; Sebright, J.L.H.; Perepezko, J. Bulk liquid undercooling and nucleation in gold. Acta Mater. 2006, 54, 4759–4769. [Google Scholar] [CrossRef]
  45. Cortella, L.; Vinet, B.; Desré, P.J.; Pasturel, A.; Paxton, A.T.; von Schilfgaarde, M. Evidence of transitory metastable phases in refractory metals solidified from highly undercooled liquids in a drop tube. Phys. Rev. Lett. 1993, 70, 1469–1472. [Google Scholar] [CrossRef] [PubMed]
  46. Ishikawa, T.; Paradis, P.F.; Fujii, R.; Saita, Y.; Yoda, S. Thermophysical property measurements of liquid and supercooled iridium by containerless methods. Int. J. Thermophys. 2005, 26, 893–904. [Google Scholar] [CrossRef]
  47. Paradis, P.F.; Ischikawa, T.; Yoda, A. Noncontact measurements of thermophysical properties of molybdenum at high temperatures. Int. J. Thermophys. 2002, 21, 555–569. [Google Scholar] [CrossRef]
  48. Paradis, P.F.; Ishikawa, T.; Korke, N. Physical properties of equilibrium and non-equilibrium liquid osmium measured by levitation technique. J. Appl. Phys. 2006, 100, 103523. [Google Scholar] [CrossRef]
  49. Fehling, J.; Scheil, E. Untersuchung der unter Kuhlbarkeit von metallschmelzen. Zeit. für Metalkunde 1962, 53, 593–600. [Google Scholar]
  50. Gragnevski, K.I.; Mullis, A.M.; Cochrane, R.F. The mechanism for spontaneous grain refinement in undercooled pure Cu melts. Mater. Sci. Eng. A 2004, 375–377, 479–484. [Google Scholar]
  51. Gragnevski, K.I.; Mullis, A.M.; Cochrane, R.F. The effect of experimental variables on the level of melt undercooling. Mater. Sci. Eng. A 2004, 375–377, 485–487. [Google Scholar]
  52. Li, D.; Eckler, K.; Herlach, D.M. Development of grain structures in highly undercooled germanium and copper. J. Cryst. Growth 1996, 160, 59–65. [Google Scholar] [CrossRef]
  53. Cortella, L.; Vinet, B. Undercooling and nucleation studies on pure refractory metals processed in the Grenoble high-drop tube. Phil. Mag. B 1995, 71, 16–21. [Google Scholar] [CrossRef]
  54. Bosio, L.; Defrain, A.; Epelboin, I. Changements de phase du gallium à la pression atmosphérique. J. Phys. Fr. 1966, 27, 61–71. (In French) [Google Scholar]
  55. Perepezko, J.H.; Sebright, J.L.; Hockel, P.G.; Wilde, G. Undercooling and solidification of atomized liquid droplets. Mater. Sci. Eng. A 2002, 326, 144–153. [Google Scholar] [CrossRef]
  56. Paradis, P.F.; Ishikawa, T.; Fujii, R.; Yoda, S. Physical properties of liquid and undercooled tungsten by levitation technique. Appl. Phys. Lett. 2005, 86, 041901. [Google Scholar] [CrossRef]
  57. Yoon, W.; Paik, J.S.; Lacourt, D.; Perepezko, J.H. The effect of pressure on phase selection during nucleation in undercooled bismuth. J. Appl. Phys. 1986, 60, 3489. [Google Scholar] [CrossRef]
  58. Payne, W.P.; Olson, J.K.; Allen, A.; Kozhevnikov, V.F.; Taylor, P.C. Sound velocity in liquid and glassy selenium. J. Non-Cryst. Sol. 2007, 353, 3254–3259. [Google Scholar]
  59. Aoyama, T.; Paradis, P.F.T.; Ishikawa, S.; Yoda, S. Observation of rapid solidfication of deeply undercooled Si melt using electrostatic levitation. Mater. Sci. Eng. A 2004, 375–377, 460–463. [Google Scholar]
  60. Beaudhuin, M.; Zaidat, K.; Duffar, T.; Lemiti, M. Silicon controlled under electromagnetic levitation. J. Mater. Sci. 2010, 45, 2218–2222. [Google Scholar] [CrossRef]
  61. Li, D.; Herlach, D.M. High undercooling of bulk molten silicon by containerless processing. Europhys. Lett. 1996, 34, 423–428. [Google Scholar] [CrossRef]
  62. Liu, R.P.; Volksmann, T.; Herlach, D.M. Undercooling of molten silicon by containerless processing. Acta Mater. 2001, 49, 439–444. [Google Scholar] [CrossRef]
  63. Menoni, C.S.; Hu, J.Z.; Spain, I. Germanium at high pressures. Phys. Rev. B 1986, 34, 362–368. [Google Scholar] [CrossRef]
  64. Kittel, C. Introduction to Solid State Physics; John Wiley & Sons: New York, NY, USA, 1967; pp. 197–223. [Google Scholar]
  65. Tong, H.Y.; Shi, F.G. Dependence of supercooling of a liquid on its overheating. J. Chem. Phys. 1997, 107, 7964–7966. [Google Scholar] [CrossRef]
  66. Rudolph, P.; Koh, H.J.; Schäfer, N.; Fukuda, T. The crystal perfection depends on the superheating of the mother phase too-experimental facts and speculations on the “melt structure” of semiconductor compounds. J. Cryst. Growth 1996, 166, 578–582. [Google Scholar] [CrossRef]
  67. Hays, C.C.; Johnson, W.L. Undercooling of bulk metallic glasses processed by electrostatic levitation. J. Non-Cryst. Solids 1999, 250–252, 596–600. [Google Scholar]
  68. Porcar, L.; de Rango, P.; Bourgault, D.; Tournier, R. Superconductors-Materials, Properties and Applications: Magnetic Texturing of High-Tc Superconductors; Gabovitch, A., Ed.; Intech: Rijeka, Croatia, 2012; pp. 171–196. [Google Scholar]

Share and Cite

MDPI and ACS Style

Tournier, R.F. Crystallization of Supercooled Liquid Elements Induced by Superclusters Containing Magic Atom Numbers. Metals 2014, 4, 359-387. https://doi.org/10.3390/met4030359

AMA Style

Tournier RF. Crystallization of Supercooled Liquid Elements Induced by Superclusters Containing Magic Atom Numbers. Metals. 2014; 4(3):359-387. https://doi.org/10.3390/met4030359

Chicago/Turabian Style

Tournier, Robert F. 2014. "Crystallization of Supercooled Liquid Elements Induced by Superclusters Containing Magic Atom Numbers" Metals 4, no. 3: 359-387. https://doi.org/10.3390/met4030359

Article Metrics

Back to TopTop