Next Article in Journal
Modeling Structural Dynamics Using FE-Meshfree QUAD4 Element with Radial-Polynomial Basis Functions
Next Article in Special Issue
Prediction of Second Melting Temperatures Already Observed in Pure Elements by Molecular Dynamics Simulations
Previous Article in Journal
Clinching of Thermoplastic Composites and Metals—A Comparison of Three Novel Joining Technologies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Building and Breaking Bonds by Homogenous Nucleation in Glass-Forming Melts Leading to Transitions in Three Liquid States

by
Robert F. Tournier
1,* and
Michael I. Ojovan
2,3
1
LNCMI-EMFL, CNRS, Université Grenoble Alpes, INSA-T, UPS, 38042 Grenoble, France
2
Department of Materials, Imperial College London, London SW7 2AZ, UK
3
Department of Radiochemistry, Lomonosov Moscow State University, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Materials 2021, 14(9), 2287; https://doi.org/10.3390/ma14092287
Submission received: 26 March 2021 / Revised: 21 April 2021 / Accepted: 26 April 2021 / Published: 28 April 2021
(This article belongs to the Special Issue Glass Science and First-Order Transitions at a Turning Point)

Abstract

:
The thermal history of melts leads to three liquid states above the melting temperatures Tm containing clusters—bound colloids with two opposite values of enthalpy +Δεlg × ΔHm and −Δεlg × ΔHm and zero. All colloid bonds disconnect at Tn+ > Tm and give rise in congruent materials, through a first-order transition at TLL = Tn+, forming a homogeneous liquid, containing tiny superatoms, built by short-range order. In non-congruent materials, (Tn+) and (TLL) are separated, Tn+ being the temperature of a second order and TLL the temperature of a first-order phase transition. (Tn+) and (TLL) are predicted from the knowledge of solidus and liquidus temperatures using non-classical homogenous nucleation. The first-order transition at TLL gives rise by cooling to a new liquid state containing colloids. Each colloid is a superatom, melted by homogeneous disintegration of nuclei instead of surface melting, and with a Gibbs free energy equal to that of a liquid droplet containing the same magic atom number. Internal and external bond number of colloids increases at Tn+ or from Tn+ to Tg. These liquid enthalpies reveal the natural presence of colloid–colloid bonding and antibonding in glass-forming melts. The Mpemba effect and its inverse exist in all melts and is due to the presence of these three liquid states.

Graphical Abstract

1. Introduction

Glass-forming melt transformations have been mainly studied, for many years, around the glass transition temperature Tg and sometimes up to the liquidus temperature Tliq. The liquid properties are often neglected because the classical nucleation equation predicts the absence of growth nuclei and nucleation phenomenon above the melting temperature. The presence of growth nuclei above Tm being known [1,2,3], an additional enthalpy is added to this equation to explain these observations. A new model of nucleation is built from the works of Turnbull’s [4] characterized by two types of homogeneous nucleation temperatures below and above Tm. The new additional enthalpy is a quadratic function of the reduced temperature θ = (T − Tm)/Tm as shown by a revised study of the maximum undercooling rate of 38 liquid elements using Vinet’s works [5,6]. A concept of two liquids is later introduced to explain the glass phase formation at Tg by an enthalpy decrease from liquid 1 to liquid 2 at this temperature. New laws minimizing the numerical coefficients of each quadratic equation are established determining the enthalpies εls(0) × ΔHm of liquid 1 and εgs(0) × ΔHm of liquid 2 for each θ value, with ΔHm being the melting enthalpy [7,8]. The thermodynamic transition at Tg is characterized by a second-order phase transition and a heat capacity jump defined by the derivative of the difference (εls(θ) − εgs(θ)) ΔHm which is equal to 1.5 ΔSm for many glass transitions with ΔSm being the melting entropy [9].
The glass transition results from the percolation of superclusters formed during cooling below Tm [10,11,12]. A thermodynamic transition characterized by critical parameters occurs by breaking bonds (configurons) and when the percolation threshold of configurons is attained [13,14,15,16,17]. Building bonds by enthalpy relaxation below Tg has for consequence the formation of a hidden undercooled phase called phase 3 with an enthalpy (εls(θ) − εgs(θ)) ΔHm equal to that of configurons with a residual bond fraction which can be overheated up to Tn+ > Tm before being melted [18]. The homogeneous nucleation temperature at Tn+ occurs in overheated liquids and is predicted for many molecular and metallic glass-forming melts.
This paper is devoted to phase transitions above Tm completing our recent work, showing that the dewetting temperatures of prefrozen and grafted layers in ultrathin films are equal to Tn+ [19]. The latent heats are exothermic or endothermic without knowing the explanation. The existence of a first-order transition is claimed for Pd42.5Ni42.5P15 and La50Al35Ni15 liquid alloys [20,21]. Our nucleation model of melting the liquid mean-range order by breaking residual bonds predicted all values of Tn+ and exothermic enthalpies at this temperature. The observation of endothermic latent heats showed the existence of three liquid states at Tm, the first one with a positive enthalpy εgs(0) × ΔHm, the second one zero, and the third one −εgs(0) × ΔHm, which is negative. The liquid is homogeneous above Tn+ when its enthalpy is equal to zero. The existence of various liquid states was also predicted without using a non-classical nucleation equation [22]. The formation temperature of a homogeneous liquid state was observed by measuring the density or the viscosity during heating and cooling, determining the point where the branching of these quantities disappears. Colloidal states were observed below this homogenization temperature and composed of thousands of atoms defining liquid heterogeneities [23,24,25,26,27]. Our objectives were to predict all these phase transitions.

2. The Homogeneous Nucleation

The Gibbs free energy change for a nucleus formation in a melt was given by Equation (1) [6,9]:
Δ G ls = ( θ ε ls ) Δ H m / V m × 4 π R 3 / 3 + 4 π R 2 σ ls
where R is the nucleus radius and following Turnbull [4], σls its surface energy, given by Equation (2), θ the reduced temperature (T − Tm)/Tm, ΔHm the melting enthalpy at Tm, and Vm the molar volume:
σ ls ( V m / N A ) 1 / 3 =   α ls Δ H m / V m
A complementary enthalpy −εls × ΔHm/Vm was introduced, authorizing the presence of growth nuclei above Tm. The classical nucleation equation was obtained for εls = 0.
The critical radius R ls * in Equation (3) and the critical thermally activated energy barrier Δ G ls * k B T in Equation (4) are calculated assuming dεls/dR = 0:
R ls * = 2 α ls ( θ ε ls ) ( V m N A ) 1 / 3
Δ G ls * k B T = 16 π Δ S m   α ls 3 3 N A k B ( 1 + θ ) ( θ ε ls ) 2
These critical parameters are not infinite at the melting temperature Tm because εls is not equal to zero. The nucleation rate J = Kvexp(− Δ G ls * k B T ) is equal to 1 when Equation (5) is respected:
Δ G ls * / k B T = ln ( K v )
The surface energy coefficient αls in Equation (2) is determined from Equations (4) and (5) and given by Equation (6):
α ls 3 = 3 N A k B   ( 1 + θ ) ( θ ε ls ) 2 16 π Δ S m ln ( K v )
The nucleation temperatures θn obtained for d α ls 3 /dθ = 0 obeys (7):
d α ls 3 / d θ   ~ ( θ n + ε ls ) ( 3 θ n + 2 ε ls ) = 0
In addition to the nucleation temperature Tn- below Tm, the existence of homogeneous nucleation up to Tn+ above Tm was confirmed by many experiments, observing the undercooling versus the overheating rates of liquid elements and CoB alloys [28,29]. This nucleation temperature could have, for consequence, the possible existence of a second melting temperature of growth nuclei above Tm and of their homogeneous nucleation at temperatures weaker than θn+.
The coefficient εls of the initial liquid called liquid 1 is a quadratic function of θ in Equation (8) [6]:
ε ls = ε ls 0 ( 1 θ 2 / θ 0 m 2 )
where θ0m is the Vogel–Fulcher–Tammann-reduced temperature leading to εls = 0 for θ = θ0m, the VFT temperature T0m of many fragile liquids being equal to ≌0.77 Tg. This quasi-universal value is known for numerous liquids including atactic polymers [30,31].
New liquid states are obtained for θ = θn+ = εls and θ = θn− = (εls − 2)/3 with Equation (7). The reduced nucleation temperatures θn− are solutions of the quadratic Equation (9):
ε ls 0 θ n 2 / θ 0 m 2 + 3 θ n + 2 ε ls 0 = 0
There is a minimum value of εls0 plotted as function of θ0m using (8) and θn− = (εls − 2)/3, determining the relation (10) between θ20m and εls0 for which the two solutions of (9) are equal in the two fragile liquids [8,32]. These values defined the temperature where the surface energy was minimum and θ20m and εls0 obeyed Equations (10) and (11):
θ 0 m 2 = 8 9 ε ls 0 4 9 ε ls 0 2 ,
ε ls ( θ = 0 ) = ε ls 0 = 1.5 θ n + 2 = a θ g + 2
The value a = 1 in the Equation (10) leads to T0m = 0.769 × Tg in agreement with many experimental values [9].
All melts and even liquid elements underwent, in addition, a glass transition because another liquid 2 existed characterized by an enthalpy coefficient εgs given by Equation (12), inducing an enthalpy change from that of liquid 1 at the thermodynamic transition at Tg [7,9,32]:
ε gs = ε gs 0 ( 1 θ 2 / θ 0 g 2 )
θ 0 g 2 = 8 9 ε gs 0 4 9 ε gs 0 2
ε gs ( θ = 0 ) = ε gs 0 = 1.5 θ n + 2 = 1.5 θ g + 2
The difference Δεlg in the Equation (15) between the coefficients εls and εgs determines the phase 3 enthalpy when the quenched liquid escapes crystallization:
Δ ε lg ( θ ) = ε ls ε gs = ε ls 0 ε gs 0 + Δ ε θ 2 ( ε ls 0 θ 0 m 2 ε gs 0 θ 0 g 2 )
The coefficient Δ ε lg ( θ )   defined a new liquid phase called phase 3 undergoing a hidden phase transition below Tg and a visible one at θn+, occurring for Δ ε lg ( θ ) = θn+, as shown by Equation (7). This transition was accompanied by an exothermic latent heat equal to Δ ε lg ( θ ) × Δ H m corresponding to about 15% of the melting heat [18]. Phase 3 was detected for the first time in supercooled water and associated with glacial phase formation [33,34,35] and recently appears as being associated with configuron formation [13,14,15,16,17,18]. The concept of configurons was initially proposed for materials with covalent bonds which can be either intact or broken [36]; then, it was extended to other systems including metallic systems based on ideas of Egami on bonds between nearest atoms in metals [16,37]. Thus, it is generically assumed that the set of bonds in condensed matter has two states; namely, the ground state corresponding to unbroken bonds and the excited state corresponding to broken bonds. The set of bonds in condensed matter is described in such a way by the statistics of a two-level system [38,39] which are separated by the energy interval Gd. The two approaches converge because the Gibbs free energy of phase 3 is equal to Gd. Phase 3 is assumed to be the configuron phase which is preserved above Tm in a liquid with medium-range order up to a temperature Tn+. Both transition temperatures Tg and Tn+ are accompanied by enthalpy or entropy changes of phase 3 and are predicted in many cases: Annealing above and below Tg, vapor deposition, formation of glacial and quasi-crystalline phases in perfect agreement with experiments. Any transformation of phase 3 changes the initial liquid enthalpy and rejuvenation at Tg < T < Tn+ does not lead to the enthalpy of the initial liquid [18].
Our new publication here was devoted to the simplest case where ultrastable glass and glacial phase are not formed. The value of θn+ was maximum in this case because all transformations below Tg and Tm modified the liquid state and decrease θn+ [19].
The heat capacity jump at Tg was equal to 1.5 × ΔHm/Tm in polymers as shown in 1960 by Wunderlich [40] and confirmed for many molecular glasses [9] (where ΔHm/Tm = ΔSm is the crystal melting entropy). The contribution of the undercooled liquid to the total heat capacity per mole is given by (16) using d Δ ε lg ( θ ) / dT :
Δ C p ( T ) = C p ( liq ) C p ( cryst ) = 2 ( T T m ) T m 2 ( Δ H m ) ( ε ls 0 θ 0 m 2 ε gs 0 θ 0 g 2 )

3. Exothermic or Endothermic Heats Observed above the Melting Temperature Tm

3.1. Exothermic Enthalpy Delivered at 688 K in Al88Ni10Y2 for Tm = 602 K

We follow data of ref. [41]. The glass transition occurs at Tg = 380 K, and the melting temperature at Tm = 602 K. In Figure 1, an annealing of 60 s at Ta = 401, 427, and 525 K increases the fraction Vf of Al-fcc precipitates up to 0.42 and decreases the volume of the amorphous phase without changing the enthalpy recovery at 688 K measured at 0.67 K/s.

3.2. Exothermic Enthalpy Delivered at Tn+ = 1622 K in (Fe71.2B24Y4.8)96Nb4

We follow data of ref. [42]. The glass transition occurs at Tg = 963 K and the melting temperature at Tm =1410 K. An enthalpy recovery occurs at 1622 K (Figure 2).

3.3. Exothermic Enthalpy Delivered at 1835 K in Ni77.5B22.5.

We follow data of ref. [24]. The glass transition occurs at Tg = 690 K and the melting temperature at Tm = 1361 K. The enthalpy recovery temperature is equal to 1835 K.
Deep transformations of eutectic liquid state are observed in Figure 3 by slow heating and aging above the melting temperature which are attributed to the formation of microdomains of 10–100 nm enriched with one of the components with prolonged relaxation time. These microdomains have an influence on the structure and properties of rapidly quenched liquid alloys [24,25]. The enthalpy recovery temperature is here the highest temperature of liquid transformation leading to its homogeneous state. A cooling from 1950 K gives rise to a homogeneous liquid leading to supercooling below Tm = 1361 K.

3.4. Exothermic Enthalpy Delivered at 1356 K in Cu47.5Zr45.1Al7.4

We follow data of ref. [43]. The glass transition occurs at Tg = 690 K and the melting temperature at Tm =1170 K. An enthalpy recovery occurs at 1356 K (Figure 4).

3.5. Endothermic Enthalpy Recovered at 1453–1475 K in a Silicate Liquid

We follow data of ref. [44]. The composition was (49.3SiO2, 15.6Al2O3, 1.8TiO2, 11.7FeO, 10.4CaO, 6.6MgO, 3.9Na2O, and 0.7K2O (wt%)). The glass transition occurred at Tg = 908 K and the melting temperature at Tm = 1313 K. The exothermic latent heat occurred at 1173 K and the amorphous fraction decline with the cycle number from 573 to 1523 K. The melting ex-tended up to 1475 K in Figure 5, and the crystallization temperature Tm occurred at 1313 K in Figure 6. The melting enthalpy recovered between Tm and Tn+ was the same all along the cycles from 2 to 21.
Figure 6 shows that Tm = 1313 K. The transition at Tn+ during continuous cooling at 20 K/mn was no longer sharp and did not have a first-order character. Crystallization occurred at the melting temperature without undercooling, showing that the nuclei were growing between Tn+ and Tm because they were formed above Tm by homogenous nucleation accompanied by an enthalpy increase. Phase 3 disappeared above Tn+ and Δεlg = 0. Crystallization was sharper and sharper during cycling from temperatures higher than Tn+, showing that the short-range order was enhanced. The enthalpy coefficient Δεlg of phase 3 grew by cooling below Tn+.

3.6. Endothermic Enthalpy Recovered at 1114 K in Zr41.2Ti13.8Cu12.5Ni10Be22.5 (Vit1)

We follow data of ref. [45]. The glass transition occurred at Tg = 625 K and the melting temperatures at Tsol = 965 K and Tliq = 1057 K (Figure 7). There was an endothermic enthalpy at T = 1114 K. The heat capacity jump at Tg was ΔCp (Tg) ≌ 21.6 J/K/g-atom. A heat capacity peak of superheated liquid after supercooling was observed during heating around T = 1114 K, accompanied by an endothermic latent heat of about 1100 J/mole. Another transition, observed by viscosity measurements, occurred at 1225 K by heating and subsequent cooling, showing that the liquid became homogeneous above this temperature [46].
Structural changes corresponding to these anomalies were still observed with in-situ synchrotron X-ray-scattering experiments in a contactless environment using an electrostatic levitator (ESL). There was an endothermic liquid–liquid transition at 1114 K during heating reinforced by the symmetrical observation of an exothermic latent heat regarding Tm = 965 K and an exothermic structural change around 816 K by supercooling.

3.7. Endothermic Enthalpy Recovered at 980–1000 K for Tm = 876–881 K in PdNiP Liquid Alloys

The heat capacities of several PdNiP alloys measured at 20 K/min are represented in Figure 8. The melting temperatures were slowly varying with composition around 880 K and an enthalpy recovery temperature was still observed around 990 K in many liquid alloys. The theoretical predictions for these liquid alloys were limited to the case of Pd42.5Ni42.5P15 [21] presented in Section 6.1 and Section 7.1.

4. Predictions of Enthalpy Recovery Temperatures at Tn+ > Tm

Equations (10)–(15) were used to calculate the enthalpy coefficients of fragile Liquids 1, 2, and 3 in Section 4.1, Section 4.2, Section 4.4, Section 4.5, and Section 4.6. Liquid Ni77.5B22.5 in 4.3 being strong, the enthalpy coefficients εls0 and εgs0 were calculated with (9) for θn− = θg, θ0g2 = 1 and θ0m2 = 4/9.

4.1. Exothermic Enthalpy Delivered at Tn+ = 688 K in Al88Ni10Y2

We follow data of ref. [41]. The enthalpy coefficients of this fragile glass-forming melt were calculated with Tg =380 K and Tm = 602 K:
Liquid   1 :   ε ls   ( θ ) = 1.63123 ( 1 θ 2 / 0.26736 )
Liquid   2 :   ϵ gs   ( θ ) = 1.44694 ( 1 θ 2 / 0.3557 )
Liquid   3 :   Δ ϵ lg ( θ ) = 0.18439 2.05237 × θ 2
The temperature Tn+ = 688 K was deduced from θn+ = Δεlgn+) = 0.14287 [48]. In Figure 1, an exothermic enthalpy peak is observed at 688 K for all samples at 0.67 K/s.

4.2. Exothermic Enthalpy Delivered at Tn+ = 1622 K in (Fe71.2B24Y4.8)96Nb4

We follow data of ref. [42]. The enthalpy coefficients of this fragile glass-forming melt were calculated with Tg = 963 K and Tm = 1410 K [48]:
Liquid   1 :     ε ls   ( θ ) = 1.61206 ( 1 θ 2 / 0.27795 )
Liquid   2 :   ϵ gs   ( θ ) = 1.41609 ( 1 θ 2 / 0.36676 )
Liquid   3 :   Δ ϵ lg ( θ ) = 0.19397 1.93329 × θ 2
The temperature Tn+ = 1622 K was deduced from θn+ = Δεlgn+) = 0.1503 [48].

4.3. Exothermic Enthalpy Delivered at Tn+ = 1835 K in Ni77.5B22.5

We follow data of ref. [24]. The enthalpy coefficients of this strong glass-forming melt were calculated with Tg = 690 K and Tm = 1410 K:
Liquid   1 :   ε ls   ( θ ) = 1.09891 ( 1 θ 2 / 0.44444 ) ,
Liquid   2 :   ϵ gs   ( θ ) = 0.51347 ( 1 θ 2 )
Liquid   3 :     Δ ϵ lg ( θ ) = 0.58553 1.958 × θ 2
The temperature Tn+ = 1835 K was deduced from θn+ = Δεlgn+) = 0.34808 [48] in agreement with Figure 3.

4.4. Exothermic Enthalpy Delivered at Tn+ = 1356 K in Cu47.5Zr45.1Al7.4

We follow data of ref. [43]. The enthalpy coefficients of this fragile glass-forming melt were calculated from Tg = 690 K and Tm = 1170 K:
Liquid   1 :   ε ls   ( θ ) = 1.5906 ( 1 θ 2 / 0.28942 )
Liquid   2 :   ϵ gs   ( θ ) = 1.3859 ( 1 θ 2 / 0.37826 )
Liquid   3 :   ϵ lg ( θ ) = 0.2047 1.83194 × θ 2
The temperature Tn+ = 1356 K and the recovered enthalpy coefficient Δεlg were deduced from θn+ = Δεlgn+) = 0.1586 [48] in agreement with Figure 4. The enthalpy coefficient Δεlg reappeared by homogeneous nucleation below Tn+ because εgsn+) was weaker than εlsn+) and liquid 1 enthalpy decreased toward that of liquid 2 at slow cooling.

4.5. Endothermic Enthalpy Recovered at Tn+ = 1470 K in a Silicate Liquid

We follow data of ref. [44]. The enthalpy coefficients of this fragile glass-forming melt were calculated with Tg = 908 K and Tm = 1313 K:
Liquid   1 :   ε ls ( θ ) = 1.69155 ( 1 θ 2 / 0.23189 )
Liquid   2 :   ε gs ( θ ) = 1.53732 ( 1 θ 2 / 0.31613 )
Phase   3 :   Δ ε lg ( θ ) = 0.15473 2.4315 × θ 2
The temperature Tn+ = 1470 K was deduced from θn+ = Δεlgn+) = 0.1195 [48] in agreement with Figure 5.

4.6. Endothermic Enthalpy Recovered at Tn+ = 1114 K in Zr41.2Ti13.8Cu12.5Ni10Be22.5 (Vit1)

We follow data of ref. [45]. The enthalpy coefficients of this fragile glass-forming melt were calculated with Tg = 625 K and Tm = 965 K [35]:
ε ls = 1.70651 × ( 1 θ 2 / 0.2226 )
ε gs = 1.4715 × ( 1 θ 2 / 0.34564 )
              Δ ε lg = 0.23501 3.409 × θ 2 .
The temperature Tn+ = 1114 K was deduced from θn+ = Δεlgn+) = 0.15407 [48] in agreement with Figure 6. The observed double transition was the consequence of the presence in the melt of nuclei, all having the same Gibbs free energy, leading to a homogenous nucleation at 818 and 1114 K as consequence of the quadratic equation of Δεlgn+) = θn+. The ordered liquid was rebuilt at Tn+ = 818 K during cooling from 1350 K with the formation in the no-man’s land of new superclusters, building a vitreous solid phase at Tg resulting of the bond number divergence. The hysteresis of viscosity disappeared at about 1225 K when the liquid is homogeneous [46]. A “colloidal” state was melted above the temperature of viscosity or density branching observed during cooling after heating [24,26,27,46]. Equation (35) was used to calculate the reduced temperature θn+ of glass-forming melt with a glass transition at θg and obeying (11) with a = 1 [48]:
  θ n + = 0.38742 × θ g
The liquidus melting temperature Tliq = 1057.5 K was deduced from Equation (35) with Tg = 625 K and Tn+ = 1225 K in perfect agreement with the experimental observation of liquidus presented in Figure 7. This finding of a second transition above Tn+ agreed with the first-order liquid–liquid transitions observed above Tn+ in Pd42.5Ni42.5P15 and La50Al35Ni15.

5. Three Liquid States above the Melting Temperature

The exothermic and endothermic transitions at Tn+ led to a liquid above Tn+ with an enthalpy coefficient Δεlg = 0. Two other liquid states existed at Tm with enthalpy coefficients equal to ±Δεlg0. The melting temperature Tm was chosen equal to Tsolidus in Figure 9. The enthalpy coefficients (±Δεlg), defined by (15) and applied to Pd42.5Ni42.5P15 in Figure 9 and in Section 7.1, were related to the enthalpy decrease and increase with temperature of these two quenched liquid states toward that of homogeneous liquid.
The homogenous liquid can be quenched along q2 (Δεlg0 = 0) in Figure 9 from above the temperature where the liquid became homogeneous, down to temperatures much weaker than Tg [49,50,51]. An enthalpy relaxation at low heating rate, equal to (−Δεlg0 × ΔHm), built the bonds of phase 3 and led by heating to the temperature where Δεlg = 0 [35]. This slow heating through Tg broke the bonds and the liquid enthalpy increases up to (+Δεlg0 × ΔHm) at Tm, producing an exothermic enthalpy at Tn+. These phenomena are observed in Figure 1, Figure 2, Figure 3 and Figure 4.
With a much higher heating rate, the enthalpy of bonds, building phase 3, did not have the time to relax below Tg, and phase 3 was not formed along the thermal path below Tg and the latent heat at Tn+ was not observed for Pd42.5Ni42.5P15 at 100 K/s, as shown in Figure 10 [21]. The liquid being frozen below Tg with Δεlg = 0 gave rise to an endothermic enthalpy at Tg due to bond breaking and the liquid returned to a homogenous state with Δεlg0 = 0 above Tn+ [10,11,12].
A quench along q1 in Figure 9 from Tm < T < Tn+ with a liquid enthalpy (+Δεlg0 × ΔHm) at Tm led to an amorphous phase with an enthalpy excess (+Δεlg0 × ΔHm). Phase 3 bonds were built during reheating and they decreased, at a low heating rate, the enthalpy coefficient from (+Δεlg0) below Tg to (−Δεlg0) at Tm, leading to an endothermic latent heat at Tn+ corresponding to crystallized nuclei melting at Tn+.
Starting heating at a very low heating rate from any liquid state led to crystallization and to a liquid enthalpy equal to (−Δεlg0 × ΔHm) at Tm.
A quench from Tm < T < Tn+ along q3 led to the enthalpy of phase 3 with crystallized nuclei being the skeleton of this phase after percolation at Tg, as shown for plastic crystals. A slow cooling led to crystallization at Tm without undercooling [19].
The endothermic and exothermic characters of the transition at Tn+ were imposed by the initial value of the liquid enthalpy after quenching and by cooling and heating rates.
Homogeneous nucleation in the liquid was expected to depend on the time of aging in the range of temperatures below and close to the homogenization temperature. The first-order liquid–liquid transitions in Pd42.5Ni42.5P15 and La50Al35Ni15 studied by [20,21] combined with our non-classical model of homogeneous nucleation shed light on these new phenomena.

6. First-Order Liquid–Liquid Transitions Observed in Pd42.5Ni42.5P15, La50Al35Ni15, and Fe2B

We follow data of ref. [21] for Pd42.5Ni42.5P15, of ref. [20] for La50Al35Ni15 and of ref. [24] for Fe2B.

6.1. Pd42.5Ni42.5P15

6.1.1. Fast Differential Scanning Calorimetry at 100 K/s

The fast differential scanning calorimetry (FDSC) heating curve at 100 K/s represented in Figure 10 and reproduced from [21] was used to determine the solidus and liquidus temperatures Tsol = 876 and Tliq = 926.5 K. A first-order liquid–liquid transition was observed at TLL = 1063 K. The sample was previously cooled from 1073 K at 40,000 K/s down to room temperature and reheated up to 1073 K, which was a temperature higher than the first-order transition observed at TLL.

6.1.2. Melting Transition Observed at 993 K above the Solidus Temperature Tsol = 876 K of Pd42.5Ni42.5P15

The samples were quenched from Tq to room temperature at a cooling rate of q = 40,000 K/s and reheated at 100 K/s up to Tq, as shown in Figure 11b [21]. There was no nucleation when cooling started from 1073 K for q > 70 K/s, while crystallization occurred for q < 7000 K/s when cooling started from 1023 K as shown in Figure 11a. The area of the crystallization peak occurring around T = 770 K in Figure 10 was plotted versus Tq in Figure 11b. The temperature T = 993 K was viewed by the authors as a liquidus temperature which was, in fact, equal to 926.5 K, as shown in Figure 9.

6.1.3. First-Order Transition Observed by 31P Nuclear Magnetic Resonance (NMR)

31P NMR was used to characterize the LLT at TLL = 1063 K above Tn+ = 993 K. The liquid alloy was first heated to 1293 K for homogenization during 30 min, and then cooled step by step to 1043 K. NMR spectra were taken isothermally after equilibrating the liquid at 1293 K at each step. The Knight shift (Ks) was determined by the ensemble average of local magnetic field around 31P nuclei, sensitive to the changes in structure, plotted in Figure 12 as a function of temperature [21]. (Ks) varied linearly above 1063 K with a slope increase of 1.76 ppm/K below 1063 K, indicating a change in the P-centered local structures at this temperature. This change was viewed as a first-order liquid–liquid transition (LLT) analogous to that observed in La50Al35Ni15 where a second change of Ks in this new liquid state was observed at lower temperatures attributed to the hysteresis of the transition [20].

6.2. La50Al35Ni15

This melt was characterized by Tg = 528 K, Tsol = 877.6 K, and Tliq = 892 K, as shown in Figure 13 ([20] Figure S1). A second liquidus temperature was found at 950 K. The temperature TLL, observed at 1033 K by measuring the 27Al Knight shift by RMN, is viewed as a first-order LLT in Figure 14. A phenomenon analogous to hysteresis led to a second transition at 1013 K.

6.3. Fe2B

The vitreous state of this compound was obtained by mechanical alloying [52]. The first-order transition occurs at TLL = 1915 K in Figure 15 with a melting temperature of 1662 K [24].

7. Predictions of First-Order Transition Temperatures by Homogenous Nucleation in Pd42.5Ni42.5P15, La50Al35Ni15 and Fe2B Melts

These first-order transitions were observed at TLL at very low cooling rates or by isothermal annealing between the melting temperature and TLL. The homogeneous liquid state characterized by Δεlg0 = 0 was stable during cooling in Figure 3 while the first-order transitions were reversible in Figure 15. The two melting temperatures Tsol and Tliq of non-congruent materials led to two nucleation temperatures Tn+.

7.1. Predictions of Transitions in Pd42.5Ni42.5P15 Melt

The temperature 993 K in Figure 11b was viewed by [21] as a liquidus temperature which was, in fact, equal to 926.5 K, as shown in Figure 10. The reduction of the enthalpy recovered by crystallization at 770 K occurred for Tq < 993 K, as shown in Figure 11b. The crystallization enthalpy at 770 K was continuously reduced without exothermic enthalpy jump equal 0.13357 × 197 = 26 in Figure 11b at 993 K. The mean-range order accompanied by exothermic enthalpy progressively reappeared by homogeneous nucleation in the liquid heated during 30 s at each temperature Tq and was completely formed at Tliq = 926.5 K because the enthalpy decrease was equal to −13.4 % at this temperature. The residual configurons melted at Tn+ = 993 K using (35) (θn+ = Δεlgn+) = 0.13357), Tm = 876 K, and Tg = 574 K. This value of Tg agreed with measurements of heat capacity of melts with similar compositions [47]. The enthalpy coefficients of Pd42.5Ni42.5P15 for the liquidus and solidus liquid states were given in Equations (36–38) using Equations (10–16):
For Tsol = 876 K and Tg = 574 K
Liquid   1 :   ε ls ( θ ) = 1.65525 ( 1 θ 2 / 0.25362 )
Liquid   2 :   ε gs ( θ ) = 1.48288 ( 1 θ 2 / 0.34081 )
Phase   3 :   Δ ε lg ( θ ) = 0.17237 2.1755 × θ 2
For TLiq = 926.45 K and Tg = 574 K
Liquid   1 :   ε ls ( θ ) = 1.61957 ( 1 θ 2 / 0.27384 )
Liquid   2 :   ε gs ( θ ) = 1.42935 ( 1 θ 2 / 0.36251 )
Phase   3 :   Δ ε lg ( θ ) = 0.19022 1.97146 × θ 2
Applying Equation (35) led to Tn+ = TLL = 1063 K in perfect agreement with Figure 12. From our analysis, a second change of Ks occurred by homogeneous nucleation in Pd42.5Ni42.5P15 at Tn+ = 993 K. This transition was not only due to the hysteresis of a first-order transition because there were two homogeneous nucleation temperatures as shown in Figure 12. This point was still confirmed in 7.2 devoted to La50Al35Ni15, where the changes of Ks occurred for two values of Tn+ because there were, in these non-congruent liquid compounds, two solid–liquid transitions characterized by solidus and liquidus temperatures.
The temperature Tn+ = TLL = 1063 K corresponded to the temperature of homogenous nucleation of colloids containing critical numbers nc of atoms with nc given by Equation (42) (see [9], Equation (48)):
n c = 8 N A k B ( 1 + Δ ε lg ) 3 27 Δ S m ( Δ ε lg ) 3 ln ( K )
where NA is the Avogadro number, kB the Boltzmann constant, ΔSm the melting entropy, and LnK ≌ 90 [5]. With Δεlg = 0.13356 and ΔSm = 8.76 J/g-atom [47], nc = 15522 at the temperature Tn+ = 993 K. With Δεlg = 0.14739 and ΔSm = 8.76 J/g-atom, nc = 11977 at the temperature Tn+ = 1063 K. Critical numbers nc, still larger, were observed in Pb-Bi liquid alloys below the temperature of liquid homogenization [27]. The number of atoms inside an elementary superatom in the homogenous liquid above 1063 K was equal to 135, with Δεlgn+) replaced in (42) by εgsn+) = 1.40526 in (42) using Equations (39) and (40). The homogenous nucleation time τ (s) for temperatures 1043 < T < 1063 K was following Equation (43) (see Figure 1d in ref. [21]):
τ   ( s ) = 5.9 × 10 3 ( 1063 T 1 ) 2.18
which led by extrapolation to τ ≌ 1.9 s at Tn+ = 993 K.
There was no growth nucleus inducing crystallization after quenching from the temperature T = 1073 K which was higher than TLL = 1063 K as shown in Figure 11a [21]. New growth nuclei were added when the melt was quenched from 1023 K, a temperature higher than Tn+ = 993 K and much higher than Tsol. Consequently, new denser nuclei growing from the colloidal state were added by homogeneous nucleation at 1023 K above 993 K. The transition at 993 K after cooling from 1073 K was due to the internal and external bond formation between colloids below 1063 K [18]. This observation agreed with the growth of nc from 11977 to 15522 between 1063 and 993 K. The transitions observed by NMR below 1063 K involved all 31P atoms and corresponded to the colloid formation through the relaxation time decrease [23]. The first-order character of this transition was observed at each step of isothermal annealing below 1063 K. The breaking of bonds inside and outside colloids occurred at the lowest temperature Tn+ during heating [18], while at the highest Tn+, a transition from colloidal state to a new homogeneous state made of elementary superatoms only organized by short-range order appeared.
The enthalpy coefficients (−Δεlg0) of phase 3 equal to those of configurons are represented in Figure 16 as a function of the temperature T (K) for Tsol and Tliq using Equations (38) and (41). The crystallization temperature occurred at the reentrant formation temperature of ultrastable glass with its enthalpy equal to −Δεlg0. This nucleation temperature opened the door to crystallization [19].

7.2. Predictions of First-Order Transitions in La50Al35Ni15 Glass-Forming Melt

We follow data of ref. [20]. The phase 3 enthalpy coefficients of La50Al35Ni15 for the liquidus and solidus liquids were given in Equations (44)–(49) for Tsol = 877.6 K, Tliq = 892 K, and Tg =528 K, and are represented in Figure 17.
For TLiq = 892 and Tg = 528 K:
Liquid   1 :     ε ls ( θ ) = 1.66143 ( 1 θ 2 / 0.25000 )
Liquid   2 :   ε gs ( θ ) = 1.42935 ( 1 θ 2 / 0.33679 )
Phase   3 :   Δ ε lg ( θ ) = 0.20404 2.2516 × θ 2
For Tsol = 877.6 and Tg = 528 K:
Liquid   1 :   ε ls ( θ ) = 1.60164 ( 1 θ 2 / 0.28357 )
Liquid   2 :   ε gs ( θ ) = 1.40246 ( 1 θ 2 / 0.37246 )
Phase   3 :   Δ ε lg ( θ ) = 0.19218 1.88273 × θ 2

7.3. Predictions of Glass Transition Temperature of Fe2B Melt

The enthalpy coefficients of the strong liquid Fe2B were calculated with Equations (9) and (35), Tm = 1662 K and Tn+ = TLL = 1915 K, given in Equations (50)–(52). Phase 3 is represented in Figure 18.
For Tm = 1662 K and Tn+ = TLL = 1915 K, Tg = 1125.7 K:
Liquid   1 :     ε ls ( θ ) = 1.60164 ( 1 θ 2 × 2.25 )
Liquid   2 :   ε gs ( θ ) = 1.1519 ( 1 θ 2 )
Phase   3 :   Δ ε lg ( θ ) = 0.19579 1.8804 × θ 2

7.4. One Liquid–Liquid Transition at Tn+ = TLL in Congruent Materials and Two in the Others

A first-order transition occurred at TLL due to the formation by cooling of colloidal state assembling elementary superatoms composed of tenths atoms bounded by short-range interactions, leading to colloids containing thousands of atoms. In the case of congruent materials, only one liquid–liquid transition was expected. The lowest and the highest temperatures Tn+ were equal and Tn+ is a first-order transition temperature equal to TLL. This is the case for Fe2B.
These colloids were similar atom clouds containing a magic atom number of atoms because they were melted by homogeneous nucleation instead of surface melting. They had a maximum radius for which their Gibbs free energy was smaller or equal to that of the melt [53].
There were two liquid–liquid transitions above the solidus and liquidus temperatures Tsol and Tliq in non-congruent materials, leading to two temperatures Tn+. The highest one was equal to TLL and related to Tliq. Above TLL, the liquid was homogeneous and atoms were only submitted to short-range order in tiny superatoms. The lowest (Tn+) was related to Tsol and was the temperature where coupling between elementary superatoms started during cooling and led to bond percolation at Tg. The lowest one was a second-order phase transition where the residual configurons were melted during heating, involving 15% of the sample volume.
A melt was only rejuvenated above Tn+ because all colloids and superatoms were disconnected.

8. Perspectives: Mpemba Effect and Bonding-Antibonding of Superatoms

8.1. Mpemba Effect and Its Inverse Relation to the Existence of Three Liquid States above the Melting Temperature

The Mpemba effect is described by a shorter time needed to crystallize a hot water system than to crystallize the same colder water system cooled down from initial lower temperatures [54]. This phenomenon was documented by Aristotle 2300 years ago [55]. The melting enthalpy of ice was ΔHm = 334 J/g with a specific heat of 4.18 J/g. Starting from a hot homogenous water, the exothermic enthalpy of formation of mean-range order below Tn+ = 295.3 K (22.1 °C) [34] was progressively equal to −0.0818 ×334 = −27.3 J/g by homogenous nucleation during slow cooling through Tn+. The value of Tn+ in water was confirmed by numerical simulations of the melting temperature of an ultrathin layer of hexagonal ice [19,56,57,58]. The water enthalpy variation being equal to 92 J/g from 22.1 to 0 °C, the temperature of 0 °C was quickly attained by the hot system because of the recovery of exothermic enthalpy. The cold water had no more available exothermic enthalpy because the formation of mean range order was much older in this water. Cooling this liquid took much more time.
The latent heat, expected at Tn+ = 22.1 °C, was not observed up to now, while Mpemba and Osborne observed this effect with a slow cooling rate of 0.01 K/s. The window of nucleation was very narrow in congruent materials because the temperature Tn+ was unique instead of extending between the two Tn+ temperatures of non-congruent substances as shown in Figure 13. At a too-high cooling rate, the liquid state, with Δεlg = 0, free of any growth nucleus, remained stable and showed undercooling. The homogeneous liquid state was stable when it escaped the formation of colloidal state at Tn+. Figure 3 showed this phenomenon in Fe77.5B22.5. On the contrary, the transition of Fe2B at Tn+ = 1915 K had a first-order character (see Figure 15) Nucleus formation started from the colloidal state and was expected to be formed at a low cooling rate.
Using the theory of nonequilibrium thermodynamics, Lu and Raz predicted a similar anomalous behavior with heating using a three-state model that we had here for all melts [59]. A cold liquid, with an enthalpy equal to (+Δεlg0 × ΔHm), obtained after building bonds below Tg, would develop an exothermic latent heat at Tn+ during heating, while a warmer liquid with an enthalpy equal to (−Δεlg0 × ΔHm) would need an endothermic enthalpy to melt its mean-range order.
The Mpemba effect and its inverse effect can be extended to many systems [59,60] and we showed that these phenomena could exist in all melts. Moreover, we assumed that analogues of Mpemba effects should occur on vitrification of liquids so that glasses would be formed quicker out of hot melts compared with melts cooled down from lower temperatures. All these new events were observable because the transition at Tn+ was a first-order transition in congruent materials [24].

8.2. Three Liquid States Associated with Bonding–Antibonding of Superatoms

In Figure 9, the enthalpy coefficient of phase 3 at Tn+ was equal to three values +Δεlg, 0, and −Δεlg depending on thermal history leading to three liquid states. The glass transitions occurred at the percolation threshold of superclusters, built by homogeneous nucleation, during the first cooling of liquids initially homogeneous. These superclusters survive in overheated colloids after their formation during the first cooling because they were melted at Tn+ by homogeneous nucleation instead of surface melting at Tm. These entities were contained in colloids with a magic atom number [61]. Thus, they were melted at Tn+ when their Gibbs free energy became equal to that of homogeneous liquid [53]. The endothermic and exothermic latent heats revealed the existence of two families of bound molecules which could be attributed to bonding and antibonding of colloids through elementary superatoms. This concept of bonding and antibonding is highly developed in bond chemistry to create new chemical structures. Bonding and antibonding of colloids could lead to higher and lower enthalpies. Two recent examples of research in this field were given [62,63]. The nature built these new superstructures in all overheated melts by homogeneous nucleation. The percolation of these superatoms led Tg to a superstructure involving 3D space in 15% of atoms [18].

9. Conclusions

Our models of homogeneous nucleation and configurons explained the formation of liquid phases above Tg with mean-range order disappearing at a temperature Tn+ much higher than the melting temperature. This transformation at Tn+ was a first-order transition in congruent materials such as Fe2B and was expected to be observable at a very low cooling rate or by homogeneous nucleation during isotherm annealing below Tn+. There were two melting temperatures in non-congruent materials called solidus and liquidus temperatures, leading to two temperatures: Tn+ starting with a unique glass transition temperature Tg. The first-order liquid–liquid transitions in Pd42.5Ni42.5P15 and La50Al35Ni15 observed with NMR at TLL = 1063 and 1033 K, respectively, occurred at the temperature Tn+ corresponding to the liquidus temperature of these two alloys. The two other second-order phase transitions, occurring at Tn+ = 993 and 1013 K respectively, were induced by the solidus temperatures.
The latent heats produced at Tn+ were exothermic or endothermic. We attributed this phenomenon to the presence of three liquid states at Tn+, with three enthalpy coefficients depending on the cooling rate and on the starting temperature of quenching. These enthalpies were equal to 0, +ΔHm × Δεlg (Tn+), and −ΔHm × Δεlg in comparison with that of a homogeneous liquid equal to zero at high temperatures. These liquids, when quenched to temperatures much weaker than Tg, were characterized by their initial enthalpy at the solidus temperature. The liquid state enthalpy after quenching was (−ΔHm × Δεlg0), or (+ΔHm × Δεlg0), or 0 depending on its initial value before quenching and on the cooling and heating rates. The enthalpy increased from −ΔHm × Δεlg0 and 0 up to +ΔHm × Δεlg0 at Tm with configuron melting. The enthalpy decreased from +ΔHm × Δεlg0 and 0 to −ΔHm × Δεlg0 at Tm, rebuilding the missing bonds. These phenomena were well described by the positive or negative variation ±ΔHm × Δεlg (Tn+), of enthalpies of bonds and configurons.
Our homogeneous nucleation model above Tm still confirmed the formation of colloids between Tn+ and Tm and at slow cooling rate, the growth of cluster-bound colloids inducing crystallization. The temperature Tn+, congruent materials being unique, was the temperature of homogenization of these melts. The highest temperature Tn+ = TLL observed in Pd42.5Ni42.5P15 and La50Al35Ni15 was a homogenization temperature of these non-congruent materials. All melts, containing atoms of different nature, were submitted to short-range order inside superatoms, being the elementary bricks building the ordered liquids and glasses.
These colloids and elementary superatoms could not be more precisely described because their magic atom number nc, and the associated enthalpy depending on nc, were unknown.
Colloids formed by homogeneous nucleation were superatoms containing magic atom numbers which were not totally melted above Tm and were fully melted by homogenous nucleation instead of surface melting at the highest temperature Tn+. They contained a critical atom number nc defined by their Gibbs free energy equal or smaller than that of the homogeneous melt. They gave rise to new molecular entities by bonding and antibonding, as shown by the opposite values of their contribution to the enthalpy at Tn+. Superstructures of elementary superatoms grew during cooling down to their percolation temperature.
The Mpemba effect and its inverse were easily predicted from this description of materials melting, leading to three stable liquid states above the melting temperature and transitions between them. The transition at Tn+ may have been not only the temperature where the mean-range order disappeared, but also a first-order transition temperature between two liquid states.

Author Contributions

Conceptualization, R.F.T. and M.I.O.; methodology, R.F.T.; software, R.F.T.; validation, R.F.T. and M.I.O.; formal analysis, R.F.T.; investigation, R.F.T.; resources, R.F.T. and M.I.O.; data curation, R.F.T.; writing—original draft preparation, R.F.T.; writing—review and editing, R.F.T. and M.I.O.; visualization, R.F.T.; supervision, R.F.T. and M.I.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data underlying this article will be shared on reasonable request from the corresponding author.

Acknowledgments

Thanks to L.N.C.M.I., Grenoble Alpes University, Imperial College London.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. De Rango, P.; Lees, M.; Lejay, P.; Sulpice, A.; Tournier, R.; Ingold, M.; Germi, P.; Pernet, M. Texturing of magnetic materials at high temperature by solidification in a magnetic field. Nature 1991, 349, 770–772. [Google Scholar] [CrossRef]
  2. 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]
  3. Porcar, L.; De Rango, P.; Bourgault, D.; Tournier, R.F. Magnetic Texturing of High-Tc Superconductors; Gabovitch, A., Ed.; BoD-Books on demand: Norderstedt, Germany, 2012; p. 171. [Google Scholar]
  4. Turnbull, D. Kinetics of solidification of supercooled liquid mercury. J. Chem. Phys. 1952, 20, 411. [Google Scholar] [CrossRef]
  5. Vinet, P.; Magnusson, L.; Frederikssen, H.; Desré, P.J. Correlations between surface and interface energies with respect to crystal nucleation. J. Colloid Interf. Sci. 2002, 255, 363–374. [Google Scholar] [CrossRef]
  6. Tournier, R.F. Presence of intrinsic growth nuclei in overheated and undercooled liquid elements. Phys. B 2007, 392, 79–91. [Google Scholar] [CrossRef]
  7. Tournier, R.F. Influence of Fermi energy equalization on crystal nucleation in glass melts. J. Alloys Comp. 2009, 483, 94–96. [Google Scholar] [CrossRef]
  8. Tournier, R.F. Thermodynamic origin of the vitreous transition. Materials 2011, 4, 869–892. [Google Scholar] [CrossRef] [Green Version]
  9. Tournier, R.F. Fragile-to-fragile liquid transition at Tg and stable-glass phase nucleation rate maximum at the Kauzmann temperature. Phys. B 2014, 454, 253–271. [Google Scholar] [CrossRef] [Green Version]
  10. Wool, R.P. Twinkling fractal theory of the glass transition. J. Polym. Sci. Part B Polym. Phys. 2008, 46, 2765–2778. [Google Scholar] [CrossRef]
  11. Wool, R.P.; Campanella, A. Twinkling fractal theory of the glass transition: Rate dependence and time-temperature superposition. J. Polym. Sci. Part B Polym. Phys. 2009, 47, 2578–2589. [Google Scholar] [CrossRef]
  12. Stanzione, J.F., III; Strawhecker, K.E.; Wool, R.P. Observing the twinkling nature of the glass transition. J. Non-Cryst. Sol. 2011, 357, 311. [Google Scholar] [CrossRef]
  13. Ojovan, M.I.; Travis, K.P.; Hand, R.J. Thermodynamic parameters of bonds in in glassy materials from viscosity temperature relationships. J. Phys. Cond. Matter. 2007, 19, 415107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Ojovan, M.I.; Lee, W.E. Connectivity and glass transition in disordered oxide systems. J. Non-Cryst. Sol. 2010, 356, 2534–2540. [Google Scholar] [CrossRef]
  15. Ojovan, M.I. Ordering and structural changes at the glass-liquid transition. J. Non-Cryst. Sol. 2013, 382, 79. [Google Scholar] [CrossRef]
  16. Ojovan, M.I. Louzguine Luzgin, D.V. Revealing structural changes at glass transition via radial distribution functions. J. Phys. Chem. B. 2020, 124, 3186–3194. [Google Scholar] [CrossRef] [PubMed]
  17. Sanditov, D.S.; Ojovan, M.I.; Darmaev, M.V. Glass transition criterion and plastic deformation of glass. Phys. B 2020, 582, 411914. [Google Scholar] [CrossRef]
  18. Tournier, R.F.; Ojovan, M.I. Undercooled Phase Behind the Glass Phase with Superheated Medium-Range Order above Glass Transition Temperature. Phys. B 2021, 602, 412542. [Google Scholar] [CrossRef]
  19. Tournier, R.F.; Ojovan, M.I. Dewetting temperatures of prefrozen and grafted layers in ultrathin films viewed as melt-memory effects. Phys. B 2021, 611, 412796. [Google Scholar] [CrossRef]
  20. Xu, W.; Sandor, M.T.; Yu, Y.; Ke, H.-B.; Zhang, H.-B.; Li, M.-Z.; Wang, M.-Z.; Liu, L.; Wu, Y. Evidence of liquid-liquid transition in glass-forming La50Al35Ni15 melt above liquidus temperature. Nat. Commun. 2015, 6, 7696. [Google Scholar] [CrossRef]
  21. Chen, E.-Y.; Peng, S.-X.; Peng, L.; Michiel, M.D.; Vaughan, G.B.M.; Yu, Y.; Yu, H.-B.; Ruta, B.; Wei, S.; Liu, L. Glass-forming ability correlated with the liquid-liquid transition in Pd42.5Ni42.5P15 alloy. Scr. Mater. 2021, 193, 117–121. [Google Scholar] [CrossRef]
  22. Tanaka, H. Liquid-liquid transition and polyamorphism. J. Chem. Phys. 2020, 153, 130901. [Google Scholar] [CrossRef] [PubMed]
  23. Popel, P.S.; Chikova, O.A.; Matveev, V.M. Metastable colloidal states of liquid metallic solutions. High Temp. Mater. Proc. 1995, 4, 219–233. [Google Scholar] [CrossRef]
  24. Popel, P.S.; Sidorov, V.E. Microheterogeneity of liquid metallic solutions and its influence on the structure and propertes of rapidly quenched alloys. Mater. Sci. Eng. 1997, A226-2289, 237–244. [Google Scholar] [CrossRef]
  25. Dahlborg, U.; Calvo-Dahlborg, M.; Popel, P.S.; Sidorov, V.R. Structure and properties of some glass-forming liquid alloys. Eur. Phys. J. 2000, B14, 639–648. [Google Scholar] [CrossRef]
  26. Manov, V.; Popel, P.; Brook-Levinson, F.; Molokanov, V.; Calvo-Dahlborg, M.; Dahlborg, U.; Sidorov, V.; Son, L.; Tarakanov, Y. Influence of the treatment of melt on the properties of amorphous materials: Ribbons, bulks and glass coated microwires. Mater. Sci. Eng. 2001, A304–A306, 54–60. [Google Scholar] [CrossRef]
  27. Popel, P.; Dahlborg, U.; Calvo-Dahlborg, M. On the existence of metastable microheterogeneities in metallic melts. IOP Conf. Ser. Mater. Sci. Eng. 2017, 192, 012012. [Google Scholar] [CrossRef]
  28. Yang, B.; Perepezko, J.H.; Schmelzer, J.W.P.; GaO, Y.; Schick, C. Dependence of crystal nucleation on prior liquid overheating by differential fast scanning calorimeter. J. Chem. Phys. 2014, 140, 104513. [Google Scholar] [CrossRef]
  29. He, Y.; Li, J.; Wang, J.; Kou, H.; Beaugnon, E. Liquid-liquid structure transition and nucleation in undercooled Co-B eutectic alloys. Appl. Phys. A 2017, 123, 391. [Google Scholar] [CrossRef]
  30. Adams, G.; Gibbs, J.H. On the temperatutre dependence of cooperative relaxation properties in glass-forming liquids. J. Chem. Phys. 1965, 43, 139. [Google Scholar] [CrossRef] [Green Version]
  31. Liu, C.-Y.; He, J.; Keunings, R.; Bailly, C. New linearized relation for the universal viscosity-temperature behavior of polymer melts. Macromolecules 2006, 39, 8867–8869. [Google Scholar] [CrossRef]
  32. Tournier, R.F. Thermodynamic and kinetic origin of the vitreous transition. Intermetallics 2012, 30, 104–110. [Google Scholar] [CrossRef]
  33. Tournier, R.F. Predicting glass-to-glass and liquid-to-liquid phase transitions in supercooled water using classical nucleation theory. Chem. Phys. 2018, 500, 45–53. [Google Scholar] [CrossRef] [Green Version]
  34. Tournier, R.F. Homogeneous nucleation of phase transformations in supercooled water. Phys. B 2020, 579, 411895. [Google Scholar] [CrossRef]
  35. Tournier, R.F. First-order transitions in glasses and melts induced by solid superclusters nucleated by homogeneous nucleation instead of surface melting. Chem. Phys. 2019, 524, 40–54. [Google Scholar] [CrossRef] [Green Version]
  36. Angell, C.A.; Rao, K.J. Configurational excitations in condensed matter and the “bond lattice”. Model for the liquid-glass transition. J. Chem. Phys. 1972, 57, 470–481. [Google Scholar] [CrossRef]
  37. Iwashita, T.; Micholson, D.M.; Egami, T. Elementary excitations and crossover phenomenon in liquids. Phys. Rev. Lett. 2013, 110, 205504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Ozhovan, M.I. Topological characteristics of bonds in SiO2 and GeO2 oxide systems at glass-liquid transition. J. Exp. Theor. Phys. 2006, 103, 819–829. [Google Scholar] [CrossRef]
  39. Benigni, P. Coupling of phase diagrams and thermochemistry. Calphad 2021, 72, 102228. [Google Scholar]
  40. Wunderlich, B. Study of the change in specific heat of monomeric and polymeric glasses during the glass transition. J. Chem. Phys. 1960, 64, 1052. [Google Scholar] [CrossRef]
  41. Kim, Y.H.; Kiraga, K.; Inoue, A.; Masumoto, T.; Jo, H.H. Crystallization and high mechanical strength of Al-based amorphous alloys. Mater. Trans. 1994, 35, 293–302. [Google Scholar] [CrossRef] [Green Version]
  42. Hu, Q.; Sheng, H.C.; Fu, M.W.; Zeng, X.R. Influence of melt temperature on the Invar effect in (Fe71.2B.024Y4.8)96Nb4 bulk metallic glasses. J. Mater. Sci. 2019, 48, 6900–6906. [Google Scholar]
  43. Jiang, H.-R.; Bochtler, B.; Riegler, S.S.; Wei, X.-S.; Neuber, N.; Frey, M.; Gallino, I.; Busch, R.; Shen, J. Thermodynamic and kinetic studies of the Cu-Zr-Al(-Sn) bulk metallic glasses. J. Alloys Comp. 2020, 844, 156126. [Google Scholar] [CrossRef]
  44. Yue., Y. Experimental evidence for the existence of an ordered structure in a silicate liquid above its liquidus temperature. J. Non-Cryst. Sol. 2004, 345–346, 523–527. [Google Scholar] [CrossRef]
  45. Wei, S.; Yang, F.; Bednarcik, J.; Kaban, I.; Shuleshova, O.; Meyer, A.; Busch, R. Liquid-liquid transition in a strong bulk metallic glass-forming liquid. Nat. Commun. 2013, 4, 2083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Way, C.; Wadhwa, P.; Busch, R. The influence of shear rate and temperature on the viscosity and fragility of the Zr41.2Ti13.8Cu12.5Ni10.0Be22.5 metallic-glass-forming liquid. Acta Mater. 2007, 55, 2977–2983. [Google Scholar] [CrossRef]
  47. Lan, L.; Ren, Y.; Wei, X.Y.; Wang, B.; Gilbert, E.P.; Shibayama, T.; Watanabe, S.; Ohnuma, M.; Wang, X.-L. Hidden amorphous phase and reentrant supercooled liquid in Pd-Ni-P metallic glass. Nat. Commun. 2017, 8, 14679. [Google Scholar] [CrossRef]
  48. Tournier, R.F. Glass phase and other multiple liquid-to-liquid transitions resulting from two-liquid competition. Chem. Phys. Lett. 2016, 665, 64–70. [Google Scholar] [CrossRef] [Green Version]
  49. Wang, L.-M.; Borick, S.; Angell, C.A. An electrospray technique for hyperquenched glass calorimetry studies: Propylene glycol and di-n-butylphthalate. J. Non-Cryst. Sol. 2007, 353, 3829–3837. [Google Scholar] [CrossRef]
  50. Hornboll, L.; Yue, Y. Enthalpy relaxation in hyperquenched glasses of different fragility. J. Non-Cryst. Sol. 2008, 354, 1832–1870. [Google Scholar] [CrossRef]
  51. Hu, L.; Zhang, C.; Yue, Y. Thermodynamic anomaly of the sub-Tg relaxation in hyperquenched metallic glasses. J. Chem. Phys. 2013, 138, 174508. [Google Scholar] [CrossRef]
  52. Passamani, F.C.; Tagarro, J.R.B.; Lareka, C.; Fernades, A.A.R. Thermal studies and magnetic properties of mechanical alloyed Fe2B. J. Phys. Cond. Mater. 2002, 14, 1975–1983. [Google Scholar] [CrossRef]
  53. Tournier, R.F. Crystallization of supercooled liquid elements induced by superclusters containing magic atom numbers. Metals 2014, 4, 359–387. [Google Scholar] [CrossRef] [Green Version]
  54. Mpemba, E.B.; Osborne, D.G. Cool? Phys. Educ. 1969, 4, 172–175. [Google Scholar] [CrossRef]
  55. Aristotle; Ross, W.D. Aristotle’s Metaphysics; Clarendon: Oxford, UK, 1923. [Google Scholar]
  56. Takaiwa, D.; Atano, I.; Koga, K.; Tanaka, H. Phase diagram of water in carbon nanotubes. PNAS 2008, 105, 39–43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Raju, M.; Van Duin, A. Phase transitions of ordered ice in graphene nanocapillaries and carbon nanotubes. Sci. Rep. 2018, 8, 3851. [Google Scholar] [CrossRef] [PubMed]
  58. Nie, G.X.; Huang, J.Y.; Huang, J.P. Melting-freezing transition of monolayer water confined by phosphorene plates. J. Phys. Chem. B 2016, 120, 9011–9018. [Google Scholar] [CrossRef]
  59. Lu, Z.; Raz, O. Nonequilibrium thermodynamics of the markovian Mpemba effect and its inverse. PNAS 2017, 114, 5883–5888. [Google Scholar] [CrossRef] [Green Version]
  60. Kumar, A.; Bechhoefer, J. Exponentially faster cooling in a colloidal system. Nature 2020, 584, 64–68. [Google Scholar] [CrossRef]
  61. Kuzmin, V.I.; Tytik, D.L.; Belashchenko, D.K.; Sirenko, A.V. Structure of silver cluster with magic numbers of atoms by data of molecular thermodynamics. Colloid. J. 2008, 70, 284–296. [Google Scholar] [CrossRef]
  62. Wu, Q.; Xu, C.; Wu, X.; Cheng, L. Evidence for the super-atom bonding from bond energies. ACS Omega. 2018, 3, 14425–14430. [Google Scholar]
  63. Koley, S.; Cui, J.Y.; Panfil, E.; Banin, U. Coupled colloidal quantum dot molecules. Account Chem. Phys. 2021, in press. [Google Scholar] [CrossRef] [PubMed]
Figure 1. DSC curves measured at 0.67 K/s of an Al88Ni10Y2 amorphous alloy aged for 60 s at different Ta. Reprinted from ref. [42], Figure 4.
Figure 1. DSC curves measured at 0.67 K/s of an Al88Ni10Y2 amorphous alloy aged for 60 s at different Ta. Reprinted from ref. [42], Figure 4.
Materials 14 02287 g001
Figure 2. (a) High-temperature DSC trace at 0.33 K/s of the master alloy and (b) the enlarged version after melting. Reprinted with permission from ref. [42], Figure 7. Copyright 2014 Springer.
Figure 2. (a) High-temperature DSC trace at 0.33 K/s of the master alloy and (b) the enlarged version after melting. Reprinted with permission from ref. [42], Figure 7. Copyright 2014 Springer.
Materials 14 02287 g002
Figure 3. Temperature dependence of the density d of Ni-22.5%B melt at slow heating after melting and time exposition for 5–20 h (•), subsequent cooling (ο), and the second heating after crystallization of the sample and repeated melting (Δ). The arrows show the “critical” temperatures at which the density instability is observed. Reprinted with permission from ref. [24], Copyright 1997 Elsevier.
Figure 3. Temperature dependence of the density d of Ni-22.5%B melt at slow heating after melting and time exposition for 5–20 h (•), subsequent cooling (ο), and the second heating after crystallization of the sample and repeated melting (Δ). The arrows show the “critical” temperatures at which the density instability is observed. Reprinted with permission from ref. [24], Copyright 1997 Elsevier.
Materials 14 02287 g003
Figure 4. Multiple DTA measurements (0.333 K/s) of Cu47.5 alloy. The first up- and down-scan cycle is well below 1350 K and the last two cycles reach 1473 K. A remarkable exothermic reaction observed at the temperature above 1350 K in the second up-scan curve is marked by a gray dashed circle. Reprinted with permission from ref. [43], Figure 9b, Copyright 2020 Elsevier.
Figure 4. Multiple DTA measurements (0.333 K/s) of Cu47.5 alloy. The first up- and down-scan cycle is well below 1350 K and the last two cycles reach 1473 K. A remarkable exothermic reaction observed at the temperature above 1350 K in the second up-scan curve is marked by a gray dashed circle. Reprinted with permission from ref. [43], Figure 9b, Copyright 2020 Elsevier.
Materials 14 02287 g004
Figure 5. The repeated DSC up-scanning to 1250 °C (Tliq + 70 °C) at 0.333 K/s. The numerals next to the graphs represent the order of the DSC up-scans. The measurements are performed in argon at the heating rate 20 °C/min. Reprinted with permission from ref. [44], Figure 2, Copyright 2004 Elsevier.
Figure 5. The repeated DSC up-scanning to 1250 °C (Tliq + 70 °C) at 0.333 K/s. The numerals next to the graphs represent the order of the DSC up-scans. The measurements are performed in argon at the heating rate 20 °C/min. Reprinted with permission from ref. [44], Figure 2, Copyright 2004 Elsevier.
Materials 14 02287 g005
Figure 6. The repeated DSC down-scanning from 1250 °C (Tn+ + 70°C). The numerals next to the graphs represent the order of the DSC down-scans. The measurements are performed in argon at the cooling rate 20 °C/min. Reprinted with permission from ref. [44], Copyright Elsevier.
Figure 6. The repeated DSC down-scanning from 1250 °C (Tn+ + 70°C). The numerals next to the graphs represent the order of the DSC down-scans. The measurements are performed in argon at the cooling rate 20 °C/min. Reprinted with permission from ref. [44], Copyright Elsevier.
Materials 14 02287 g006
Figure 7. Cp measured upon heating at 30 K/min for the amorphous sample (upper) and once-melted crystallized sample (lower) (vertically shifted for clarity). Reprinted from ref. [45], Figure 1b.
Figure 7. Cp measured upon heating at 30 K/min for the amorphous sample (upper) and once-melted crystallized sample (lower) (vertically shifted for clarity). Reprinted from ref. [45], Figure 1b.
Materials 14 02287 g007
Figure 8. DSC data during heating for Pd-Ni-P metallic glasses. The heating rate is 20 K/min. The curves have been shifted vertically for clarity. Reprinted from ref. [47], Figure S3.
Figure 8. DSC data during heating for Pd-Ni-P metallic glasses. The heating rate is 20 K/min. The curves have been shifted vertically for clarity. Reprinted from ref. [47], Figure S3.
Materials 14 02287 g008
Figure 9. The enthalpy coefficients of Pd42.5Ni42.5P15 versus T (K). Tn+ = 993 K; Tsolidus = 876 K; Tg = 574 K; T3 = 527.3 K. The enthalpy coefficients (±Δεlg) given by (15) versus T (K). Three quenching lines q1, q2, and q3 along +Δεlg0 = 0.17237, 0, and −Δεlg0.
Figure 9. The enthalpy coefficients of Pd42.5Ni42.5P15 versus T (K). Tn+ = 993 K; Tsolidus = 876 K; Tg = 574 K; T3 = 527.3 K. The enthalpy coefficients (±Δεlg) given by (15) versus T (K). Three quenching lines q1, q2, and q3 along +Δεlg0 = 0.17237, 0, and −Δεlg0.
Materials 14 02287 g009
Figure 10. DSC with heating rate of 100 K/s. A typical FDSC heating curve of the sample obtained from Tq = 1073 K and q = 40,000 K/s. The enthalpy of crystallization is denoted as ΔH. The inset shows the temperature protocol of the FDSC experiments. Temperatures Tsol, Tliq, and TLL added. Reprinted with permission from ref. [21], Figure 3a, Copyright 2021 Elsevier.
Figure 10. DSC with heating rate of 100 K/s. A typical FDSC heating curve of the sample obtained from Tq = 1073 K and q = 40,000 K/s. The enthalpy of crystallization is denoted as ΔH. The inset shows the temperature protocol of the FDSC experiments. Temperatures Tsol, Tliq, and TLL added. Reprinted with permission from ref. [21], Figure 3a, Copyright 2021 Elsevier.
Materials 14 02287 g010
Figure 11. The content of amorphous phase as a function of cooling rate q− of the samples obtained from Tq = 1073 and 1023 K, respectively (a). The crystallized fraction characterized by ΔH is evaluated by FDSC as shown in Figure 10b. Area of the second exothermic peak ΔH as a function of Tq as shown in Figure 10. (b) Area of the exothermic peak ΔH as a function of Tq with a cooling rate of 40,000 K/s. Reprinted with permission from ref. [21], Figure 3b and Figure S1, Copyright 2021 Elsevier.
Figure 11. The content of amorphous phase as a function of cooling rate q− of the samples obtained from Tq = 1073 and 1023 K, respectively (a). The crystallized fraction characterized by ΔH is evaluated by FDSC as shown in Figure 10b. Area of the second exothermic peak ΔH as a function of Tq as shown in Figure 10. (b) Area of the exothermic peak ΔH as a function of Tq with a cooling rate of 40,000 K/s. Reprinted with permission from ref. [21], Figure 3b and Figure S1, Copyright 2021 Elsevier.
Materials 14 02287 g011
Figure 12. The changes of the Knight shift Ks at different undercooled temperatures Tiso after quenching the melt from 1173 K. The solid and open symbols represent the initial and equilibrium Ks, respectively. Reprinted with permission from ref. [21], Figure 1c, Copyright 2021 Elsevier.
Figure 12. The changes of the Knight shift Ks at different undercooled temperatures Tiso after quenching the melt from 1173 K. The solid and open symbols represent the initial and equilibrium Ks, respectively. Reprinted with permission from ref. [21], Figure 1c, Copyright 2021 Elsevier.
Materials 14 02287 g012
Figure 13. DSC trace of as-cast La50Al35Ni15 BMG. The DSC curve obtained at a heating rate of 10 K/min. Liquidus temperature (Tliq) indicated by red arrows. The two liquidus temperatures (Tn+) and solidus temperature (Tsol) are added. Reprinted from ref. [20], Figure S1.
Figure 13. DSC trace of as-cast La50Al35Ni15 BMG. The DSC curve obtained at a heating rate of 10 K/min. Liquidus temperature (Tliq) indicated by red arrows. The two liquidus temperatures (Tn+) and solidus temperature (Tsol) are added. Reprinted from ref. [20], Figure S1.
Materials 14 02287 g013
Figure 14. Temperature dependence of 27Al knight shift Ks during continuous cooling and reheating in the temperature interval of 973–1143 K. The upper red dashed line represents the fitting curves of Ks versus T with a slope of 0.22 ppm/K and the lowest blue dashed line represents the fitting curve with a slope of 0.33 p.p.m./K. The two fitting curves intersect at 1033 K. Three characteristic temperatures indicated by black arrows 1033, 1013, and 1008 K. Reprinted from ref. [20], Figure 1b.
Figure 14. Temperature dependence of 27Al knight shift Ks during continuous cooling and reheating in the temperature interval of 973–1143 K. The upper red dashed line represents the fitting curves of Ks versus T with a slope of 0.22 ppm/K and the lowest blue dashed line represents the fitting curve with a slope of 0.33 p.p.m./K. The two fitting curves intersect at 1033 K. Three characteristic temperatures indicated by black arrows 1033, 1013, and 1008 K. Reprinted from ref. [20], Figure 1b.
Materials 14 02287 g014
Figure 15. Temperature dependence of the density d of Fe-26.4 at %B and Fe-33.3 at %B melts at heating after melting (•) and subsequent cooling (ο). Arrow shows the anomaly linked with structural transformation in a liquid compound. Reprinted with permission from ref. [24], Figure 8, Copyright 1997 Elsevier.
Figure 15. Temperature dependence of the density d of Fe-26.4 at %B and Fe-33.3 at %B melts at heating after melting (•) and subsequent cooling (ο). Arrow shows the anomaly linked with structural transformation in a liquid compound. Reprinted with permission from ref. [24], Figure 8, Copyright 1997 Elsevier.
Materials 14 02287 g015
Figure 16. Enthalpy coefficients of liquidus and solidus melts versus T (K). Tliq = 926.5 K, Tn+ = 1063 K, Tsol = 876 K, and its Tn+ = 993 K. Crystallization at the nucleation temperature T3 = 748.4 K of phase 3 in solidus melt instead of T3 = 774.2 K in liquidus melt. In the liquidus melt, TLL = Tn+ = 1063 K.
Figure 16. Enthalpy coefficients of liquidus and solidus melts versus T (K). Tliq = 926.5 K, Tn+ = 1063 K, Tsol = 876 K, and its Tn+ = 993 K. Crystallization at the nucleation temperature T3 = 748.4 K of phase 3 in solidus melt instead of T3 = 774.2 K in liquidus melt. In the liquidus melt, TLL = Tn+ = 1063 K.
Materials 14 02287 g016
Figure 17. La50Al35Ni15 enthalpy coefficients of liquidus and solidus melts. Tsol = 877.6 K; Tn+ = 1013 K; Tliq = 892 K; Tn+ = TLL = 1033 K. The enthalpy coefficients of ultrastable phase 3 are (−0.19918) for the solidus and (−0.20404) for the liquidus. The two melts have the same Tg = 574 K.
Figure 17. La50Al35Ni15 enthalpy coefficients of liquidus and solidus melts. Tsol = 877.6 K; Tn+ = 1013 K; Tliq = 892 K; Tn+ = TLL = 1033 K. The enthalpy coefficients of ultrastable phase 3 are (−0.19918) for the solidus and (−0.20404) for the liquidus. The two melts have the same Tg = 574 K.
Materials 14 02287 g017
Figure 18. Enthalpy coefficient Δ ε lg (T) of Fe2B Phase 3 and configurons. T3 = 903 K, Tg = 1125.7 K, Tm = 1662 K, Tn+ = TLL = 1915 K.
Figure 18. Enthalpy coefficient Δ ε lg (T) of Fe2B Phase 3 and configurons. T3 = 903 K, Tg = 1125.7 K, Tm = 1662 K, Tn+ = TLL = 1915 K.
Materials 14 02287 g018
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tournier, R.F.; Ojovan, M.I. Building and Breaking Bonds by Homogenous Nucleation in Glass-Forming Melts Leading to Transitions in Three Liquid States. Materials 2021, 14, 2287. https://doi.org/10.3390/ma14092287

AMA Style

Tournier RF, Ojovan MI. Building and Breaking Bonds by Homogenous Nucleation in Glass-Forming Melts Leading to Transitions in Three Liquid States. Materials. 2021; 14(9):2287. https://doi.org/10.3390/ma14092287

Chicago/Turabian Style

Tournier, Robert F., and Michael I. Ojovan. 2021. "Building and Breaking Bonds by Homogenous Nucleation in Glass-Forming Melts Leading to Transitions in Three Liquid States" Materials 14, no. 9: 2287. https://doi.org/10.3390/ma14092287

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop