Acessibilidade / Reportar erro

On the Heat Capacity of Pure Elements and Phases

Abstract

Recently, a model was proposed to predict cv as a function of temperature from the absolute zero to the melting temperature applied. This solution was based on critical grain nucleation to determine the volume, which contains the total number of modes for a particular equilibrium and non-equilibrium state to calculate the density of state (DoS), which is strongly dependent on the nucleus radius for both pure element and compound. Electronic and rotational energies were regarded for both elements and compounds in this formulation. The anomalies associated with cv can be easily considered in terms of their entropies, independent of their nature, as a local change in the DoS. Comparisons of cv for elements and compounds are performed against Thermodynamics software simulations and experimental data.

Keywords:
Molar heat capacity; Density of State; Elements and compounds; Computational Thermodynamics; Anomalies of heat capacity


1. Introduction

The molar heat capacity for the solid-state of matter is important thermophysical properties for many branches of physics and engineering. There are two methods available for its calculation: (i) for high temperatures, in which empirical formulae based on integrals and experimental coefficients are normally used to calculate the molar specific heat as a polynomial function of temperature11 Abu-Eishah SI, Haddad Y, Solieman A, Bajbouj A. A new correlation for the specific heat of metals, metal oxides and metal fluorides as a function of temperature. Lat Am Appl Res. 2004;34:257-65. and (ii) with the use of Computational Thermodynamics packages and databases to be numerically determined for a specific class of materials22 Sundman B, Kattner UR, Sigli C, Stratmann M, Le Tellier R, Palumbo M, et al. The OpenCalphad thermodynamic software interface. Comput Mater Sci. 2016;125:188-96.. For lower temperatures, the realm of the quantum pronounced effects, historically, Einstein33 Einstein A. Die plancksche theorie der strahlung und die theorie der spezifischen Wärme. Ann Phys. 1907;327(1):180-90. modeled the atoms in a solid as independent harmonic oscillators vibrating at the same frequency, thereby modeling the density of state as a delta function. This simple density of state sometimes provides a good correlation with experimental heat capacity measurements at high temperatures, failing at low temperatures. Debye44 Debye P. Zur theorie der spezifischen Wärmen. Ann Phys. 1912;344(14):789-839.,55 Ashcroft NW, Mermin ND. Solid state physics. 1. ed. New York: Cengage Learning; 2011. p. 491-598. otherwise modeled the vibrations in a solid as normal modes of a continuous elastic body, which corroborates well for long-wavelength vibrations that do not depend on the detailed atomic character of the solid and do conform better experimental scatter to lower temperatures, but failing for many materials with a gap in the density of state66 Schliesser JM, Woodfield BF. Development of a Debye heat capacity model for vibrational modes with a gap in the density of states. J Phys Condens Matter. 2015;27(28):285402.,77 Costa EDM, Lemes NHT, Alves MO, Braga JP. Phonon density of states from the experimental heat capacity: an improved distribution function for solid aluminum using an inverse framework. J Mol Model. 2014;20(8):2360.. Another approach is applying ab initio calculations to predict several thermodynamics properties and the molar heat capacity88 DeTar D. Theoretical ab initio calculation of entropy, heat capacity and heat content. J Phys Chem A. 1998;102(26):5128-41.,99 Li Z, Wang C, Kang W, Li C, Zhang P. Temperature and compression effects on electron heat capacity and electron-phonon coupling in aluminum and beryllium: insights from ab initio simulations. Plasma. 2015;22. http://dx.doi.org/10.1063/1.4935843.
http://dx.doi.org/10.1063/1.4935843...
. Debye’s model does not consider rotational1010 Schroeder DV. An introduction to thermal physics. 1st ed. New York: Addison-Wesley Professional; 1999. 409 p., electronic1111 Mizutani U, Kamiya A. Electronic specific heat measurements for quasicrystals and Frank-Kasper crystals in Mg-AI-Ag, Mg-Al-Cu, Mg-Al-Zn, Mg-Ga-Zn and Al-Li-Cu alloy systems. J Phys Condens Matter. 1991;3(21):3711-8., and magnetic1212 Inden G. Computer calculation of the free energy contribution due to chemical and/or magnetic ordering. In: Proc. Project Meeting CALPHAD; 1976; Dusseldorf. Proceedings. Duesseldorf: Düsseldorf Inst.; 1976. pp. 1-13.

13 Hillert M, Jarl M. A model for alloying in ferromagnetic metals. Calphad. 1978;2(3):227-38.
-1414 Chuang Y, Schmid R, Chang YA. Magnetic contributions to the thermodynamic functions of pure Ni, Co and Fe. Metall Trans, A, Phys Metall Mater Sci. 1985;16(2):153-65. energies contributions to the molar specific heat. The magnetic contribution to the molar heat capacity usually is empirical formulae to account for contributions of Curie, Neel, and Schottky transition anomalies1515 Li Z, Mao H, Selleby M. Thermodynamic modeling of pure Co accounting two magnetic states for the Fcc phase. J Phase Equilibria Diffus. 2018;39(5):502-9.. They calculated the Gibbs-Thomson coefficient for the equilibrium and the non-equilibrium solidification of Al-Cu-Si-Mg alloys as a function of Si content. By observing a different set of Gibbs-Thomson coefficients values for equilibrium and non-equilibrium conditions, they defined this coefficient for alloy by the limiting case1616 Ferreira IL, Garcia A. The application of numerical and analytical approaches for the determination of thermophysical properties of Al–Si–Cu–Mg alloys. Contin Mech Thermodyn. 2020;32:1231-44.. Ferreira et al.1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418. derived a model for pure elements and compounds, regarding the critical radius expressed in terms of the temperature drop ΔTr by means of the correlation between the solid-liquid surface tension σsl and the bulk melting entropy by unit volume ΔS, which writes in terms of the Gibbs-Thomson coefficient Γ. Consequently, a relation between bulk lattice and the surface lattice interatomic spacings, i.e., the reciprocal lattice in both networks and density of state (DoS), and the total number of atoms in the volume and a correspondent density of atoms n limited by nucleation conditions were proposed. Authors regarded the electronic and rotational energies contributions to the molar heat capacity cv.

In this paper, calculations are performed for molar heat capacities of pure elements and phases, compared with the Thermo-Calc Software simulations and with experimental data.

2. Numerical Approach and Analytical Models

The thermal energy of materials is stored in several forms, such as translational, vibrational, rotational, electronic, and magnetic. In a solid element and compound, the translational contribution can be neglected1010 Schroeder DV. An introduction to thermal physics. 1st ed. New York: Addison-Wesley Professional; 1999. 409 p.. The magnetic anomalies are generally modeled empirically1111 Mizutani U, Kamiya A. Electronic specific heat measurements for quasicrystals and Frank-Kasper crystals in Mg-AI-Ag, Mg-Al-Cu, Mg-Al-Zn, Mg-Ga-Zn and Al-Li-Cu alloy systems. J Phys Condens Matter. 1991;3(21):3711-8.

12 Inden G. Computer calculation of the free energy contribution due to chemical and/or magnetic ordering. In: Proc. Project Meeting CALPHAD; 1976; Dusseldorf. Proceedings. Duesseldorf: Düsseldorf Inst.; 1976. pp. 1-13.

13 Hillert M, Jarl M. A model for alloying in ferromagnetic metals. Calphad. 1978;2(3):227-38.

14 Chuang Y, Schmid R, Chang YA. Magnetic contributions to the thermodynamic functions of pure Ni, Co and Fe. Metall Trans, A, Phys Metall Mater Sci. 1985;16(2):153-65.
-1515 Li Z, Mao H, Selleby M. Thermodynamic modeling of pure Co accounting two magnetic states for the Fcc phase. J Phase Equilibria Diffus. 2018;39(5):502-9.. Considering all contributions to the thermal heat capacity, it provides,

c v = c v T r a n s + c v V i b + c v R o t + c v e + c v M a g (1)

In this paper, applying Equation 1 to a solid element or compound, and the discussion about the magnetic contribution cvMag is postponed. By now, the thermal molar heat capacity expressed in terms of its components is given by

c v = c v V i b + c v R o t + c v e (2)

The Gibbs-Thomson coefficient describes for pure elements the melting temperature depression ΔTm K, based on the solid-liquid interface tension σsl N.m1 and on bulk melting entropy by unit volume ΔS J.K1.m3. The Gibbs-Thomson is responsible for the effect of melting temperature drop66 Schliesser JM, Woodfield BF. Development of a Debye heat capacity model for vibrational modes with a gap in the density of states. J Phys Condens Matter. 2015;27(28):285402.,77 Costa EDM, Lemes NHT, Alves MO, Braga JP. Phonon density of states from the experimental heat capacity: an improved distribution function for solid aluminum using an inverse framework. J Mol Model. 2014;20(8):2360., which is associated with the surface free energy, due to elements bonds in the surface presenting a larger lattice interatomic spacing concerning that of the bulk. Let’s consider an isolated solid particle of diameter d in its liquid, the Gibbs-Thomson equation for the structural melting point depression can be expressed by1616 Ferreira IL, Garcia A. The application of numerical and analytical approaches for the determination of thermophysical properties of Al–Si–Cu–Mg alloys. Contin Mech Thermodyn. 2020;32:1231-44.:

Δ T m d = 4 γ s l T m b u l k Δ H d = n 4 γ s l Δ S d ζ σ s l γ s l 4 γ s l Δ S d ζ 4 σ s l Δ S d ζ 4 Γ d ζ (3)

where n and ζ are surface energy/surface tension relation parameters associated with the knowledge that crystals exhibit a surface stress that gives rise to small but detectable strains in the interior of the crystal, assumed as ζ=1m and n=σslγsl, Tmbulk is the bulk melting temperature K, ΔH is the latent heat of melting per unit volume J.m3 and d is the spherical solid-phase diameter m, respectively.

According to the definition presented, the Gibbs-Thomson coefficient Γ can be expressed as,

Γ = σ s l Δ S = σ s l T m b u l k Δ H (4)

The same does not occur to alloys. Jácome et al.1818 Jácome PAD, Landim MC, Garcia A, Furtado AF, Ferreira IL. The application of computational thermodynamics and a numerical model for the determination of surface tension and Gibbs Thomson coefficient of aluminum based alloys. Thermochim Acta. 2011;523(1-2):142-9., while applying several microstructural evolution models for cellular and dendric growth to binary aluminum alloys, concluded that the Gibbs-Thomson of the solvent was not adequate to predict the evolution of high solute content alloys. Consider an alloy given by Al-50at%Ni, which one solvent, ΓAl or ΓNi should be adequate to be carried out in the microstructural evolution models? They provide different results for the primary and secondary dendrite arm spacing. Consequently, the authors defined the alloy Gibbs-Thomson coefficient ΓEq able to encompass the full range of alloy concentration,

Γ E q = σ s l Δ S E q = σ s l T L b u l k Δ H E q (5)

where σsl is the alloy solid-liquid surface tension, ΔSEq and the alloy solidification entropy, TLbulk is the liquidus temperature, and finally, ΔHEq is the alloy latent heat per unit volume J.m3.

As alloy solidification hardly occurs in equilibrium governed by the Lever Rule or in absolute non-equilibrium by the absence of solute back-diffusion in solid governed by the Scheil equation. Many authors contributed to the estimation and calculation of back-diffusion parameter β. In 1981, Clyne and Kurz1919 Clyne TW, Kurz W. Solute redistribution during solidification with rapid solid state diffusion. Metall Trans, A, Phys Metall Mater Sci. 1981;12(6):965-71. derived an approximate relation,

β = 2 α 1 e 1 / α e 1 / 2 α (6)

where α is the standard Fourier diffusion number. This proposed back diffusion treatment uses the Clyne and Kurz1919 Clyne TW, Kurz W. Solute redistribution during solidification with rapid solid state diffusion. Metall Trans, A, Phys Metall Mater Sci. 1981;12(6):965-71. correction of the Brody and Flemings2020 Brody HB, Flemings MC. Solute redistribution during solidification. Trans Metall Soc AIME. 1966;236:615-24.. In 2001, a general and exact description of the back-diffusion parameter was derived, according to Voller2121 Voller VR. On a general back-diffusion parameter. J Korean Cryst Growth Cryst Technol. 2001;226(4):562-8.. The model used in this paper to calculate the back-diffusion parameter β is the one derived by Voller.

What is the connection between equilibrium and non-equilibrium solidification for alloy the Gibbs-Thomson coefficient? It’s common knowledge that in the equilibrium solidification, higher latent heat is expected when compared to that in non-equilibrium conditions, as demonstrated in the work of Ferreira and Garcia1616 Ferreira IL, Garcia A. The application of numerical and analytical approaches for the determination of thermophysical properties of Al–Si–Cu–Mg alloys. Contin Mech Thermodyn. 2020;32:1231-44. and presented in Figure 1 for the alloys example above.

Figure 1
Equilibrium and non-equilibrium latent heat and latent heat per gram as a function of temperature for Al-50at%Ni.

Figure 1 shows heat and latent heat per gram as a function of the temperature solidification range. The curve corresponding to the highest values of heat is associate with infinite diffusion in solid. The lowest curve represents the absence of back-diffusion in the solid phase. Between the curves is the so-called finite back-diffusion condition1919 Clyne TW, Kurz W. Solute redistribution during solidification with rapid solid state diffusion. Metall Trans, A, Phys Metall Mater Sci. 1981;12(6):965-71.

20 Brody HB, Flemings MC. Solute redistribution during solidification. Trans Metall Soc AIME. 1966;236:615-24.
-2121 Voller VR. On a general back-diffusion parameter. J Korean Cryst Growth Cryst Technol. 2001;226(4):562-8..

In 2019, Ferreira and Garcia1616 Ferreira IL, Garcia A. The application of numerical and analytical approaches for the determination of thermophysical properties of Al–Si–Cu–Mg alloys. Contin Mech Thermodyn. 2020;32:1231-44. calculated the Gibbs-Thomson coefficient for the equilibrium and the non-equilibrium solidification of Al-Cu-Si-Mg alloys as a function of Si content. By observing a different set of Gibbs-Thomson coefficients values for equilibrium and non-equilibrium conditions, they defined this coefficient for alloy by the limiting case. Expecting a greater undercooling for non-equilibrium solidification conditions. Under infinite diffusion in the solid (β=1) and no-back-diffusion in the solid (β=0)

Γ E q = σ s l Δ S E q = σ s l T m b u l k Δ H E q (7a)
Δ T E q d = 4 Γ E q d (7b)

and a non-equilibrium one as,

Γ N o n E q = σ s l Δ S N o n E q = σ s l T m b u l k Δ H N o n E q (8a)
Δ T N o n E q d = 4 Γ N o n E q d (8b)

Ferreira and Garcia1616 Ferreira IL, Garcia A. The application of numerical and analytical approaches for the determination of thermophysical properties of Al–Si–Cu–Mg alloys. Contin Mech Thermodyn. 2020;32:1231-44. assumed the solid-solid phase nucleation, based on several observations of equilibrated boundary groove shapes at multiple crystallographic orientations influencing the crystal-melt interfacial energy2222 Erol M, Marasli N, Böyük U. Determination of anisotropy of Crystal-melt interfacial energy from the observed grain boundary groove shapes at multiple orientations. Surf Rev Lett. 2009;16(4):579-88.,2323 Prokofjev SI. Estimations of grain-boundary surface tension in elemental solids. J Mater Sci. 2017;52(8):4265-77., where the Gibbs-Thomson coefficient definition could be generalized for solid-phase transformations by carrying out the grain boundary surface tension σGB to predict the solid phase transformation temperature drop ΔTTd, that is,

Γ T S P T E q = σ G B Δ S S P T N o n = σ G B T T b u l k Δ H S P T N o n (9a)
Δ T T S P T E q d 4 Γ T E q d (9b)

and,

Γ T S P T N o n E q = σ G B Δ S S P T N o n E q = σ G B T T b u l k Δ H S P T N o n E q (10a)
Δ T T d 4 Γ T S P T N o n E q d (10b)

where, ΓTEq and ΓTNonEq is the Gibbs-Thomson coefficient of equilibrium and non-equilibrium solid phase transformation K.m, σGB is the interface tension of the grain-boundary N.m1,ΔSSPTEq and ΔSSPTNonEq are the bulk effective entropies of the solid phase transformation per unit volume for the equilibrium and non-equilibrium in J.m3.K1, respectively.ΔHSPTEq and ΔHSPTNonEqare the bulk heats of the solid phase transformation per unit volume J.m3 for the equilibrium and non-equilibrium conditions, respectively. TTbulk is the bulk solid transformation temperature K. The eutectic and eutectoid reactions are both examples of such solid transformations. The grain boundary surface tension σGB is greater than the solid/liquid surface σsl, as far as the solid-solid interface barrier energy for the nucleation of a new solid phase is greater than that of the solid-liquid.

In the case of an element, for a given nucleus to be stable and to grow, it must have at least a radius r greater than or even equal to the critical radius rC, i.e., rrC. This critical radius can be expressed in terms of the temperature drop ΔTr through the correlation between the solid-liquid surface tension σsl and the bulk melting entropy by unit volume ΔS, which can be written in terms of the Gibbs-Thomson coefficient Γ.

Δ T r r C = 2 Γ r (11)

There is a relation between bulk lattice and the surface network parameters. Consequently, the reciprocal lattice in both networks and density of state (DoS) which is dependent on the nucleus radius. The density of state Dω for a given grain of volume regarding the critical nucleation radius, is defined as

D ω = ω 2 2 π 2 ν 3 (12)

where ω is the frequency, ν is the speed of sound in the solid phase. For a total number of atoms N in the volume and a correspondent density of atoms n, these variables can be expressed as,

N = n (13)

The first Brillouin zone is exchanged by an integral over a sphere of radius kD, containing precisely N wave vectors allowed. As a volume of space k by wave vector is e requires,

2 π 3 N = 4 π k D 3 3 (14)

Then, the density of atoms n can be obtained as,

n = k D 3 6 π 2 = 1 6 π 2 k B Θ D ν 3 (15)

As observed in Equation 15, the compound fundamental frequencies are expressed as a function of the linear combination of Debye’s temperatures of elements i, that is,

ω D = k B Θ D (16)

where, ΘD is the Debye‘s temperature of element, kB and are the constant of Boltzmann and Planck, respectively.

The electronic contribution to cve is written in terms of the phonon energy cvVib as the following,

c v e c v V i b = 5 24 π 3 Z Θ D 3 T 2 T m b u l k (17)

where, Z is the valence of the element, Tmbulk is the melting temperature of element K and T is the absolute temperature K.

In 2019, Ferreira et al.1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418. considered the following approach for the rotational energy,

E R o t = 5 4 2 J J + 1 M ¯ r 2 J (18)

where, J is the rotational level corresponding to integer J=0,1,2,3,, r and M¯ are the atomic radius and the molar mass, respectively. The rotational contribution cvRot to molar heat capacity can be derived as,

c v R o t = 5 4 R 3 k B 2 ω D T + Θ D 2 J J + 1 M ¯ r 2 J . m o l 1 . K 1 (19)

where, R is the universal gas constant J.mol1.K1, ωD is the maximum admissible frequency known as Debye’s frequency.

Debye’s temperature for pure elements is tipically found in the literature55 Ashcroft NW, Mermin ND. Solid state physics. 1. ed. New York: Cengage Learning; 2011. p. 491-598.. The Equation 17 addition of the electronic and Equation 19 of the rotational contributions to cv , it provides,

c v = 1.0 + D ω 9 N a k B T Θ D 3 0 T Θ D x 4 e x e x 1 2 d x 1 + c v e + n + 1 / 2 9.0 c v R o t + 1 E ρ D i a E D i a ρ R T 3 Θ D T m 2 (20)

Let’s analyze the effects of surface and the bulk reciprocal lattices, for a grain of radius greater than the critical concerning the Brillouin zone, whose reciprocal lattice do not coincide for bulk and surface, due to atoms arrangement in both lattices. Another important aspect is the difference in the solute concentration in the bulk and in the surface regions that contributes to change the reciprocal lattice.

The density of state DCompωComp for a grain of a compound of volume , with a certain critical nucleation radius is defined as

D C o m p ω C o m p = ω C o m p 2 2 π 2 ν 3 (21)

where ωComp is the frequency, ν is the speed of sound in the solid compound. For a total number of atoms N in the volume and a correspondent density of atoms n, these variables can be expressed as

N = n (22)

The first Brillouin zone is exchanged by an integral over a sphere of radius kD, containing precisely N wave vectors allowed. As a volume of space k by wave vector is e requires

2 π 3 N = 4 π k D 3 3 (23)

Then, the density of atoms n can be obtained as

n = k D 3 6 π 2 = 1 6 π 2 k B Θ D , C o m p ν 3 (24)

and,

Θ D , C o m p = i = 1 n x i Θ D , i (25)

As observed in Equation 14, the compound fundamental frequencies are expressed as a function of the linear combination of Debye’s temperatures of elements i, that is,

ω D , C o m p = k B Θ D , C o m p (26)

where, ΘD,i is the Debye‘s temperature of element i, ΘD,Comp is the compound Debye‘s temperature, kB and are the constant of Boltzmann and Planck, respectively.

The undercooling for a critical grain of a volume can be written for solid-liquid nucleation

Γ C o m p = σ s l C o m p Δ S C o m p = σ s l T L b u l k Δ H C o m p (27a)
Δ T r C = 2 Γ C o m p r C (27b)

and for solid-solid nucleation as

Γ G C o m p = σ G B C o m p Δ S T C o m p = σ G B T T b u l k Δ H T C o m p (27c)
Δ T r C G = 2 Γ T C o m p r C G (27d)

where, ΓComp is the solid-liquid nucleation Gibbs-Thomson coefficient K.m, ΓGComp is the solid-solid nucleation Gibbs-Thomson coefficient K.m, rC is the critical solid-liquid nucleation grain radius m, rCG is the critical solid-solid nucleation grain radius m.

The element i electronic contribution to cve is written in terms of the phonon energy cvVib as the following,

c v e , i c v V i b = 5 24 π 3 Z i Θ D , i 3 T 2 T m , i b u l k (28)

where, Zi is the valence of element i, Tm,ibulk is the melting temperature of element i K and T is the absolute temperature K.

Thus, the total electronic contribution cve,i to the electronic molar heat capacity cve,Comp, can be expressed as

c v e , C o m p = i = 1 n x i c v e , i + i = 1 n j > i n x i x j c v e , i c v e , j + i = 1 n x i c v e , i (29)

where, xi is the molar fraction of element i.

In 2019, Ferreira et al.1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418., regarded the following approach for the rotational energy for each element i,

E R o t , i = 5 4 2 J i J i + 1 M ¯ i r i 2 J (30)

where, Ji is the rotational level corresponding to integer J=0,1,2,3,, ri and M¯i are the atomic radius and the molar mass of element i, respectively. The rotational contribution cvRot to molar heat capacity can be derived as,

c v R o t = 5 4 R 3 k B 2 ω D , C o m p T + Θ D , C o m p 2 i = 1 n x i J i J i + 1 M ¯ i r i 2 J . m o l 1 . K 1 (31)

where, R is the universal gas constant J.mol1.K1, ωD,Comp is the maximum admissible frequency known as Debye’s frequency.

The modified Debye’s equation for compounds is derived on the Neumann-Koop principal. In this model proposed by Ferreira et al.1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418., the Authors neglected magnetic anomalies such as Curie, Neel, and Schottky. The obtained final equation molar heat capacity for a compound is given by,

c v = 1.0 + D C o m p ω C o m p 9 N a k B T Θ D , C o m p 3 0 T Θ D , C o m p x 4 e x e x 1 2 d x 1 + c v e + n + 1 / 2 9.0 c v R o t + 1 E i ρ D i a E D i a ρ i R T 3 Θ D , A l l o y T L 2 (32)

2.1. Magnetic contribution due to anomalies

Specific heat can be characterized by the fundamental excitations involved in the phase transition2424 Souza M, Paupitz R, Seridonio A, Lagos R. Specific heat anomalies in solids described by a multilevel model. Braz J Phys. 2016;46(2):206-12.. According to the authors, phase transitions involving spin, charge, lattice, and orbital degrees of freedom, which under certain circumstances for a given level of lattice thermal vibration associated with special band structure, would define a state of matter, the so-called superconducting state. The concurrence among those phase transitional mechanisms would easily explain how a superconducting state would be achieved at higher temperatures by increasing the pressure2525 Snider E, Dasenbrock-Gammon N, McBride R, Debessai M, Vindana H, Vencatasamy K, et al. Room-temperature superconductivity in a carbonaceous sulfur hydride. Nature. 2020;586(7829):373-7., which is also a variable of state, such as the solute composition2626 Larsson S. Effect of pressure on superconducting properties. J Supercond Nov Magn. 2015;28(6):1693-8.. According to Souza et al.2424 Souza M, Paupitz R, Seridonio A, Lagos R. Specific heat anomalies in solids described by a multilevel model. Braz J Phys. 2016;46(2):206-12., entropy change is associated with a first-order transition, no matter its nature, and can be directly obtained by integrating specific heat over the temperature T. It means that entropy plays the most important role in predicting anomalies. As Ferreira et al.1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418. stated the transformation (transitional) entropies change the Density of State Dω. Consequently, the entropy must be provided, as far as the nucleation entropies of solid-liquid and solid-solid transformations needed to be provided to determine the critical volume, which encompasses the total number of modes.

3. Results and Discussion

The physical properties of elements used to calculate the molar heat capacity are provided in Table 1 55 Ashcroft NW, Mermin ND. Solid state physics. 1. ed. New York: Cengage Learning; 2011. p. 491-598.,2828 Zalden P, Siegert KS, Rols S, Fischer HE, Schlich F, Hu T, et al. Specific heat of (GeTe)x(Sb2Te3)1-x phase-change materials: the impact of disorder and anharmonicity. Chem Mater. 2014;26(7):2307-12..

Table 1
Thermophysical properties of elements.

Figures 2-5 presents the molar heat capacities for pure Al, Fe, Zr, and Be and the experimental data. Debye’s model predictions are present as a reference for Ferreira’s model calculations1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418.. Instead of using the sound velocities as an approximation for the last term of Equation 21, the value of Young’s modulus and the correspondent density was applied to this term. The model proposed1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418., agrees for low and high temperatures with experimental data for Al2929 Giauque W, Meads J. The heat capacities and entropies of Aluminum and Copper from 15 to 300°K. J Chem Phys. 1941;63:423-32.,3030 Brooks CR, Bingham RE. The specific heat of aluminum from 330 to 890 K and contributions from the formation of vacancies and anharmonic effects. J Phys Chem Solids. 1968;29(9):1553-60., as far as with Thermo-Calc simulations for equilibrium.

Figure 2
Comparison of the molar heat capacity of pure Al by applying Debye, Thermo-Calc, the present approach, and experimental results of Giauque and Meads2929 Giauque W, Meads J. The heat capacities and entropies of Aluminum and Copper from 15 to 300°K. J Chem Phys. 1941;63:423-32. and Brooks and Bingham3030 Brooks CR, Bingham RE. The specific heat of aluminum from 330 to 890 K and contributions from the formation of vacancies and anharmonic effects. J Phys Chem Solids. 1968;29(9):1553-60..
Figure 3
Comparison of the molar heat capacity of pure Fe by applying Debye, Thermo-Calc, and Ferreira et al.1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418., and Valencia and Quested3131 Valencia JJ, Quested P. Thermophysical properties. In: ASM International. ASM handbook: casting. Ohio: ASM International; 2008. p. 468-81. (vol. 15)., ( A ) equilibrium, and, ( B ) non-equilibrium.
Figure 4
Comparison of the molar heat capacity of pure Zr by carrying out Debye, Thermo-Calc, and Ferreira’s model.
Figure 5
Comparison of the molar heat capacity of pure Be by carrying out Debye, Thermo-Calc, Ferreira’s model, and experimental data for low-temperature Cristescu and Simon3232 Cristescu S, Simon F. Specific heat of beryllium, germanium, and hafnium at low temperatures. Z Phys Chem, B Chem Elem Proz Aufbau Mater. 1934;25:273. and high-temperature Howe et al.3333 Howe SD, O’Brien RC, Ambrosi RM, Gross B, Katalenich J, Sailer L, et al. The Mars Hopper: an impulse-driven, long range, long-lived mobile platform utilizing in-situ Martian resources. Acta Astronaut. 2011;69(11-12):1050-6..

In the case of Fe, equilibrium, and non-equilibrium predictions are shown in Figures 3a and Figure 3b, respectively. For low temperature, T300K, the experimental data are those of Kelley3434 Kelley KK. The specific heat of pure iron at low temperatures. J Chem Phys. 1943;11(1):16-8., Euken and Werth3535 Eucken A, Werth H. Die spezifische wärme einiger metalle und metallegierungen bei tiefen temperaturen. Zeits f anorg Allgem Chemie. 1930;188:152-72., and Simon and Swain3636 Simon F, Swain RL. Untersuchungen über die spezifische Wärme bei tiefen temperaturen. Z Phys Chem, B Chem Elem Proz Aufbau Mater. 1935;28B:189-98.. For high temperatures, T>300K, Valencia and Quested3131 Valencia JJ, Quested P. Thermophysical properties. In: ASM International. ASM handbook: casting. Ohio: ASM International; 2008. p. 468-81. (vol. 15). compiled experimental data from three sources from Smithells3737 Smithells CJ. General physical properties. In: Brandes EA, Brook GB, editors. Metals reference book. 7th ed. Amsterdam: Butterworth-Heinemann; 1992., Mills3838 Mills KC. Recommended values of thermophysical properties for selected commercial alloys. Cambridge: Woodhead Publishing; 2002., and Gaskell3939 Gaskell DR. Introduction to metallurgical thermodynamics. New York: McGraw-Hil; 1973. p. 497-501.. The experimental data determined by Awbery and Griffiths4040 Awbery JH, Griffiths E. The thermal capacity of pure iron. Proc R Soc Lond A Math Phys Sci. 1940;174(956):1-15. fits the Thermo-Calc calculations for the equilibrium condition. In Figure 3a a good agreement is observed for low and high temperatures for all experimental data, Thermo-Calc simulations, and Ferreira’s model1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418., except the dataset compilation found in Valencia and Quested3131 Valencia JJ, Quested P. Thermophysical properties. In: ASM International. ASM handbook: casting. Ohio: ASM International; 2008. p. 468-81. (vol. 15).. As Valencia’s compilation dataset provides values below those expected for the equilibrium, it is probably determined under non-equilibrium conditions. Figure 2b presents predictions of Ferreira’s model for non-equilibrium conditions, even assuming that a non-equilibrium condition is established since the beginning of data measurements of the three datasets, which is probably not true, Ferreira’s model fits the data accordingly.

Figure 4 presents the molar specific heat for Zr as a function of temperature from 0 K to 900 K, for Debye’s model, Ferreira’s model predictions for fundamental (n=0) and fourth (n=4) normal modes and Thermo-Calc calculations. The experimental scatter for high temperatures is that found in Pankratz and Mrazek4141 Pankratz LB, Mrazek RV. Thermodynamic properties of elements and oxides. Washington: US Bureau of Mines; 1982., and Valencia and Quested3131 Valencia JJ, Quested P. Thermophysical properties. In: ASM International. ASM handbook: casting. Ohio: ASM International; 2008. p. 468-81. (vol. 15)., whose data is compiled from Smithells3737 Smithells CJ. General physical properties. In: Brandes EA, Brook GB, editors. Metals reference book. 7th ed. Amsterdam: Butterworth-Heinemann; 1992.. A good agreement can be noticed among the experimental datasets and Ferreira’s model calculation.

Figure 5 shows the comparison of predictions for molar specific heat capacity considering Ferreira’s model and Thermo-Calc. The experiments for low temperature, Cristescu and Simon3232 Cristescu S, Simon F. Specific heat of beryllium, germanium, and hafnium at low temperatures. Z Phys Chem, B Chem Elem Proz Aufbau Mater. 1934;25:273., and high temperatures, Howe et al.3333 Howe SD, O’Brien RC, Ambrosi RM, Gross B, Katalenich J, Sailer L, et al. The Mars Hopper: an impulse-driven, long range, long-lived mobile platform utilizing in-situ Martian resources. Acta Astronaut. 2011;69(11-12):1050-6. agree with the theoretical predictions of Ferreira’s model and Thermo-Calc.

Figure 6 presents the calculation of molar specific heat capacity of pure Nickel as a function of temperature, comparison with Thermo-Calc and, the experimental data found in Desai4242 Desai PD. Thermodynamic properties of nickel. Int J Thermophys. 1987;8(6):763-80.. The magnetic transition ferromagnetic/paramagnetic of Ni can be captured if the calculation of molar specific heat capacity is performed by the composition of equivalent wavevectors at lattice points n’s, using n = {0, 1, 3, 7, 10, 13, 10, 7, 3, 1, 0} the model fits the experimental scatter. It seems transformation pass through the sequence of equivalent wavevectors, k+2π/a, where a is the lattice spacing at the current temperature, and its anomalies found near the Curie point, as observed by Kohlhaas et al.4343 Kohlhaas R, Dünner PH, Schmitz-Pranghe N. Über die Temperaturabhangigkeit der Gitterparameter von Eisen, Kobalt und Nickel im bereich hoher temperaturen. Z Angew Phys. 1967;23:245-9..

Figure 6
Comparison of the molar heat capacity of pure Ni by applying Debye, Thermo-Calc, Ferreira’s model, and experimental data from Desai4242 Desai PD. Thermodynamic properties of nickel. Int J Thermophys. 1987;8(6):763-80..

Figures 7-9 present the molar heat capacities for AL2CU, AL3MG2, and AL3NI2. Debye’s model’s predictions are present only as a reference for Ferreira’s model calculations1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418.. Instead of using the sound velocities as an approximation for the last term of Equation 32, the value of Young’s modulus and the correspondent density was applied to this term. From the point of view of nucleation of a compound, for a nucleating stable grain concerning the Brillouin zone, another aspect that plays an important role is the difference in the solute concentration in the bulk and the surface regions contributes to change both reciprocal lattices. In the case of the AL2CU phase, calculations are performed by applying Thermo-Calc, and TTAL7 database, Thermo-Calc TCAPI5 (Thermo-Calc Application Programming Interface version 5) interface using C++ language routine to perform calculations of specific heat by applying the TCMP2 database, and, finally, Thermo-Calc and COST-507 databases. The first two simulations seem to agree with each other for the whole temperature range. Thermo-Calc using COST-507 seems to deviate from equilibrium but still provides good results slightly. Ferreira’s model prediction fits Thermo-Calc and TCAPI simulations for all databases until temperature reaches 450 K. Beyond, this temperature agrees better with the Thermo-Calc (TTAL7) and TCAPI5 (TCMP2) calculations. In Figure 8, calculations were made for the AL3MG2 phase for the equilibrium conditions, where Thermo-Calc, TCAPI, and Ferreira’s model simulations agreed for the whole temperature range. In the Al3Ni2 intermetallic phase, only Thermo-Calc using the TTAL7 database, and TCAPI5 routine by utilizing TCMP2 calculation are performed. Ferreira’s model results behave equally with the specific molar heat capacity predicted of Thermo-Calc and TCAPI calculations for the whole range, as shown in Figure 9.

Figure 7
Comparison of the molar heat capacity of the AL2CU phase by carrying out Debye, Thermo-Calc, and Thermo-Calc TCAPI5 simulations and the Ferreira’s model1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418..
Figure 8
Comparison of the molar heat capacity of the AL3MG2 phase by carrying out Debye, Thermo-Calc, and Thermo-Calc TCAPI5 simulations and the Ferreira’s model1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418..
Figure 9
Comparison of the molar heat capacity of the AL3NI2 phase by carrying out Debye, Thermo-Calc, and Thermo-Calc TCAPI5 simulations and Ferreira‘s model1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418..

Figure 10 presents the calculation for the molar specific heat of GeTe compound from 100 to 900 K by applying Ferreira’s model and comparing it with two sets of experimental data found in Zalden et al.2828 Zalden P, Siegert KS, Rols S, Fischer HE, Schlich F, Hu T, et al. Specific heat of (GeTe)x(Sb2Te3)1-x phase-change materials: the impact of disorder and anharmonicity. Chem Mater. 2014;26(7):2307-12.. A very interesting physical behavior can be observed, as around 650 K there’s a transition from the second n=1 normal mode to the fundamental normal mode n=0 captured by Ferreira’s model’s predictions. Calculations with other Pure metals and phases have shown similar behavior.

Figure 10
Comparison of the molar heat capacity of Germanium Telluride (GeTe) by carrying out Debye, Ferreira’s model1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418., and experimental data Zalden et al.2828 Zalden P, Siegert KS, Rols S, Fischer HE, Schlich F, Hu T, et al. Specific heat of (GeTe)x(Sb2Te3)1-x phase-change materials: the impact of disorder and anharmonicity. Chem Mater. 2014;26(7):2307-12..

The critical volume , of radius rrC, which depends on the nucleation kinetics, as demonstrated by Equation 11 and Equation 12, provide the total number of modes N, under nucleation kinetics restrains. For the lack of nucleation data, the perovskite K2La2Ti3O10 will not be calculated in this paper as many thermophysical properties couldn’t be found in the literature. Figure 11 represents only the first dataset of Sankovich et al.4444 Sankovich AM, Markin AV, Smirnova NN, Zvereva IA. Heat capacity and the thermodynamic properties of layered perovskite-like oxides K2La2Ti3O10 and K2Nd2Ti3O10. Russ J Phys Chem. 2019;93(3):407-16..

Figure 11
Comparison of the molar heat capacity of perovskite K2La2Ti3O10 by carrying out Debye, Ferreira et al.1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418. model and first dataset of Sankovich et al.4444 Sankovich AM, Markin AV, Smirnova NN, Zvereva IA. Heat capacity and the thermodynamic properties of layered perovskite-like oxides K2La2Ti3O10 and K2Nd2Ti3O10. Russ J Phys Chem. 2019;93(3):407-16..

4. Conclusion

The model proposed previously by Ferreira et al.1717 Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418. the density of state (DoS) is a function of the nucleation parameters, which influence the reciprocal lattices in the bulk and in the surface of the grain to determine the total number of modes, and consequently, the correct predictions of the Density of State. The model successfully predicts the molar heat capacity's behavior from absolute zero to high temperatures to the melting point for elements and compounds. Furthermore, the formulation encompasses solid-solid grain nucleation, as long as the Gibbs-Thomson coefficient, which is also defined in terms of the solid-solid nucleation, to predict, for instance, the molar specific heat capacity of perlite. A composition of equivalent wavevectors successfully predicted the magnetic transition of Ni. An interesting aspect is the transition mode of the wavevector observed in GeTe, which behavior can also be verified in pure metals, as far as many other substances. The lack of trustworthy thermophysical properties at nucleation temperature (fusion temperature) for perovskite K2La2Ti3O10 did not allow the author to extend the predictions for higher temperatures, as far as Ferreira’s model does not have any adjustment parameters, it encompasses only physical properties. Nevertheless, it has predicted the behavior of perovskite's heat capacity correctly where Debye’s model fails, as it has no information of the critical volume that contains the total number of modes for a particular equilibrium and non-equilibrium state. Anomalies observed in the molar specific heat capacity, such as thermal, magnetic, configurational transitions, and electronic, can be treated by combining equivalent wavevectors and the Density of State (DoS). This would also explain how the change of other thermodynamic variables of state, such as composition and pressure, can dislocate the superconducting state's critical temperature to higher values.

5. Acknowledgments

The authors acknowledge the financial support provided by FAPERJ (The Scientific Research Foundation of the State of Rio de Janeiro), CAPES (Coordenação de Aperfeiçoamento de Pessoal de Nível Superior - Brasil - Finance Code 001) and CNPq (National Council for Scientific and Technological Development).

6. References

  • 1
    Abu-Eishah SI, Haddad Y, Solieman A, Bajbouj A. A new correlation for the specific heat of metals, metal oxides and metal fluorides as a function of temperature. Lat Am Appl Res. 2004;34:257-65.
  • 2
    Sundman B, Kattner UR, Sigli C, Stratmann M, Le Tellier R, Palumbo M, et al. The OpenCalphad thermodynamic software interface. Comput Mater Sci. 2016;125:188-96.
  • 3
    Einstein A. Die plancksche theorie der strahlung und die theorie der spezifischen Wärme. Ann Phys. 1907;327(1):180-90.
  • 4
    Debye P. Zur theorie der spezifischen Wärmen. Ann Phys. 1912;344(14):789-839.
  • 5
    Ashcroft NW, Mermin ND. Solid state physics. 1. ed. New York: Cengage Learning; 2011. p. 491-598.
  • 6
    Schliesser JM, Woodfield BF. Development of a Debye heat capacity model for vibrational modes with a gap in the density of states. J Phys Condens Matter. 2015;27(28):285402.
  • 7
    Costa EDM, Lemes NHT, Alves MO, Braga JP. Phonon density of states from the experimental heat capacity: an improved distribution function for solid aluminum using an inverse framework. J Mol Model. 2014;20(8):2360.
  • 8
    DeTar D. Theoretical ab initio calculation of entropy, heat capacity and heat content. J Phys Chem A. 1998;102(26):5128-41.
  • 9
    Li Z, Wang C, Kang W, Li C, Zhang P. Temperature and compression effects on electron heat capacity and electron-phonon coupling in aluminum and beryllium: insights from ab initio simulations. Plasma. 2015;22. http://dx.doi.org/10.1063/1.4935843
    » http://dx.doi.org/10.1063/1.4935843
  • 10
    Schroeder DV. An introduction to thermal physics. 1st ed. New York: Addison-Wesley Professional; 1999. 409 p.
  • 11
    Mizutani U, Kamiya A. Electronic specific heat measurements for quasicrystals and Frank-Kasper crystals in Mg-AI-Ag, Mg-Al-Cu, Mg-Al-Zn, Mg-Ga-Zn and Al-Li-Cu alloy systems. J Phys Condens Matter. 1991;3(21):3711-8.
  • 12
    Inden G. Computer calculation of the free energy contribution due to chemical and/or magnetic ordering. In: Proc. Project Meeting CALPHAD; 1976; Dusseldorf. Proceedings. Duesseldorf: Düsseldorf Inst.; 1976. pp. 1-13.
  • 13
    Hillert M, Jarl M. A model for alloying in ferromagnetic metals. Calphad. 1978;2(3):227-38.
  • 14
    Chuang Y, Schmid R, Chang YA. Magnetic contributions to the thermodynamic functions of pure Ni, Co and Fe. Metall Trans, A, Phys Metall Mater Sci. 1985;16(2):153-65.
  • 15
    Li Z, Mao H, Selleby M. Thermodynamic modeling of pure Co accounting two magnetic states for the Fcc phase. J Phase Equilibria Diffus. 2018;39(5):502-9.
  • 16
    Ferreira IL, Garcia A. The application of numerical and analytical approaches for the determination of thermophysical properties of Al–Si–Cu–Mg alloys. Contin Mech Thermodyn. 2020;32:1231-44.
  • 17
    Ferreira IL, Castro JA, Garcia A. Determination of heat capacity of pure metals, compounds and alloys by analytical and numerical methods. Thermochim Acta. 2019;682:178418.
  • 18
    Jácome PAD, Landim MC, Garcia A, Furtado AF, Ferreira IL. The application of computational thermodynamics and a numerical model for the determination of surface tension and Gibbs Thomson coefficient of aluminum based alloys. Thermochim Acta. 2011;523(1-2):142-9.
  • 19
    Clyne TW, Kurz W. Solute redistribution during solidification with rapid solid state diffusion. Metall Trans, A, Phys Metall Mater Sci. 1981;12(6):965-71.
  • 20
    Brody HB, Flemings MC. Solute redistribution during solidification. Trans Metall Soc AIME. 1966;236:615-24.
  • 21
    Voller VR. On a general back-diffusion parameter. J Korean Cryst Growth Cryst Technol. 2001;226(4):562-8.
  • 22
    Erol M, Marasli N, Böyük U. Determination of anisotropy of Crystal-melt interfacial energy from the observed grain boundary groove shapes at multiple orientations. Surf Rev Lett. 2009;16(4):579-88.
  • 23
    Prokofjev SI. Estimations of grain-boundary surface tension in elemental solids. J Mater Sci. 2017;52(8):4265-77.
  • 24
    Souza M, Paupitz R, Seridonio A, Lagos R. Specific heat anomalies in solids described by a multilevel model. Braz J Phys. 2016;46(2):206-12.
  • 25
    Snider E, Dasenbrock-Gammon N, McBride R, Debessai M, Vindana H, Vencatasamy K, et al. Room-temperature superconductivity in a carbonaceous sulfur hydride. Nature. 2020;586(7829):373-7.
  • 26
    Larsson S. Effect of pressure on superconducting properties. J Supercond Nov Magn. 2015;28(6):1693-8.
  • 27
    George PK, Thompson ED. The Debye temperature of Nickel from 0 to 300K. J Phys Chem Solids. 1967;28(12):2539-44.
  • 28
    Zalden P, Siegert KS, Rols S, Fischer HE, Schlich F, Hu T, et al. Specific heat of (GeTe)x(Sb2Te3)1-x phase-change materials: the impact of disorder and anharmonicity. Chem Mater. 2014;26(7):2307-12.
  • 29
    Giauque W, Meads J. The heat capacities and entropies of Aluminum and Copper from 15 to 300°K. J Chem Phys. 1941;63:423-32.
  • 30
    Brooks CR, Bingham RE. The specific heat of aluminum from 330 to 890 K and contributions from the formation of vacancies and anharmonic effects. J Phys Chem Solids. 1968;29(9):1553-60.
  • 31
    Valencia JJ, Quested P. Thermophysical properties. In: ASM International. ASM handbook: casting. Ohio: ASM International; 2008. p. 468-81. (vol. 15).
  • 32
    Cristescu S, Simon F. Specific heat of beryllium, germanium, and hafnium at low temperatures. Z Phys Chem, B Chem Elem Proz Aufbau Mater. 1934;25:273.
  • 33
    Howe SD, O’Brien RC, Ambrosi RM, Gross B, Katalenich J, Sailer L, et al. The Mars Hopper: an impulse-driven, long range, long-lived mobile platform utilizing in-situ Martian resources. Acta Astronaut. 2011;69(11-12):1050-6.
  • 34
    Kelley KK. The specific heat of pure iron at low temperatures. J Chem Phys. 1943;11(1):16-8.
  • 35
    Eucken A, Werth H. Die spezifische wärme einiger metalle und metallegierungen bei tiefen temperaturen. Zeits f anorg Allgem Chemie. 1930;188:152-72.
  • 36
    Simon F, Swain RL. Untersuchungen über die spezifische Wärme bei tiefen temperaturen. Z Phys Chem, B Chem Elem Proz Aufbau Mater. 1935;28B:189-98.
  • 37
    Smithells CJ. General physical properties. In: Brandes EA, Brook GB, editors. Metals reference book. 7th ed. Amsterdam: Butterworth-Heinemann; 1992.
  • 38
    Mills KC. Recommended values of thermophysical properties for selected commercial alloys. Cambridge: Woodhead Publishing; 2002.
  • 39
    Gaskell DR. Introduction to metallurgical thermodynamics. New York: McGraw-Hil; 1973. p. 497-501.
  • 40
    Awbery JH, Griffiths E. The thermal capacity of pure iron. Proc R Soc Lond A Math Phys Sci. 1940;174(956):1-15.
  • 41
    Pankratz LB, Mrazek RV. Thermodynamic properties of elements and oxides. Washington: US Bureau of Mines; 1982.
  • 42
    Desai PD. Thermodynamic properties of nickel. Int J Thermophys. 1987;8(6):763-80.
  • 43
    Kohlhaas R, Dünner PH, Schmitz-Pranghe N. Über die Temperaturabhangigkeit der Gitterparameter von Eisen, Kobalt und Nickel im bereich hoher temperaturen. Z Angew Phys. 1967;23:245-9.
  • 44
    Sankovich AM, Markin AV, Smirnova NN, Zvereva IA. Heat capacity and the thermodynamic properties of layered perovskite-like oxides K2La2Ti3O10 and K2Nd2Ti3O10 Russ J Phys Chem. 2019;93(3):407-16.

Publication Dates

  • Publication in this collection
    12 Mar 2021
  • Date of issue
    2021

History

  • Received
    18 Nov 2020
  • Reviewed
    12 Jan 2021
  • Accepted
    26 Jan 2021
ABM, ABC, ABPol UFSCar - Dep. de Engenharia de Materiais, Rod. Washington Luiz, km 235, 13565-905 - São Carlos - SP- Brasil. Tel (55 16) 3351-9487 - São Carlos - SP - Brazil
E-mail: pessan@ufscar.br