Next Article in Journal
Acid-Modified Clays for the Catalytic Obtention of 5-Hydroxymethylfurfural from Glucose
Previous Article in Journal
Photodegradation of Fipronil by Zn-AlPO4 Materials Synthesized by Non-Hydrolytic Sol–Gel Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Hill Equation for Solid Specific Heat Capacity Calculation

1
School of Engineering & Technology, Western Carolina University, Cullowhee, NC 28783, USA
2
Department of Chemical Engineering, University of Colorado Boulder, Boulder, CO 80309, USA
3
Department of Chemical Engineering, Florida State University, Tallahassee, FL 32310, USA
*
Author to whom correspondence should be addressed.
ChemEngineering 2022, 6(4), 56; https://doi.org/10.3390/chemengineering6040056
Submission received: 20 June 2022 / Revised: 6 July 2022 / Accepted: 12 July 2022 / Published: 15 July 2022

Abstract

:
The Hill Equation and Hill Coefficient have been used extensively in biochemistry for the description of noncovalent binding. Previously, the Hill Coefficient was correlated with the Gibbs free energy, which suggests that the Hill Equation might be extensible to covalent binding phenomena. To evaluate this possibility, the Hill Equation was compared to the Debye Model and Einstein Solid in the calculation of heat capacity for 53 covalent solids, which included stainless steels and refractory ceramics. Hill Equation specific heat predictions showed a standard error of 0.37 J/(mole⋅Kelvin), whereas errors from the Debye Model and Einstein Solid were higher at 0.45 J/(mole⋅Kelvin) and 0.81 J/(mole⋅Kelvin), respectively. Furthermore, the Hill Equation is computationally efficient, a feature that can accelerate industrial chemical process simulation(s). Given its speed, simplicity, and accuracy, the Hill Equation likely offers an alternative means of specific heat calculation in chemical process models.

1. Introduction

Chemical engineers commonly rely on computationally efficient and readily implemented thermodynamic expressions to simulate, design, scale, and control process equipment [1,2,3,4,5]. Thus, although complex ab initio methods are now available for accurate thermodynamic predictions, especially for solid specific heat, classical thermodynamic models continue to pervade the process industries [4,5,6,7,8,9,10,11,12]. Notably, despite the current availability of both ab initio and classical thermodynamic methods for solid specific heat estimation, thermodynamic databanks still recommend empirical piecewise polynomials for heat capacity calculation [13,14]. These empirical models are likely inaccurate outside their source data and they require the implementation of ungainly lookup tables and case structures in process simulations. To find a feasible alternative for use in industrial process models we explored the accuracy of classical thermodynamic expressions for the prediction of heat capacity across 53 inorganic compounds and technical solids, which included stainless steels and refractory ceramics. Specifically, heat capacity predictions from two canonical equations, the Debye Model and Einstein Solid, were compared to results from the Hill Equation of biochemistry and an empirical Multilinear Model [15,16,17]. Although use of the Hill Equation in this role was novel, evidence herein shows that this facile mathematical expression can outperform alternatives. Thus, the Hill Equation presents a possible new tool for specific heat determination in industrial process simulation.
The mathematical description of solid heat capacity must capture temperature effects, which are sigmoidal. Figure 1 shows typical changes in specific heat across temperatures for a generic solid [18]. At absolute zero, solid heat capacity vanishes as molecular motion freezes. Near melting a maximal specific heat (Cvmax) is achieved as all aspects of a solid lattice are excited. If molecules within a solid are viewed as simple harmonic oscillators, the Dulong–Petit limit predicts Cvmax = 3R at incipient melting [19]. Classical expressions for heat capacity calculation respect this limit, although varied vibrational modes (phonons) and electronic effects within a solid lead to departures from Dulong and Petits’ theory [15,16,18].
The extremum and sigmoidal behavior shown in Figure 1 are hallmarks of expressions that successfully calculate constant-volume specific heat. We hypothesized that the Hill Equation from biochemistry could capture these features. To adapt this equation for specific heat, concentration within the Hill Equation was substituted with temperature:
C v ( T ) 3 R = T n K + T n
where R is the gas constant, 3R is the maximum heat capacity for a classical three-dimensional solid lattice (Figure 1), T is absolute temperature, K is the dissociation constant, and n is the Hill Coefficient (variable definitions are available in Appendix A). Prior derivations that established an equivalence between intensive properties in molecular aggregates support the replacement of molar concentration, which is usually observed within the Hill Equation, by temperature [21,22]. For example, a similar substitution is seen across the Hill Equation and Langmuir isotherm, which share an identical functional form, but depend on molar concentration or pressure, respectively.
The Hill Equation, as posed, satisfies the idealized Dulong–Petit limit [19]:
lim T C v ( T ) = 3 R
For departures from ideality due to electronic, magnetic, and cooperative modes of thermal energy, the constant 3R was allowed to vary:
C v ( T ) = C v max T n K + T n
Thus, three parameters fit the Hill Equation to specific heat data: Cvmax, K, and n (Appendix A). In solid specific heat estimation, it was expected that n would always exceed unity (n > 1), which is indicative of cooperative binding in noncovalent systems [23].
The Hill Equation was compared to the Einstein Solid, which is derived from the microcanonical ensemble [16]:
C v ( T ) 3 R = ( ε R T ) 2 e ε / ( R T ) ( e ε / ( R T ) 1 ) 2
where ε is a quantitized energy of solid vibration and all other variables are described in Appendix A. For departures from ideality due to electronic, magnetic, and cooperative modes of thermal energy, the constant 3R was allowed to vary:
C v ( T ) = C v max ( ε R T ) 2 e ε / ( R T ) ( e ε / ( R T ) 1 ) 2  
Thus, two parameters fit the Einstein Solid to specific heat data: Cvmax and ε.
The Einstein Solid often underestimates low-temperature specific heat (<100 K) [24], which prompted development of the Debye Model [15]:
C v ( T ) 3 R = 3 ( T T D ) 3 0 T D / T x 4 e x ( e x 1 ) 2 d x  
where TD is the Debye temperature, indicative of the highest energy frequency supported by the solid lattice, and all other variables are as previously described. For departures from ideality due to electronic, magnetic, and cooperative modes of thermal energy the constant 3R was allowed to vary:
C v ( T ) = C v max 3 ( T T D ) 3 0 T D / T x 4 e x ( e x 1 ) 2 d x  
Thus, two parameters fit the Debye Model to specific heat data: Cvmax and TD.
Typically, constant-pressure specific heat, not constant-volume specific heat, is engineering relevant (Figure 1). The two quantities are related as follows [20]:
C p ( T ) = C v ( T ) + α 2 ( T ) V ( T ) K T ( T ) T  
where α is the coefficient of isobaric thermal expansion, V is the mole volume, and KT is the isothermal bulk modulus (Appendix A). As has been pursued previously, the last term of Equation (8) was approximated as a linear function with slope m [25]:
C p ( T ) = C v ( T ) + m T  
In all cases, slope m provided an additional variable for fitting specific heat data.
The Hill Equation, Einstein Solid, and Debye model were compared to an empirical model, termed the Multilinear Model, for heat capacity calculation [17]. The Multilinear Model was shown to be superior to simple polynomial fits for a wide range of solids [17]. It features four fitted parameters: a, b, c, and d:
C p ( T ) = a T b e c T e d / T  
As described in the Results, each specific heat equation was fit to the same dataset drawn from 53 covalent solids. Therein, the Hill Equation performed favorably when compared to alternatives. The Discussion attempts to explain the Hill Equation’s applicability to specific heat calculation, which may have a fundamental basis in theory. However, at this time the expression is, at most, potentially a semi-empirical model for specific heat calculation.

2. Materials and Methods

Presumably, only accurate parameter estimates for the Hill Equation, Einstein Solid, Debye Model, and Multilinear model match these equations to experimental data. To assess heat capacity prediction by each formula, parameters were fit to data for 53 inorganic solids, alloys, and technical ceramics from information in literature publications and the NIST and DIPPR databanks [13,14,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45]. Fitted parameters were discovered through minimization of the L2 norm between model predictions and experimental heat capacity results from temperatures that spanned absolute zero to 3000 Kelvin. Specifically, MATLAB’s fmincon algorithm was used to find Einstein Solid, Debye Model, and Hill Equation parameters described in the Introduction. To initialize parameter estimates, facets of the specific heat formulas were exploited. For example, for all formulae m were approximated at high temperature prior to numerical optimization:
lim T d C v d T m  
Similarly, high-temperature data provided an estimate of Cvmax:
lim T C v C v max  
The ultimate values of parameter estimates for all models are available in the publication Supplement (Table S1). Multilinear Model parameters were fit by linear regression as previously described [17].

3. Results

Experimental data for 53 inorganic and technical solids were mined from thermochemical tables and literature reports (Table S1, https://zenodo.org/record/6779574, accessed on 11 July 2022), data which were fit to the Debye Model, Einstein Solid, Hill Equation, and Multilinear Model. Table S1 lists the results and Figure 2 plots experimental heat capacity against calculated heat capacity for all 53 solids. The correspondence between experiment and prediction appeared strong for the putative fundamental models, especially the Hill Equation, but was variable for the empirical Multilinear Model (Figure 2). Figure 3 and Figure 4 show example fits for technical metals and refractory ceramics. Generally, the Einstein Solid underestimated heat capacity at low temperatures (<100 K), as has been previously observed, while the Hill Equation overestimated specific heat in this range [24]. To further parse model fidelity across the temperature range, absolute relative errors were evaluated:
A R E ( T i ) = | C p ( T i ) C p , exp ( T i ) C p , exp ( T i ) |  
where Cp,exp is the experimentally measured specific heat at temperature Ti. The average absolute error was determined for each solid over T ≤ 100 K and T > 100 K. Figure 5 shows the results, wherein the Debye Model best minimized relative errors below 100 K (Figure 5A). However, the Hill Equation best minimized errors at temperatures above 100 K (Figure 5C). Furthermore, the Hill Equation also showed the lowest standard error of the estimate across the entire temperature range (Figure 5B), where the standard estimate error was:
S E = i = 1 k ( C p ( T i ) C p , exp , ( T i ) ) 2 k l  
for k experimental observations and l fitted parameters [17]. The Hill Equation showed an average standard error of 0.37 J/(mole·Kelvin) across all the 53 examined solids. Average standard error of the Debye Model and Einstein Solid predictions were 0.45 J/(mole·Kelvin) and 0.81 J/(mole·Kelvin), respectively. Figure 6 shows that the correlation coefficient, R2, was strong for all the putative fundamental models, but was inconsistent for the empirical Multilinear Model. Overall, the Hill Equation outperformed alternative representations of heat capacity at temperatures above 100 K and was highly accurate across all temperatures.
Prior work has shown a correlation between the Hill coefficient and Gibbs free energy of interaction in noncovalent binding [23]. Under Richard’s Rule, the standard enthalpy of fusion underpins Gibb’s free energy in covalent solid binding [18]. A correlation between the enthalpy of fusion and the Hill Coefficient could imply a fundamental basis for use of the Hill Equation in solid specific heat estimation. To explore this possibility, the discovered Hill Coefficients were plotted against the enthalpy of fusion for alkaline metals and alkaline halides. Figure 6 shows the results. A linear trend was observed within the alkaline metals and alkaline halide groups on the periodic table, although not across alkaline metals and alkaline halides generally.

4. Discussion

The Hill Equation was fit to experimental heat capacity data for 53 inorganic compounds and technical solids. At common chemical process temperatures (>100 K [46]), the Hill Equation predicted specific heat capacity more accurately than the Debye Model, the Einstein Solid, and the Multilinear Model. At lower temperatures the Debye Model proved superior. However, the Debye Model requires the evaluation of a numerical integral, which can slow process simulation. Table 1 shows that use of the Debye Model quintuples computation time relative to the Hill Equation. Such time inflation can add additional computational burden to industrial-process engineering problems, which are already resource-intensive, frequently time-sensitive, and often NP-hard [47,48,49]. Thus, applications for the analytic Hill Equation likely exist at elevated temperatures whenever computational efficiency is essential.
Whether the Hill Equation for heat capacity harbors a fundamental basis in physics and statistical thermodynamics is unclear. Prior authors showed a correlation between the Hill Coefficient and Gibb’s free energy when molecules noncovalently interact [23]. Based on this result, the energy of covalent fusion likely also correlates with the Hill Coefficient [18]. Figure 6 shows that this proves true within groups on the periodic table, where a linear correlation is evident between the Hill Coefficient and heat of fusion. The Hill Coefficient reflects cooperativity in binding phenomena, where large n indicates that multiple molecules must simultaneously coordinate to bind and relax [23]. Thus, n may reflect the degrees of freedom that coordinate and/or excite a solid’s unit cell. Although nontraditional, this perspective allows a derivation that supports Hill Equation use for solid specific heat calculation (Appendix B). Additionally, the Hill Equation is closely related to the logistic equation [50]. Thus, the Hill Equation for heat capacity might be reconciled by a statistical and/or populations view of solid lattice behavior. However, currently a rigorous derivation for the Hill Equation in specific heat calculation is lacking. Based on current evidence the Hill Equation represents, at most, an analytic and efficient expression for heat capacity calculation that is likely useful in chemical-process simulation.
Figure 6. Enthalpy of fusion and Hill Coefficient correlations for the alkali metals and alkali halides. Fusion energies were taken from the CRC Handbook [51].
Figure 6. Enthalpy of fusion and Hill Coefficient correlations for the alkali metals and alkali halides. Fusion energies were taken from the CRC Handbook [51].
Chemengineering 06 00056 g006

5. Conclusions

Four different specific heat equations were fit to the same dataset drawn from 53 covalent solids. Therein, the Hill Equation performed favorably when compared to established alternatives, especially above 100 Kelvin. Below 100 Kelvin, the Debye Model remains the best descriptor of solid heat capacity, but requires the evaluation of a numerical integral. Conversely, the simpler Hill Equation calculates substantially faster than Debye computations (Table 1), but remains highly accurate. In heat capacity prediction, Hill Equation standard error of the estimate (0.37 J·kelvin−1·mole−1) was similar to the Debye Equation (0.45 J·kelvin−1·mole−1). Thus, the Hill Equation is likely useful as a fast and reliable expression for heat capacity calculation in applied industrial-process simulations. However, without a rigorous theoretical basis, currently the Hill Equation is, at most, potentially a semi-empirical model of specific heat.

Supplementary Materials

The following supporting information can be downloaded at https://zenodo.org/record/6779574 (accessed on 11 July 2022).

Author Contributions

Validation, J.C.T.; investigation, S.C.R.; supervision, B.A.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Variable definitions used in equations herein:
SymbolUnitsDefinition
αKelvin−1coefficient of isobaric thermal expansion
a fitted parameter of the empirical multilinear model
AREunitlessabsolute relative error
b fitted parameter of the empirical multilinear model
c fitted parameter of the empirical multilinear model
Cp(T)J/(mole·Kelvin)constant pressure temperature dependent solid heat capacity
Cp,exp(Ti)J/(mole·Kelvin)specific heat observed in experiment at temperature Ti
CvmaxJ/(mole·Kelvin)constant-volume solid heat capacity as near melting
Cv(T)J/(mole·Kelvin)constant-volume temperature-dependent solid heat capacity
d fitted parameter of the empirical multilinear model
D[Joules]ndeabsorption constant
εJoules/molequantitized energy level within the Einstein Solid
kintegernumber of datapoints
kBJoules/KelvinBoltzmann constant
KKelvinndissociation constant
KTPascalsisothermal bulk modulus
lintegernumber of fitted parameters in a given heat capacity model
mJ/(mole·Kelvin2)fitted slope
nunitlessHill Coefficient
Qunitlesspartition function
RJ/(mole·Kelvin)gas constant
TKelvinabsolute temperature
TDKelvinDebye temperature
TfusionKelvinTemperature of solid fusion (phase change)
TiKelvinThe ith temperature in an experimental dataset
UJoules/moleinternal energy
V(T)m3/moletemperature-dependent mole volume
SEJ/(mole·Kelvin)standard error of the estimate

Appendix B

A basic derivation for specific heat from the Hill Equation borrows liberally from kinetic theory, statistical thermodynamics, and continuum mechanics. Within the context of kinetic theory, the Hill Coefficient n can be viewed as degrees of freedom for the equipartition of energy that excites a solid’s unit cell:
n k B T 2  
where kB is the Boltzmann Constant and T is absolute temperature. For simplicity, we consider the energy needed to excite two unit cells, the minimal realization of a solid:
n k B T  
This energy can be absorbed or released by the solid C:
n ( k B T ) + C ( k B T ) n C  
Within a continuum solid this energetic equilibrium can be is summarized by a deabsorption coefficient:
D = [ C ] [ k B T ] n [ ( k B T ) n C ]  
The sum of states within the continuum solid is partitioned between excited and relaxed sites:
[ C ] Q = [ C ] + [ ( k B T ) n C ] = [ C ] ( 1 + [ k B T ] n D )  
where Q is a partition function. This interpretation of the partition function Q, sometimes called the “binding polynomial,” aligns with established published analysis [52]:
Q = 1 + ( k B T ) n D  
It can be used within the definition of internal energy at constant volume from statistical thermodynamics:
U = k B T 2 ln { Q } T  
where the temperature derivative of U gives specific heat:
C v = U T = 2 k B T ln { Q } T + k B T 2 T ( ln { Q } T )  
If the rightmost higher derivative is considered negligible, Equation (A8) simplifies:
C v 2 k B T ln { Q } T  
The Hill Equation results when partition function Q (Equation (A6)) is subjected to the approximate derivative for Cv (Equation (A9)):
C v = 2 k B T ln { 1 + ( k B T ) n / D } T = 2 n k B ( k B T ) n D + ( k B T ) n  
Equation (A10) is from an extensive analysis based on a minimal solid composed of two-unit cells. The intensive heat capacity of a single-unit cell is presumably half:
C v = n k B ( k B T ) n D + ( k B T ) n  
Multiplication by Avogadro’s number gives:
C v = n R ( k B T ) n D + ( k B T ) n  
Finally, rearrangement gives:
C v = n R T n D / k B n + T n  
where D/kn = K and this division lends the dissociation constant units of Kelvinn. The postulated Hill Equation for heat capacity calculation (Equation (1)) arises for n = 3 degrees of freedom:
C v = 3 R T n K + T n  
This derivation for the Hill Equation in specific heat is speculative. However, conceptually it provides some justification for the Hill Equation in this role. A more rigorous derivation is likely needed to truly rationalize the Hill Equation’s use for solid heat capacity.

References

  1. Chen, C.-C.; Mathias, P.M. Applied Thermodynamics for Process Modeling. Am. Inst. Chem. Eng. AIChE J. 2002, 48, 194. [Google Scholar] [CrossRef]
  2. Corriou, J.-P.; Assaf, J.-C. Special Issue on “Chemical Process Design, Simulation and Optimization”; MDPI: Basel, Switzerland, 2020; Volume 8, p. 1596. [Google Scholar]
  3. Asprion, N.; Bortz, M. Process Modeling, Simulation and Optimization: From Single Solutions to a Multitude of Solutions to Support Decision Making. Chem. Ing. Tech. 2018, 90, 1727–1738. [Google Scholar] [CrossRef]
  4. Rhodes, C.L. The Process Simulation Revolution: Thermophysical Property Needs and Concerns. J. Chem. Eng. Data 1996, 41, 947–950. [Google Scholar] [CrossRef]
  5. Hendriks, E.; Kontogeorgis, G.M.; Dohrn, R.; de Hemptinne, J.-C.; Economou, I.G.; Zilnik, L.F.; Vesovic, V. Industrial Requirements for Thermodynamics and Transport Properties. Ind. Eng. Chem. Res. 2010, 49, 11131–11141. [Google Scholar] [CrossRef]
  6. Wang, Y.; Shang, S.-L.; Fang, H.; Liu, Z.-K.; Chen, L.-Q. First-Principles Calculations of Lattice Dynamics and Thermal Properties of Polar Solids. NPJ Comput. Mater. 2016, 2, 1–10. [Google Scholar] [CrossRef]
  7. Nath, P.; Plata, J.J.; Usanmaz, D.; Al Orabi, R.A.R.; Fornari, M.; Nardelli, M.B.; Toher, C.; Curtarolo, S. High-Throughput Prediction of Finite-Temperature Properties Using the Quasi-Harmonic Approximation. Comput. Mater. Sci. 2016, 125, 82–91. [Google Scholar] [CrossRef] [Green Version]
  8. Huang, L.-F.; Lu, X.-Z.; Tennessen, E.; Rondinelli, J.M. An Efficient Ab-Initio Quasiharmonic Approach for the Thermodynamics of Solids. Comput. Mater. Sci. 2016, 120, 84–93. [Google Scholar] [CrossRef] [Green Version]
  9. Mathias, P.M. The Gibbs–Helmholtz Equation in Chemical Process Technology. Ind. Eng. Chem. Res. 2016, 55, 1076–1087. [Google Scholar] [CrossRef]
  10. Kang, J.W.; Diky, V.; Chirico, R.D.; Magee, J.W.; Muzny, C.D.; Abdulagatov, I.; Kazakov, A.F.; Frenkel, M. Quality Assessment Algorithm for Vapor−Liquid Equilibrium Data. J. Chem. Eng. Data 2010, 55, 3631–3640. [Google Scholar] [CrossRef]
  11. Kontogeorgis, G.M.; Coutsikos, P. Thirty Years with EoS/GE Models What Have We Learned? Ind. Eng. Chem. Res. 2012, 51, 4119–4142. [Google Scholar] [CrossRef] [Green Version]
  12. Economou, I.G.; de Hemptinne, J.-C.; Dohrn, R.; Hendriks, E.; Keskinen, K.; Baudouin, O. Industrial Use of Thermodynamics Workshop: Round Table Discussion on 8 July 2014. Chem. Eng. Res. Des. 2014, 92, 2795–2796. [Google Scholar] [CrossRef]
  13. Chase, M.W. National Information Standards Organization (US). NIST-JANAF Thermochemical Tables; American Chemical Society: Washington, DC, USA, 1998; Volume 9. [Google Scholar]
  14. Thomson, G.H. The DIPPR® Databases. Int. J. Thermophys. 1996, 17, 223–232. [Google Scholar] [CrossRef]
  15. Debye, P. Zur Theorie Der Spezifischen Wärmen. Ann. Phys. 1912, 344, 789–839. [Google Scholar] [CrossRef] [Green Version]
  16. Einstein, A. Die Plancksche Theorie Der Strahlung Und Die Theorie Der Spezifischen Wärme. Ann. Phys. 1907, 327, 180–190. [Google Scholar] [CrossRef] [Green Version]
  17. Abu-Eishah, S.I.; 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, 257–265. [Google Scholar]
  18. Grimvall, G. Thermophysical Properties of Materials; Elsevier: Amsterdam, The Netherlands, 1999; ISBN 0-08-054286-7. [Google Scholar]
  19. Dulong, P.L.; Petit, A.-T. Recherches sur Quelques Points Importans de la Theorie de la Chaleur; Annales de Chimie et de Physique: Paris, France, 1819. [Google Scholar]
  20. Smith, J.M.; Van Ness, H.C.; Abbott, M.M. Introduction to Chemical Engineering Thermodynamics; McGraw Hill Chemical Engineering Series 632; McGraw Hill: New York, NY, USA, 1996. [Google Scholar]
  21. Wyman, J.; Gill, S.J.; Colosimo, A. On Thermal Transitions in Biological Macromolecules. Biophys. Chem. 1979, 10, 363–369. [Google Scholar] [CrossRef]
  22. Gill, S.J.; Richey, B.; Bishop, G.; Wyman, J. Generalized Binding Phenomena in an Allosteric Macromolecule. Biophys. Chem. 1985, 21, 1–14. [Google Scholar] [CrossRef]
  23. Cattoni, D.I.; Chara, O.; Kaufman, S.B.; González Flecha, F.L. Cooperativity in Binding Processes: New Insights from Phenomenological Modeling. PLoS ONE 2015, 10, e0146043. [Google Scholar] [CrossRef] [Green Version]
  24. Gopal, E. Specific Heats at Low Temperatures; Springer: Berlin, Germany, 2012; ISBN 1-4684-9081-8. [Google Scholar]
  25. Fei, Y.; Saxena, S.K. An Equation for the Heat Capacity of Solids. Geochim. Cosmochim. Acta 1987, 51, 251–254. [Google Scholar] [CrossRef]
  26. Collocott, S.J.; White, G.K. Thermal Expansion and Heat Capacity of Some Stainless Steels and FeNi Alloys. Cryogenics 1986, 26, 402–405. [Google Scholar] [CrossRef]
  27. Touloukian, Y.S.; Buyco, E.H. Thermophysical Properties of Matter—The TPRC Data Series, Volume 4. Specific Heat-Metallic Elements and Alloys; Thermophysical and Electronic Properties Information Analysis Center, 1971. Available online: https://onlinelibrary.wiley.com/doi/abs/10.1002/bbpc.19710751034 (accessed on 11 July 2022).
  28. Clark, J.; Tye, R. Thermophysical Properties Reference Data for Some Key Engineering Alloys. High Temp. High Press. 2003, 35, 1–14. [Google Scholar] [CrossRef]
  29. Corsan, J.M.; Mitchem, N.I. The Specific Heat of Fifteen Stainless Steels in the Temperature Range 4K-30K. Cryogenics 1979, 19, 11–16. [Google Scholar] [CrossRef]
  30. Corsan, J.M.; Mitchem, N.I. Specific Heat of Stainless Steels Between 4 and 300 K. In Proceedings of the 6th International Cryogenic Engineering Conference, Grenoble, France, 11–14 May 1976; pp. 342–344. [Google Scholar]
  31. Lee, S.H.; Kim, S.W.; Kang, K.H. Effect of Heat Treatment on the Specific Heat Capacity of Nickel-Based Alloys. Int. J. Thermophys. 2006, 27, 282–292. [Google Scholar] [CrossRef]
  32. Douglasr, G.T.F.T.B.; McCoskeYr, R.E.; Ginnings, D.C. Thermal Properties of Aluminum Oxide From 0 to 1200 K. J. Res. Natl. Bur. Stand. 1956, 57, 67–82. [Google Scholar]
  33. Kelley, K.K.; Moore, G.E. The Specific Heats at Low Temperatures of Anhydrous Chlorides of Calcium, Iron, Magnesium and Manganese1. J. Am. Chem. Soc. 1943, 65, 1264–1267. [Google Scholar] [CrossRef]
  34. Todd, S.S. Heat Capacities at Low Temperatures and Entropies of Magnesium and Calcium Fluorides. J. Am. Chem. Soc. 1949, 71, 4115–4116. [Google Scholar] [CrossRef]
  35. Roberts, L.M. The Atomic Heats of Lithium, Sodium and Potassium between 1.5 and 20 K. Proc. Phys. Soc. Sect. B 1957, 70, 744. [Google Scholar] [CrossRef]
  36. Raman, C.V. The Specific Heats of the Alkali Halides and Their Spectroscopic Behaviour. In Proceedings of the Indian Academy of Sciences-Section A; Springer: Berlin/Heidelberg, Germany, 1962; Volume 56, pp. 60–69. [Google Scholar]
  37. Berg, W.T.; Morrison, J.A. The Thermal Properties of Alkali Halide Crystals. I. The Heat Capacity of Potassium Chloride, Potassium Bromide, Potassium Iodide and Sodium Iodide between 2.8 and 270 K. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1957, 242, 467–477. [Google Scholar]
  38. Southard, J.C.; Nelson, R.A. Low Temperature Specific Heats. IV. The Heat Capacities of Potassium Chloride, Potassium Nitrate and Sodium Nitrate1. J. Am. Chem. Soc. 1933, 55, 4865–4869. [Google Scholar] [CrossRef]
  39. Kirkham, A.J.; Yates, B. The Adiabatic Measurement of the Specific Heats of Potassium Chloride and Rubidium Chloride at Low Temperatures. Cryogenics 1968, 8, 381–385. [Google Scholar] [CrossRef]
  40. Barron, T.H.K.; Berg, W.T.; Morrison, J.A. On the Heat Capacity of Crystalline Magnesium Oxide. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1959, 250, 70–83. [Google Scholar]
  41. Robie, R.A.; Hemingway, B.S. Thermodynamic Properties of Minerals and Related Substances at 298.15 K and 1 Bar (105 Pascals) Pressure and at Higher Temperatures; US Government Printing Office: Washington, DC, USA, 1995.
  42. Archer, D.G. Thermodynamic Properties of the NaCl+ H2O System l. Thermodynamic Properties of NaCl (Cr). J. Phys. Chem. Ref. Data 1992, 21, 1–21. [Google Scholar] [CrossRef] [Green Version]
  43. Kirkham, A.J.; Yates, B. The Low-Temperature Specific Heats and Related Thermodynamic Properties of Sodium Fluoride and Caesium Bromide. J. Phys. C Solid State Phys. 1968, 1, 1162. [Google Scholar] [CrossRef]
  44. Munro, R.G. Material Properties of a Sintered α-SiC. J. Phys. Chem. Ref. Data 1997, 26, 1195–1203. [Google Scholar] [CrossRef]
  45. Smith, S.J.; Stevens, R.; Liu, S.; Li, G.; Navrotsky, A.; Boerio-Goates, J.; Woodfield, B.F. Heat Capacities and Thermodynamic Functions of TiO2 Anatase and Rutile: Analysis of Phase Stability. Am. Mineral. 2009, 94, 236–243. [Google Scholar] [CrossRef]
  46. Broughton, J. Process Utility Systems: Introduction to Design, Operation, and Maintenance; IChemE: New York, NY, USA, 1994; ISBN 0-85295-483-2. [Google Scholar]
  47. Charpentier, J.-C. Among the Trends for a Modern Chemical Engineering, the Third Paradigm: The Time and Length Multiscale Approach as an Efficient Tool for Process Intensification and Product Design and Engineering. Chem. Eng. Res. Des. 2010, 88, 248–254. [Google Scholar] [CrossRef]
  48. Allgöwer, F.; Zheng, A. Nonlinear Model Predictive Control; Birkhäuser: Basel, Switzerland, 2012; Volume 26, ISBN 3-0348-8407-9. [Google Scholar]
  49. Grossmann, I. Enterprise-wide Optimization: A New Frontier in Process Systems Engineering. AIChE J. 2005, 51, 1846–1857. [Google Scholar] [CrossRef]
  50. Barlow, R.; Blake, J.F. Hill Coefficients and the Logistic Equation. Trends Pharmacol. Sci. 1989, 10, 440–441. [Google Scholar] [CrossRef]
  51. Lide, D.R. CRC Handbook of Chemistry and Physics; CRC Press: Boca Raton, FL, USA, 2004; Volume 85, ISBN 0-8493-0485-7. [Google Scholar]
  52. Dill, K.A.; Bromberg, S.; Stigter, D. Molecular Driving Forces: Statistical Thermodynamics in Biology, Chemistry, Physics, and Nanoscience; Garland Science: New York, NY, USA, 2010; ISBN 0-203-80907-6. [Google Scholar]
Figure 1. The general characteristics of temperature-dependent specific heat in solids. At absolute zero the constant-volume heat capacity curve, Cv(T), originates and grows monotonically through the Debye Temperature (TD). A solid is often characterized by its Debye Temperature where its highest frequency vibrations are presumably excited [18]. The constant-volume heat capacity curve terminates when a solid melts at its temperature of fusion (Tfusion). The solid’s maximal constant-volume specific heat is observed (Cvmax) at this location. Constant-pressure heat capacity, not constant-volume heat capacity, is usually engineering-relevant [20]. As shown, the constant-pressure curve Cp(T) consistently exceeds the constant-volume curve Cv(T).
Figure 1. The general characteristics of temperature-dependent specific heat in solids. At absolute zero the constant-volume heat capacity curve, Cv(T), originates and grows monotonically through the Debye Temperature (TD). A solid is often characterized by its Debye Temperature where its highest frequency vibrations are presumably excited [18]. The constant-volume heat capacity curve terminates when a solid melts at its temperature of fusion (Tfusion). The solid’s maximal constant-volume specific heat is observed (Cvmax) at this location. Constant-pressure heat capacity, not constant-volume heat capacity, is usually engineering-relevant [20]. As shown, the constant-pressure curve Cp(T) consistently exceeds the constant-volume curve Cv(T).
Chemengineering 06 00056 g001
Figure 2. Accuracy of the specific heat models, where a 1:1 correspondence reflects accurate model predictions. For comparative purposes, predicted specific heat (Cp) and experimental specific heat (Cp,exp) were normalized by specific heat at each solid’s Debye Temperature, Cp(TD), which yields unitless graphs. (A) Debye Model training data (x-axis) versus Debye Model predictions (y-axis) across all 53 solids. (B) Hill Equation training data (x-axis) versus Hill Equation predictions (y-axis) across all 53 solids. (C) Einstein Solid training data (x-axis) versus Einstein Solid predictions (y-axis) across all 53 solids. (D) Multilinear Model training data (x-axis) versus Multilinear Model predictions (y-axis) across all 53 solids.
Figure 2. Accuracy of the specific heat models, where a 1:1 correspondence reflects accurate model predictions. For comparative purposes, predicted specific heat (Cp) and experimental specific heat (Cp,exp) were normalized by specific heat at each solid’s Debye Temperature, Cp(TD), which yields unitless graphs. (A) Debye Model training data (x-axis) versus Debye Model predictions (y-axis) across all 53 solids. (B) Hill Equation training data (x-axis) versus Hill Equation predictions (y-axis) across all 53 solids. (C) Einstein Solid training data (x-axis) versus Einstein Solid predictions (y-axis) across all 53 solids. (D) Multilinear Model training data (x-axis) versus Multilinear Model predictions (y-axis) across all 53 solids.
Chemengineering 06 00056 g002
Figure 3. Model predictions for the specific heat of example metal alloys.
Figure 3. Model predictions for the specific heat of example metal alloys.
Chemengineering 06 00056 g003
Figure 4. Model predictions for the specific heat of example technical and refractory ceramics.
Figure 4. Model predictions for the specific heat of example technical and refractory ceramics.
Chemengineering 06 00056 g004
Figure 5. Evaluation metrics of specific the heat models. Each vertical bar is from data analysis on a single covalent solid from the 53 compounds that were fit to specific heat expressions. The results of Debye Model fits to all 53 solids are shown leftmost in each graph, whereas the Multilinear Model (Multi) fits to all 53 solids are shown rightmost in each graph, with Einstein and Hill Equation fits nested left and nested right within each graph. (A) Average absolute relative error (AARE) of each model fit to each solid for data from temperatures below 100 Kelvin, unitless. Low AARE reflects a better model fit. (B) Standard error in J/(mole⋅Kelvin) of each model fit to each solid across all available temperature data. Low standard error reflects a better model fit. (C) Average absolute relative error (AARE) of each model fit to each solid for data from temperatures above 100 Kelvin, unitless. Low AARE reflects a better model fit. (D) Correlation coefficient of each model fit to each solid. Unity correlation reflects a better model fit.
Figure 5. Evaluation metrics of specific the heat models. Each vertical bar is from data analysis on a single covalent solid from the 53 compounds that were fit to specific heat expressions. The results of Debye Model fits to all 53 solids are shown leftmost in each graph, whereas the Multilinear Model (Multi) fits to all 53 solids are shown rightmost in each graph, with Einstein and Hill Equation fits nested left and nested right within each graph. (A) Average absolute relative error (AARE) of each model fit to each solid for data from temperatures below 100 Kelvin, unitless. Low AARE reflects a better model fit. (B) Standard error in J/(mole⋅Kelvin) of each model fit to each solid across all available temperature data. Low standard error reflects a better model fit. (C) Average absolute relative error (AARE) of each model fit to each solid for data from temperatures above 100 Kelvin, unitless. Low AARE reflects a better model fit. (D) Correlation coefficient of each model fit to each solid. Unity correlation reflects a better model fit.
Chemengineering 06 00056 g005
Table 1. Relative heat capacity computation time for compiled C code on an Intel i5-4690K CPU. Debye integration was by the Trapezoidal Rule with 1 Kelvin granularity.
Table 1. Relative heat capacity computation time for compiled C code on an Intel i5-4690K CPU. Debye integration was by the Trapezoidal Rule with 1 Kelvin granularity.
Debye5.06
Einstein0.96
Hill1
Multi0.99
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rowe, S.C.; Chubukov, B.A.; Telotte, J.C. A Hill Equation for Solid Specific Heat Capacity Calculation. ChemEngineering 2022, 6, 56. https://doi.org/10.3390/chemengineering6040056

AMA Style

Rowe SC, Chubukov BA, Telotte JC. A Hill Equation for Solid Specific Heat Capacity Calculation. ChemEngineering. 2022; 6(4):56. https://doi.org/10.3390/chemengineering6040056

Chicago/Turabian Style

Rowe, Scott C., Boris A. Chubukov, and John C. Telotte. 2022. "A Hill Equation for Solid Specific Heat Capacity Calculation" ChemEngineering 6, no. 4: 56. https://doi.org/10.3390/chemengineering6040056

Article Metrics

Back to TopTop