Abstract
Many chemical and physical equilibrium conditions can be determined from minimizing the Gibbs free energies of the system. Efficient analytical representations of the entropy and Gibbs free energy of carbonyl sulfide remain elusive in the communality of science and engineering. Here, we report two analytical representations of the entropy and Gibbs free energy for carbonyl sulfide, and the prediction procedures only involve six molecular constants of the carbonyl sulfide molecule. In the temperature range from 300 to 6000 K, the average relative deviations of the predicted molar entropy and reduced Gibbs free energy values of carbonyl sulfide from the National Institute of Standards and Technology database are arrived at 0.150 and 0.189%, respectively.
1. Introduction
Carbonyl sulfide (COS or OCS, hereafter COS) is a colorless and toxic organic sulfur compound, which widely exists in natural gas, petroleum gas, water gas, and semiwater gas.1−8 COS also originates from the volcanic activity9 and biomass burning.10,11 COS passes through the same diffusion pathway into the leaf chloroplasts as carbon dioxide (CO2) and is consumed by the enzyme carbonic anhydrase.12−18 Plants do not produce COC, and plant leaves directly consume COS. Observations of COS have appeared as an efficient tool for understanding terrestrial carbon uptake and plant physiology. As natural gas has economic and ecological advantages, this leads it to become more attractive for many countries. In order to transport natural gas and utilize commercially, it is usually necessary to remove acid gases, including CO2 and hydrogen sulfide (H2S), and impurities such as COS and mercaptans. Emission of toxic COS without treatment in coal gas, biogas, tail gas, and other industrial gas will yield air pollution and can also lead to equipment corrosion and poisoning the catalysts. Therefore, the industry has developed strict specifications for the transportation and utilization of sulfur compounds, limiting the total content of sulfur to 35–50 ppm. To prevent equipment damage and catalyst empoisoning in various processes such as methane reforming and oil refining, the efficient removal of COS can be achieved by adsorption using aqueous solutions of various alkanolamines.19 The removal of COS and reducing its emission have aroused much attention.19−30 The knowledge of Gibbs free energy plays an important role for understanding adsorbate–adsorbent interactions. A higher negative value of the change in the Gibbs free energy implies a more energetically favorable adsorption process. The accurate prediction of the Gibbs free energy change in an adsorption process is useful for the design and fabrication of suitable adsorbents.
In order to carry out optimization of COS recovery and emissions, one requires accurate and reliable thermochemical property data on COS.31−35 An accurate and efficient analytical representation of the Gibbs free energy for COS can afford direct and rapid pathways to determine the change in Gibbs free energy and minimization of Gibbs free energy for a system containing COS. However, obtaining efficient analytical representations of the entropy and Gibbs free energy for COS remains a perennial challenging task in science and engineering community. Based on the effective descriptions of molecular internal vibrations in terms of several improved oscillators, including the improved Tietz oscillator, improved Manning–Rosen oscillator, and Rosen–Morse oscillator, some authors have successfully predicted thermochemical properties of some gaseous diatomic substances.36−52 In the previous work,53 our group investigated two analytical representations of the entropy and Gibbs free energy of carbon dioxide (CO2) which is a triatomic linear symmetric molecule. Here, we further extend our treatment approach proposed in ref (53) to the triatomic linear nonsymmetric molecule COS. Span and Wagner54 suggested that all the thermodynamic properties of a pure substance can be obtained by combining the Helmholtz free energy and its derivatives. Although Lemmon and Span55 proposed an empirical analytical representation for the ideal-gas part of the Helmholtz free energy of COS, it contains 13 adjustable coefficients which were determined by fitting many experimental data points. Here, we report two efficient analytical representations to our knowledge of the molar entropy and Gibbs free energy of COS from the first principles. To check the effectiveness of the proposed prediction models, a comparison of the predicted values with the available data from the National Institute of Standards and Technology (NIST) database56 is also performed.
2. Results and Discussion
In order to test the proposed models, calculations of the molar entropy and reduced Gibbs free energy are performed and compared to available data from the NIST data,56 where six molecular constants of the COS molecule are collected from the work made by Lochte-Holtgreven et al.,57 and Robinson58 and Tubergen et al.:59De = 76 kcal/mol–1, reCO = 1.16 Å, reCS = 1.56 Å, ωes = 859 cm–1, ωea = 2062 cm–1, and ωeb = 521 cm–1. The reduced Gibbs free energy is represented as Gr = −(G – H298.15)/T; here, H298.15 denotes the molar enthalpy value of COS at 298.15 K. The molar Gibbs free energy determined from expression 14 for COS is −34 825.05 J·mol–1 at 298 K. The NIST value of reduced molar Gibbs free energy is 231.6 J·mol–1·K–1 at 298 K.56 By the aid of these two Gibbs free energy values, the molar enthalpy value of COS is determined as 34 191.75 J·mol–1 at 298.15 K. The calculated entropy and reduced Gibbs free energy values of COS from expressions 13 and 14, in the temperature range of 300–6000 K at the pressure of 1 bar, are presented in Figure 1, in which the NIST data56 are also depicted. The predicted values agree with less than 0.150 and 0.189% average relative deviations, respectively, with the NIST data. These deviation values tell us that the proposed models behave very well in reproducing the entropy and Gibbs free energy values of COS. The NIST data are taken from the NIST-JANAF Thermochemical Tables compiled by Chase.60 The values in the NIST-JANAF Tables are determined by using plenty of theoretical or experimental spectroscopic constants. In the future, we will expand the applications of the proposed prediction models under the conditions of different pressures.
Figure 1.

Temperature variation of the thermodynamic properties of COS for (A) molar entropy and (B) molar-reduced Gibbs free energy.
The conventional prediction treatment of entropy and Gibbs free energy values of substances involves a large number of experimental spectroscopy data. The present procedure involves only six molecular constant of the COS molecule and calculations begin by inputting directly the molecular constant values to the analytical representations. This merit can help engineers to acquire directly and quickly thermochemical property data of COS. Because acid gas COS is toxic and corrosive, the experimental thermochemical data are scarce. The reason is that performing experimental measurements on COS is fraught with difficulties as breath of it can injure people at concentrations researchers sometimes encounter. The present prediction models provide available computational means as the alternatives to obtain the required thermochemical property data.
3. Conclusions
We present two analytical representations for the entropy and Gibbs free energy calculations of COS. The proposed procedure merely rely on experimental values of six molecular constants of the COS molecule. The average relative deviations of molar entropy and reduced molar Gibbs free energy values for COS from the NIST database are 0.150 and 0.189%, respectively. The agreement of the predictions with the NIST data is very good. The new prediction strategy for the entropy and Gibbs free energy can be adapted to other substances such as methane.
4. Computational Methods
The COS molecule is a linear nonsymmetric triatomic molecule. As shown in Figure 2, the COS molecule contains internal symmetric stretching vibration, antisymmetric stretching vibration, and a degenerate pair of bending vibrations. Our strategy to treat these vibrations is to choose the improved Tietz oscillator36 and harmonic oscillator to represent symmetric vibration and other vibrations, respectively. An expression for the vibrational partition function subject to the symmetric stretching mode of COS can be derived based on the utilization of the improved Tietz oscillator36 and reaches the following form from eq 14 of ref (37)
![]() |
1 |
where De denotes
the energy to dissociate COS into CO and S, k denotes
the Boltzmann constant, T denotes the absolute temperature,
,
,
,
,
, and
; here, h is the Planck
constant, μOS represents the reduced mass of oxygen and sulfur
atoms, reOS = reCO + reCS, reCO represents the C–O bond length, reCS represents the C–S bond length, and q is
a dimensionless adjustable parameter. The ± signs are taken as
the plus and minus for q < 0 and q > 0, respectively. The parameter α is given by
Here,
, c is the speed of light,
and ωes represents the symmetric stretching frequency.
The symbol erfi stands for the imaginary error function. The factor
e1/13 stands for the contributions of the excited states
of the COS molecule to the vibrational partition function. The symbol vmax is the highest vibrational quantum number,
and
![]() |
Here, [x] stands for the greatest integer, which is less than x for noninteger x.
Figure 2.

Schematic diagram of the vibration modes of the COS molecule.
We attribute the antisymmetric stretching vibration and bending vibrations of the COS molecule to harmonic vibrations. The corresponding vibrational partition function can be represented by37
| 2 |
where Θν = hcωe/k, ωe stands for the equilibrium frequency.
The basic thermodynamic formula employed in deriving analytical representations of the entropy and Gibbs free energy of COS are the following thermodynamic relationships
| 3 |
| 4 |
From eqs 1, 3, and 4, we can obtain the molar vibrational entropy and Gibbs free energy subject to the contributions of the symmetric stretching vibration of COS and represent them as follows, respectively
![]() |
5 |
![]() |
6 |
where R is the universal gas constant. From eqs 2–4, we arrive at the following two expressions for the molar vibrational entropy and Gibbs free energy accompanied subject to contributions of the antisymmetric stretching vibration and two bending vibrations of COS
| 7 |
| 8 |
where Θva = hcωea/k and Θv = hcωeb/k; here, ωea and ωeb stand for the antisymmetric stretching vibration and bending vibration frequencies of COS, respectively.
The molar entropies subject to contributions of translation and rotation of COS can be represented by,50 respectively
| 9 |
| 10 |
where
, mCOS refers
to the mass of the COS molecule, ICOS represents
the rotational inertia of the COS molecule, and P represents the gas pressure. Refering to the literature,51 the molar Gibbs free energies subject to contributions
of translation and rotation of COS can be written as follows, respectively
| 11 |
| 12 |
With the help of eqs 5–12, we obtain analytical representations of the total molar entropy and Gibbs free energy of COS as follows, respectively
| 13 |
| 14 |
Because there are only six basic molecular constants involved in our calculation models, including De, reCO, reCS, ωes, ωea, and ωeb, predictions of the molar entropy and Gibbs free energy of COS would become direct and simple through the proposed analytical representations 13 and 14. In order to see what effect the temperature would have on the predicted thermochemical properties of COS, we calculate the molar entropy and Gibbs free energy values of COS at different temperatures.
Acknowledgments
This work was supported by the Key Program of National Natural Science Foundation of China under grant no. 51534006 and the Sichuan Province Foundation of China for Fundamental Research Projects under grant no. 2018JY0468.
The authors declare no competing financial interest.
References
- Hinderaker G.; Sandall O. C. Absorption of carbonyl sulfide in aqueous diethanolamine. Chem. Eng. Sci. 2000, 55, 5813–5818. 10.1016/s0009-2509(00)00420-6. [DOI] [Google Scholar]
- Andersen W. C.; Abdulagatov A. I.; Bruno T. J. The ASTM copper strip corrosion test: Application to propane with carbonyl sulfide and hydrogen sulfide. Energy Fuels 2003, 17, 120–126. 10.1021/ef020145m. [DOI] [Google Scholar]
- Chu X.; Li W.; Li B.; Chen H. Sulfur transfers from pyrolysis and gasification of direct liquefaction residue of Shenhua coal. Fuel 2008, 87, 211–215. 10.1016/j.fuel.2007.04.014. [DOI] [Google Scholar]
- Yi H.; Wang H.; Tang X.; Ning P.; Yu L.; He D.; Zhao S. Effect of calcination temperature on catalytic hydrolysis of COS over CoNiAl catalysts derived from hydrotalcite precursor. Ind. Eng. Chem. Res. 2011, 50, 13273–13279. 10.1021/ie200489t. [DOI] [Google Scholar]
- Chen L.; Bhattacharya S. Sulfur emission from Victorian brown coal under pyrolysis, oxy-fuel combustion and gasification conditions. Environ. Sci. Technol. 2013, 47, 1729–1734. 10.1021/es303364g. [DOI] [PubMed] [Google Scholar]
- Du Q.; Zhang C.; Mu Y.; Cheng Y.; Zhang Y.; Liu C.; Song M.; Tian D.; Liu P.; Liu J.; Xue C.; Ye C. An important missing source of atmospheric carbonyl sulfide: Domestic coal combustion. Geophys. Res. Lett. 2016, 43, 8720–8727. 10.1002/2016gl070075. [DOI] [Google Scholar]
- Zhang Z.; Li Z.; Cai N. Formation of reductive and corrosive gases during air-staged combustion of blends of anthracite/sub-bituminous coals. Energy Fuels 2016, 30, 4353–4362. 10.1021/acs.energyfuels.6b00599. [DOI] [Google Scholar]
- Ma H.; Zhou L.; Ma S.; Wang Z.; Cui Z.; Zhang W.; Li J. Reaction mechanism for sulfur species during pulverized coal combustion. Energy Fuels 2018, 32, 3958–3966. 10.1021/acs.energyfuels.7b03868. [DOI] [Google Scholar]
- Rasmussen R. A.; Khalil M. A. K.; Dalluge R. W.; Penkett S. A.; Jones B. Carbonyl sulfide and carbon disulfide from the eruptions of Mount St. Helens. Science 1982, 215, 665–667. 10.1126/science.215.4533.665. [DOI] [PubMed] [Google Scholar]
- Crutzen P. J.; Heidt L. E.; Krasnec J. P.; Pollock W. H.; Seiler W. Biomass burning as a source of atmospheric gases CO, H2, N2O, NO, CH3CJ and COS. Nature 1979, 282, 253–256. 10.1038/282253a0. [DOI] [Google Scholar]
- Whelan M. E.; Min D.-H.; Rhew R. C. Salt marsh vegetation as a carbonyl sulfide (COS) source to the atmosphere. Atmos. Environ. 2013, 73, 131–137. 10.1016/j.atmosenv.2013.02.048. [DOI] [Google Scholar]
- Commane R.; Meredith L. K.; Baker I. T.; Berry J. A.; Munger J. W.; Montzka S. A.; Templer P. H.; Juice S. M.; Zahniser M. S.; Wofsy S. C. Seasonal fluxes of carbonyl sulfide in a midlatitude forest. Proc. Natl. Acad. Sci. U.S.A. 2015, 112, 14162–14167. 10.1073/pnas.1504131112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Montzka S. A.; Calvert P.; Hall B. D.; Elkins J. W.; Conway T. J.; Tans P. P.; Sweeney C. On the global distribution, seasonality, and budget of atmospheric carbonyl sulfide (COS) and some similarities to CO2. J. Geophys. Res. 2007, 112, D09302. 10.1029/2006jd007665. [DOI] [Google Scholar]
- Campbell J. E.; Carmichael G. R.; Chai T.; Mena-Carrasco M.; Tang Y.; Blake D. R.; Blake N. J.; Vay S. A.; Collatz G. J.; Baker I.; Berry J. A.; Montzka S. A.; Sweeney C.; Schnoor J. L.; Stanier C. O. Photosynthetic control of atmospheric carbonyl sulfide during the growing season. Science 2008, 322, 1085–1088. 10.1126/science.1164015. [DOI] [PubMed] [Google Scholar]
- Sun W.; Maseyk K.; Lett C.; Seibt U. Litter dominates surface fluxes of carbonyl sulfide in a Californian oak woodland. J. Geophys. Res.: Biogeosci. 2016, 121, 438–450. 10.1002/2015jg003149. [DOI] [Google Scholar]
- Kooijmans L. M. J.; Maseyk K.; Seibt U.; Sun W.; Vesala T.; Mammarella I.; Kolari P.; Aalto J.; Franchin A.; Vecchi R.; Valli G.; Chen H. Canopy uptake dominates nighttime carbonyl sulfide fluxes in a boreal forest. Atmos. Chem. Phys. 2017, 17, 11453–11465. 10.5194/acp-17-11453-2017. [DOI] [Google Scholar]
- Sun W.; Kooijmans L. M. J.; Maseyk K.; Chen H.; Mammarella I.; Vesala T.; Levula J.; Keskinen H.; Seibt U. Soil fluxes of carbonyl sulfide (COS), carbon monoxide, and carbon dioxide in a boreal forest in southern Finland. Atmos. Chem. Phys. 2018, 18, 1363–1378. 10.5194/acp-18-1363-2018. [DOI] [Google Scholar]
- Sun W.; Maseyk K.; Lett C.; Seibt U. Stomatal control of leaf fluxes of carbonyl sulfide and CO2 in a Typha freshwater marsh. Biogeosciences 2018, 15, 3277–3291. 10.5194/bg-15-3277-2018. [DOI] [Google Scholar]
- Lee S. C.; Snodgrass M. J.; Park M. K.; Sandall O. C. Kinetics of removal of carbonyl sulfide by aqueous monoethanolamine. Environ. Sci. Technol. 2001, 35, 2352–2357. 10.1021/es0017312. [DOI] [PubMed] [Google Scholar]
- Kim J.; Do J. Y.; Park N.-K.; Hong J.-P.; Kang M. Adsorption/desorption behavior of carbonyl sulfide gas on Scheelite type MWO4 adsorbent. Sep. Purif. Technol. 2018, 207, 58–67. 10.1016/j.seppur.2018.06.040. [DOI] [Google Scholar]
- Williams B. P.; Young N. C.; West J.; Rhodes C.; Hutchings G. J. Carbonyl sulphide hydrolysis using alumina catalysts. Catal. Today 1999, 49, 99–104. 10.1016/s0920-5861(98)00413-1. [DOI] [Google Scholar]
- Rhodes C.; Riddel S. A.; West J.; Williams B. P.; Hutchings G. J. The low-temperature hydrolysis of carbonyl sulfide and carbon disulfide: a review. Catal. Today 2000, 59, 443–464. 10.1016/s0920-5861(00)00309-6. [DOI] [Google Scholar]
- Wang L.; Wang S.; Yuan Q. Removal of carbon disulfide via coupled reactions on a bi-functional catalyst: Experimental and modeling results. Chemosphere 2007, 69, 1689–1694. 10.1016/j.chemosphere.2007.06.015. [DOI] [PubMed] [Google Scholar]
- Li W.; Shudong W.; Quan Y. Removal of carbonyl sulfide at low temperature: Experiment and modeling. Fuel Process. Technol. 2010, 91, 777–782. 10.1016/j.fuproc.2010.02.013. [DOI] [Google Scholar]
- Wang X.; Qiu J.; Ning P.; Ren X.; Li Z.; Yin Z.; Chen W.; Liu W. Adsorption/desorption of low concentration of carbonyl sulfide by impregnated activated carbon under micro-oxygen conditions. J. Hazard. Mater. 2012, 229-230, 128–136. 10.1016/j.jhazmat.2012.05.084. [DOI] [PubMed] [Google Scholar]
- Wang Y.; Chen S.; Chen H.; Yuan X.; Zhang Y. Removal of carbonyl sulfide from CO2 stream using AgNO3-modified NaZSM-5. J. Energy Chem. 2013, 22, 902–906. 10.1016/s2095-4956(14)60270-9. [DOI] [Google Scholar]
- Zhao S.; Yi H.; Tang X.; Song C. Low temperature hydrolysis of carbonyl sulfide using Zn-Al hydrotalcite-derived catalysts. Chem. Eng. J. 2013, 226, 161–165. 10.1016/j.cej.2013.04.039. [DOI] [Google Scholar]
- Ma Y.; Wang X.; Ning P.; Cheng C.; Xu K.; Wang F.; Bian Z.; Yan S. Conversion of COS by corona plasma and the effect of simultaneous removal of COS and dust. Chem. Eng. J. 2016, 290, 328–334. 10.1016/j.cej.2016.01.052. [DOI] [Google Scholar]
- Chiche D.; Schweitzer J.-M. Investigation of competitive COS and HCN hydrolysis reactions upon an industrial catalyst: Langmuir-Hinshelwood kinetics modeling. Appl. Catal., B 2017, 205, 189–200. 10.1016/j.apcatb.2016.12.002. [DOI] [Google Scholar]
- Mi J.; Chen X.; Zhang Q.; Zheng Y.; Xiao Y.; Liu F.; Au C.-t.; Jiang L. Mechanochemically synthesized MgAl layered double hydroxide nanosheets for efficient catalytic removal of carbonyl sulfide and H2S. Chem. Commun. 2019, 55, 9375–9378. 10.1039/c9cc03637g. [DOI] [PubMed] [Google Scholar]
- Zarei S.; Ganji H.; Sadi M.; Rashidzadeh M. Kinetic modeling and optimization of Claus reaction furnace. J. Nat. Gas Sci. Eng. 2016, 31, 747–757. 10.1016/j.jngse.2016.03.086. [DOI] [Google Scholar]
- Zarei S.; Ganji H.; Sadi M.; Rashidzadeh M. Thermo-kinetic modeling and optimization of the sulfur recovery unit thermal stage. Appl. Therm. Eng. 2016, 103, 1095–1104. 10.1016/j.applthermaleng.2016.05.012. [DOI] [Google Scholar]
- Zarei S. Life cycle assessment and optimization of Claus reaction furnace through kinetic modeling. Chem. Eng. Res. Des. 2019, 148, 75–85. 10.1016/j.cherd.2019.06.005. [DOI] [Google Scholar]
- Rahman R. K.; Ibrahim S.; Raj A. Oxidative destruction of monocyclic and polycyclic aromatic hydrocarbon (PAH) contaminants in sulfur recovery units. Chem. Eng. Sci. 2016, 155, 348–365. 10.1016/j.ces.2016.08.027. [DOI] [Google Scholar]
- Rahman R. K.; Ibrahim S.; Raj A. Multi-objective optimization of sulfur recovery units using a detailed reaction mechanism to reduce energy consumption and destruct feed contaminants. Comput. Chem. Eng. 2019, 128, 21–34. 10.1016/j.compchemeng.2019.05.039. [DOI] [Google Scholar]
- Jia C.-S.; Diao Y.-F.; Liu X.-J.; Wang P.-Q.; Liu J.-Y.; Zhang G.-D. Equivalence of the Wei potential model and Tietz potential model for diatomic molecules. J. Chem. Phys. 2012, 137, 014101. 10.1063/1.4731340. [DOI] [PubMed] [Google Scholar]
- Jia C.-S.; Wang C.-W.; Zhang L.-H.; Peng X.-L.; Zeng R.; You X.-T. Partition function of improved Tietz oscillators. Chem. Phys. Lett. 2017, 676, 150–153. 10.1016/j.cplett.2017.03.068. [DOI] [Google Scholar]
- Jia C.-S.; Zhang L.-H.; Wang C.-W. Thermodynamic properties for the lithium dimer. Chem. Phys. Lett. 2017, 667, 211–215. 10.1016/j.cplett.2016.11.059. [DOI] [Google Scholar]
- Song X.-Q.; Wang C.-W.; Jia C.-S. Thermodynamic properties for the sodium dimer. Chem. Phys. Lett. 2017, 673, 50–55. 10.1016/j.cplett.2017.02.010. [DOI] [Google Scholar]
- Jia C.-S.; Zhang L.-H.; Peng X.-L.; Luo J.-X.; Zhao Y.-L.; Liu J.-Y.; Guo J.-J.; Tang L.-D. Prediction of entropy and Gibbs free energy for nitrogen. Chem. Eng. Sci. 2019, 202, 70–74. 10.1016/j.ces.2019.03.033. [DOI] [Google Scholar]
- Ocak Z.; Yanar H.; Salti M.; Aydogdu O. Relativistic spinless energies and thermodynamic properties of sodium dimer molecule. Chem. Phys. 2018, 513, 252–257. 10.1016/j.chemphys.2018.08.015. [DOI] [Google Scholar]
- Khordad R.; Ghanbari A. Analytical calculations of thermodynamic functions of lithium dimer using modified Tietz and Badawi-Bessis-Bessis potentials. Comput. Theor. Chem. 2019, 1155, 1–8. 10.1016/j.comptc.2019.03.019. [DOI] [Google Scholar]
- Buchowiecki M. On the exclusion of the negative contribution to the molecular partition function. Mol. Phys. 2019, 117, 1640–1644. 10.1080/00268976.2018.1552802. [DOI] [Google Scholar]
- Wang J.-F.; Peng X.-L.; Zhang L.-H.; Wang C.-W.; Jia C.-S. Entropy of gaseous boron monobromide. Chem. Phys. Lett. 2017, 686, 131–133. 10.1016/j.cplett.2017.08.047. [DOI] [Google Scholar]
- Jia C.-S.; Wang C.-W.; Zhang L.-H.; Peng X.-L.; Tang H.-M.; Liu J.-Y.; Xiong Y.; Zeng R. Predictions of entropy for diatomic molecules and gaseous substances. Chem. Phys. Lett. 2018, 692, 57–60. 10.1016/j.cplett.2017.12.013. [DOI] [Google Scholar]
- Jia C.-S.; Wang C.-W.; Zhang L.-H.; Peng X.-L.; Tang H.-M.; Zeng R. Enthalpy of gaseous phosphorus dimer. Chem. Eng. Sci. 2018, 183, 26–29. 10.1016/j.ces.2018.03.009. [DOI] [Google Scholar]
- Jiang R.; Jia C.-S.; Wang Y.-Q.; Peng X.-L.; Zhang L.-H. Prediction of enthalpy for the gases CO, HCl, and BF. Chem. Phys. Lett. 2019, 715, 186–189. 10.1016/j.cplett.2018.11.044. [DOI] [Google Scholar]
- Jia C.-S.; You X.-T.; Liu J.-Y.; Zhang L.-H.; Peng X.-L.; Wang Y.-T.; Wei L.-S. Prediction of enthalpy for the gases Cl2, Br2, and gaseous BBr. Chem. Phys. Lett. 2019, 717, 16–20. 10.1016/j.cplett.2019.01.001. [DOI] [Google Scholar]
- Jiang R.; Jia C.-S.; Wang Y.-Q.; Peng X.-L.; Zhang L.-H. Prediction of Gibbs free energy for the gases Cl2, Br2, and HCl. Chem. Phys. Lett. 2019, 726, 83–86. 10.1016/j.cplett.2019.04.040. [DOI] [Google Scholar]
- Jia C.-S.; Zeng R.; Peng X.-L.; Zhang L.-H.; Zhao Y.-L. Entropy of gaseous phosphorus dimer. Chem. Eng. Sci. 2018, 190, 1–4. 10.1016/j.ces.2018.06.009. [DOI] [Google Scholar]
- Peng X.-L.; Jiang R.; Jia C.-S.; Zhang L.-H.; Zhao Y.-L. Gibbs free energy of gaseous phosphorus dimer. Chem. Eng. Sci. 2018, 190, 122–125. 10.1016/j.ces.2018.06.027. [DOI] [Google Scholar]
- Tang B.; Wang Y.-T.; Peng X.-L.; Zhang L.-H.; Jia C.-S. Efficient predictions of Gibbs free energy for the gases CO, BF, and gaseous BBr. J. Mol. Struct. 2020, 1199, 126958. 10.1016/j.molstruc.2019.126958. [DOI] [Google Scholar]
- Wang J.; Jia C. S.; Li C. J.; Peng X. L.; Zhang L. H.; Liu J. Y.. Thermodynamic properties for carbon dioxide. ACS Omega 2019, 10.1021/acsomega.9b02488. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Span R.; Wagner W. A New Equation of State for Carbon Dioxide Covering the Fluid Region from the Triple-Point Temperature to 1100 K at Pressures up to 800 MPa. J. Phys. Chem. Ref. Data 1996, 25, 1509–1596. 10.1063/1.555991. [DOI] [Google Scholar]
- Lemmon E. W.; Span R. Short fundamental equations of state for 20 industrial fluids. J. Chem. Eng. Data 2006, 51, 785–850. 10.1021/je050186n. [DOI] [Google Scholar]
- National Institute of Standards and Technology (NIST) . NIST Chemistry WebBook, NIST Standard Reference Database Number 69. http://webbook.nist.gov/chemistry/, 2017.
- Lochte-Holtgreven W.; Bawn C. E. H.; Eastwood E. Photochemical dissociation of carbonyl sulphide. Nature 1932, 129, 869–870. 10.1038/129869c0. [DOI] [Google Scholar]
- Robinson D. Z. The Experimental Determination of the Intensities of Infrared Absorption Bands. IV. Measurements of the Stretching Vibrations of OCS and CS2. J. Chem. Phys. 1951, 19, 881–886. 10.1063/1.1748401. [DOI] [Google Scholar]
- Tubergen M. J.; Lavrich R. J.; McCargar J. W. Infrared spectrum and group theoretical analysis of the vibrational modes of carbonyl sulfide. J. Chem. Educ. 2000, 77, 1637–1639. 10.1021/ed077p1637. [DOI] [Google Scholar]
- Chase M. W.NIST-JANAF Thermochemical Tables. J. Phys. Chem. Ref. Data 1998, Monograph No. 9. [Google Scholar]





