Skip to main content
Log in

Extracting meaningful standard enthalpies and entropies of activation for surface reactions from kinetic rates

  • Published:
Reaction Kinetics, Mechanisms and Catalysis Aims and scope Submit manuscript

Abstract

While analyses based on the Arrhenius kinetic model have been successfully applied since its introduction, the Arrhenius model neglects to describe pre-exponential factor in a scientifically meaningful fashion. Since the 1930s, transition-state theory (TST), has met with success in interpreting the pre-exponential factor’s value, allowing a standard entropy of activation to be estimated. However, analyses based on TST’s assumptions have been applied inconsistently in the literature, particularly in corrosion science, leading to difficulty in comparison of standard entropy of activations from different studies. In this work, the foundational principles of TST and standard states are discussed and a standard method to apply TST in analyzing rates of surface reactions is recommended. When full details are not available, reacting species’ concentrations should be normalized to the concentration of active surface sites. For corrosion reactions, conversion relationships are given to convert from units of corrosion rate to surface reaction rate, consistent with TST. This method is dubbed a surface reactant equi-density approximation. Application of this standard to reported data results in adjustments of the standard entropy of activation between − 65 and +50 J/mol K and brings reported entropies into a narrower range of values.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

Data availability

All data generated or analyzed during this study are included in this published article.

Notes

  1. Strictly speaking, the Arrhenius equation is an empirical equation, and only provides direct physico-chemical knowledge when the kinetics are reflective of a single chemical elementary step. When those two conditions are not met, physico-chemical knowledge cannot be directly inferred with confidence from parameters extracted by fitting to the Arrhenius form. The energetic span model can help assess whether the kinetics observed are reflective of an elementary step, and in general microkinetic modeling may be required to assess the degree of rate control of individual elementary steps. For systems where the molecules are well-mixed and where a single chemical elementary step has a high degree of rate control, an empirical fitting to the Arrhenius form provides physico-chemical information. See [6,7,8,9,10,11].

  2. For example for A + B → C, writing a thermodynamic equilibrium constant with transition states and rearranging for [TS] we find: \(\kappa \frac{{k_{B} T}}{h}\left[ {\text{TS}} \right] = \kappa \frac{{k_{B} T}}{h}\left( {K_{TS}^\circ } \right)^{ - 1} \left[ A \right]\left[ B \right]*\frac{{\left[ {\text{TS}} \right]^\circ }}{\left[ A \right]^\circ \left[ B \right]^\circ } = k*\left[ A \right]\left[ B \right]\).

  3. The authors note that TST can also be applied using a statistical mechanical formulation based on solely a Boltzmann distribution and absolute energy difference. That formulation does not require the use of standard states and will not be presented here (the absolute energy formulation can be applied for special cases when one additionally knows the number of particles rather than solely concentrations and is thus not practical for most experimental systems). See Refs. [2, 15, 19].

  4. This is a common approximation during applications of TST. When application of TST is over small temperature ranges, this assumption is likely to hold true, but will become less valid as the temperature range is increased.

  5. An elementary step is almost never greater than second order. For example, A + B → D and A2 → E are each second order. A + B + C → D and A2 + B → E are each third order and would generally not be a permissible for elementary steps. A pre-equilibrium that results in an apparent third order reaction due to substitution of concentrations is not unusual (as described in chemical kinetics and physical chemistry textbooks), but the distinction is important as it will affect the units of the rate constant. If a practitioner believes they have a third order elementary step, they should consult a chemical kinetics specialist.

  6. As noted under Eq. 3, the units of the intrinsic rate constant will be dictated by the concentration of the transition state, but can be scaled arbitrarily. This is among the reasons why standardization of units for rates is desirable. For example, with an elementary step of A + B → C, the intrinsic rate constant has units of [TS]/ut but the phenomenological (“observed”) rate constant would have units sets by the rate. Consider the situation where the transition state is a surface species and has concentration units mol m−2, B is a type of site and has concentration units of mol m−2, and A is a solute with concentration units of mol m−3. d[A]/dt and d[B]/dt will have different units, and thus rate equations with different units for phenomenological rate constant. From Eq. 3, the intrinsic rate constant will have units of mol m−2 s−1. If monitoring d[B]/dt in mol m−2 s−1, the phenomenological rate will have the same units as the transition state conversion rate. However, if monitoring d[A]/dt, the phenomenological rate will have different units from the transition state conversion rate. This unit difference manifests in the phenomenological rate constant having a scaling factor relative to the intrinsic rate constant due to different units in [A]° and [TS]°. In practice, this affects the pre-exponential and the entropy of activation. As it may not be practical for all users to account for this factor, simply stating the elementary step and type of transition state assumed and using common units across studies will make any ‘errors’ systematic, allowing entropies of activation to be directly comparable. Standardization will also facilitate the obtaining of truly intrinsic values by specialists later.

  7. For example, converting between columns 1 and 6 is done with, \(\frac{{C^\circ_{{{\text{A/cm}}^{2} }} }}{{C^\circ_{{{\text{mg/cm}}^{2} {\text{s}}}} }} = \frac{{F n_{e} }}{{1000 MM_{surface} }}\), where \(MM_{surface}\) is the molar mass of the surface (g/mol surface), F is the Faraday constant (9.6485 × 104 C/mol electrons), \(n_{e}\) is the number of electrons transferred in the reaction (or the mol e/mol rxn), and the factor of 1000 converts the result in mg from g. The units of this factor are C/mg, which is required to convert the rate constant from a standard state of 1 C/cm2 to 1 mg/cm2.

  8. This picture also assumes that there is an elementary step which has a high degree of rate control a single atom or an ensemble of atoms at the surface changing bonding or re-arranging. In this case, the elementary step is a unimolecular surface species rearrangement (occurring within that ensemble of atoms, which can be treated as an 'embedded molecule' within the surface) and the transition state is also a surface species (effectively, the rest of the surface is like a solvation environment around that ensemble of atoms).

  9. Note that the true activation energy is not theoretically different in the Arrhenius and TST formulations (it is defined as a local derivative). Rather, the empirical estimate of the activation is slightly different because of the temperature dependence of A in TST [33, 60].

References

  1. Laidler KJ, King MC (1983) Development of transition-state theory. J Phys Chem 87(15):2657–2664. https://doi.org/10.1021/j100238a002

    Article  CAS  Google Scholar 

  2. Laidler KJ, Meiser JH, Sanctuary BC (2002) Physical chemistry, 4th edn. W.H Freeman, Brooks Cole

    Google Scholar 

  3. Eyring H (1935) The activated complex and the absolute rate of chemical reactions. Chem Rev 17(1):65–77. https://doi.org/10.1021/cr60056a006

    Article  CAS  Google Scholar 

  4. Eyring H (1935) The activated complex in chemical reactions. J Chem Phys 3(2):107–115. https://doi.org/10.1063/1.1749604

    Article  CAS  Google Scholar 

  5. Schmitz G, Lente G (2020) Fundamental concepts in chemical kinetics. ChemTexts 6(1):1

    Article  CAS  Google Scholar 

  6. Kozuch S, Shaik S (2011) How to conceptualize catalytic cycles? The Energetic span model. Acc Chem Res 44(2):101–110. https://doi.org/10.1021/ar1000956

    Article  CAS  PubMed  Google Scholar 

  7. Kozuch S (2012) A refinement of everyday thinking: the energetic span model for kinetic assessment of catalytic cycles. Comput Mol Sci 2(5):795–815. https://doi.org/10.1002/wcms.1100

    Article  CAS  Google Scholar 

  8. Savara A, Rossetti I, Chan-Thaw CE, Prati L, Villa A (2016) Microkinetic modeling of benzyl alcohol oxidation on carbon-supported palladium nanoparticles. ChemCatChem 8(15):2482–2491. https://doi.org/10.1002/cctc.201600368

    Article  CAS  Google Scholar 

  9. Matera S, Schneider WF, Heyden A, Savara A (2019) Progress in accurate chemical kinetic modeling, simulations, and parameter estimation for heterogeneous catalysis. ACS Catal 9(8):6624–6647. https://doi.org/10.1021/acscatal.9b01234

    Article  CAS  Google Scholar 

  10. Stegelmann C, Andreasen A, Campbell CT (2009) Degree of rate control: how much the energies of intermediates and transition states control rates. J Am Chem Soc 131(23):8077–8082. https://doi.org/10.1021/ja9000097

    Article  CAS  PubMed  Google Scholar 

  11. Campbell CT (2017) The degree of rate control: a powerful tool for catalysis research. ACS Catal 7(4):2770–2779. https://doi.org/10.1021/acscatal.7b00115

    Article  CAS  Google Scholar 

  12. Matta CF, Massa L, Gubskaya AV, Knoll E (2011) Can one take the logarithm or the sine of a dimensioned quantity or a unit? Dimensional analysis involving transcendental functions. J Chem Educ 88(1):67–70. https://doi.org/10.1021/ed1000476

    Article  CAS  Google Scholar 

  13. Bérces T (1996) The reaction path in chemistry: current approaches and perspectives. React Kinet Catal Lett 58(2):417–419. https://doi.org/10.1007/BF02067053

    Article  Google Scholar 

  14. Lente G (2015) Deterministic kinetics in chemistry and systems biology: the dynamics of complex reaction networks. Springerbriefs in molecular science. Springer International Publishing, Cham. https://doi.org/10.1007/978-3-319-15482-4

  15. Truhlar DG (2015) Transition state theory for enzyme kinetics. Arch Biochem Biophys 582:10–17. https://doi.org/10.1016/j.abb.2015.05.004

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Truhlar DG, Garrett BC, Klippenstein SJ (1996) Current status of transition state theory. J Phys Chem 100:12771–12800

    Article  CAS  Google Scholar 

  17. Atkins PW (1997) Physical chemistry. Macmillan Higher Education, New York

    Google Scholar 

  18. Shpan’ko I, Sadovaya I (2018) Enthalpy-entropy compensation effect and other aspects of isoparametricity in reactions between trans-2,3-bis(3-bromo-5-nitrophenyl)oxirane and arenesulfonic acids. Reac Kinet Mech Cat 123(2):473–484. https://doi.org/10.1007/s11144-017-1340-6

    Article  CAS  Google Scholar 

  19. Savara A (2013) Standard states for adsorption on solid surfaces: 2D gases, surface liquids, and langmuir adsorbates. J Phys Chem C 117(30):15710–15715. https://doi.org/10.1021/jp404398z

    Article  CAS  Google Scholar 

  20. Savara A, Schmidt CM, Geiger FM, Weitz E (2009) Adsorption entropies and enthalpies and their implications for adsorbate dynamics. J Phys Chem C 113(7):2806–2815. https://doi.org/10.1021/jp806221j

    Article  CAS  Google Scholar 

  21. IUPAC (2019) Compendium of chemical terminology, 2nd Ed. (the “Gold Book”); the standard state. Blackwell Scientific Publications. https://doi.org/10.1351/goldbook.S05925. Accessed October 17 2019

  22. Robinson PJ (1978) Dimensions and standard states in the activated complex theory of reaction rates. Textbook Errors 55(8):509–510

    CAS  Google Scholar 

  23. Division IPaBC (2007) Quantities, units and symbols in physical chemistry, 3rd edn. RCS Publishing, Cambridge

    Google Scholar 

  24. Savara A (2016) Comment on “equilibrium constants and rate constants for adsorbates: two-dimensional (2D) ideal gas, 2D ideal lattice gas, and ideal hindered translator models”. J Phys Chem C 120(36):20478–20480. https://doi.org/10.1021/acs.jpcc.6b07553

    Article  CAS  Google Scholar 

  25. CODATA (1974) CODATA guidelines on reporting data for chemical kinetics. CODATA, Warrenton, VA

    Google Scholar 

  26. Sutton JE, Lorenzi JM, Krogel JT, Xiong Q, Pannala S, Matera S, Savara A (2018) Electrons to reactors multiscale modeling: catalytic CO oxidation over RuO2. ACS Catal 8(6):5002–5016. https://doi.org/10.1021/acscatal.8b00713

    Article  CAS  Google Scholar 

  27. Sutton JE, Danielson T, Beste A, Savara A (2017) Below-room-temperature C-H bond breaking on an inexpensive metal oxide: methanol to formaldehyde on CeO2(111). J Phys Chem Lett 8(23):5810–5814. https://doi.org/10.1021/acs.jpclett.7b02683

    Article  CAS  PubMed  Google Scholar 

  28. Bray JM, Schneider WF (2014) First-principles thermodynamic models in hetergeneous catalysis. In: Asthagiri A, Janik MJ (eds) Computational catalysis. Royal Society of Chemistry, Camnridge

    Google Scholar 

  29. Danielson T, Hin C, Savara A (2016) Generalized adsorption isotherms for molecular and dissociative adsorption of a polar molecular species on two polar surface geometries: perovskite (100) (Pm-3m) and fluorite (111) (Fm-3m). J Chem Phys 145(6):064705. https://doi.org/10.1063/1.4960508

    Article  CAS  Google Scholar 

  30. Amzel LM (1997) Loss of translational entropy in binding, folding, and catalysis. Proteins 28(2):144–149

    Article  CAS  Google Scholar 

  31. Yu YB, Privalov PL, Hodges RS (2001) Contribution of translational and rotational motions to molecular association in aqueous solution. Biophys J 81(3):1632–1642. https://doi.org/10.1016/S0006-3495(01)75817-1

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Siebert X, Amzel LM (2004) Loss of translational entropy in molecular associations. Proteins 54(1):104–115. https://doi.org/10.1002/prot.10472

    Article  CAS  PubMed  Google Scholar 

  33. Mills I, Cvitas T, Klaus H, Nikola K, Kuchitsu K (1993) Quantities, units and symbols in physical chemistry, 2nd edn. Blackwell Science, Oxford

    Google Scholar 

  34. Behpour M, Ghoreishi SM, Gandomi-Niasar A, Soltani N, Salavati-Niasari M (2009) The inhibition of mild steel corrosion in hydrochloric acid media by two Schiff base compounds. J Mater Sci 44(10):2444–2453. https://doi.org/10.1007/s10853-009-3309-y

    Article  CAS  Google Scholar 

  35. Shetty KS, Shetty AN (2015) Studies on corrosion behavior of 6061 Al—15 vol. pct. SiC(p) composite in HCl medium by electrochemical techniques. Surf Eng Appl Electrochem 51(4):374–381. https://doi.org/10.3103/s1068375515040134

    Article  Google Scholar 

  36. Loto RT, Loto CA (2018) Corrosion behaviour of S43035 ferritic stainless steel in hot sulphate/chloride solution. J Mater Res Technol 7(3):231–239. https://doi.org/10.1016/j.jmrt.2017.07.004

    Article  CAS  Google Scholar 

  37. Sebhaoui J, El Bakri Y, El Aoufir Y, Anouar EH, Guenbour A, Nasser AA, Mokhtar Essassi E (2019) Synthesis, NMR characterization, DFT and anti-corrosion on carbon steel in 1 M HCl of two novel 1,5-benzodiazepines. J Mol Struct 1182:123–130. https://doi.org/10.1016/j.molstruc.2019.01.037

    Article  CAS  Google Scholar 

  38. Aoufir YE, Lgaz H, Bourazmi H, Kerroum Y, Ramli Y, Guenbour A, Salghi R, El-Hajjaji F, Hammouti B, Oudda H (2016) Quinoxaline derivatives as corrosion inhibitors of carbon steel in hydrochloric acid media: electrochemical, DFT and monte carlo simulation studies. J Mater Environ Sci 7(12):4330–4347

    Google Scholar 

  39. Kahled KF, Elhabib OA, El-mghraby A, Ibrahim OB, Ibrahim MAM (2010) Inhibitive effect of thiosemicarbazone derivative on corrosion of mild steel in hydrochloric acid solution. J Mater Environ Sci 1(3):139–150

    Google Scholar 

  40. Nautiyal P, Subramanian KA, Dastidar MG (2014) Kinetic and thermodynamic studies on biodiesel production from Spirulina platensis algae biomass using single stage extraction–transesterification process. Fuel 135:228–234. https://doi.org/10.1016/j.fuel.2014.06.063

    Article  CAS  Google Scholar 

  41. Zhang Y, Zhao R, Sanchez-Sanchez M, Haller GL, Hu J, Bermejo-Deval R, Liu Y, Lercher JA (2019) Promotion of protolytic pentane conversion on H-MFI zeolite by proximity of extra-framework aluminum oxide and Brønsted acid sites. J Catal 370:424–433. https://doi.org/10.1016/j.jcat.2019.01.006

    Article  CAS  Google Scholar 

  42. Everett DH (1987) Application of thermodynamics to interfacial phenomena. Pure Appl Chem. https://doi.org/10.1351/pac198759010045

    Article  Google Scholar 

  43. Everett DH (1986) Reporting data on adsorption from solution at the solid/solution interface (Recommendations 1986). Pure Appl Chem. https://doi.org/10.1351/pac198658070967

    Article  Google Scholar 

  44. Everett DH (1972) Manual of symbols and terminology for physicochemical quantities and units, appendix II: definitions, terminology and symbols in colloid and surface chemistry. Pure Appl Chem. https://doi.org/10.1351/pac197231040577

    Article  Google Scholar 

  45. Everett DH (1965) Thermodynamics of adsorption from solution. Part 2.—imperfect systems. Trans Faraday Soc 61:2478–2495. https://doi.org/10.1039/tf9656102478

    Article  CAS  Google Scholar 

  46. Everett DH (1964) Thermodynamics of adsorption from solution. Part 1.—perfect systems. Trans Faraday Soc 60:1803–1813. https://doi.org/10.1039/tf9646001803

    Article  CAS  Google Scholar 

  47. Gumma S, Talu O (2010) Net adsorption: a thermodynamic framework for supercritical gas adsorption and storage in porous solids. Langmuir 26(22):17013–17023. https://doi.org/10.1021/la102186q

    Article  CAS  PubMed  Google Scholar 

  48. Corti DS, Kerr KJ, Torabi K (2011) On the interfacial thermodynamics of nanoscale droplets and bubbles. J Chem Phys 135(2):024701. https://doi.org/10.1063/1.3609274

    Article  CAS  PubMed  Google Scholar 

  49. Eriksson JC (1971) Thermodynamics of surface-phase systems: VI. On the rigorous thermodynamics of insoluble surface films. J Colloid Interface Sci 37(4):659–667. https://doi.org/10.1016/0021-9797(71)90344-4

    Article  CAS  Google Scholar 

  50. Eriksson JC (1969) Thermodynamics of surface phase systems: V. Contribution to the thermodynamics of the solid-gas interface. Surface Sci 14(1):221–246. https://doi.org/10.1016/0039-6028(69)90056-9

    Article  CAS  Google Scholar 

  51. Xie Z, Yan B, Kattel S, Lee JH, Yao S, Wu Q, Rui N, Gomez E, Liu Z, Xu W, Zhang L, Chen JG (2018) Dry reforming of methane over CeO2—supported Pt-Co catalysts with enhanced activity. Appl Catal B 236:280–293. https://doi.org/10.1016/j.apcatb.2018.05.035

    Article  CAS  Google Scholar 

  52. Schmidt CM, Savara A, Weitz E, Geiger FM (2007) Enthalpy and entropy of acetone interacting with degussa P25 TiO2 determined by chemical ionization mass spectrometry. J Phys Chem C 111(23):8260–8267. https://doi.org/10.1021/jp068324i

    Article  CAS  Google Scholar 

  53. Doyle PJ, Raiman SS, Rebak R, Terrani KA (2017) Characterization of the hydrothermal corrosion behavior of ceramics for accident tolerant fuel cladding. In: Jackson JH, Paraventi D, Wright M (eds) Proceedings of the 18th international conference on environmental degradation of materials in nuclear power systems—water reactors. The Minerals, Metals & Materials Society, pp 269–280. https://doi.org/10.1007/978-3-319-68454-3_23

  54. Terrani KA (2018) Accident tolerant fuel cladding development: promise, status, and challenges. J Nucl Mater 501:13–30. https://doi.org/10.1016/j.jnucmat.2017.12.043

    Article  CAS  Google Scholar 

  55. Terrani KA, Pint BA, Parish CM, Silva CM, Snead LL, Katoh Y (2014) Silicon carbide oxidation in steam up to 2 MPa. J Am Ceram Soc 97(8):2331–2352. https://doi.org/10.1111/jace.13094

    Article  CAS  Google Scholar 

  56. Terrani KA, Yang Y, Kim YJ, Rebak R, Meyer HM, Gerczak TJ (2015) Hydrothermal corrosion of SiC in LWR coolant environments in the absence of irradiation. J Nucl Mater 465:488–498. https://doi.org/10.1016/j.jnucmat.2015.06.019

    Article  CAS  Google Scholar 

  57. Moss SJ, Coady CJ (1963) Potential-energy surfaces and transition state theory. J Chem Educ 60(6):455–461

    Article  Google Scholar 

  58. Campbell CT, Sprowl LH, Árnadóttir L (2016) Equilibrium constants and rate constants for adsorbates: two-dimensional (2D) ideal gas, 2D ideal lattice gas, and ideal hindered translator models. J Phys Chem C 120(19):10283–10297. https://doi.org/10.1021/acs.jpcc.6b00975

    Article  CAS  Google Scholar 

  59. Campbell CT, Sprowl LH, Árnadóttir L (2016) Reply to “comment on ‘equilibrium constants and rate constants for adsorbates: two-dimensional (2D) ideal gas, 2D ideal lattice gas, and ideal hindered translator models’”. J Phys Chem C 120(36):20481–20482. https://doi.org/10.1021/acs.jpcc.6b07756

    Article  CAS  Google Scholar 

  60. IUPAC (2019) Compendium of chemical terminology, 2nd edn. (the “Gold Book”); Activation energy. Blackwell Scientific Publications, Oxford. http://goldbook.iupac.org/terms/view/A00102. Accessed 28 Oct 2019

  61. Benensky KM (2018) Evaluation of novel refractory carbide matrix fuels for nuclear thermal propulsion. Paper presented at the Nuclear and Emerging Technolgies for Space,

  62. Kandziolka MV, Kidder MK, Gill L, Wu Z, Savara A (2014) Aromatic–hydroxyl interaction of an alpha-aryl ether lignin model-compound on SBA-15, present at pyrolysis temperatures. Phys Chem Chem Phys 16(44):24188–24193. https://doi.org/10.1039/C4CP02633K

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Thanks to John Stahl for his assistance in this work. This work was supported by the U.S. Department of Energy, Office of Nuclear Energy, Advanced Fuels Campaign (P.J.D and S.S.R) and supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, Chemical Sciences, Geosciences, and Biosciences Division (A.S.). This manuscript has been authored by UT-Battelle, LLC under Contract No. DE-AC05-00OR22725 with the U.S. Department of Energy. The United States Government retains and the publisher, by accepting the article for publication, acknowledges that the United States Government retains a non-exclusive, paid-up, irrevocable, world-wide license to publish or reproduce the published form of this manuscript, or allow other to do so, for United States Government purposes.  The Department of Energy will provide public access to these results of federally sponsored research in accordance with the DOE Public Access Plan (http://energy.gov/downloads/doe-public-access-plan).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Peter J. Doyle.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix 1: Examples of the use of standard states in TST

Most textbooks initially teach TST in the context of gas phase reactions, and we follow that approach here as it makes some of the concepts easier to explain. Consider a gas phase reaction of A(g)\(\to\) B(g) + C(g), where the activity coefficients are 1, the standard state for each gas is considered an ideal gas at 1 bar at a given temperature. For typographical convenience, a double sided arrow (\(\leftrightarrow\)) will be used rather than the equilibrium symbol (⇋) in subscripts of the following terms and equations. The standard equilibrium constant, \(K_{{{\text{A}} \leftrightarrow {\text{B}} + {\text{C}}}}^\circ\), will be related to the kinetic equilibrium constant, \(K_{{{\text{A}} \leftrightarrow {\text{B}} + {\text{C}}}}\),and rate constants \(\frac{{{\text{k}}_{\text{f}} }}{{{\text{k}}_{\text{r}} }}\) as follows:

$$K_{{{\text{A}} \leftrightarrow {\text{B}} + {\text{C}}}}^\circ = K_{{{\text{A}} \leftrightarrow {\text{B}} + {\text{C}}}} *\frac{{\left( {{\text{c}}_{\text{f}}^\circ } \right)^{ - 1} }}{{\left( {{\text{c}}_{\text{r}}^\circ } \right)^{ - 1} }} = \frac{{{\text{k}}_{\text{f}} }}{{{\text{k}}_{\text{r}} }}\frac{{\left( {{\text{c}}_{\text{f}}^\circ } \right)^{ - 1} }}{{\left( {{\text{c}}_{\text{r}}^\circ } \right)^{ - 1} }}$$
(22)

In conventional TST, a quasi-equilibrium is established between the reactants in the transition state (the activated state). This is the origin of the \(\frac{{\left( {{\text{c}}_{\text{R}}^\circ } \right)^{ - 1} }}{{\left( {{\text{c}}_{\text{TS}}^\circ } \right)^{ - 1} }}\) term in Eq. 3, [22] which has the standard states of the reactants in the transition state. This term is part of the Arrhenius pre-exponential. A correct accounting of units (including the standard states), allows extracting of the standard entropy of activation and standard enthalpy of activation from the rate constant from Eq. 3. In the forward reaction of A(g)\(\to\) B(g) + C(g), if we assume activity coefficients of 1 and that the transition state consists of a single gas molecule, then:

$$k_{{{\text{A}} \to {\text{B}} + {\text{C}}}} = \kappa \frac{{k_{B} T}}{h}e^{{ - \frac{{\Delta^{\ddag } G^\circ }}{RT}}} \frac{{\gamma_{\text{TS}} }}{{\gamma_{\text{R}} }}\frac{{(c_{TS}^\circ )^{ - 1} }}{{\left( {c_{R}^\circ } \right)^{ - 1} }} = \kappa \frac{{k_{B} T}}{h}e^{{ - \frac{{\Delta^{\ddag } H^\circ }}{RT} + \frac{{\Delta^{\ddag } S^\circ }}{R}}} \frac{{\left( {{\text{P}}_{\text{TS}}^\circ } \right)^{ - 1} }}{{\left( {{\text{P}}_{\text{A}}^\circ } \right)^{ - 1} }}$$
(23)

For such cases where the transition state and reactant are both a single gas molecule, the standard state terms cancel and this is one reason the term is often omitted in textbooks. The IUPAC recommended value for \(P^\circ\) is 1 bar, and the standard states would cancel for Eq. 5: \(\frac{{\left( {{\text{P}}_{\text{TS}}^\circ } \right)^{ - 1} }}{{\left( {{\text{P}}_{\text{A}}^\circ } \right)^{ - 1} }} = \frac{P^\circ }{P^\circ }\). However, the \(c^\circ\) terms do not always cancel, and the authors recommend including them explicitly in both the numerator and denominator prior to cancellation, to avoid misleading readers.

An example where the terms would not cancel is for the bimolecular reaction of D(g) + E(g)\(\to\) F(g), \(\frac{{(c_{TS}^\circ )^{ - 1} }}{{(c_{R}^\circ )^{ - 1} }} = \frac{{\left( {{\text{P}}_{\text{TS}}^\circ } \right)^{ - 1} }}{{\left( {{\text{P}}_{\text{E}}^\circ } \right)^{ - 1} \left( {{\text{P}}_{\text{E}}^\circ } \right)^{ - 1} }} = \frac{P^\circ P^\circ }{P^\circ }\). Also consider the case of adsorption to a surface with site designated as S, and a single gas molecule A adsorbing: A + S \(\to\) A − S, where A − S is the bound state. If we assume that the transition state is like the adsorbed state (immobile and with geometry similar to A − S), then with activity coefficients of 1 we arrive at right-hand side of the following situation:

$$k_{{{\text{A}} + {\text{S}} \to {\text{A}} - {\text{S}}}} = \kappa \frac{{k_{B} T}}{h}e^{{ - \frac{{\Delta^{\ddag } G^\circ }}{RT}}} \frac{{\gamma_{\text{TS}} }}{{\gamma_{\text{R}} }}\frac{{(c_{TS}^\circ )^{ - 1} }}{{\left( {c_{R}^\circ } \right)^{ - 1} }} = \kappa \frac{{k_{B} T}}{h}e^{{ - \frac{{\Delta^{\ddag } H^\circ }}{RT} + \frac{{\Delta^{\ddag } S^\circ }}{R}}} \frac{{\left( {{{\theta }}_{\text{TS}}^\circ } \right)^{ - 1} }}{{\left( {{\text{P}}_{\text{A}}^\circ } \right)^{ - 1} \left( {{{\theta }}_{{{\text{e}} - {\text{sites}}}}^\circ } \right)^{ - 1} }}$$
(24)

Here relative coverage has been chosen (further explained below) for the concentration units in the standard state of the immobile transition state (\({{\theta }}_{\text{TS}}^\circ\)), and also for the empty sites \(\left( {{{\theta }}_{\text{e-sites}}^\circ } \right)\). Again, the units for the standard states do not cancel, and it becomes more important that the units are carefully accounted for when extracting thermodynamics quantities. In contrast, for an elementary chemical reaction step of molecular desorption from a surface, the terms generally will cancel. For solutes, IUPAC recommends using a value of 1 mol solute/L solvent or 1 mol solute/kg solvent, [33] establishing standard states from which standard entropies can be calculated on a common scale [15, 57].

Appendix 2: Standard states of adsorbates on reacting surfaces

For surface-adsorbed species, it is possible to describe the activity of the adsorbates and of the surface using either an absolute coverage (σads) or a relative coverage (\({{\theta }}_{\text{ads}}\)). For an absolute coverage, the units are relatively clear, such as 1.39 × 1016 molecules m−2 [19]. However, for a relative coverage, defined relative to Langmuir (or lattice) adsorption sites, such as (molecules adsorbed)/(lattice sites), it is common to report them as unitless, such as \({{\theta }}_{\text{ads}}\) = 0.34. This lack of specified units is misleading as the term is not truly unitless: a relative concentration is still a type of concentration, defined on a particular scale. In fact, carrying the units is necessary to directly compare results between studies because there are two types of monolayers (lattice-saturated versus geometric [20]), and also because the denominator is sometimes the saturation coverage, leading to three different possible relative scales even for immobile adsorbates.

The authors thus recommend distinguishing the three types using the following acronyms for the different relative coverage units: relative to lattice (rtL) when using (n molecules adsorbed)/(m total lattice sites of that type), relative to saturate coverage (rtS) when using (n molecules adsorbed)/(nmax saturation coverage), and relative to geometric area (rtG) when using (aggregate area of molecules adsorbed)/(geometric surface area). In this case, the three quantities of \({{\theta }}_{\text{ads}}\) = 0.34 rtL, \({{\theta }}_{\text{ads}}\) = 0.34 rtS, and \({{\theta }}_{\text{ads}}\) = 0.34 rtG each have different meanings, and it is possible to convert between them if one knows the number of sites per unit area. For immobile adsorbates, it has been recommended to set the standard state of the adsorbates to being equal to that of the empty sites \({{\theta }}_{\text{ads}}^\circ = {{\theta }}_{\text{e-sites}}^\circ\), as these terms then cancel [19].

For the 2-D gas adsorbates (including 2D gas transition states), \({{\sigma }}_{\text{ads}}^\circ = 1.39 \times 10^{ - 7} {\text{mol/m}}^{2}\) has been recommended [19], though other recommendations do exist [19, 24, 58, 59]. The above two choices (for \({{\theta }}_{\text{ads}}^\circ = {{\theta }}_{\text{e-sites}}^\circ\) and \({{\sigma }}_{\text{ads}}^\circ = 1.39 \times 10^{ - 7} {\text{mol/m}}^{2}\)) are based upon enabling standard entropies to be easily compared between different systems to gain insights about the dynamics of the molecule.

It should be noted that an immobile adsorbate and a 2-D gas have different types of phases and different activity units. Therefore, they cannot share the same kind of standard state, and to discriminate between these two cases thus typically requires doing the calculation twice (once assuming an immobile adsorbate model, once assuming a 2-D gas adsorbate model) [20], resulting in two different values for the standard entropy (one for each model) from a single experiment. The experimentally calculated values from the two different models can then be compared to the theoretical values for each model to see which model has a better match between experiment and theory [9]. For a pure solid or liquid phase reacting with the gas phase in stoichiometric reactions (such as dissolution/sublimation of a sugar crystal) the activity and standard state both take a value of 1. For solutes in solution, the IUPAC recommended standard state is 1 mol/L or 1 mol solute/kg solvent [33].

Appendix 3: Effect of using transition state theory on the estimate of the activation energy

Consider a system investigated at two temperatures, \(T_{i}\) and \(T_{f}\) with experimental reaction rates \(Rate_{i}\) and \(Rate_{f}\), respectively, and unit concentrations. Assuming \({{\Delta }}^{\ddag } H^\circ\) and \(\Delta^{\ddag } S^\circ\) to be constant across the temperature range and that \(P{{\Delta }}^{\ddag } V^\circ = 0\), the activation energy may be empirically expressed simply as Eqs. 25 and 26 for the Arrhenius and TST estimations, respectively.Footnote 9 Thus, a plot of ln(k) vs 1/T is not entirely linear for TST, while it is assumed to be linear in an Arrhenius plot.

$$\frac{{\ln \left( {\frac{{Rate_{i} }}{{Rate_{f} }}} \right)T_{f} T_{i} R}}{{T_{i} - T_{f} }} = E_{{a_{\text{Arr - plot,Emp}} }}$$
(25)
$$\frac{{\ln \left( {\frac{{Rate_{i} T_{f} }}{{Rate_{f} T_{i} }}} \right)T_{f} T_{i} R}}{{T_{i} - T_{f} }} = E_{{a_{TST,Emp} }}$$
(26)

Subtracting Eq. 26 from Eq. 25 yields the upper limit for difference between the empirical activation energy estimate from each type of linearization, Eq. 27:

$${{\Delta }}E_{{a,Emp{ - }Error}} = E_{{a_{{Arr{ - }plot,Emp}} }} - E_{{a_{TST,Emp} }} = \frac{{\ln \left( {\frac{{T_{i} }}{{T_{f} }}} \right)T_{f} T_{i} R}}{{T_{i} - T_{f} }}$$
(27)

This difference term is plotted over a wide range of initial and final temperatures (against the temperature difference of the initial and final data points) in Fig. 2.

Fig. 2
figure 2

Upper limit for the difference in the estimation of the activation energy between an Arrhenius-plot and TST as function of the temperature range of \({\text{T}}_{\text{i}}\) to \({\text{T}}_{\text{f}}\) using Eq. 27. Arrows and lines represent a constant Tf and Ti, respectively, as labeled on the plot. Changes in the x-axis following one of these lines represents changes in the variable term. For example, the arrow closest to both axes, Tf = 500 K, corresponds to Ti = 0 at the x-axis and Ti = 500 K where Tf − Ti = 0 (y-intercept). (Color figure online)

Several points are notable from this figure. First, the maximum difference between these theories can be substantial, up to ~ 20 kJ/mol. However, this is only valid for extremely high temperature studies going to up to 3500 K and beginning at 1500 K. While some systems, such as those for nuclear space propulsion [61] would certainly find such high temperature ranges relevant, most reaction systems are studied well below these conditions. Even forthcoming Generation IV nuclear power applications are being designed for upper temperatures near 1300 K, for the most aggressive systems [61]. In these systems the maximum deviation is significantly lower, near 10–12 kJ/mol. Depending on the system and data qualty, this difference may be within the experimental error of the data. Second, as inspection of Eq. 27 demonstrates, a lower reaction temperature significantly decreases the difference between the estimates. The effect of temperature range is so impactful, that a study ranging from 100 to 2000 K would report a difference of only 3 kJ/mol in the estimated activation energy. Over ranges of several hundred K or more, other errors may become dominant (for example, the common approximation that the standard enthalpy and standard entropy are independent of temperature may become inaccurate). Thus, while some difference between the theories is inescapable, in most systems experimental errors will be larger than this temperature dependence change. From a practical perspective, even a few kJ/mol can make a difference [62], and Eq. 27 can be used to go from activation energy estimates obtained from the end-points in an Arrhenius-plot to a more accurate TST estimate. The above considerations also suggest that a better choice is Eq. 8 or a modified Arrhenius plot of the form, where in the below equation \(u_{T}\) has the units of temperature.

$$\ln \left( {\frac{k}{T}\frac{{u_{T} }}{{u_{k} }}} \right) = \ln \left( {\frac{{k_{B} u_{T} }}{{h u_{k} }}e^{{\frac{{{{\Delta }}^{\ddag } S^\circ }}{R} + 1 - \frac{{P{{\Delta }}^{\ddag } V^\circ }}{RT}}} \frac{{(c_{TS}^\circ )^{ - 1} }}{{\left( {c_{R}^\circ } \right)^{ - 1} }}} \right) - \frac{{E_{a} }}{RT}$$
(28)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Doyle, P.J., Savara, A. & Raiman, S.S. Extracting meaningful standard enthalpies and entropies of activation for surface reactions from kinetic rates. Reac Kinet Mech Cat 129, 551–581 (2020). https://doi.org/10.1007/s11144-020-01747-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11144-020-01747-2

Keywords

Navigation