Next Article in Journal
Hydrodeoxygenation of Levulinic Acid to γ-Valerolactone over Mesoporous Silica-Supported Cu-Ni Composite Catalysts
Previous Article in Journal
Gas Sensors Based on Single-Wall Carbon Nanotubes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Comparison between 1,2-Dihydropyridine and 1,4-Dihydropyridine on Hydride-Donating Ability and Activity

The State Key Laboratory of Elemento-Organic Chemistry, College of Chemistry, and Collaborative Innovation Center of Chemical Science and Engineering, Nankai University, Tianjin 300071, China
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(17), 5382; https://doi.org/10.3390/molecules27175382
Submission received: 21 July 2022 / Revised: 16 August 2022 / Accepted: 19 August 2022 / Published: 24 August 2022
(This article belongs to the Section Organic Chemistry)

Abstract

:
In this paper, detailed comparisons of the driving force in thermodynamics and intrinsic force in the kinetics of 1,2-dihydropyridine and 1,4-dihydropyridine isomers of PNAH, HEH, and PYH in hydride transfer reactions are made. For 1,2-PNAH and 1,4-PNAH, the values of the thermodynamic driving forces, kinetic intrinsic barriers, and thermo-kinetic parameters are 60.50 and 61.90 kcal/mol, 27.92 and 26.34 kcal/mol, and 44.21 and 44.12 kcal/mol, respectively. For 1,2-HEH and 1,4-HEH, the values of the thermodynamic driving forces, kinetic intrinsic barriers, and thermo-kinetic parameters are 63.40 and 65.00 kcal/mol, 31.68 and 34.96 kcal/mol, and 47.54 and 49.98 kcal/mol, respectively. For 1,2-PYH and 1,4-PYH, the order of thermodynamic driving forces, kinetic intrinsic barriers, and thermo-kinetic parameters are 69.90 and 72.60 kcal/mol, 33.06 and 25.74 kcal/mol, and 51.48 and 49.17 kcal/mol, respectively. It is not difficult to find that thermodynamically favorable structures are not necessarily kinetically favorable. In addition, according to the analysis of thermo-kinetic parameters, 1,4-PNAH, 1,2-HEH, and 1,4-PYH have a strong hydride-donating ability in actual chemical reactions.

1. Introduction

Dihydropyridine (DHP) is the active structural core of a wide variety of natural products, drugs, and functional materials [1]. Of the five possible isomers, only 1,2 and 1,4-DHP have been studied in depth [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16]. Among them, 1,4-DHP is closest to NAD(P)H coenzyme; its biological applications are particularly extensive [17]. The active center of many drugs, such as nifedipine and amlodipine, is 1,4-DHP. Because it contains a chiral center, 1,2-DHP is mostly used as an important raw material for the active skeleton of natural alkaloids such as ibogaine, dioscorine, and the antiviral drug oseltamivir phosphate [18,19]. As organic hydride ion donors, both 1,2-DHP isomer and 1,4-DHP isomer have the same one-step hydride transfer mechanism when reacting with some negative ions, such as acridine perchlorate, 4-acetamido-2,2,6,6-tetramethyl-1-oxopiperidinium perchlorate, and 9-phenyl-2,3-dihydroxanthylium perchlorate (AcrH+ClO4, Tempo+ClO4 and PhXn+ClO4) [20,21,22,23,24]. An interesting question arises here: what is the difference between 1,2-DHP and 1,4-DHP in their hydride-donating ability?
To answer the above question, we chose three types of usual manmade NADH analogues, phenyl-1,4-dihydronicotinamide (1,4-PNAH) and phenyl-1,2-dihydronicotinamide (1,2-PNAH), N-CH3-1,2-Hantzsch (1,2-HEH) and N-CH3-1,4 Hantzsch (1,4-HEH), and 1-phenoxyacyl-1,2-dihydropyridine (1,2-PYH) and 1-phenoxyacyl-1,4-dihydropyridine (1,4-PYH), as the research objects (Scheme 1). Furthermore, in this work, bond dissociation free energy [ΔG°(XH)] as the thermodynamic driving force was used to discuss the hydride-donating ability of the above NADH analogues in terms of thermodynamics [25]. It is well-known that the thermodynamic driving force for the self-exchange transfer reaction of hydride ions is 0 kcal/mol (XH + X+→X+ + XH). Therefore, the activation free energy of the self-exchange reaction (ΔGXH/X) is used to describe the kinetic intrinsic barrier of the hydride donor. This indicates the kinetic intrinsic barrier of the compound itself in the chemical reaction. Thermo-kinetic parameters [ΔG°(XH)] are used to describe the actual hydride-donating ability of NADH analogues in chemical reactions [26]. It should be noted that in previous research reports of our group, we combined thermodynamic parameter [ΔG°(XH)] and kinetic parameter (ΔGXH/X) to define a new compound’s intrinsic physical parameter, which was named the thermo-kinetic parameter [ΔG°(XH)] [27]. According to the definition of the thermo-kinetic parameter (Equation (3)), the ΔG°(XH) value is determined by the value of ΔG°(XH) and the value of ΔGXH/X, and its value reflects the actual hydride-donating ability of the compound in a hydride transfer reaction. The larger the ΔG°(XH) value, the weaker the hydride-donating ability, and the smaller the ΔG°(XH) value, the stronger the hydride-donating ability.

2. Results

Two dihydrogen isomers of PNAH, HEH, and PYH were synthesized according to the literature methods and were identified by 1H NMR; the detailed data are listed in the Supporting Information [20,22,28,29,30,31]. The enthalpy change of the two dihydrogen isomers reacting with hydride acceptors was determined in acetonitrile using an isothermal titration calorimeter (CSC-4200 ITC) at 298 K as described previously (Figure 1) [32]. All kinetic tests were monitored in 298 K dry and anaerobic acetonitrile using an Applied Photophysics SX.18MV-R stopped-flow apparatus (Figure 2). The second rate constant (k2), activation free energies (ΔGXH/Y), and molar free energies ΔG°(XH/Y) of the three group reactions are listed in Table 1; see also Scheme 1. According to the data in Table 1; Table 2; and Equations (1)–(3), the thermodynamic driving forces, self-exchange reaction activation energies, and thermo-kinetic parameters values of 1,2/4-PNAH, 1,2/4-HEH, and 1,2/4-PYH are easily obtained (Table 3).
ΔG° = ΔG°H-D(XH) + ΔG°H-A(Y+)
ΔGXH/Y = ΔG°(XH) + ΔG°(Y+)
ΔG°(XH) = 1/2[ΔGXH/X + ΔG°(XH)]

3. Discussion

3.1. Analysis of Thermodynamic Driving Forces of 1,2/4-PNAH, 1,2/4-HEH, and 1,2/4-PYH as Hydride Donors in Acetonitrile at 298 K

As shown in Table 3, obviously, the heterolytic bond dissociation free energies of 1,2-PNAH and 1,4-PNAH, 1,2-HEH and 1,4-HEH, and 1,2-PYH and 1,4-PYH were 60.50 and 61.90 kcal/mol, 63.40 and 65.00 kcal/mol, and 69.90 and 72.60 kcal/mol, respectively. Whether it is 1,2-DHP or 1,4-DHP isomers, the order of ΔG°(XH) is PYH > HEH > PNAH. This indicates that PNAH has the best hydride-donating ability in thermodynamics, and the positive ion salt of PYH is commonly used as a hydride ion acceptor. In addition, the ΔG°(XH) of all the 1,4-DHP isomers was larger than that of the 1,2-DHP isomers, which indicates that the 1,2-DHP isomers have a better hydride-donating ability in thermodynamics.

3.2. Analysis of Kinetic Intrinsic Barriers of 1,2/4-PNAH, 1,2/4-HEH, and 1,2/4-PYH as Hydride Donors in Acetonitrile at 298 K

As shown in Table 3, the ΔGXH/X values of 1,2-PNAH and 1,4-PNAH, 1,2-HEH and N-1,4-HEH, and 1,2-PYH and 1,4-PYH were 27.92 and 26.34 kcal/mol, 31.68 and 34.96 kcal/mol, and 33.06 and 25.74 kcal/mol, respectively. For 1,2-DHP isomers, the order of ΔGXH/X is 1,2-PYH > 1,2-HEH > 1,2-PNAH. It is shown that 1,2-PNAH has the best hydride-donating ability, then 1,2-HEH, and 1,2-PYH has the worst hydride-donating ability in dynamics. However, for the 1,4-DHP isomers, the order of hydride-donating ability changed. The order of ΔGXH/X was 1,4-HEH > 1,4-PNAH > 1,4-PYH, which indicates that the alteration of the structure has a great effect on ΔGXH/X. The rule of the ΔGXH/X of the three groups of isomers was also different. The order of ΔGXH/X for the three groups of isomers was 1,2-PNAH > 1,4-PNAH, 1,2-HEH < 1,4-HEH, and 1,2-PYH > 1,4-PYH, respectively, which indicates that the effect of structure on ΔGXH/X is not a single rule. In addition, this also shows that the laws of the hydride-donating ability of three group dihydrogen isomers in kinetics and thermodynamics are almost completely different. Therefore, it is unscientific to use a single thermodynamic or kinetic parameter to analyze the hydride-donating ability of a compound.

3.3. Analysis of Thermo-Kinetic Parameters of 1,2/4-PNAH, 1,2/4-HEH, and 1,2/4-PYH as Hydride Donors in Acetonitrile at 298 K

As shown in Table 3, the ΔG°(XH) values of 1,2-PNAH and 1,4-PNAH, 1,2-HEH and 1,4-HEH, and 1,2-PYH and 1,4-PYH were 44.21 and 44.12 kcal/mol, 47.54 and 49.98 kcal/mol, and 51.48 and 49.17 kcal/mol, respectively. For 1,2-DHP isomers, the order of ΔG°(XH) was 1,2-PYH > 1,2-HEH > 1,2-PNAH. This indicates that 1,2-PNAH is the best hydride donor in actual hydride transfer reactions, then 1,2-HEH, and 1,2-PYH is the worst hydride donor in actual hydride transfer reactions. Additionally, for 1,4-DHP isomers, the order of ΔG°(XH) is 1,4-HEH > 1,4-PYH > 1,4-PNAH, which indicates that 1,4-PNAH is the best hydride donor in chemical reactions, then 1,4-PYH, and 1,4-HEH is the worst hydride donor in chemical reactions.
Furthermore, the order of ΔG°(XH) for the three groups of isomers is 1,2-PNAH > 1,4-PNAH, 1,2-HEH < 1,4-HEH and 1,2-PYH > 1,4-PYH, respectively. For PNAH, the thermodynamic driving force of 1,4-PNAH was 1.4 kcal/mol larger than that of 1,2-PNAH, and the kinetic intrinsic barrier was 1.58 kcal/mol smaller than that of 1,2-PNAH. Therefore, the difference in thermodynamic driving force shows that 1,4-PNAH has a smaller thermo-kinetic parameter, which indicates that 1,4-PNAH is a better hydride donor. For HEH, the thermodynamic driving force of 1,4-HEH was 1.6 kcal/mol larger than that of 1,2-HEH, and the kinetic intrinsic barrier was 3.28 kcal/mol larger than that of 1,2-HEH, which indicates that the difference in the thermodynamic driving force and the difference in the kinetic intrinsic barrier together mean that 1,2-HEH has a smaller thermo-kinetic parameter. This also means that 1,2-HEH has a better hydride-donating ability. Additionally, for PYH, the thermodynamic driving force of 1,4-PYH was 2.7 kcal/mol larger than that of 1,2-PYH, and the kinetic intrinsic barrier was 7.32 kcal/mol smaller than that of 1,2-PYH. This shows that the difference in the kinetic intrinsic barrier means that 1,4-PYH has a smaller thermo-kinetic parameter and a better hydride-donating ability. The internal reasons affecting the order of the thermo-kinetic parameters of the above three groups of compounds are dominated by the kinetic intrinsic barrier, co-dominated by a kinetic intrinsic barrier and thermodynamic driving force, and dominated by the thermodynamic driving force, respectively. In addition, it also shows that there is no linear relationship between the active site and the thermo-kinetic parameter of the dihydrogen isomers, and it is not advisable to infer a hydride-donating ability based on the structure of dihydrogen isomers. It is not difficult to find that the above analysis results are different from the results of the thermodynamic analysis and kinetic intrinsic barrier analysis. Therefore, the method of judging the hydride-donating ability only by the thermodynamic analysis and kinetic intrinsic barrier analysis is wrong.

4. Conclusions

In this work, we compared the hydride-donating ability of three groups of 1,2-DHP and 1,4-DHP dihydrogen isomers using thermodynamic parameters, kinetic parameters, and thermo-kinetic parameters. When describing the hydride-donating ability of NADH analogues, it is necessary to combine kinetic and thermodynamic parameters into thermo-kinetic parameter analysis instead of using a single experimental result. The order of the actual hydride-donating ability of the 1,2-DHP isomers and the 1,4-DHP isomers described by thermo-kinetic parameters is 1,2-PNAH > 1,2-HEH > 1,2-PYH and 1,4-PNAH > 1,4-PYH > 1,4-HEH, respectively. For the three groups of dihydrogen isomers, the actual hydride-donating ability described by thermo-kinetic parameters is 1,4-PNAH > 1,2-PNAH, 1,2-HEH > 1,4-HEH, and 1,4-PYH > 1,2-PYH, respectively. This indicates that there is no fixed linear relationship between the isomers’ structure and the actual hydride-donating ability. This method is not only applicable when describing hydride-donating abilities but also in describing hydrogen-donating abilities, proton-donating abilities, electron-donating abilities, etc.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27175382/s1. The detailed synthesis method of the compound are shown in SI. Detailed 1H NMR and 13C NMR data of typical compounds are shown in SII [20,22,28,29,30,31].

Funding

This work was supported by the National Natural Science Foundation of China (grant no. 21672111, 21472099, 21390400, and 21102074) and the 111 Project (B06005).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Liu, Z.; He, J.H.; Zhang, M.; Shi, Z.J.; Tang, H.; Zhou, X.Y.; Tian, J.J.; Wang, X.C. Borane-Catalyzed C3-Alkylation of Pyridines with Imines, Aldehydes, or Ketones as Electrophiles. J. Am. Chem. Soc. 2022, 144, 4810–4818. [Google Scholar] [CrossRef] [PubMed]
  2. Zhou, J.; Yu, J.-S.; Wu, W.-B.; Zhang, Z.-H.; Mu, B.-S. Recent Advances in Synthesis of Chiral 1,2-Dihydropyridines. Acta Chim. Sin. 2021, 79, 685–693. [Google Scholar] [CrossRef]
  3. Zhu, X.Q.; Zhang, J.Y.; Cheng, J.P. Negative Kinetic Temperature Effect on the Hydride Transfer from NADH Analogue BNAH to the Radical Cation of N-Benzylphenothiazine in Acetonitrile. J. Org. Chem. 2006, 71, 7007–7015. [Google Scholar] [CrossRef]
  4. Zhu, X.Q.; Li, H.R.; Li, Q.; Ai, T.; Lu, J.Y.; Yang, Y.; Cheng, J.P. Determination of the C4-H Bond Dissociation Energies of NADH Models and Their Radical Cations in Acetonitrile. Chem. Eur. J. 2003, 9, 871–880. [Google Scholar] [CrossRef] [PubMed]
  5. Li, Y.; Zhu, X.Q. Theoretical Prediction of Activation Free Energies of Various Hydride Self-Exchange Reactions in Acetonitrile at 298 K. ACS Omega 2018, 3, 872–885. [Google Scholar] [CrossRef]
  6. Fukuzumi, S.; Suenobu, T.; Patz, M.; Hirasaka, T.; Itoh, S.; Fujitsuka, M.; Ito, O. Selective One-Electron and Two-Electron Reduction of C60 with NADH and NAD Dimer Analogues via Photoinduced Electron Transfer. J. Am. Chem. Soc. 1998, 120, 8060–8068. [Google Scholar] [CrossRef]
  7. Fukuzumi, S.; Inada, O.; Suenobu, T. Mechanisms of Electron-Transfer Oxidation of NADH Analogues and Chemiluminescence. Detection of the Keto and Enol Radical Cations. J. Am. Chem. Soc. 2003, 125, 4808–4816. [Google Scholar] [CrossRef]
  8. Fukuzumi, S.; Fujioka, N.; Kotani, H.; Ohkubo, K.; Lee, Y.M.; Nam, W. Mechanistic Insights into Hydride-Transfer and Electron-Transfer Reactions by a Manganese(IV)-Oxo Porphyrin Complex. J. Am. Chem. Soc. 2009, 131, 17127–17134. [Google Scholar] [CrossRef]
  9. Peterson, R.L.; Himes, R.A.; Kotani, H.; Suenobu, T.; Tian, L.; Siegler, M.A.; Solomon, E.I.; Fukuzumi, S.; Karlin, K.D. Cupric superoxo-mediated intermolecular C-H activation chemistry. J. Am. Chem. Soc. 2011, 133, 1702–1705. [Google Scholar] [CrossRef]
  10. Zhang, J.; Yang, J.D.; Cheng, J.P. Diazaphosphinanes as hydride, hydrogen atom, proton or electron donors under transition-metal-free conditions: Thermodynamics, kinetics, and synthetic applications. Chem. Sci. 2020, 11, 3672–3679. [Google Scholar] [CrossRef]
  11. Vala, R.M.; Patel, D.M.; Sharma, M.G.; Patel, H.M. Impact of an aryl bulky group on a one-pot reaction of aldehyde with malononitrile and N-substituted 2-cyanoacetamide. RSC Adv. 2019, 9, 28886–28893. [Google Scholar] [CrossRef] [PubMed]
  12. Sharma, M.G.; Vala, R.M.; Patel, H.M. Pyridine-2-carboxylic acid as an effectual catalyst for rapid multi-component synthesis of pyrazolo[3,4-b]quinolinones. RSC Adv. 2020, 10, 35499–35504. [Google Scholar] [CrossRef] [PubMed]
  13. Sharma, M.G.; Vala, R.M.; Patel, D.M.; Lagunes, I.; Fernandes, M.X.; Padrón, J.M.; Ramkumar, V.; Gardas, R.L.; Patel, H.M. Anti-Proliferative 1,4-Dihydropyridine and Pyridine Derivatives Synthesized through a Catalyst-Free, One-Pot Multi-Component Reaction. ChemistrySelect 2018, 3, 12163–12168. [Google Scholar] [CrossRef]
  14. Patel, D.M.; Sharma, M.G.; Vala, R.M.; Lagunes, I.; Puerta, A.; Padron, J.M.; Rajani, D.P.; Patel, H.M. Hydroxyl alkyl ammonium ionic liquid assisted green and one-pot regioselective access to functionalized pyrazolodihydropyridine core and their pharmacological evaluation. Bioorg. Chem. 2019, 86, 137–150. [Google Scholar] [CrossRef]
  15. Sharma, M.G.; Rajani, D.P.; Patel, H.M. Green approach for synthesis of bioactive Hantzsch 1,4-dihydropyridine derivatives based on thiophene moiety via multicomponent reaction. R Soc. Open Sci. 2017, 4, 170006. [Google Scholar] [CrossRef]
  16. A Khedkar, S.; B Auti, P. 1, 4-Dihydropyridines: A Class of Pharmacologically Important Molecules. Med. Chem. 2014, 14, 282–290. [Google Scholar] [CrossRef]
  17. Yang, J.D.; Chen, B.L.; Zhu, X.Q. New Insight into the Mechanism of NADH Model Oxidation by Metal Ions in Nonalkaline Media. J. Phys. Chem. B 2018, 122, 6888–6898. [Google Scholar] [CrossRef]
  18. Sharma, V.K.; Singh, S.K. Synthesis, utility and medicinal importance of 1,2- & 1,4-dihydropyridines. RSC Adv. 2017, 7, 2682–2732. [Google Scholar] [CrossRef]
  19. Sharma, M.G.; Pandya, J.; Patel, D.M.; Vala, R.M.; Ramkumar, V.; Subramanian, R.; Gupta, V.K.; Gardas, R.L.; Dhanasekaran, A.; Patel, H.M. One-Pot Assembly for Synthesis of 1,4-Dihydropyridine Scaffold and Their Biological Applications. Polycycl. Aromat. Compd. 2019, 41, 1495–1505. [Google Scholar] [CrossRef]
  20. Zhu, X.Q.; Cao, L.; Liu, Y.; Yang, Y.; Lu, J.Y.; Wang, J.S.; Cheng, J.P. Thermodynamics and kinetics of the hydride-transfer cycles for 1-aryl-1,4-dihydronicotinamide and its 1,2-dihydroisomer. Chem. Eur. J. 2003, 9, 3937–3945. [Google Scholar] [CrossRef]
  21. Xia, K.; Shen, G.B.; Zhu, X.Q. Thermodynamics of various F420 coenzyme models as sources of electrons, hydride ions, hydrogen atoms and protons in acetonitrile. Org. Biomol. Chem. 2015, 13, 6255–6268. [Google Scholar] [CrossRef] [PubMed]
  22. Shen, G.B.; Xia, K.; Li, X.T.; Li, J.L.; Fu, Y.H.; Yuan, L.; Zhu, X.Q. Prediction of Kinetic Isotope Effects for Various Hydride Transfer Reactions Using a New Kinetic Model. J. Phys. Chem. A 2016, 120, 1779–1799. [Google Scholar] [CrossRef] [PubMed]
  23. Zhu, X.Q.; Zhang, M.T.; Yu, A.; Wang, C.H.; Cheng, J.P. Hydride, Hydrogen Atom, Proton, and Electron Transfer Driving Forces of Various Five-Membered Heterocyclic Organic Hydrides and Their Reaction Intermediates in Acetonitrile. J. Am. Chem. Soc. 2008, 130, 2501–2516. [Google Scholar] [CrossRef]
  24. Yuasa, J.; Yamada, S.; Fukuzumi, S. A Mechanistic Dichotomy in Scandium Ion-Promoted Hydride Transfer of an NADH Analogue: Delicate Balance between One-Step Hydride-Transfer and Electron-Transfer Pathways. J. Am. Chem. Soc. 2006, 128, 14938–14948. [Google Scholar] [CrossRef] [PubMed]
  25. Fu, Y.H.; Shen, G.B.; Wang, K.; Zhu, X.Q. Comparison of Thermodynamic, Kinetic Forces for Three NADH Analogues to Release Hydride Ion or Hydrogen Atom in Acetonitrile. ChemistrySelect 2021, 6, 8007–8010. [Google Scholar] [CrossRef]
  26. Fu, Y.H.; Wang, K.; Shen, G.B.; Zhu, X.Q. Quantitative comparison of the actual antioxidant activity of Vitamin C, Vitamin E, and NADH. J. Phys. Org. Chem. 2022, 35, e4358. [Google Scholar] [CrossRef]
  27. Fu, Y.-H.; Shen, G.-B.; Li, Y.; Yuan, L.; Li, J.-L.; Li, L.; Fu, A.-K.; Chen, J.-T.; Chen, B.-L.; Zhu, L.; et al. Realization of Quantitative Estimation for Reaction Rate Constants Using only One Physical Parameter for Each Reactant. ChemistrySelect 2017, 2, 904–925. [Google Scholar] [CrossRef]
  28. Bobbitt, J.M. Oxoammonium Salts. 6. 4-Acetylamino-2,2,6,6-tetramethylpiperidine-1-oxoammonium Perchlorate: A Stable and Convenient Reagent for the Oxidation of Alcohols. Silica Gel Catalysis. J. Org. Chem. 1998, 63, 9367–9374. [Google Scholar] [CrossRef]
  29. Comins, D.L.; Abdullah, A.H. Synthesis of l-Acyl-1,4-dihydropyridines via Copper Hydride Reduction of 1-Acylpyridinium Salts. J. Org. Chem. 1984, 49, 3392–3394. [Google Scholar] [CrossRef]
  30. Zhu, X.Q.; Deng, F.H.; Yang, J.D.; Li, X.T.; Chen, Q.; Lei, N.P.; Meng, F.K.; Zhao, X.P.; Han, S.H.; Hao, E.J.; et al. A classical but new kinetic equation for hydride transfer reactions. Org. Biomol. Chem. 2013, 11, 6071–6089. [Google Scholar] [CrossRef]
  31. Zhu, X.Q.; Tan, Y.; Cao, C.T. Thermodynamic Diagnosis of the Properties and Mechanism of Dihydropyridine-Type Compounds as Hydride Source in Acetonitrile with “Molecule ID Card”. J. Phys. Chem. B 2010, 114, 2058–2075. [Google Scholar] [CrossRef] [PubMed]
  32. Zhao, H.; Li, Y.; Zhu, X.Q. Thermodynamic Parameters of Elementary Steps for 3,5-Disubstituted 1,4-Dihydropyridines To Release Hydride Anions in Acetonitrile. ACS Omega 2018, 3, 13598–13608. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. 1,2-DHP and 1,4-DHP isomers hydride transfer reaction equation.
Scheme 1. 1,2-DHP and 1,4-DHP isomers hydride transfer reaction equation.
Molecules 27 05382 sch001
Figure 1. Isothermal titration calorimetry (ITC) for the reaction heat of 1,2−PYH with PhXn+ in CH3CN at 298 K. Titration was conducted by adding 10 μL of 1,2−PYH (2.34 mM) every 600 s into the CH3CN containing PhXn+ (ca.10 mM).
Figure 1. Isothermal titration calorimetry (ITC) for the reaction heat of 1,2−PYH with PhXn+ in CH3CN at 298 K. Titration was conducted by adding 10 μL of 1,2−PYH (2.34 mM) every 600 s into the CH3CN containing PhXn+ (ca.10 mM).
Molecules 27 05382 g001
Figure 2. Time profile of the UV absorbance at 372 nm due to PhXn+ for the reactions of PhXn+ (0.1 mM) with 1,2-PYH (2 mM) in CH3CN at 298 K.
Figure 2. Time profile of the UV absorbance at 372 nm due to PhXn+ for the reactions of PhXn+ (0.1 mM) with 1,2-PYH (2 mM) in CH3CN at 298 K.
Molecules 27 05382 g002
Table 1. Second-order rate constants (k2), activation free energies (ΔG), and molar free energy change (ΔG°) of oxidations of dihydropyridine compounds in CH3CN at 298 K.
Table 1. Second-order rate constants (k2), activation free energies (ΔG), and molar free energy change (ΔG°) of oxidations of dihydropyridine compounds in CH3CN at 298 K.
PNAH + AcrH+HEH + TEMPO+PYH + PhXn+
1,2-PNAH1,4-PNAH1,2-HEH1,4-HEH1,2-PYH1,4-PYH
k2a10.0612.301.44 × 1052.34 × 1039.50 × 10−14.76 × 101
ΔGb16.0515.9610.4112.8517.4715.16
ΔG° c−15.70−14.30−37.30−35.70−21.70−19.00
a k2 (M−1 s−1) is the second-order rate constant of the hydride transfer in CH3CN at 298 K. The uncertainty is smaller than 5%. b Derived from Eyring equation (T = 298 K), the unit is kcal/mol. c ΔG° is equal to the corresponding reaction heat; the latter was measured by titration calorimetry in CH3CN at 298 K. The data given in kcal/mol are the average values of at least three independent runs.
Table 2. Bond formation free energy and thermo-kinetic parameters of hydride acceptor used in this work (T = 298 K).
Table 2. Bond formation free energy and thermo-kinetic parameters of hydride acceptor used in this work (T = 298 K).
CompoundsΔG°(Y+) aΔG≠o(Y+) b
AcrH+−76.2−28.16
TEMPO+−100.7−37.13
PhXn+−91.6−34.01
a,b Reference from [30]; the unit is kcal/mol.
Table 3. Thermodynamic driving forces [ΔG°(XH)], self-exchange reaction activation energies [ΔGXH/X], and thermo-kinetic parameters [ΔG°(XH)] of 1,2/4-PNAH, 1,2/4-HEH, and 1,2/4-PYH as hydride donors at 298 K in CH3CN at 298 K.
Table 3. Thermodynamic driving forces [ΔG°(XH)], self-exchange reaction activation energies [ΔGXH/X], and thermo-kinetic parameters [ΔG°(XH)] of 1,2/4-PNAH, 1,2/4-HEH, and 1,2/4-PYH as hydride donors at 298 K in CH3CN at 298 K.
1,2-PNAH1,4-PNAH1,2-HEH1,4-HEH1.2-PYH1,4-PYH
ΔG°(XH)60.5061.9063.4065.0069.9072.60
ΔGXH/X27.9226.3431.6834.9633.0625.74
ΔG°(XH)44.2144.1247.5449.9851.4849.17
The unit is kcal/mol.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, J.-Y.; Zhu, X.-Q. Comparison between 1,2-Dihydropyridine and 1,4-Dihydropyridine on Hydride-Donating Ability and Activity. Molecules 2022, 27, 5382. https://doi.org/10.3390/molecules27175382

AMA Style

Zhang J-Y, Zhu X-Q. Comparison between 1,2-Dihydropyridine and 1,4-Dihydropyridine on Hydride-Donating Ability and Activity. Molecules. 2022; 27(17):5382. https://doi.org/10.3390/molecules27175382

Chicago/Turabian Style

Zhang, Jin-Ye, and Xiao-Qing Zhu. 2022. "Comparison between 1,2-Dihydropyridine and 1,4-Dihydropyridine on Hydride-Donating Ability and Activity" Molecules 27, no. 17: 5382. https://doi.org/10.3390/molecules27175382

Article Metrics

Back to TopTop