DFT calculations and dynamic NMR revealed the coalescent NMR phenomena of the 6/6/6/9 tetracyclic merosesquiterpenoids with an unprecedented 9,15-dioxatetracyclo[8.5.3.04,17.014,18]octadecane core skeleton

Hanqi Zhang Biao Gao Yuanyuan Feng Guijuan Zheng Zhijun Liu Lichun Kong Junjun Liu Haji Akber Aisa Guangmin Yao

Citation:  Hanqi Zhang, Biao Gao, Yuanyuan Feng, Guijuan Zheng, Zhijun Liu, Lichun Kong, Junjun Liu, Haji Akber Aisa, Guangmin Yao. DFT calculations and dynamic NMR revealed the coalescent NMR phenomena of the 6/6/6/9 tetracyclic merosesquiterpenoids with an unprecedented 9,15-dioxatetracyclo[8.5.3.04,17.014,18]octadecane core skeleton[J]. Chinese Chemical Letters, 2025, 36(9): 111234. doi: 10.1016/j.cclet.2025.111234 shu

DFT calculations and dynamic NMR revealed the coalescent NMR phenomena of the 6/6/6/9 tetracyclic merosesquiterpenoids with an unprecedented 9,15-dioxatetracyclo[8.5.3.04,17.014,18]octadecane core skeleton

English

  • The functionalized nine-membered ring plays an important role in the bioactivities of natural products, and has been found in meroterpenoids [1], sesquiterpenoids [2], diterpenoids [3,4], and alkaloids from plants [5], fungi, and marine organisms. Neocarzinostatin (Fig. S1 in Supporting information), a polypeptide containing an enediynyl nine-membered core, has been used to treat various human cancers in the clinic [6]. α-Viniferin and protoxenicin A (Fig. S1) bearing a nine-membered ring exhibit their potential as lead compounds for anti-inflammatory and anticancer drugs, respectively [7]. The unique architectural features and potent medicinal properties of such natural products have attracted more and more attention from synthetic and medicinal chemists.

    Ericaceae plants are widely used as traditional Chinese medicines and folk medicines, and many structurally diverse meroterpenoids [8], diterpenoids [9,10], triterpenoids [11], and flavonoids [11], have been reported. In a continuing to search for structurally novel natural products from Ericaceae plants, two pairs of novel 6/6/6/9 tetracyclic merosesquiterpenoid enantiomers featuring an unprecedented 9,15-dioxatetracyclo[8.5.3.04,17.014,18]octadecane skeleton (Fig. 1), namely dauroxonanols A (1) and B (2), were isolated from the leaves of Rhododendron dauricum, a famous traditional Chinese medicine Man-Shan-Hong. The nuclear magnetic resonance (NMR) spectra of 1 showed very broad resonances, and 13C NMR spectrum of 1 exhibited only 13 instead of 22 carbon resonances. These coalescent NMR phenomena made a great challenge to elucidate the structure of 1 using NMR data analysis.

    Figure 1

    Figure 1.  Chemical structures of (±)-dauroxonanols A (1) and B (2) and the nomenclature of the unprecedented ring system.

    The 1H NMR spectrum of dauroxonanol A (1) acquired in chloroform-d at room temperature (Fig. 2A and Table S1 in Supporting information) displayed two extremely broad resonances for an aromatic proton (δH 6.48, br. s, H-6) and an oxymethine (δH 4.04, br. s, H-15α), along with normal resonances for four methyl groups (δH 1.26, s, H3-17; 1.25, s, H3-18; 1.34, s, H3-20; 2.21, s, H3-21), an aromatic proton (δH 6.30, br. s, H-8), and an olefinic proton (δH 4.73, s, H-19a). 13C NMR spectrum of 1 only revealed 13 recognizable carbon resonances (Fig. 2B and Table S2 in Supporting information) instead of 22 carbon resonances assigned by its HRESIMS ion at m/z 365.2079 (calcd. for C22H30O3Na, 365.2093). To obtain high-resolution data, the NMR spectra of 1 were acquired in pyridine-d5 and methanol-d4, respectively, and increasing the number of scans (ns) from 1000 (1 h) to 2000 (2 h). However, none of them showed sharp NMR resonances (Figs. 2A and B, Fig. S2 in Supporting information).

    Figure 2

    Figure 2.  1H (400 MHz, A) and 13C (100 MHz, B) NMR spectra of 1 at 298 K with broad signals. Key 1H–1H COSY and HMBC correlations of 1 (C). ORTEP drawing of the crystal structure of 1 with the ellipsoid contour at 15% probability level (D). 1H–1H COSY, HMBC, and NOESY correlations of 2 (E).

    Heteronuclear single quantum correlation (HSQC) data assigned ten protonated carbons and two methylenes with broad carbon resonances at δC 37.3 (C-3) and 23.5 (C-10). 1H–1H correlation spectroscopy (COSY) spectrum only showed one spin system for H2-9/H2-10 (Fig. 2C). The connections of C-3/C-9/C-20 to the oxygenated tertiary carbon C-2, C-15/C-17/C-18 to the oxygenated tertiary carbon C-16, and C-21/C-6/C-8 to C-7 were defined by HMBC correlations (Fig. 2C). Except for these, there were no more useful NMR data to determine the structure of 1.

    Finally, the structure of 1 was established as a 6/6/6/9 tetracyclic meroterpenoid with an unprecedented 9,15-dioxatetracyclo[8.5.3.04,17.014,18]octadecane scaffold by single crystal X-ray diffraction analysis (Fig. 2D and Fig. S3 in Supporting information).

    Since Rhododendron meroterpenoids have been reported as scalemic mixtures [8,12], 1 having a large specific rotation ([α]D25 −41) was subjected to chiral high performance liquid chromatography (HPLC) analysis (Fig. S4A and Table S3 in Supporting information), revealing that 1 is a scalemic mixture in a ratio of 1:4, differing from the crystallographic data. Subsequent chiral HPLC separation yielded a pair of enantiomers, (+)-1 ([α]D25 +67) and (−)-1 ([α]D25 −67), and their absolute configurations were established as 2S, 4S, 11R, 15R and 2R, 4R, 11S, 15S, respectively, by ECD calculations (Fig. S4B) [[13], [14]–15].

    The crystal structure (Fig. S5 in Supporting information) and the P21 space group of (−)-1 from a mixed solvent of MeOH/H2O (v/v, 20:1) were found to be identical to those of 1 from methanol (Fig. 2D). Thus, the crystal of 1 previously obtained from methanol should be the major enantiomer (−)-1. The absolute configuration of (−)-1 was further confirmed by the Flack parameter of 0.00(9) [16]. While, crystals of (+)-1 for X-ray diffraction were not obtained.

    The structure of dauroxonanol B (2) was established as a 15-epimer of 1 by the comprehensive spectroscopic data analyses (Fig. 2E). Detailed structural elucidation and chiral separation of 2 were described in Figs. S6–S8 and Tables S4–S6 (Supporting information).

    Dauroxonanols A (1) and B (2) represent the first 6/6/6/9 tetracyclic merosesquiterpene skeleton bearing an unprecedented 9,15-dioxatetracyclo[8.5.3.04,17.014,18] octadecane core, and their biosynthetic pathways are proposed in Scheme S1 (Supporting information).

    NMR spectra are usually the average behavior of all conformations of molecules in solvents under certain conditions. To interpret the coalescent NMR phenomenon of 1, a conformational analysis was carried out [[17], [18]–19], revealing the presence of 20 low-energy conformers in 1 (Fig. S9 and Tables S7 and S8 in Supporting information). These conformers mainly differ in the conformation of the oxonane ring moiety and could be classified into three distinct groups: skewed chair-boat, twist chair-boat, and twist chair-chair.

    To investigate the conformational exchange of the oxonane ring, the relaxed potential energy surface scan along the torsion angle ωC11–C12–C13–C14 was conducted at B3LYP/6-31G(d, p) level using Gaussian 09 program [13,20], starting from conformer SCB1. Results (Fig. S10 in Supporting information) revealed the inversion involves two steps: SCB1 migrates to TCC1 through TS1, and then, TCC1 transitions to TCB1 through TS2. Subsequent optimization and Gibbs free energies calculations at the M06-2X/def2-TZVP level in chloroform (Fig. 3 and Table S9 in Supporting information) revealed a lower energy barrier $\left(\Delta G^{\ddagger}\right)$ of 15.86 kcal/mol from SCB1 to TCB1 [21]. Thus, SCB1 and TCB1 readily interconvert in chloroform at 298 K.

    Figure 3

    Figure 3.  DFT calculated free energy profiles for the pathways of the conformational changes in 1 and 2 at the M06-2X/def2-TZVP level in chloroform at 298 K. The key states along the pathways were obtained by full optimization based on the potential energy surface scan in Fig. S10.

    Generally, when a molecule possesses multiple conformers and their interconversion rates are fast, the observed NMR signals represent an average from each conformer and manifest as sharp peaks [17]. Thus, the broadening and missing NMR signals of 1 indicated that its conformers interconvert at a rate slower than but still at a rate relatively close to the NMR timescale [17,22,23]. To promote the interconversion rates of conformers of 1, the temperature-rising NMR experiments of 1 in pyridine-d5 and chloroform-d were performed. In pyridine-d5 (Fig. 4A), the peak pattern of H-15 was changed from a broad singlet at 298, 303, and 313 K to be a triplet-like at 323 K, and a normal sharp triplet at 333 K. At 343 K, H-6 showed a very broad peak at δH 7.09. Similar dynamic effects occurred in chloroform-d (Figs. S11 and S12 in Supporting information), in which H-4β, H-6, H-15α, H-19, C-3, C-13–C-15, and C-19 showed broad signals at 326 K. With increasing temperature, the NMR spectra of 1 in pyridine-d5 and chloroform-d showed sharper and higher resolution compared to those at 298 K. Therefore, higher temperatures may accelerate the interconversion between different conformers of 1, thereby improving the resolution of NMR spectrum. However, not all peaks of 1 still could be observed even at 373 K, due to the limitations of the interconversion of conformers.

    Figure 4

    Figure 4.  Temperature-dependent 1H NMR (600 MHz) spectra of 1 in pyridine-d5 (298–373 K, A) and in chloroform-d (213–293 K, B), respectively. Broadening or missing NMR signals in 1 in 298 K (C). The experimental (213 K) and calculated 1H (D) and 13C (E) NMR chemical shifts differences between conformers SCB1 and TCB1.

    When conformational equilibrium is slow, more than one conformer may be observed in the NMR spectrum [2426]. To observe the coexisting conformers, the temperature-decreasing NMR experiments of 1 in chloroform-d (Fig. 4B) were carried out. At 283 K, the 1H NMR spectrum of 1 still showed broadening and missing signals as those at 293 and 298 K. From 273 K, the 1H NMR spectrum started to exhibit distinct resonances for two conformers. As temperature decreases, the 1H NMR spectrum shows sharper and higher resolution signals. At 213 K, the 1H NMR spectrum of 1 displayed well-resolved signals for two independent conformers in a ratio of 1:0.88 (Fig. S13 in Supporting information), and their NMR data (Tables S1 and S2 in Supporting information) were completely assigned by HSQC and HMBC data analysis. For H-6 (Fig. 4B), two distinct signals (Δν = 143.8 Hz) were observed below 253 K, and they were broadening from 263 K to 283 K, and finally coalesced to be a very broad peak at 293 K. The free energy of activation $\Delta G_{\mathrm{c}}{ }^{\ddagger}$ and the rate constant kc of the process of transformation between two conformers at the coalescence temperature Tc = 293 K were calculated to be 13.8 kcal/mol and 319.5 s‒1, respectively, using the Eyring equation [27]. The free energies of activation $\Delta G_{\mathrm{c}}{ }^{\ddagger}$ of H-4β, H-6, H-8, H-15, H-19a, H-19b, H-20, and H-21, ranging from 13.3 kcal/mol to 14.0 kcal/mol (Table 1), were found to be very close to the calculated energy barrier $\Delta G^{\ddagger}$ of 15.86 kcal/mol from SCB1 to TCB1 (Fig. 3). Their rate constants k, ranging from 12.1 to 723.6 s‒1, fall within the intermediate-exchange regime of 1 < k < 104 s‒1, and some hydrogens coalesced around 293 K. This is why some peaks are broadening and missing in the 1H NMR spectrum at 298 K.

    Table 1

    Table 1.  Representative activation parameters of 1.
    DownLoad: CSV
    NucleiΔν (Hz)Tc (K)kc (s‒1)$\Delta G_{\mathrm{c}}{ }^{\ddagger}$ (kcal/mol)
    H-4β325.7293723.613.3
    H-6143.8293319.513.8
    H-85.4626312.114.0
    H-15164.8293366.113.7
    H-19a179.6293399.113.7
    H-19b208.0293462.113.5
    H-2016.627336.814.0
    H-2127.127360.313.7

    Larger NMR chemical shift differences between two major conformers SCB1 and TCB1 would hinder the observation of the NMR signals of 1 (Fig. 4C). To further prove this deduction, the NMR data of SCB1 and TCB1 were calculated at the mPW1PW91/6–311+G(2d, p) level using the gauge independent atomic orbital (GIAO) method [10,28,29]. The calculated NMR data of conformers SCB1 and TCB1 well matched the experimental ones at 213 K (Tables S10 and S11 in Supporting information) with much smaller root mean square deviation (RMSD) values (0.07–0.14 ppm for 1H and 2.00–2.07 ppm for 13C) than the expected precision limits (0.16 ppm for 1H and 2.45 ppm for 13C) [30]. The calculated 1H NMR chemical shift differences of H2–3, H-4, H-6, H2–10, H-11, H2–13, H2–14, and H2–19 exceeded 0.1 ppm (Fig. 4D and Table S12 in Supporting information), and the calculated 13C NMR chemical shift differences of C-3, C-4, C-10–C-15, and C-19 between two conformers SCB1 and TCB1 exceeded 1.8 ppm (Fig. 4E and Table S13 in Supporting information). These notable chemical shift differences indicated an NMR distinction between two conformers of 1 (Fig. 4C). In contrast, the chemical shift differences of C-2, C-4a, C-8a, C-5, C-6–C-9, C-17, C-18, C-20, and C-21 were less than 1.0 ppm, and exhibited sharp and normal carbon resonances. Thus, the dynamic conformation inversion of the oxonane ring and the large NMR differences between conformers are responsible for the coalescent NMR characteristics of 1. Compared to 1, dauroxonanol B (2) showed relatively normal NMR resonances except for C-13, which was too broad and too weak to be observed (Fig. S14 in Supporting information). Conformational analysis and density functional theory (DFT) calculations elucidated the transformation of the oxonane ring conformation in 2 from a twisted chair-boat to a skewed chair-boat, with relatively low energy barriers at 12.84 kcal/mol (Fig. 3 and Table S14 in Supporting information). Furthermore, 13C NMR calculations revealed a larger chemical shift difference (∆δC 8.8 ppm) of C-13 between SBC1 and TBB1 (Table S15 in Supporting information), potentially accounting for the broadening and weak C-13 resonance in the 13C NMR spectrum of 2. Interestingly, the energy barrier of 2 ($\Delta G^{\ddagger}$ = 12.84 kcal/mol) was found to be lower than 1 ($\Delta G^{\ddagger}$ = 15.86 kcal/mol), indicating faster interconversion rates between conformers of 2. This observation could explain the higher resolution of the NMR signals in 2 compared to 1.

    Nine-membered rings are important structural blocks of bioactive natural products [2426]. Due to the flexible nine-membered ring, the intermediate exchange rate often manifests itself in broadened NMR linewidths. Thus, it is difficult to determine their structures based on NMR data analysis. This study provided a new example to elucidate the structures of natural products with a novel 9,15-dioxatetracyclo[8.5.3.04,17.014,18]octadecane skeleton.

    (+)/(−)-Dauroxonanols A (1) and B (2) displayed significant α-glucosidase inhibitory activities with half maximal inhibitory concentration (IC50) values ranging from 90.68 µmol/L to 363.98 µmol/L (Table S16 and Fig. S16 in Supporting information), which were 2–8 times more potent than the positive control acarbose (IC50 = 714.75 µmol/L). Molecular docking revealed the binding modes of (+)/(−)-1 and 2 with α-glucosidase. Detailed structure-activity relationship analysis and molecular docking studies (Fig. S17 in Supporting information) were described in Supporting Information.

    In conclusion, dauroxonanols A (1) and B (2), two pairs of novel merosesquiterpenoid enantiomers possessing an unprecedented 9,15-dioxatetracyclo[8.5.3.04,17.014,18] octadecane skeleton, were isolated from R. dauricum. NMR spectra of 1 showed very broad and missing resonances, leading to a great challenge to elucidate its structure using NMR data analysis. DFT calculations, conformational analysis, and dynamic NMR study revealed that the coalescent NMR phenomena of 1 and 2 result from the conformational changes of the flexible oxonane ring in 1 and 2. These results enriched the structural diversity of merosesquiterpenoids with a flexible nine-membered ring.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Hanqi Zhang: Software, Methodology, Investigation, Formal analysis, Data curation. Biao Gao: Methodology, Investigation, Data curation. Yuanyuan Feng: Investigation, Data curation. Guijuan Zheng: Writing – review & editing, Funding acquisition, Conceptualization. Zhijun Liu: Software, Data curation. Lichun Kong: Data curation. Junjun Liu: Methodology, Investigation. Haji Akber Aisa: Supervision, Resources, Conceptualization. Guangmin Yao: Writing – review & editing, Validation, Supervision, Project administration, Funding acquisition, Formal analysis, Conceptualization.

    This work was supported by the National Natural Science Foundation of China (Nos. 22207036, 22277034, 22477034, and 22107033), Interdisciplinary Research Program of Huazhong University of Science and Technology (No. 2023JCYJ037), and International Cooperation Project of Hubei Provincial Key R&D Plan (No. 2023EHA040). We are grateful to the Analytical and Testing Center at Huazhong University of Science and Technology for IR, ECD, and single crystal X-ray diffraction data collection, Medical Subcenter at Huazhong University of Science and Technology for NMR data acquisition. The computation is completed in the HPC Platform of Huazhong University of Science and Technology.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2025.111234.


    1. [1]

      H.Y. Lou, Y.N. Li, P. Yi, et al., Org. Lett. 22 (2020) 6903–6906. doi: 10.1021/acs.orglett.0c02434

    2. [2]

      M. Pereira, T. Silva, L. Lopes, et al., Molecules 17 (2012) 14046–14057. doi: 10.3390/molecules171214046

    3. [3]

      S. Li, L. Gan, Y.J. Tian, et al., Bioorg. Chem. 114 (2021) 105222. doi: 10.1016/j.bioorg.2021.105222

    4. [4]

      J. Liu, Q. Tang, J. Huang, et al., J. Org. Chem. 87 (2022) 9806–9814. doi: 10.1021/acs.joc.2c00784

    5. [5]

      J.M. Jiang, D. Xia, X.L. Zhu, et al., Tetrahedron 114 (2022) 132782. doi: 10.1016/j.tet.2022.132782

    6. [6]

      J. Fujita, M. Taniguchi, C. Hashizume, et al., Mol. Pharmacol. 101 (2022) 322–333. doi: 10.1124/molpharm.121.000379

    7. [7]

      T. Huber, R.E. Wildermuth, T. Magauer, Chemistry 24 (2018) 12107–12120. doi: 10.1002/chem.201705919

    8. [8]

      Q. Shi, T.T. Li, Y.M. Wu, et al., Phytochemistry 180 (2020) 112524. doi: 10.1016/j.phytochem.2020.112524

    9. [9]

      G. Zheng, L. Huang, Y. Feng, et al., Bioorg. Chem. 142 (2024) 106928. doi: 10.1016/j.bioorg.2023.106928

    10. [10]

      Y. Feng, S. Zha, H. Zhang, et al., Chin. Chem. Lett. 34 (2023) 107742. doi: 10.1016/j.cclet.2022.107742

    11. [11]

      C. Ye, M. Jin, C.S. Jin, et al., Nat. Prod. Res. 35 (2021) 1331–1339. doi: 10.1080/14786419.2019.1648455

    12. [12]

      G.H. Huang, Z. Hu, C. Lei, et al., J. Nat. Prod. 81 (2018) 1810–1818. doi: 10.1021/acs.jnatprod.8b00273

    13. [13]

      H.W. Yan, Y.N. Yang, X. Zhang, et al., Bioorg. Chem. 128 (2022) 106091. doi: 10.1016/j.bioorg.2022.106091

    14. [14]

      X. Zhang, J. Li, K.Z. Lu, et al., Chin. Chem. Lett. 35 (2024) 109456. doi: 10.1016/j.cclet.2023.109456

    15. [15]

      Q. Tan, R.Z. Fan, W.C. Yang, et al., Chin. Chem. Lett. 35 (2024) 109390. doi: 10.1016/j.cclet.2023.109390

    16. [16]

      R.W. Hooft, L.H. Straver, A.L. Spek, J. Appl. Crystallogr. 41 (2008) 96–103. doi: 10.1107/S0021889807059870

    17. [17]

      T. Mitsuhashi, T. Kikuchi, S. Hoshino, et al., Org. Lett. 20 (2018) 5606–5609. doi: 10.1021/acs.orglett.8b02284

    18. [18]

      H.W. Yan, L.H. Zhao, X. Zhang, et al., J. Nat. Prod. 84 (2021) 2981–2989. doi: 10.1021/acs.jnatprod.1c00830

    19. [19]

      RDKit: Open-source cheminformatics. http://www.rdkit.org.

    20. [20]

      G.Y. Xia, B.B. Xiao, L.Y. Wang, et al., Chin. Chem. Lett. 34 (2023) 108073. doi: 10.1016/j.cclet.2022.108073

    21. [21]

      W.Y. Liu, X.X. Lei, W.J. Wang, et al., Chin. Chem. Lett. 36 (2025) 110478. doi: 10.1016/j.cclet.2024.110478

    22. [22]

      L.C. Forster, G.K. Pierens, J.K. Clegg, et al., J. Nat. Prod. 83 (2020) 714–719. doi: 10.1021/acs.jnatprod.9b01051

    23. [23]

      H.X. Liu, K. Chen, Y. Yuan, et al., Org. Biomol. Chem. 14 (2016) 7354–7360. doi: 10.1039/C6OB01215A

    24. [24]

      A.B. Holmes, R.P. McGeary, Nine-membered rings, in: A.R. Katritzky, C.W. Rees, E.F.V. Scriven (Eds.), Comprehensive Heterocyclic Chemistry Ⅱ, Pergamon, Oxford, 1996, pp. 737–788.

    25. [25]

      D.O. Tymoshenko, Nine-membered rings, in: A.R. Katritzky, C.A. Ramsden, E.F.V. Scriven, R.J.K. Taylor (Eds.), Comprehensive Heterocyclic Chemistry Ⅲ, Elsevier, Oxford, 2008, pp. 547–611.

    26. [26]

      A.M.P. Koskinen, Nine-membered rings, in: D.S. Black, J. Cossy, C.V. Stevens (Eds.), Comprehensive Heterocyclic Chemistry Ⅳ, Elsevier, Oxford, 2022, pp. 537–590.

    27. [27]

      H. Shanan-Atidi, K.H. Bar-Eli, J. Phys. Chem. Lett. 74 (1970) 961–963. doi: 10.1021/j100699a054

    28. [28]

      G. Zheng, K. Lv, H. Wang, et al., J. Am. Chem. Soc. 145 (2023) 3196–3203. doi: 10.1021/jacs.2c12963

    29. [29]

      X.J. Wang, J.L. Xin, H. Xiang, et al., Chin. Chem. Lett. 35 (2024) 109682. doi: 10.1016/j.cclet.2024.109682

    30. [30]

      L. Grauso, Y. Li, S. Scarpato, et al., Org. Lett. 22 (2020) 78–82. doi: 10.1021/acs.orglett.9b03964

  • Figure 1  Chemical structures of (±)-dauroxonanols A (1) and B (2) and the nomenclature of the unprecedented ring system.

    Figure 2  1H (400 MHz, A) and 13C (100 MHz, B) NMR spectra of 1 at 298 K with broad signals. Key 1H–1H COSY and HMBC correlations of 1 (C). ORTEP drawing of the crystal structure of 1 with the ellipsoid contour at 15% probability level (D). 1H–1H COSY, HMBC, and NOESY correlations of 2 (E).

    Figure 3  DFT calculated free energy profiles for the pathways of the conformational changes in 1 and 2 at the M06-2X/def2-TZVP level in chloroform at 298 K. The key states along the pathways were obtained by full optimization based on the potential energy surface scan in Fig. S10.

    Figure 4  Temperature-dependent 1H NMR (600 MHz) spectra of 1 in pyridine-d5 (298–373 K, A) and in chloroform-d (213–293 K, B), respectively. Broadening or missing NMR signals in 1 in 298 K (C). The experimental (213 K) and calculated 1H (D) and 13C (E) NMR chemical shifts differences between conformers SCB1 and TCB1.

    Table 1.  Representative activation parameters of 1.

    NucleiΔν (Hz)Tc (K)kc (s‒1)$\Delta G_{\mathrm{c}}{ }^{\ddagger}$ (kcal/mol)
    H-4β325.7293723.613.3
    H-6143.8293319.513.8
    H-85.4626312.114.0
    H-15164.8293366.113.7
    H-19a179.6293399.113.7
    H-19b208.0293462.113.5
    H-2016.627336.814.0
    H-2127.127360.313.7
    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  50
  • HTML全文浏览量:  8
文章相关
  • 发布日期:  2025-09-15
  • 收稿日期:  2025-01-14
  • 接受日期:  2025-04-17
  • 修回日期:  2025-04-11
  • 网络出版日期:  2025-04-18
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章