Hong-Xin Liuab,
Kai Chenc,
Gui-Hua Tangd,
Yun-Fei Yuana,
Hai-Bo Tan*a and
Sheng-Xiang Qiu*a
aProgram for Natural Product Chemical Biology, Key Laboratory of Plant Resources Conservation and Sustainable Utilization, South China Botanical Garden, Chinese Academy of Sciences, Guangzhou 510650, China. E-mail: sxqiu@scbg.ac.cn; tanhaibo@scbg.ac.cn; Fax: +86-20-37081190; Tel: +86-20-37081190
bState Key Laboratory of Applied Microbiology Southern China, Guangdong Provincial Key Laboratory of Microbial Culture Collection and Application, Guangdong Open Laboratory of Applied Microbiology, Guangdong Institute of Microbiology, Guangzhou 510070, China
cSchool of Chemistry and Chemical Engineering, South China University of Technology, Guangzhou 510640, China
dSchool of Pharmaceutical Sciences, Sun Yat-sen University, Guangzhou 510006, China
First published on 11th May 2016
Tomentodiones A (1) and B (2), a pair of C-11′ epimers of caryophyllene-conjugated phloroglucinols with an unprecedented skeleton, were isolated from the leaves of Rhodomyrtus tomentosa. Their structures were elucidated through the application of extensive spectroscopic measurements with the absolute configuration of 1 determined by single-crystal X-ray diffraction analysis and electronic circular dichroism (ECD) calculations. The biogenetic pathways of 1 and 2 were proposed to involve an intermolecular, inverse electron demand Diels–Alder cycloaddition reaction as the key step, and their biomimetic total synthesis was accomplished.
| No. | 1 | 2 | ||
|---|---|---|---|---|
| δH | δC | δH | δC | |
| 1 | 2.05 (m) | 52.1, CH | 1.48 (m) | 58.1, CH |
| 2α | 1.75 (m) | 33.0, CH2 | 1.59 (m) | 24.1, CH2 |
| 2β | 1.45 (m) | 1.37 (m) | ||
| 3α | 1.68 (m) | 36.3, CH2 | 1.58 (m) | 36.1, CH2 |
| 3β | 1.65 (m) | 1.73 (m) | ||
| 4 | 83.0, C | 86.3, C | ||
| 5 | 2.10 (m) | 40.3, CH | 1.75 (m) | 39.8, CH |
| 6α | 2.09 (m) | 37.4, CH2 | 2.04 (m) | 46.2, CH2 |
| 6β | 1.90 (m) | 1.38 (m) | ||
| 7α | 2.48 (m) | 35.3, CH2 | 2.13 (m) | 35.3, CH2 |
| 7β | 2.10 (m) | 2.34 (m) | ||
| 8 | 152.2, C | 150.8, C | ||
| 9 | 2.40 (m) | 42.4, CH | 2.41 (m) | 42.3, CH |
| 10a | 1.79 (m) | 22.1, CH2 | 1.66 (m) | 25.2, CH2 |
| 10b | 1.40 (m) | 1.66 (m) | ||
| 11 | 33.5, C | 34.4, C | ||
| 12 | 0.99 (s) | 30.3, CH3 | 0.96 (s) | 29.7, CH3 |
| 13 | 0.98 (s) | 22.2, CH3 | 0.97 (s) | 21.8, CH3 |
| 14 | 1.06 (s) | 20.9, CH3 | 1.35 (s) | 25.4, CH3 |
| 15a | 4.91 (s) | 110.7, CH2 | 4.91 (s) | 111.0, CH2 |
| 15b | 4.88 (s) | 4.90 (s) | ||
| 1′ | 197.5, C | 197.2, C | ||
| 2′ | 55.2, C | 55.4, C | ||
| 3′ | 213.9, C | 213.9, C | ||
| 4′ | 47.6, C | 47.6, C | ||
| 5′ | 170.3, C | 170.3, C | ||
| 6′ | 113.6, C | 116.4, C | ||
| 7′ | 1.35 (s) | 26.0, CH3 | 1.32 (s) | 25.4, CH3 |
| 8′ | 1.33 (s) | 23.3, CH3 | 1.33 (s) | 23.3, CH3 |
| 9′ | 1.39 (s) | 24.0, CH3 | 1.33 (s) | 24.9, CH3 |
| 10′ | 1.32 (s) | 25.4, CH3 | 1.35 (s) | 25.4, CH3 |
| 11′ | 2.28 (m) | 34.5, CH | 2.88 (m) | 28.3, CH |
| 12′a | 1.82 (m) | 38.8, CH2 | 1.39 (m) | 40.9, CH2 |
| 12′b | 1.33 (m) | 1.10 (m) | ||
| 13′ | 1.44 (m) | 25.4, CH | 1.43 (m) | 26.6, CH |
| 14′ | 0.84 (d, 6.6) | 24.0, CH3 | 0.82 (d, 6.6) | 24.6, CH3 |
| 15′ | 0.75 (d, 6.6) | 24.0, CH3 | 0.97 (d, 6.6) | 21.6, CH3 |
A COSY experiment revealed the presence of two fragments (Fig. 2): a (C-14′/C-15′/C-13′/C-12′/C-11′/C-5/C-6/C-7), and b (C-3/C-2/C-1/C-9/C-10), indicating the presence of an isobutyl group. The HMBC correlations between the methine proton signal at δH 2.28 with C-6′ (113.6), C-5′, C-4 (δC 83.0), and C-5 (δC 40.3) suggested that the isobutyl group was attached to C-11′ of the syncarpic acid moiety. In addition, based on fragments a and b, the HMBC correlations between the two methylene protons at δH 4.91/4.88 and C-7 (δC 35.3), C-8 (δC 152.2), and C-9 indicated that the two fragments are connected through the terminal double bond. Furthermore, the HMBC correlations of the methine at δH 2.40 with C-1 (δC 52.1), C-2 (δC 33.0), C-7 (δC 35.3), and C-11 (δC 33.5), and based on the correlations between the gem-dimethyl group signals at δH 0.99/0.98 with the 13C signals at C-1, C-10, and C-11, suggested that 1 contained a caryophyllene moiety considering an earlier report.15 The only uncertainty for the plane structure of 1 was one remaining indices of hydrogen deficiency. Based on the molecular formula, the remaining oxygen atom was assigned to bridge between C-5′ and C-4, which resulted in the formation of a dihydropyran ring thereby completely accounting for the hydrogen deficiency requirements. In light of the aforementioned evidence, the plane structure of 1 was concluded to be as shown.
The relative configuration of 1 was determined through interpretation of the proton coupling constants and NOESY correlations (Fig. 3). The relative stereochemistry of C-1 and C-9 in the caryophyllene unit was deduced to be identical to that of natural β-caryophyllene. This conclusion was drawn from the informative NOE correlations observed between H-9/H-2α, which suggested that the two protons were cofacial, and were arbitrarily assigned as having an α-orientation. Whereas the observed NOE interactions between H-5/H-1 and H-5/H-2β indicated that H-5 was β-oriented. Finally, the correlation between H3-14/H-11′ revealed that the isobutyl group at C-11′ was also β-oriented.
Conclusive evidence confirming the stereochemistry was obtained from a single-crystal X-ray diffraction experiment. After many attempts with different solvent combinations, a single-crystal of compound 1 suitable for X-ray crystallography was obtained. The X-ray crystallographic measurements were completed using CuKα with a Flack parameter of 0.2 (3) that clarified the molecular structure of 1 (Fig. 4). In addition, the X-ray diffraction experiment also provided some conformational information regarding the ring conjunctions within 1. With respect to the caryophyllene moiety, the dimethylcyclobutane ring was trans-fused with the seven-membered ring, whereas the isobutyl group had the same orientation as H-1 in the dimethylcyclobutane ring.
Compound 2 was obtained as a colorless oil. The molecular formula was assigned as C30H46O3 based on the positive ion mode HRESIMS ([M + H]+ m/z 455.3530, calcd 455.3525), indicating the presence of eight indices of hydrogen deficiency. A close comparison of its 1H and 13C NMR spectra (Table 1) with those of 1 suggested that they share the same plane structure (Table 1), which was re-confirmed by the HMBC correlations, such as those from H3-12 to C-1, C-10, C-11; H3-13 to C-1, C-10, C-11; H3-14 to C-3, C-4, C-5; H-15 to C-7, C-8, C-9; and H-11′ to C-4, C-5′. However, differences were detected for the chemical shifts between 1 and 2, particularly those associated with the chiral centers at C-1, C-5, and C-11 [C-1 (δC 52.1, δH 2.05), C-5 (δC 40.3, δH 2.10), and C-11′ (δC 34.5, δH 2.28) for 1], [C-1 (δC 58.1, δH 1.48), C-5 (δC 39.8, δH 1.75), and C-11′ (δC 28.3, δH 2.88) for 2], possibly caused by the strong spatial steric interactions due to a different orientation of the bulky functionality. In addition, all of the proton signals close to C-11′ were shifted to higher field, implying that compounds 1 and 2 were epimers in view of the ‘shielding effect’ of the isobutyl group at C-11′. The relative configuration of 2 was further confirmed by the NOESY spectrum (Fig. 3), wherein the NOE correlations of H-1/H-5, H-5/H-2β, and H-2α/H-9 definitively established the relative configuration of the sesquiterpenoid moiety in 2 as being identical to that in 1. The observed correlations of H-11′/H-5, H-12′/H3-14 indicated that the isobutyl moiety was α-oriented. Therefore, compound 2 was an epimer at C-11′ of 1 and was given the trivial name of tomentodione B.
Considering that tomentodiones A (1) and B (2) are the first representatives of this natural product class with an unprecedented molecular framework that bear a stereogenic center at C-11′, it is not feasible to assign their absolute configuration by empirical comparison of the electronic circular dichroism (ECD) spectrum with that of any structurally similar, configurationally established analogue. In such circumstance, quantum chemical calculations to predict the electronic circular dichroism (ECD) appeared to be the alternative approach. ECD calculations were performed according to standard protocols with slight modifications (see ESI† for details). In summary, conformational analysis was carried out using MacroModel 2010 to locate the conformers with low energies. All of the conformations within 10 kcal mol−1 of the lowest were optimized at the B3LYP/6-31G(d) level of theory.16,17 For those within 3.0 kcal mol−1 relative to the minimum after DFT optimization, the electronic transitions and rotational strengths were calculated at the same theoretical level through time-dependent, density-functional theory (TDDFT) calculations in Gaussian 09. The solvent effect in methanol was considered using the SMD solvent model.18 The Boltzmann-population-weighted calculated ECD curves were generated by SpecDis software.19 Subsequently, the calculated ECD spectra of 1 and 2 were compared with the experimental results, as depicted in Fig. 5, which revealed a good agreement between the experimental and calculated ECD curves. As a consequence, the absolute configurations of C-11′ in 1 and 2 were assigned as S and R, respectively.
![]() | ||
| Fig. 5 Experimental and calculated ECD spectra of 1 and 2 in methanol (1_exp and 2_exp refer to the experimental from results of 1 and 2, and 1_calc and 2_calc means the results calculation). | ||
Tomentodiones A (1) and B (2) are two novel meroterpenoids biogenetically derived through the conjugation between a β-caryophyllene moiety and a β-triketone, resulting in the formation of a new, unique skeleton. To account for this biotransformation, a plausible biogenetic route was proposed involving an intermolecular, inverse electron demand, hetero-Diels–Alder cycloaddition reaction as the key step (Scheme 1). Briefly, leptospermone, a metabolite frequently isolated from Myrtaceous plants, is a postulated precursor.20 Selective reduction and dehydroxylation would transform it to the key intermediate 5, which can undergo a Diels–Alder [4 + 2] cycloaddition reaction with β-caryophyllene to afford both 1 and 2, respectively, due to different regioselectivity options.
Based on this speculation and analysis of the associated biogenetic transformation, the biomimetic total synthesis of tomentodiones A (1) and B (2) was conducted (Scheme 2). With the acetylphloroglucinol 3 as the starting material, the syncarpic acid 4 was readily accessed in 67% yield through C-methylation and retro-Friedel–Crafts acylation.21 The proline-catalyzed Knoevenagel condensation further transformed 4 to the key intermediate 5 in almost quantitative yield, and was successfully characterized by NMR. As expected, the biomimetic Diels–Alder cycloaddition occurred smoothly by treatment of 5 and (−)-β-caryophyllene neat or in refluxing toluene to achieve the total synthesis of tomentodiones A (1) and B (2) in excellent efficiency (62% or 57% yield). Gratifyingly, 1 and 2 were identical spectroscopically and chromatographically with the isolated natural products. Moreover, their physical data including melting points and optical rotation data were also in good agreement with those recorded for naturally occurring ones. The aforementioned results further confirmed their absolute configurations and validated the quantum chemical calculations. The effectiveness of this reaction, in spite of the required conditions, may also provide positive evidence that a similar transformation could explain the biosynthesis of these metabolites.
Compounds 1 and 2 were evaluated for their antimicrobial activity against the bacteria Escherichia coli and Staphylococcus aureus. They were devoid of significant activity, but exhibited marginal inhibitory potency at a concentration of 100 μM.
:
0 → 0
:
100, v/v) to afford seven fractions (A–G). Fraction B (30 g) (n-hexane/ethyl acetate, 95
:
5, v/v) was further chromatographed over RP-C18 silica gel chromatography (MeOH/H2O, 95
:
5 → 100
:
0, v/v) to obtain sub-fractions B1 and B2. Sub-fractions B1 (2.5 g) was separated by a Sephadex LH-20 (MeOH) chromatography, followed by repeated silica gel chromatography (n-hexane/ethyl acetate, 50
:
1 → 10
:
1, v/v) to afford compound 1 (250 mg) and 2 (280 mg).
ε) 268.9 (3.07) nm; IR (KBr) νmax 3675, 2953, 1716, 1651, 1464, 1382, 889 cm−1; 1H- (500 MHz) and 13C- (125 MHz) NMR data were summarized in Table 1; ESIMS: positive ion mode [M + H]+, m/z 455; HRESIMS: positive ion mode [M + H]+, m/z 455.3459 (calcd for C30H47O3, 455.3525).
ε) 267.8 (3.74) nm; IR (KBr) νmax 3675, 2953, 1715, 1617, 1492, 1382, 890 cm−1; 1H- (500 MHz) and 13C- (125 MHz) NMR data were summarized in Table 1; ESIMS: positive ion mode [M + H]+, m/z 455; HRESIMS: positive ion mode [M + H]+, m/z 455.3530 (calcd for C30H47O3, 455.3525).
027 reflections in −11 ≤ h ≤ 11, −12 ≤ k ≤ 10, −31 ≤ l ≤ 31, measured in the range 6.72° ≤ θ ≤ 133.84°, completeness θmax = 66.92%, GOOF = 1.024, final R indices [I > 2σ(I)]: R1 = 0.0461, wR2 = 0.1177, final R indices (all data): R1 = 0.0514, wR2 = 0.1232, Flack parameter 0.2(3), largest difference peak and hole = 0.24 and −0.23 e Å−3. Data were collected on an Agilent Xcalibur Nova single-crystal diffractometer using Cu Kα radiation. The full-matrix least-squares calculation was used to refine the crystal structure. The reported crystallographic data (CCDC 1431823) in this paper for the structure of tomentodione A (1) has been deposited in the Cambridge Crystallographic Data Centre.†
:
1 to 10
:
1) to afford 1 (112 mg, 26%), as well as a mixture of 2 and by-products (270 mg) as a colorless solid. The mixture was further purified using a 10 cm C18 reversed-phase column with 100% MeOH as eluent to give 2 (133 mg, 31%). Both products were identical in all aspects to the natural isolates. 1: colorless crystals; mp 81–82 °C; [α]20D +38 (c 0.1, MeOH); 2: colorless oil; [α]20D +19 (c 0.2, MeOH). 1H- (500 MHz) and 13C- (125 MHz) NMR spectral data were identical to those of the natural compounds.
:
1 to 10
:
1) to yield 1 (127 mg, 28%), as well as a mixture of 2 and by-products (300 mg) as a colorless solid. The mixture was further purified using a 10 cm C18 reversed-phase column with 100% MeOH as eluent to afford 2 (146 mg, 34%). The products were identical in all aspects to the natural isolates.
Footnote |
| † Electronic supplementary information (ESI) available. CCDC 1431823. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6ra08776k |
| This journal is © The Royal Society of Chemistry 2016 |