Next Article in Journal
Glycosaminoglycans and Their Mimetics
Next Article in Special Issue
A Direct Method for β-Selective Glycosylation with an N-Acetylglucosamine Donor Armed by a 4-O-TBDMS Protecting Group
Previous Article in Journal
The Potential of Cyclodextrins as Novel Active Pharmaceutical Ingredients: A Short Overview
Previous Article in Special Issue
Diphenylcarbene Protected by Four ortho-Iodine Groups: An Unusually Persistent Triplet Carbene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Modeling of the Bioactivation of an Organic Nitrate by a Thiol to Form a Thionitrate Intermediate

1
Department of Chemistry, School of Science, Tokyo Institute of Technology, 2-12-1 Ookayama, Meguro-ku, Tokyo 152-8551, Japan
2
Department of Chemistry, Graduate School of Science, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan
*
Author to whom correspondence should be addressed.
Molecules 2017, 22(1), 19; https://doi.org/10.3390/molecules22010019
Submission received: 22 November 2016 / Revised: 20 December 2016 / Accepted: 23 December 2016 / Published: 25 December 2016
(This article belongs to the Special Issue Cutting-Edge Organic Chemistry in Japan)

Abstract

:
Thionitrates (R–SNO2) have been proposed as key intermediates in the biotransformation of organic nitrates that have been used for the clinical treatment of angina pectoris for over 100 years. It has been proposed and widely accepted that a thiol would react with an organic nitrate to afford a thionitrate intermediate. However, there has been no example of an experimental demonstration of this elementary chemical process in organic systems. Herein, we report that aryl- and primary-alkyl-substituted thionitrates were successfully synthesized by the reaction of the corresponding lithium thiolates with organic nitrates by taking advantage of cavity-shaped substituents. The structure of a primary-alkyl-substituted thionitrate was unambiguously established by X-ray crystallographic analysis.

1. Introduction

Organic nitrates such as nitroglycerin and isosorbide-5-mononitrate, have been widely used for the treatment of angina pectoris and congestive heart failure since the first clinical application of nitroglycerin in the late 19th century. They are regarded as nitric oxide (NO) prodrugs, and it has been assumed that they are converted to NO or an NO congener, which leads to vasorelaxation [1,2,3,4]. Although the mechanism of biotransformation of organic nitrates is still not fully understood, the potential chemical mechanism is a thiol-dependent pathway, for which several reaction pathways have been proposed as shown in Scheme 1 [5,6]. Importantly, all of them have the same first step in common, i.e., the formation of a thionitrate by the reaction of a cysteine thiol (Cys-SH) in protein with an organic nitrate. However, very little chemical evidence has been available for this elementary chemical transformation [7,8]. Recently, a thionitrate intermediate, which was formed in the active site of aldehyde dehydrogenase-2 by soaking with nitroglycerin, has been observed by protein X-ray crystallographic analysis and also detected by mass spectrometry [9]. In the artificial systems, thionitrates are highly reactive species and easily undergo bimolecular decomposition (Scheme 2) [10,11,12]. Hence, model studies on the chemical processes involving R–SNO2 have been difficult and scarcely investigated [12,13,14].
In the course of our studies on biologically relevant highly reactive species containing sulfur and selenium, we have developed various cavity-shaped substituents and demonstrated that these substituents are very effective for kinetic stabilization of highly reactive species that otherwise undergo facile bimolecular decomposition [13,14,15,16,17,18,19,20,21,22]. Herein, we report a model study on the bioactivation of an organic nitrate by a thiol, which is the fundamental chemical process in the thiol-dependent biotransformation of nitrates. By taking advantage of the cavity-shaped aromatic and primary-alkyl substituents (Figure 1), denoted as Bpq and BpqCH2, respectively, we demonstrated that thionitrates are formed by the reaction of the corresponding thiolates with organic nitrates. The first crystallographic analysis of a primary-alkyl-substituted thionitrate is also presented.

2. Results and Discussion

Several stable thionitrates have been synthesized and isolated by taking advantage of kinetic stabilization utilizing appropriate bulky substituents such as tertiary alkyl groups [10,11,16] and cavity-shaped aromatic groups [17,23]. We previously reported the synthesis of a stable aryl thionitrate, BpqSNO2 (1), by the reaction of the corresponding S-nitrosothiol with an excess amount of t-BuONO or N2O4 [17]. Thus, we first employed the Bpq group for a model study on the bioactivation of organic nitrates. Treatment of thiol 2 bearing a Bpq group with an excess amount of isoamyl nitrate in C6D6 resulted in no reaction (Scheme 3). To increase the nucleophilicity of the sulfur, 2 was deprotonated with n-BuLi to generate lithium thiolate 3, which was then allowed to react with an excess amount of isoamyl nitrate in benzene. This reaction sequence successfully afforded thionitrate 1 (64%) together with thiol 2 (36%). This is the first experimental demonstration of the formation of a thionitrate by the reaction of a thiolate with an organic nitrate.
As a model compound for naturally occurring cysteine-derived thionitrates (Cys-SNO2), a primary-alkyl-substituted thionitrate is considered to be the most relevant. However, the steric demands of usual primary-alkyl groups are too small to protect such reactive species, and there has been no example of the synthesis of a thionitrate bearing a primary-alkyl group. Meanwhile, we have recently developed an effective primary-alkyl steric protection group, a BpqCH2 group (Figure 1), with a cavity-shaped framework and succeeded in the isolation of reactive species such as a sulfenic acid [18], a sulfenyl iodide [19], a selenenic acid [20], and a selenenyl iodide [21] by utilizing this substituent. Thus, a model study by utilizing the BpqCH2 group was examined. When lithium thiolate 5, prepared by the reaction of BpqCH2SH (4) with n-BuLi, was treated with an excess amount of isoamyl nitrate in benzene, the corresponding thionitrate 6 was formed almost quantitatively and isolated as stable colorless crystals in 74% yield after recrystallization (Scheme 4). Characterization of 6 was performed by NMR and IR spectroscopies and elemental analysis. In the IR spectrum of 6, asymmetric and symmetric NO2 vibrational bands were observed at 1531 and 1297 cm−1, respectively, which are almost identical to those of tertiary-alkyl-substituted thionitrates [16]. The 1H-NMR spectrum of 6 exhibited a singlet due to the methylene protons in the CH2SNO2 moiety at 5.10 ppm, which is shifted downfield by more than 1 ppm relative to that of thiol 4 (3.91 ppm). In the 13C-NMR spectrum of 6, the methylene carbon resonates at 41.3 ppm, which is also shifted downfield in comparison with 4 (30.5 ppm). The reaction of thiolate 5 with isosorbide-5-mononitrate (ISMN), which is widely used as a drug in treatment of angina pectoris, also afforded thionitrate 6 (Scheme 5), albeit in low yield (4%). Since thiol 4 was recovered in 95% yield in this reaction, it is probable that most of 5 was protonated and deactivated by the OH group in ISMN, resulting in the low yield of 6.
The structure of 6 was unambiguously established by X-ray crystallographic analysis (Figure 2), presenting the first example of the crystallographic analysis of a primary-alkyl-substituted thionitrate. The CH2SNO2 moiety is incorporated within the cavity and effectively protected from the bimolecular decomposition by the peripheral moiety of the substituent. Selected structural parameters of 6 are summarized in Table 1, together with those of the aryl-substituted thionitrates 1 [14] and the tertiary-alkyl-substituted thionitrate 7 [16] reported by us as well as those of the aryl derivative 8 [24] reported by Itoh et al. In the crystalline state, 6 features disorder of the SNO2 moiety with respect to a mirror plane perpendicular to the central benzene ring in the ratio of 0.8:0.2, both of which possess a similar structure. The structural parameters of the major component are discussed in the following. The CH2SNO2 moiety is well incorporated within the cavity of the Bpq group. The C–S–N–O dihedral angles are 4.8(5)° and –176.0(4)°, showing that the C–SNO2 moiety is almost coplanar. The C–S bond length (1.803(3) Å) in 6 is slightly longer than those in the aromatic derivatives 1 (1.7651(15) Å) and 8 (1.764(2) Å) and similar to that of the tertiary-alkyl derivative 7 (1.789(6) Å). The C–S–N bond angles of 6 (100.5(2)°) are similar to those of 1 (100.47(7)°) and 8 (99.75(10)°) yet much smaller than that of 7 (107.9(5)°). The enlargement of the C–S–N bond angle was observed only in 7, which is probably due to the steric repulsion between the SNO2 moiety and m-terphenyl groups. The structure of the primary-alkyl-substituted compound 6 shown here is expected to serve as reference data for X-ray crystallographic analysis of protein–SNO2.

3. Materials and Methods

Unless otherwise stated, all operations were performed in a MBraun UNIlab glovebox an argon atmosphere, in a Miwa 1ADB-3KTG glovebox under a nitrogen atmosphere, or by using high-vacuum and standard Schlenk techniques under an argon atmosphere. Benzene (anhydrous) was purchased from Kanto Chemical (Tokyo, Japan) and distilled from benzophenone ketyl prior to use. Other chemicals were purchased from commercial sources and used as received. 1H-NMR spectra were recorded on a JEOL ECX-500, a JEOL ECX-400, or a JEOL ECS-400 spectrometer (JEOL, Tokyo, Japan), and the chemical shifts of 1H are referenced to the residual proton signal of CDCl3 (δ 7.25). 13C-NMR spectra were recorded on a JEOL ECX-500 or a JEOL ECX-400 spectrometer (JEOL, Tokyo, Japan), and the chemical shifts of 13C are referenced to the signal of CDCl3 (δ 77.0). All spectra were assigned with the aid of DEPT (distorsionless enhancement by polarization transfer), COSY (correlated spectroscopy), HMQC (heteronuclear multi quantum correlation), and HMBC (heteronuclear multiple bond correlation) NMR experiments. IR spectra were recorded on a JASCO FT/IR-4100 (JASCO, Tokyo, Japan) by utilizing a JASCO ATR Pro550S unit. Melting points were measured with a Yanaco MP-S3 (Yanaco, Tokyo, Japan)and are uncorrected.

3.1. The Reaction of Thiol 2 with Isoamyl Nitrate

To a solution of thiol 2 (3.5 mg, 3.9 µmol) in C6D6 (0.6 mL) was added isoamyl nitrate (30 µL, 0.23 mmol) at ambient temperature. The reaction was monitored by 1H-NMR spectroscopy in a screw-capped NMR tube, and no reaction was observed for 2 h.

3.2. Synthesis of Thionitrate 1

n-BuLi (1.6 mol/L in hexane, 0.30 mL, 0.48 mmol) was diluted with benzene (3 mL). To a solution of thiol 2 (18.3 mg, 20 µmol) in benzene (6 mL) was added the diluted n-BuLi solution (0.15 mL, 22 µmol). The reaction mixture was stirred for 30 min at ambient temperature, and then isoamyl nitrate (30 µL, 0.23 mmol) was added at ambient temperature. After 10 min, the solvent was removed in vacuo, and the residue was washed with hexane to afford the mixture of thionitrate 1 [17] and thiol 2. The yields of 1 and 2 were estimated to be 64% and 36%, respectively, by 1H-NMR spectroscopy.

3.3. Synthesis of Thionitrate 6

In a glovebox under an argon atmosphere, to a solution of thiol 4 (20.3 mg, 22.1 µmol) [18] in benzene (6.6 mL) in a Schlenk flask with a J-young valve was added n-BuLi (56 mM in hexane, 434 µL, 24 µmol) at ambient temperature. After 1 h, the flask was carefully sealed and transferred to a glovebox under a nitrogen atmosphere. To the reaction mixture was added isoamyl nitrate (27 µL, 0.20 mmol) at ambient temperature. After 30 min, the resulting yellow solution was evaporated in vacuo, followed by recrystallization from hexane to afford 6 (14.2 mg, 16 µmol, 74%) as colorless crystals.
6: colorless crystals; mp 203.5–204.4 °C (dec); 1H-NMR (400 MHz, CDCl3) δ 1.07 (d, J = 6.9 Hz, 24H), 1.13 (d, J = 6.9 Hz, 24H), 2.80 (sept, J = 6.9 Hz, 8H), 5.10 (s, 2H), 7.06 (t, J = 1.5 Hz, 2H), 7.14–7.22 (m, 12H), 7.27-7.50 (m, 7H); 13C-NMR (126 MHz, CDCl3) δ 24.0 (q), 24.1 (q), 30.6 (d), 41.3 (t), 122.5 (d), 128.0 (d), 128.4 (d), 128.6 (d), 129.8 (d), 130.5 (d), 138.5 (s), 139.8 (s), 141.1 (s), 143.9 (s), 146.67 (s), 146.72 (s); IR (ATR) υ 1531 cm−1 (asym-NO2), 1297 cm−1 (sym-NO2). Anal. Calcd for C67H79NOS: C, 83.61; H, 8.27, N, 1.46; S, 3.33. Found: C, 83.37; H, 8.38; N, 1.33; S, 3.20.

3.4. Formation of 6 by the Reaction with Isosorbide-5-Mononitrate (ISMN)

In a glovebox under an argon atmosphere, to a solution of thiol 4 (18.5 mg, 20.2 µmol) in benzene (6 mL) in a Schlenk flask with a J-young valve was added n-BuLi (56 mM in hexane, 436 µL, 24 µmol) at ambient temperature. After 1 h, the flask was carefully sealed and transferred to a glovebox under a nitrogen atmosphere. To the reaction mixture was added a solution of ISMN (4.3 mg, 22 µmol) in benzene (1 mL) at ambient temperature. After 30 min, the reaction mixture was evaporated in vacuo. The 1H-NMR analysis of the crude mixture confirmed the formation of 6 (4%) together with 4 (95%).

3.5. X-Ray Crystallography

Single crystals of 6·1.5CHCl3 were grown in their hexane-CHCl3 solution. A colorless crystal of 6·1.5CHCl3 was mounted on a loop. All measurements were made on a Rigaku/MSC Mercury CCD (Charge Coupled Device) diffractometer (Rigaku, Tokyo, Japan) with graphite-monochromated Mo-Kα radiation (λ = 0.71070 Å) at –153 °C. The structures were solved by the direct method and refined by full-matrix least squares on F2 using SHELXL 97 [25]. Crystal Data for 6·1.5CHCl3 (M = 1141.42 g/mol): triclinic, space group P-1, a = 11.6005(15) Å, b = 15.8742(14) Å, c = 18.923(2) Å, α = 99.399(3)°, β = 105.607(4)°, γ = 100.004(3)°, V = 3223.0(6) Å3, Z = 2, Dcalc = 1.176 g/cm3, 20857 reflections measured, 11040 unique (Rint = 0.0148). The final R1 was 0.0941 (I > 2σ(I)) and wR2 was 0.2475 (all data). CCDC 1518311 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html (or from the CCDC, 12 Union Road, Cambridge CB2 1EZ, UK; Fax: +44 1223 336033; E-mail: [email protected])

4. Conclusions

Formation of a thionitrate by the reaction of a thiolate with an organic nitrate was experimentally demonstrated for the first time by taking advantage of molecular cavities. The X-ray crystallographic analysis of a primary-alkyl-substituted thionitrate was also established. The present results provide chemical corroboration to the formation of a thionitrate intermediate proposed for the first step in the thiol-mediated bioactivation of an organic nitrate, which has long been suggested but not evidenced. Further investigations on the modeling of the bioactivation processes of organic nitrates are currently in progress.

Supplementary Materials

The following are available online at https://www.mdpi.com/1420-3049/22/1/19/s1.

Acknowledgments

This work was partly supported by a Grant-in-Aid for Scientific Research (B) (No. 15H03776), a Grant-in-Aid for Scientific Research on Innovative Areas “Stimuli-responsive Chemical Species for the Creation of Functional Molecules” (No. 15H00930), and a Grant-in-Aid for Scientific Research (C) (No. 26410038) from the Ministry of Education, Culture, Sports, Science, and Technology of Japan.

Author Contributions

Tsukasa Sano, Keiichi Shimada, and Yohei Aoki performed the experimental works; Takayuki Kawashima, Shohei Sase, and Kei Goto conceived and designed the project; Tsukasa Sano, Shohei Sase, and Kei Goto wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fung, H.L. Biochemical mechanism of nitroglycerin action and tolerance: Is this old mystery solved? Ann. Rev. Pharmacol. Toxicol. 2004, 44, 67–85. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, Z.Q.; Stamler, J.S. Bioactivation of nitroglycerin by the mitochondrial aldehyde dehydrogenase. Trends Cardiovasc. Med. 2006, 16, 259–265. [Google Scholar] [CrossRef] [PubMed]
  3. Mayer, B.; Beretta, M. The enigma of nitroglycerin bioactivation and nitrate tolerance: News, views and troubles. Br. J. Pharmacol. 2008, 155, 170–184. [Google Scholar] [CrossRef] [PubMed]
  4. Daiber, A.; Münzel, T. Organic nitrate therapy, nitrate tolerance, and nitrate-induced endothelial dysfunction: emphasis on redox biology and oxidative stress. Antioxid. Redox Signaling 2015, 23, 899–942. [Google Scholar] [CrossRef] [PubMed]
  5. Thatcher, R.G.J.; Weldon, H. NO problem for nitroglycerin: Organic nitrate chemistry and therapy. Chem. Soc. Rev. 1998, 27, 331–337. [Google Scholar] [CrossRef]
  6. Wenzl, M.V.; Beretta, M.; Griesberger, M.; Russwurm, M.; Koesling, D.; Schmidt, K.; Mayer, B.; Gorren, A.C.F. Site-directed mutagenesis of aldehyde dehydrogenase-2 suggests three distinct pathways of nitroglycerin biotransformation. Mol. Pharmacol. 2011, 80, 258–266. [Google Scholar] [CrossRef] [PubMed]
  7. Yeates, R.A.; Laufen, H.; Leitold, M. The reaction between organic nitrates and sulfhydryl compounds. A possible model system for the activation of organic nitrates. Mol. Pharmacol. 1985, 28, 555–559. [Google Scholar] [PubMed]
  8. Yeates, R.A. Possible mechanism of activation of soluble guanylate-cyclase by organic nitrates. Drug Res. 1992, 42, 1314–1317. [Google Scholar]
  9. Lang, B.S.; Gorren, A.C.F.; Oberdorfer, G.; Wenzl, M.V.; Furdui, C.M.; Poole, L.B.; Mayer, B.; Gruber, K. Vascular bioactivation of nitroglycerin by aldehyde dehydrogenase-2. J. Biol. Chem. 2012, 287, 38124–38134. [Google Scholar] [CrossRef] [PubMed]
  10. Oae, S.; Shinhama, K. Organic thionitrites and related substances—A review. Org. Prep. Proced. Int. 1983, 15, 165–198. [Google Scholar] [CrossRef]
  11. Oae, S.; Shinhama, K.; Fujimori, K.; Kim, Y.H. Physical-properties and various reactions of thionitrites and related substances. Bull. Chem. Soc. Jpn. 1980, 53, 775–784. [Google Scholar] [CrossRef]
  12. Artz, J.D.; Yang, K.X.; Lock, J.; Sanchez, C.; Bennett, B.M.; Thatcher, G.R.J. Reactivity of thionitrate esters: Putative intermediates in nitrovasodilator activity. Chem. Commun. 1996, 927–928. [Google Scholar] [CrossRef]
  13. Goto, K.; Shimada, K.; Furukawa, S.; Miyasaka, S.; Takahashi, Y.; Kawashima, T. Formation of a stable sulfenic acid by hydrolysis of a thionitrate and a sulfenyl bromide. Chem. Lett. 2006, 35, 862–863. [Google Scholar] [CrossRef]
  14. Goto, K.; Yoshikawa, S.; Ideue, T.; Sase, S. Transnitrosation from a stable thionitrate to an amine with concomitant formation of a sulfenic acid. J. Sulfur Chem. 2013, 34, 705–710. [Google Scholar] [CrossRef]
  15. Goto, K. Synergy of reactivity and stability in nanoscale molecular architectures. In Synergy in Supramolecular Chemistry; Nabeshima, T., Ed.; CRC Press: Boca Raton, FA, USA, 2015; pp. 191–218. [Google Scholar]
  16. Goto, K.; Hino, Y.; Kawashima, T.; Kaminaga, M.; Yano, E.; Yamamoto, G.; Takagi, N.; Nagase, S. Synthesis and crystal structure of a stable S-nitrosothiol bearing a novel steric protection group and of the corresponding S-nitrothiol. Tetrahedron Lett. 2000, 41, 8479–8483. [Google Scholar] [CrossRef]
  17. Goto, K.; Hino, Y.; Takahashi, Y.; Kawashima, T.; Yamamoto, G.; Takagi, N.; Nagase, S. Synthesis, structure, and reactions of the first stable aromatic S-nitrosothiol bearing a novel dendrimer-type steric protection group. Chem. Lett. 2001, 30, 1204–1205. [Google Scholar] [CrossRef]
  18. Ishihara, M.; Abe, N.; Sase, S.; Goto, K. Synthesis, structure, and reactivities of a stable primary-alkyl-substituted sulfenic acid. Chem. Lett. 2015, 44, 615–617. [Google Scholar] [CrossRef]
  19. Sase, S.; Aoki, Y.; Abe, N.; Goto, K. Stable sulfenyl iodide bearing a primary alkyl steric protection group with a cavity-shaped framework. Chem. Lett. 2009, 38, 1188–1189. [Google Scholar] [CrossRef]
  20. Sase, S.; Kakimoto, R.; Goto, K. Synthesis of a stable selenoaldehyde by self-catalyzed thermal dehydration of a primary-alkyl-substituted selenenic acid. Angew. Chem. Int. Ed. 2015, 54, 901–904. [Google Scholar] [CrossRef] [PubMed]
  21. Sase, S.; Kakimoto, R.; Kimura, R.; Goto, K. Synthesis of a stable primary-alkyl-substituted selenenyl iodide and its hydrolytic conversion to the corresponding selenenic acid. Molecules 2015, 20, 21415–21420. [Google Scholar] [CrossRef] [PubMed]
  22. Goto, K.; Sonoda, D.; Shimada, K.; Sase, S.; Kawashima, T. Modeling of the 5′-deiodination of thyroxine by iodothyronine deiodinase: Chemical corroboration of a selenenyl iodide intermediate. Angew. Chem. Int. Ed. 2010, 49, 545–547. [Google Scholar] [CrossRef] [PubMed]
  23. Itoh, M.; Takenaka, K.; Okazaki, R.; Takeda, N.; Tokitoh, N. The first stable aromatic S-nitrosothiol: Synthesis, structure and reactivity. Chem. Lett. 2001, 30, 1206–1207. [Google Scholar] [CrossRef]
  24. Itoh, M.; Takenaka, K.; Okazaki, R. Synthesis, structure, and reactivity of stable aromatic S-nitrothiols bearing bulky substituents. Nippon Joshi Daigaku Kiyo Rigakubu 2004, 12, 33–38. [Google Scholar]
  25. Sheldrick, G.M. SHELXL-97, Computer Program for Crystal Structure Refinement; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  • Sample Availability: Not available.
Scheme 1. Proposed thiol-dependent pathways of biotransformation of organic nitrates.
Scheme 1. Proposed thiol-dependent pathways of biotransformation of organic nitrates.
Molecules 22 00019 sch001
Scheme 2. Bimolecular decomposition of thionitrates.
Scheme 2. Bimolecular decomposition of thionitrates.
Molecules 22 00019 sch002
Figure 1. Molecular cavities utilized for the present model study.
Figure 1. Molecular cavities utilized for the present model study.
Molecules 22 00019 g001
Scheme 3. Reactions of thiol 2 with isoamyl nitrate.
Scheme 3. Reactions of thiol 2 with isoamyl nitrate.
Molecules 22 00019 sch003
Scheme 4. Synthesis of thionitrate 6 by the reaction of thiolate 5 with isoamyl nitrate.
Scheme 4. Synthesis of thionitrate 6 by the reaction of thiolate 5 with isoamyl nitrate.
Molecules 22 00019 sch004
Scheme 5. Formation of 6 by the reaction of thiolate 5 with ISMN.
Scheme 5. Formation of 6 by the reaction of thiolate 5 with ISMN.
Molecules 22 00019 sch005
Figure 2. Oak Ridge thermal ellipsoid plot (ORTEP) drawing of 6 (50% probability).
Figure 2. Oak Ridge thermal ellipsoid plot (ORTEP) drawing of 6 (50% probability).
Molecules 22 00019 g002
Table 1. Selected bond lengths and angles for thionitrates.
Table 1. Selected bond lengths and angles for thionitrates.
Molecules 22 00019 i001
6 (this work)1 [14]7 [16]8 [24]
Bond lengths (Å)
S(1)–N(1)1.780 (6) 1.7898 (17)1.746 (9)1.795 (2)
N(1)–O(1)1.205 (8)1.2180 (19)1.229 (9)1.215 (3)
N(1)–O(2)1.211 (6)1.213 (2)1.239 (9)1.228 (3)
C(1)–S(1)1.803 (3) 1.7651 (15)1.789 (6)1.764 (2)
Bond angles (°)
S(1)–N(1)–O(1)121.7 (4)120.25 (12)114.1 (9)121.19 (17)
S(1)–N(1)–O(2)114.1 (9)113.28 (13)119.7 (9)113.48 (18)
O(1)–N(1)–O(2)124.00 (57)126.44 (17)126.2 (11)125.3 (2)
C(1)–S(1)–N(1)100.5 (2)100.47 (7)107.9 (5)99.75 (10)

Share and Cite

MDPI and ACS Style

Sano, T.; Shimada, K.; Aoki, Y.; Kawashima, T.; Sase, S.; Goto, K. Modeling of the Bioactivation of an Organic Nitrate by a Thiol to Form a Thionitrate Intermediate. Molecules 2017, 22, 19. https://doi.org/10.3390/molecules22010019

AMA Style

Sano T, Shimada K, Aoki Y, Kawashima T, Sase S, Goto K. Modeling of the Bioactivation of an Organic Nitrate by a Thiol to Form a Thionitrate Intermediate. Molecules. 2017; 22(1):19. https://doi.org/10.3390/molecules22010019

Chicago/Turabian Style

Sano, Tsukasa, Keiichi Shimada, Yohei Aoki, Takayuki Kawashima, Shohei Sase, and Kei Goto. 2017. "Modeling of the Bioactivation of an Organic Nitrate by a Thiol to Form a Thionitrate Intermediate" Molecules 22, no. 1: 19. https://doi.org/10.3390/molecules22010019

Article Metrics

Back to TopTop