Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Apr 1.
Published in final edited form as: J Mol Struct. 2010 Apr 1;968(1-3):13–18. doi: 10.1016/j.molstruc.2010.01.012

X-ray, kinetics and DFT studies of photochromic substituted benzothiazolinic spiropyrans

Satish Kumar a, Kriya Velasco, Alison McCurdy b,*
PMCID: PMC2850124  NIHMSID: NIHMS170635  PMID: 20383273

Abstract

Photochromic molecules have the potential to find utility in a wide variety of applications including photoswitchable binding and optical memory. This work explores the relationship between photochromism and structural parameters such as particular bond lengths for this class of compounds for which very few crystal structures have been published. Photochemical kinetics, Density Functional Theory (DFT) and X-ray crystallography were used to study the benzothiazolinic spiropyran 3-methyl-6-nitro-3′-methylspiro-[2H-l-benzopyran-2,2′-benzothiazoline]. A second benzothiazolinic spiropyran 3-methyl-8-methoxy-6-nitro-3′-methylspiro-[2H-l-benzopyran-2,2′-benzothiazoline] was synthesized and subjected to photochemical and computational studies. Selected structural and photochemical data for these, related benzothiazolinic spirooxazines and spiropyrans, and related thiazolidinic spiropyrans are compared. Both benzothiazolinic spiropyrans exhibit photochromic properties that are influenced by substituents, solvent, and temperature. The crystallographic Cspiro-O bond distance of 3-methyl-6-nitro-3′-methylspiro-[2H-l-benzopyran-2,2′-benzothiazoline] that has been shown to correlate with photochromic properties is 1.458 Å. The crystallographic Cspiro-O bond distance matches that of the structure generated by DFT calculations exactly. The effect of substituents on calculated bond lengths and photochemical parameters was determined. For this class of compounds, both X-ray geometry and DFT optimized geometry may be used to predict photochromism, but not degree of photocolorability.

Keywords: photochromism, X-ray, benzothiazolinic spiropyran

1. Introduction

Photochromic spiropyrans have attracted considerable attention recently owing to their diverse applications, such as in optical memory devices, light filters, drug delivery systems, and biological calcium signal mimics [16]. These spiropyrans isomerize reversibly between two structures (closed, colorless and open, colored) in response to light as shown in Figure 1 for 3-methyl-6-nitro-3′-methylspiro-[2H-l-benzopyran-2,2′-benzothiazoline] (1). In many cases, thermal (dark) processes compete with the photochemical conversion of the less stable to the more stable form. Photochromism of spiropyran compounds has been reviewed extensively by Bertelson [78]. The Cspiro-O bond length and substituents have been correlated with photochromic properties of spiropyrans and spirooxazines [913]. Specifically, for spiropyrans and spirooxazines, such as those shown in Figure 2, the Cspiro-O bond lengths are elongated by 0.03 – 0.06 Å from the normal Csp3-O bond length in 6-membered heterocycles of 1.41 – 1.43 Å. The corresponding Cspiro-N1 bond lengths are found to be shorter than those found in 5-membered heterocycles of 1.47–1.48 Å. For spirooxazines, better photocolorability correlates with longer Cspiro-O bond lengths. These correlations are thought to reflect the degree of interaction between the lone pair on one heteroatom bound to the spiro carbon with the σ* antibonding orbital of the polar Cspiro-O bond that breaks during photochemical ring opening [12].

Figure 1.

Figure 1

Photochromism of the benzothiazolinic spiropyran 1; the colorless closed form 1 is interconverted with the colored open form.

Figure 2.

Figure 2

Representative spiropyran compounds with correlations between photochromism and Cspiro-O bond length. Left: spiroindolinopyran; Right: spirooxazine.

Our interest in this subject derives from the fact that suitably functionalized spiropyran compounds have potential for application in mimicking calcium oscillations in biological systems [1415]. While spirobisbenzopyran compounds do not have appropriate photophysical properties at room temperature, benzothiazolinic spiropyran compounds offer a significant improvement. In order to understand the photochromic properties of benzothiazolinic spiropyrans, we have studied some photochemical properties and determined the crystal structure of compound 1. The related compound 2 was also synthesized and characterized. The synthesis of 1 and 2 were first reported by Guglielmetti et al. [1617]. However, no crystal structure was reported. In fact, no crystal structure for a closed form of a benzothiazolinic spiropyran has been reported in the literature. The structure of the open merocyanine form of a benzothiazolinic spiropyran was reported by Miler-Srenger et al.[18]. In this work, the crystal structure of 1 (Figure 5) is compared with the equilibrium geometry obtained using density functional calculations shown in Figure S4 (B3LYP/6-31G*) as well as certain structural reported and calculated features of related photochromic and nonphotochromic spiropyran scaffolds shown below.

Figure 5.

Figure 5

Ortep view [22] of 3-methyl-6-nitro-3′-methylspiro-[2H-l-benzopyran-2,2′-benzothiazoline] with displacement ellipsoids at the 50% probability level.

2. Experimental and Computational

The compounds 2, 7 and 8 were synthesized according to the procedure reported in Guglielmetti et al. [19]. The procedure was adapted to synthesize compound 1 as described below and in Scheme 1.

Scheme 1.

Scheme 1

Compound 1 (60% yield). A mixture of 8 (0.28 g, 0.80 mmol) and of 5-nitrosalicylaldehyde (0.14 g, 0.83 mmol) were added to a 50 mL round bottom flask. The reaction flask was evacuated and filled with N2 (gas). Ethanol (15 mL) and piperidine (0.1 mL) were added to the reaction flask. The reaction mixture was refluxed for 1 hour. The solvent was evaporated and impure product was purified using flash column chromatography (10% Ethyl acetate in hexane) to give 1 (0.12 g, 60%) as a pure product. 1H NMR (CDCl3, 400MHz, δ): 8.05 (dd, J=8.88 Hz, 1H), 7.98 (d, J=2.24 Hz, 1H), 7.20–7.14 (m, 2H), 6.91–6.84 (m, 2H), 6.84–6.64 (m, 2H), 3.04 (s, 3H), 2.10 (s, 3H); 13C NMR (CDCl3, 400MHz, δ): 155.8, 144.6, 142.4, 131.6, 125.9, 125.6, 125.0, 122.1, 121.7, 121.6, 121.0, 120.5, 117.1, 114.8, 108.0, 30.5, 19.62. HRMS (m/z): Calcd,: 326.0725; found: 326.0725. Crystals suitable for single crystal X-ray determinations were grown by keeping a solution of 1 in THF for several weeks in the dark.

Compound 2 (51% yield). 1H NMR (CDCl3, 400MHz, δ): 7.68 (d, J = 3.2 Hz, 1H), 7.63 (d, J = 3.6 Hz, 1H), 7.154 (m, 2H), 6.86 (t, J = 10 Hz, 1H), 6.66 (d, J = 10 Hz, 1H), 6.59 (d, J =1.60 Hz, 1H), 3.81 (s, 3H), 3.07 (s, 3H), 2.07 (d, J= 2.0 Hz, 3H); 13C NMR (CDCl3, 400MHz, δ): 148.7, 145.4, 144.7, 142.0, 131.7, 125.8, 122.2, 121.5, 121.0, 119.8, 114.9, 114.3, 107.8, 107.6, 56.5, 30.4, 19.4..

Photochemistry

A 30 mL solution of 1 (3.68 × 10−5 M) or 2 (6.17 × 10−5 M) was prepared in the appropriate solvent in a 3-neck flask containing a stir bar. The flask was cooled to −30 °C in a thermostatted cryobath. Light from a 200W Mercury Xenon lamp, filtered by water and an Oriel #59800, 330nm bandpass filter, was directed via a 3 foot fiber optic bundle into the top of a 3-neck flask while the solution was stirred. UV-Vis observations were made using a Hellma 10mm immersion probe connected to a Varian Cary 50 spectrophotometer. Rates are reported as averages of three runs, except for 1 in isopropanol and acetonitrile, in which the reaction was performed once and the rate was averaged from five different wavelengths.

Computation

The program Spartan ‘04 was used. AM1 was used to generate the initial geometry for 1. The compound 1 was then subjected to geometry optimization (B3LYP/6-31G*), and the resulting geometry was used to determine single point energy. Cartesian coordinates and energies are listed in Table S6. For compound 2, AM1 was used to generate conformers. These conformers were then subjected to geometry optimization (B3LYP/6-31G*), and the resulting geometries were used to determine single point energies. Cartesian coordinates and energies are listed in Tables S7–S9.

Crystal Data

C17H14N2O3S Monoclinic, P21/n
Mr 326.36 Z = 4
MoKα radiation T = 100 (2) K
a = 9.1899 (4) Å V = 1522.24(11) Å3
b = 11.8526 (5) Å 0.29 × 0.28 × 0.28 mm3
c = 13.9833 (6) Å μ = 0.229 mm−1
β= 91.9440 (10)° Density = 1.424 Mg/m3

Data collection

Bruker SMART 1000 25987 measured reflections
Absorption correction: None 6105 Independent reflections
Tmax = 0.9386 I > 2σ(I), 4660 reflections
Tmin = 0.9365 Rint = 0.0777

Refinement

R1 (w=1/σ2(Fo2)) = 0.0443 S = 1.689
wR2 = 0.0872 Δρmax=0.621eÅ−3
Parameters 264 Δρmin= −0.356 e. Å−3
Restraint 0 H atom unrestrained

Data collection

Bruker SMART v5.630; cell refinement: Bruker SAINT v6.45A; data reduction: Bruker SAINT v6.45A; program(s) used to solve structure: SHELXS97 [20]; program(s) used to refine structure: SHELXL97 [21]; molecular graphics: Bruker XP v5.1; Ortep3 [22]; Mercury [23]; software used to prepare material for publication: Bruker XCIF v6.12.

Crystallographic data have been deposited at the CCDC, 12 Union Road, Cambridge CB2 1EZ, UK and copies can be obtained on request, free of charge, by quoting the publication citation and the deposition number 632419.

3. Results and Discussion

3.1 Synthesis and Characterization of 1 and 2

As detailed in the experimental section, the known targets 1 and 2 were prepared from a quaternized ethylbenzothiazole. Addition of piperidine generated the enamine base, which was condensed with the appropriately substituted salicylaldehyde to make the spiropyran. The reported 1H spectrum (for 2) and13C spectra (for 1 and 2) are consistent with those determined in this work [24].

3.2 Photochemical studies

At several temperatures, solutions of 1 and 2 in a variety of solvents were irradiated with UV light, and subsequently were allowed to return to their original colorless form in the dark. Compound 1 was found to exhibit weak photochromism at room temperature only in polar solvents (isopropanol, but not in toluene or THF). Photochromism in 1 was observed in several solvents (THF, acetonitrile, isopropanol) tested at lower temperatures to slow thermal closure rates (Table 1). Closure rates had been reported earlier only for photoirradiation in toluene at 25°C. After irradiation at low temperature, the solution is colored, as shown in Figure 4, and then thermally bleaches to a colorless solution over time. Thermal fading rates are slower in more polar solvents, due to solvent stabilization of the polar zwitterionic open forms. The addition of an electron-donating group (2) has the effect of increasing the rate of thermal closure, as the nucleophilicity of the phenolic oxygen is enhanced, or as the stability of the zwitterionic form is decreased.

Table 1.

First order rate constants of thermal bleaching at −30°

Solvent 2-propanol CH3CN THF Toluene
kT min−1 (1) 5.28 ±0.29 × 10−3 6.00 ±0.81 × 10−3 7.40 ±0.54 × 10−3 Insoluble photoproduct
kT min−1 (2) Negligible Negligible 1.11 ±0.19 × 10−2 4.14 ±1.33 × 10−2

Figure 4.

Figure 4

Absorption spectra of an irradiated solution of spiropyran 1 in THF (3.68 × 10−5 M) at −30°C as it thermally (without light) reverts from open to closed form.

3.3 Crystallography and Intermolecular Contacts

The target compound 1 crystallized in the P21/n space group from THF as yellow crystals (Figure 5). One interesting feature of the structure is the presence of several short intermolecular distances (Table 2 and Figure 6). For example, edge to face interactions are observed between aromatic systems, such as the intermolecular short contact observed between H11-C4, displaced from the centroid (H11-centroid distance is 3.166 Å) [25]. In addition, there is a C-H…pi interaction between the methyl H17A and C4 (H17A-centroid distance is 3.307Å). The crystal lattice is further stabilized by the presence of short intermolecular contacts between carbon atoms. For example, the C3 – C10 distance is 3.391 Å and the C10 – C11 distance is 3.343 Å. Similar short contacts were reported in the crystal structure of a related compound 3-ethyl-8-methoxy-6-nitro-2H-1-benzopyran-2-spiro-2′-(3′-methylthiazolidine) (4) by Miler-Srenger et al. [26]. However, the crystal structures of two benzothiazolinic spirooxazines (3-methyl-2,3-dihydro-l,3-benzothiazolespiro-2,2′-(3-methyl-2H-phenanthr[9,10-b] [ 1,4]oxazine) (5), 3-methyl-2,3-dihydro-1,3-benzothiazolespiro-2,2′-(3-methyl-2H-naphth[2, I-b][ 1,4]-oxazine) (6) reported earlier by Sun et al. [26] show no short intermolecular contacts, possibly due to the absence of electron donating or withdrawing substituents. Compound 2 did not crystallize well.

Table 2.

Short intermolecular contacts.

Atom1 Atom2 Length Length-VdW
C3 C10 3.391 −0.009
C4 H11 2.798 −0.102
H17A C4 2.859 −0.041
C10 C11 3.343 −0.057

Figure 6.

Figure 6

Crystal packing diagram (Mercury [23]) of 1 showing short contacts.

3.4 Computational Studies

Examination of various bond distances and bond angles reveals that in the solid state, the molecule 1 adopts a geometry very close to that obtained using DFT calculations (Tables 3, S1, S11, and S12). This suggests that the weak C-H…pi interactions do not affect the geometry significantly for 1. For example, the bond distance Cspiro-O1 (C8-O1) is 1.458 Ǻ in both structures. This bond distance is of particular interest, due to its potential for use in prediction of photochromic properties. For the related spiropyrans 4, 5, and 6, the calculated bond lengths are slightly less accurate, but are within 0.01 Å. The calculated Cspiro-N bond lengths are within 0.01 Å of experimental values for 1 and 4, but do not accurately reproduce crystallographic distances for spirooxazines 5 and 6.

Table 3.

Calculated bond lengths and crystallographic bond lengths.

Erel (kcal/mole)a Cspiro-O Bond Length (calc) Cspiro-N Bond Length (calc) Cspiro-O Bond length (crystal) Cspiro-N Bond length (crystal)
1 - 1.458 Å 1.437 Å 1.458 Å 1.434 Å
2-a 0 1.454 Å 1.437 Å - -
2-b 1.689 1.458 Å 1.436 Å - -
2-c 2.595 1.462 Å 1.440 Å - -
3-a 0 1.437 Å 1.443 Å - -
3-b 0.652 1.440 Å 1.442 Å - -
4 - 1.462 Å 1.436 Å 1.469 Å [26] 1.421 Å [26]
5 - 1.436 Å 1.438 Å 1.422 Å [27] 1.402 Å [27]
6 - 1.436 Å 1.438 Å 1.430 Å [27] 1.409 Å [27]
Saturated heterocycles [12] - - - 1.41–1.43 Å 1.47–1.48 Å
a

Relative to that of the lowest energy conformer

3.5 Structural Features of and Photochromism and Thermochromism

As for other classes of spiropyrans and spirooxazines, efforts to correlate substituent effects with photochromism have been pursued extensively for benzothiazolinic spiropyrans by several investigators such as Vandewyer et al. [31] and Kister et al. [32]. The presence of an electron-withdrawing group in position 6 and an electron-donating group in position 8 of the benzopyran ring system is often required for good photochromic properties. Compound 1 has one nitro substituent at position 6, and it is weakly photochromic at room temperature. By comparison, a second benzothiazolinic spiropyran (2) with both nitro and methoxy substituents was synthesized and characterized. This compound did not form suitable crystals to obtain a crystal structure, however photochemical and computational studies were pursued. Compound 2 was photochromic, and using the measure of colorability, or the steady state absorbance in the visible region upon irradiation, it was found to be more efficient at ring opening than 1 in THF at −30°C. This substituent effect in a different solvent and at a different temperature is also documented by other investigators (Table 4). Compound 3 was reported to be not photochromic [16]. The slight shortening of the Cspiro-O bond length with the added methoxy group (1 versus 2) is consistent with that found for the same substituent changes in photochromic spiroindolinopyrans [9]. For spiropyrans 14, as the Cspiro-O bond increases in length, the Cspiro-N bond decreases, which is consistent with what is observed crystallographically for indolinospiropyrans.

Table 4.

Bond lengths and photochemical properties for 1 and related spiropyran compounds.

Cspiro-O (Å) Cspiro-N (Å) Colorability = (A0/cL); (solvent, T) kT
1 1.458 1.434 4916 (THF, −30°) 700 (toluene, 25°) [28] 1.11 × 10−4 s−1 (THF, −30°) 1.8 s−1 (toluene, 25°) [26]
2 1.454a 1.437a 8995 (THF, −30°) 3700 (toluene, 25°) [28] 1.85 × 10−4 s−1 (THF, −30°) 6.5 s−1 (toluene, 25°) [28]
3 1.437a 1.443a 0b -
4 1.469 [26] 1.421 2300 (toluene, 25°) [29] 76 s−1 (toluene, 25°) [30]
5 1.422 [27] 1.402 0b -
6 1.430 [27] 1.409 0b -
a

Calculated using DFT for the most stable isomer (Table 3).

b

Not photochromic

3.5.1 Cspiro-N1 bond length

A crystallographic Cspiro-N1 bond length of 1.434 is observed for benzothiazolinic spiropyran 1, which is slightly longer than the reported value 1.421 Å for the related thiazolidinic spiropyran 4 and also longer than those found for the benzothiazolinic spirooxazines 5 and 6 (Table 3). Calculated bond lengths for Cspiro-N1 in benzothiazolinic spiropyrans 2 and 3 are close to that observed for 1. Taken together, bond lengths in all six compounds are significantly shorter than C(sp3)-N(sp3) bond lengths found in five-membered heterocycles. A shorter Cspiro-N1 bond is an indication of donation of lone pair electrons from N1 to Cspiro, which would stabilize the cation that develops on Cspiro when the compound is in the open form. There does not appear to be a correlation between Cspiro-N1 bond length and either photochromism or the extent of photocolorability for the series of compounds 16 (Table 4). It is possible, however, that the uncertainty of the calculated Cspiro-N1 bond lengths in the spirooxazines masks any trend that may exist.

The crystallographic bond angles around the spiro carbon atom of 1 are found to be between 106.15 to 114.01°. This distortion from the tetrahedral angle and the short Cspiro-N1 bond length indicate that the spiro carbon possesses a degree of sp2 character. The sum of the bond angles at N1 was found to be 348.21°, which lies between the sum for sp3 hybridized bond angles (327°) and the sum for sp2 hybridized bond angles (360°). The sum of the bond angles, as well as the shorter Cspiro-N distance described above, indicates that N1 has some sp2 character to stabilize a developing positive charge on Cspiro, suggesting a possibility of thermochromism. This possibility was also reported by Sun et al. [27] for their nonphotochromic benzothiazolinic spirooxazines 5 and 6. Thermochromism was not observed for 1 at room temperature, or up to 50°C in isopropanol.

3.5.2 Cspiro-O bond length

As discussed in the introduction, a longer Cspiro-O bond is associated with photochromism in spiropyrans and spirooxazines. In addition, for spirooxazines, this bond length correlates with the degree of photocolorability. The extent to which either photochromism or the degree of photocolorability of thiazolidinic and benzothiazolinic spiropyrans and benzothiazolinic spirooxazines may also be predicted by Cspiro-O bond length was explored using data collected and reported for compounds 16.

Taking together both the crystallographic and computational distances in Table 4, it appears that compounds with Cspiro-O bond distances greater than or equal to 1.454 Å (4, 1, 2) are photochromic, while compounds with Cspiro-O bond distances less than or equal to 1.437 Å (3, 6, 5) are not photochromic. Within the uncertainty of calculated bond Cspiro-O bond distances of up to 0.01 Å (Table 3), this cutoff is consistent with that observed for non benzothiazolinic or thiazolidinic spiropyrans and spirooxazines. In those classes of photochromic compounds, Cspiro-O bond distances were found to be greater than or equal to 1.463 Å [9]. It is interesting to note that the extent of colorability for compounds 16 does not correlate with longer Cspiro-O bonds. Clearly, many factors contribute to the colorability of a compound, including the effect of substituents on the nature of the electronic transitions that lead to the photoexcited state, rates of thermal closure, and irradiation wavelength-dependent quantum yields for photochemical ring opening.

4. Conclusions

Few crystallographic studies have been published for this subclass of benzothiazolinic spiropyrans, and no crystal structure of the closed form of any benzothiazolinic spiropyran has been reported. Therefore, correlations between experimental molecular geometry and photochromism or the degree of photocolorability for this subclass of spiropyrans are not yet well substantiated. In this work, we confirm that for benzothiazolinic or thiazolidinic spiropyrans and spirooxazines, the crystallographic bond distances such as Cspiro-O can be used as an indicator of photochromism, but not for the degree of photocolorability. Moreover, for compounds such as these, the crystal structure and the computational results are similar for Cspiro-O bonds. Therefore, computation may also serve as a useful tool to predict photochromism from ground state structures for benzothiazolinic or thiazolidinic spiropyrans and spirooxazines.

Supplementary Material

01

Figure 3.

Figure 3

Structures of related spiropyrans (1, 2, 3, 4) and spirooxazines (5, 6).

Acknowledgments

We thank Dr. Michael W. Day and Lawrence M. Henling (Caltech) for performing single-crystal X-ray crystallographic studies. Thanks to Josh Cohen for assistance in the preparation of this manuscript. This work was supported by a grant from the National Institutes of Health (NIGMS MBRS S06 GM08101).

Appendix A. Supplementary Data

CCDC 632419 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via External link http://www.ccdc.cam.ac.uk/conts/retrieving.html (or from the Cambridge Crystallographic Data Centre, 12, Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033). Supplementary data associated with this article can be found, in the online version, at

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Gust D, Moore TA, Moore AL. Chem Commun. 2006:1169. doi: 10.1039/b514736k. [DOI] [PubMed] [Google Scholar]
  • 2.Morrison H, editor. Biological Applications of Photochemical Switches. John Wiley and Sons; New York: 1993. [Google Scholar]
  • 3.Feringa B. Molecular Switches. Wiley-VCH; New York: 2003. [Google Scholar]
  • 4.Lu NT, Nguyen VN, Kumar S, McCurdy A. J Org Chem. 2005;70:9067. doi: 10.1021/jo051449j. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Berkovic G, Krongauz V, Weiss V. Chem Rev. 2000;100:1741. doi: 10.1021/cr9800715. [DOI] [PubMed] [Google Scholar]
  • 6.Raymo FM, Tomasulo M. Chem Eur J. 2006;12:3186. doi: 10.1002/chem.200501178. [DOI] [PubMed] [Google Scholar]
  • 7.Bertelson RC. In: Organic Photochromic and Thermochromic Compounds. Crano JC, Guglielmetti RJ, editors. Chapter 1 Plenum Press; New York: 1999. [Google Scholar]
  • 8.Bertelson RC. In: Photochromism. Brown GH, editor. Chapter 3 John Wiley & Sons; New York: [Google Scholar]
  • 9.Nakatsu K, Yoshioka H. Kwansei Gakuin Univ Natural Sci Rev. 1996;1:21. [Google Scholar]
  • 10.Millini R, Del Piero G, Allegrini P, Malatesta V, Castaldi G. Acta Cryst. 1993;C49:1205. [Google Scholar]
  • 11.Sheng Y, Leszczynski J, Garcia A, Rosario R, Gust D, Springer J. J Phys Chem B. 2004;108:16233. [Google Scholar]
  • 12.Crano JC, RC . Organic Photochromic and Thermochromic Compounds: Physicochemical Studies, Biological Applications, and Thermochromism (Topics in Applied Chemistry) Vol. 2. Plenum Press; New York: 1999. Chapter 7. [Google Scholar]
  • 13.Simkin BY, Makarov SP, Furmanova NG, Karaev KS, Minkin YI. J Heterocycl Chem. 1984;20:602. [Google Scholar]
  • 14.McCurdy A, Kawaoka AM, Thai H, Yoon SC. Tetrahedron Lett. 2001;42:7763. [Google Scholar]
  • 15.Kumar S, Hernandez D, Hoa B, Lee Y, Yang JS, McCurdy A. Org Lett. 2008;10:3761. doi: 10.1021/ol801406b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Guglielmetti R, Metzger J. Bull Soc Chim Fr. 1967;8:2824. [Google Scholar]
  • 17.Guglielmetti R, Vernin G, Metzger J. Bull Soc Chim Fr. 1967;19:2713. [Google Scholar]
  • 18.Miler-Srenger E, Guglielmetti R. J Chem Res (s) 1982;145:1980. [Google Scholar]
  • 19.Guglielmetti R, Meyer R, Dupuy C. J Chem Ed. 1973;50:413. [Google Scholar]
  • 20.Sheldrick GM. Acta Cryst. 1990;A46:467. [Google Scholar]
  • 21.Sheldrick GM. SHELXS97 and SHELXL97. University of Gottingen; Germany: 1997. [Google Scholar]
  • 22.Farrugia LJ. J Appl Cryst. 1997;30:565. [Google Scholar]
  • 23.Macrae CG, Edgington RR, McCabe P, Pidcock E, Shields GP, Taylor R, Towler M, van de Streek J. J Appl Cryst. 2006;39:453. [Google Scholar]
  • 24.Samat AM, Guglielmetti RJ, Martin G. Org Magn Resonance. 1976;8:62. [Google Scholar]
  • 25.Burguete MI, Bolte M, Frias JC, Garcia-Espana E, Luis SV, Miravet JF. Tetrahedron. 2002;58:4179. [Google Scholar]
  • 26.Miler-Srenger E, Guglielmetti R. Acta Cryst. 1984;C40:2050. [Google Scholar]
  • 27.Sun X, Liang Y, Fan M, Knobbe ET, Holt EM. Acta Cryst. 1997;C53:823. [Google Scholar]
  • 28.Gaude D, Gautron R, Guglielmetti R, Dufly JC. Bull Soc Chim Fr. 1981. p. 14. [Google Scholar]
  • 29.Maguet M, Garnier F, Guglielmetti R. J Chem Res (s) 1982;6:1519. [Google Scholar]
  • 30.Gaude D, Gautron R, Maguet M, Guglielmetti R. Helv Chim Acta. 1983;66:342. [Google Scholar]
  • 31.Vandewyer PH, Hoefnagels J, Smets G. Tetrahedron. 1969;25:3251. [Google Scholar]
  • 32.Kister J, Pommier HP, Metzger J, Guglielmetti R. Can J Chem. 1984;62:2372. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

01

RESOURCES