research communications\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Synthesis and structure of 2,4,6-tri­cyclo­butyl-1,3,5-trioxane

CROSSMARK_Color_square_no_text.svg

aA.V.Topchiev Institute of Petrochemical Synthesis, Russian Academy of Sciences, 29 Leninsky prospect, 119991, Moscow, Russian Federation, bN.D. Zelinsky Institute of Organic Chemistry, Russian Academy of Sciences, 47, Leninsky prospect, 119991, Moscow, Russian Federation, and cA.N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of, Sciences, 28 Vavilova Str, Moscow, 119991, Russian Federation
*Correspondence e-mail: sshorunov@ips.ac.ru

Edited by V. Khrustalev, Russian Academy of Sciences, Russia (Received 11 April 2019; accepted 31 July 2019; online 3 October 2019)

The synthesis and structure of 2,4,6,-tri­cyclo­butyl-1,3,5-trioxane, C15H24O3 1, is described. It was formed in 39% yield during the work-up of the Swern oxidation of cyclo­butyl­methanol and may serve as a stable precursor of the cyclo­butane carbaldehyde. The mol­ecule of 1 occupies a special position (3.m) located at the center of its 1,3,5-trioxane ring. The latter is in a chair conformation, with the symmetry-independent O and C atoms deviating by 0.651 (4) Å from the least-squares plane of the other atoms of the trioxane ring. All three cyclo­butane substituents, which have a butterfly conformation with an angle between the two planes of 25.7 (3)°, are in the cis conformation relative to the 1,3,5-trioxane ring. Inter­molecular C—H⋯O inter­actions between the 1,3,5-trioxane rings consolidate the crystal structure, forming stacks along the c-axis direction. The crystal studied was refined a as a racemic twin.

1. Chemical context

Aliphatic aldehydes are well known to form trimers – derivatives of 1,3,5-trioxane. Some of these trimers are commercially-available, stable forms of the corresponding aldehydes, from which monomers can be regenerated by simple heating. Examples are trioxane and paraldehyde. For our research in the area of polymer chemistry, we need regular amounts of cyclo­butane carbaldehyde. Besides this, cyclo­butane carbaldehyde is a versatile synthetic building block for the incorporation of the cyclo­butane ring into organic mol­ecules in medicinal chemistry and drug discovery (Deaton et al., 2008[Deaton, D. N., Gao, E. N., Graham, K. P., Gross, J. W., Miller, A. B. & Strelow, J. M. (2008). Bioorg. Med. Chem. Lett. 18, 732-737.]; Inagaki et al., 2011[Inagaki, F., Sugikubo, K., Oura, Y. & Mukai, C. (2011). Chem. Eur. J. 17, 9062-9065.]). However, it is known to be unstable and prone to polymerization and decomposition by various routes (Slobodin & Blinova, 1953[Slobodin, Ya. M. & Blinova, M. V. (1953). Zhur. Obshch. Khim. 23, 1994-1998.]; Roquitte & Walters, 1962[Roquitte, B. C. & Walters, W. D. (1962). J. Am. Chem. Soc. 84, 4049-4052.]; Funke & Cerfontain, 1976[Funke, C. V. & Cerfontain, H. (1976). J. Chem. Soc. Perkin Trans. 2, pp. 826-827.]). Bearing all that in mind, we assumed that the trimeric form of cyclo­butane carbaldehyde should have a long shelf life and might find application as a stable equivalent of cyclo­butane carbaldehyde for long-term storage, with the possibility of performing a thermal breakdown back to the monomer once required. In this paper we describe our attempted synthesis of 2,4,6,-tri­cyclo­butyl-1,3,5-trioxane, 1, the trimer of cyclo­butane carbaldehyde.

One of the most straightforward routes to cyclo­butane carbaldehyde is the oxidation of the commercially available cyclo­butyl­methanol (Yusubov et al., 2007[Yusubov, M. S., Gilmkhanova, M. P., Zhdankin, V. V. & Kirschning, A. (2007). Synlett, 4, 563-566.]; Deaton et al. 2008[Deaton, D. N., Gao, E. N., Graham, K. P., Gross, J. W., Miller, A. B. & Strelow, J. M. (2008). Bioorg. Med. Chem. Lett. 18, 732-737.]). Further treatment with calcium chloride (Slobodin & Blinova, 1953[Slobodin, Ya. M. & Blinova, M. V. (1953). Zhur. Obshch. Khim. 23, 1994-1998.]) should afford the desired trimer. By employing Swern oxidation conditions, we were able to obtain the desired aldehyde monomer in 39% yield. (Fig. 1[link]). It was found that freshly distilled cyclo­butane carbaldehyde solidifies quickly upon short storage at room temperature, forming a colourless solid material readily soluble in organic solvents, especially non-polar ones. No calcium chloride was required for the trimerization, contrary to what was observed by Slobodin and co-workers. It should be noted that the melting point of our trimer was 391–393 K, close to that reported by Slobodin (391.5–393.5 K) while Funke and co-workers observed the melting to proceed at 397–398 K (Funke & Cerfontain, 1976[Funke, C. V. & Cerfontain, H. (1976). J. Chem. Soc. Perkin Trans. 2, pp. 826-827.]). Analysis of the synthesized compound by means of 1H and 13C NMR spectroscopy revealed the presence of unchanged cyclo­butane rings along with the absence of the aldehyde group [no downfield signals of the aldehyde proton (ca 9–12 ppm) in the 1H NMR spectrum and no signals of the aldehyde carbon (ca 180-220 ppm) in the 13C NMR spectrum]. The most downfield signal in the 1H NMR spectrum was a doublet at 4.81 ppm and the most downfield signal in the 13C NMR spectrum was observed at 103.3 ppm. The number of signals and their integral intensities (in the 1H NMR spectrum) demonstrated that, aside from the disappearance of the aldehyde group, no further degradation of the parent mol­ecule had taken place. We made an assumption that the desired trimer had formed. However, dimeric, trimeric, tetra­meric and higher cyclic forms or a linear polymer could also form. It should be noticed that the NMR technique cannot distinguish between these forms, since all the integral intensities and coupling constants would be the same, unless a large macrocycle is formed, where conformational effects would make some equivalent signals become non-equivalent. The mass spectrum of 1 solved the question and proved the structure to be trimeric. However, neither NMR nor mass spectrometry can provide information about the assignment of the cyclo­butyl groups with respect to the trioxane ring. It is known that paraldehyde – the trimer of acetaldehyde – exists in the form of two structural isomers: the cis isomer with all three methyl groups being in equatorial positions with respect to the trioxane ring and the trans isomer with one methyl group in an axial position and the two methyl groups in equatorial positions (Carpenter & Brockway,1936[Carpenter, D. C. & Brockway, L. O. (1936). J. Am. Chem. Soc. 58, 1270-1273.]; Kewley, 1970[Kewley, R. (1970). Can. J. Chem. 48, 852-855.]). These structures have the smallest 1,3-diaxial strain and are the only ones observed. In our case, with more bulky cyclo­butyl groups, one might expect that only the cis isomer would be formed.

[Figure 1]
Figure 1
Synthesis of 2,4,6-tri­cyclo­butyl-1,3,5-trioxane, 1.

The mol­ecular structure of 2,4,6-tri­cyclo­butyl-1,3,5-trioxane, 1, has been elucidated by X-ray diffraction analysis on single crystals grown from methanol. In summary, 2,4,6,-tri­cyclo­butyl-1,3,5-trioxane, 1, was formed in 39% yield during the work-up of the oxidation reaction of cyclo­butyl­methanol with DMSO and oxalyl chloride (Swern oxidation). It is a trimeric form of the cyclo­butane carbaldehyde. The aldehyde itself is unstable and quickly polymerizes even at room temperature. The described trimer could probably serve as a stable form of the cyclo­butane carbaldehyde. The compound is also of potential inter­est as a solid fuel.

[Scheme 1]

2. Structural commentary

The mol­ecule of 1 (Fig. 2[link]) occupies a special position (3.m) occurring in the center of its 1,3,5-trioxane ring. The latter is in a chair conformation with the symmetry-independent atoms O1 and C1 deviating by 0.651 (4) Å from the least-squares plane of the other ring atoms. All three cyclo­butane substituents, which show a butterfly conformation with an angle between the two planes of 25.7 (3)°, are in a cis conformation relative to the 1,3,5-trioxane ring.

[Figure 2]
Figure 2
Mol­ecular structure of 1 with displacement ellipsoids drawn at the 50% probability level. Symmetry codes: (A) y − x, −x, z; (B) −y, x − y, z; (C) y, x, z; (D) −y + x, −y, z; (E) −x, −x + y, z.

3. Supra­molecular features

Infinite stacks of mol­ecules of 1 along the c-axis direction consolidate the crystal structure (Fig. 3[link]) through C1—H1A⋯O1iv inter­actions between adjacent 1,3,5-trioxane rings (Table 1[link]).

Table 1
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
C1—H1A⋯O1iv 1.00 2.52 3.514 (3) 177
Symmetry code: (iv) [y, -x+y, z+{\script{1\over 2}}].
[Figure 3]
Figure 3
A fragment of the infinite stacks formed by mol­ecules of 1 along the c-axis direction. Hydrogen atoms except those of the 1,3,5-trioxane rings are omitted for clarity. Red dashed lines represent inter­molecular C—H⋯O inter­actions (Table 1[link]).

4. Database survey

The cisciscis configuration of the cyclo­butane rings observed in 1 seems to be typical for 2,4,6-tris­ubstituted-1,3,5-trioxanes, as follows from the search for such compounds in the Cambridge Structural Database (CSD, Version 5.40, November 2018 update; Groom et al, 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]). Indeed, only in 2,4,6-tris­(tri­chloro­meth­yl)-1,3,5-trioxane (refcode PRCHLA; Hay & Mackay, 1980[Hay, D. G. & Mackay, M. F. (1980). Acta Cryst. B36, 2367-2371.]) out of the 31 entries found, is one of the three substituents in a trans conformation with the other two are in a cis conformation. Another compound (MUKCEC; Arias-Ugarte et al., 2015[Arias-Ugarte, R., Wekesa, F. S. & Findlater, M. (2015). Tetrahedron Lett. 56, 2406-2411.]) consists of as a superposition of both cisciscis and transciscis conformations as a result of the disorder of one of the subsitutents. In the latter conformation, the central ring is in a boat conformation as the equatorially oriented cyclobutane rings become axially oriented.

5. Synthesis and crystallization

General experimental remarks

The synthesis of 1 was carried out under a purified argon atmosphere. The 1H and 13C NMR spectra were recorded on a Varian MercuryPlus 300 (300 MHz) spectrometer using CDCl3 as solvent. Di­chloro­methane, tri­ethyl­amine and DMSO were distilled over calcium hydride. Methanol was distilled over magnesium turnings. Oxalyl chloride was distilled over phospho­rus pentoxide. Cyclo­butyl­methanol is commercially available and was used without further purification.

Synthesis of compound 1

To a stirred solution of oxalyl chloride (13.5 g, 0.11 mol) in 110 ml of dry di­chloro­methane was added dropwise at 195K the solution of dry di­methyl­sulfoxide (16.5 g, 0.21 mol) in 30 ml of dry di­chloro­methane followed by the solution of cyclo­butyl­methanol (8.1 g, 0.09 mol) in 110 ml of dry di­chloro­methane. The reaction mixture was stirred at 195 K for 1 h and then neat tri­ethyl­amine (48.2 g, 0.47 mol) was added. The cooling bath was removed and when the reaction mixture had returned to room temperature, 100 ml of water was added and the reaction mixture was further stirred for 10 min. The flask content was then transferred to a separatory funnel, the bottom organic phase was collected and the aqueous phase was extracted with di­chloro­methane (2 × 100 ml). The combined organic phase was washed with 10% HCl (3 × 100 ml), water (2 × 100 ml), dried with anhydrous magnesium sulfate, filtered and the solvent was removed on the rotary evaporator. The residue was distilled under atmospheric pressure and the fraction boiling in the range 399–393 K was collected. The yield was 3.09 g (39%) and the product solidified after standing at room temperature overnight, qu­anti­tatively forming compound 1. 1H NMR (300 MHz, CDCl3): δ 1.65–2.16 (6H, m), 2.55 (1H, dt, J = 15.7; 7.7 Hz), 4.81 (1H, d, J = 5.6 Hz). 13C NMR (75 MHz, CDCl3): δ 18.5; 22.4; 38.1; 103.3. Mass spectrum m/z: 252 ([M+]), 169, 85, 67.

Colourless crystals of 1 suitable for X-ray analysis were obtained by dissolving the crude solid material (0.05 g) in 5 ml of warm methanol and keeping the resulting solution in a vial tightly stoppered with a plug of cotton wool for two days at room temperature.

6. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 2[link]. The hydrogen atoms were positioned geometrically (C—H = 1.00 and 0.99 Å for CH and CH2 groups, respectively) and refined using a riding model with Uiso(H) = 1.2Ueq(C). The crystal studied was refined as a racemic twin with a BASF of 0.146.

Table 2
Experimental details

Crystal data
Chemical formula C15H24O3
Mr 252.34
Crystal system, space group Hexagonal, P63cm
Temperature (K) 120
a, c (Å) 9.9966 (12), 7.9461 (10)
V3) 687.68 (19)
Z 2
Radiation type Mo Kα
μ (mm−1) 0.08
Crystal size (mm) 0.30 × 0.25 × 0.20
 
Data collection
Diffractometer Bruker APEXII DUO CCD area detector
Absorption correction Multi-scan (SADABS; Bruker, 2008[Bruker (2008). APEX2, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.])
Tmin, Tmax 0.701, 0.746
No. of measured, independent and observed [I > 2σ(I)] reflections 7859, 671, 645
Rint 0.026
(sin θ/λ)max−1) 0.681
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.041, 0.108, 1.03
No. of reflections 671
No. of parameters 35
No. of restraints 1
H-atom treatment H-atom parameters constrained
Δρmax, Δρmin (e Å−3) 0.36, −0.22
Absolute structure Refined as an inversion twin
Computer programs: APEX2 and SAINT (Bruker, 2008[Bruker (2008). APEX2, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXT (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]) and SHELXL2014/6 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]).

Supporting information


Computing details top

Data collection: APEX2 (Bruker, 2008); cell refinement: SAINT (Bruker, 2008); data reduction: SAINT (Bruker, 2008); program(s) used to solve structure: SHELXT (Sheldrick, 2015a); program(s) used to refine structure: SHELXL2014/6 (Sheldrick, 2015b); molecular graphics: SHELXL2014/6 (Sheldrick, 2015b); software used to prepare material for publication: SHELXL2014/6 (Sheldrick, 2015b).

2,4,6-Tricyclobutyl-1,3,5-trioxane top
Crystal data top
C15H24O3Dx = 1.219 Mg m3
Mr = 252.34Mo Kα radiation, λ = 0.71073 Å
Hexagonal, P63cmCell parameters from 7859 reflections
a = 9.9966 (12) Åθ = 3–30°
c = 7.9461 (10) ŵ = 0.08 mm1
V = 687.68 (19) Å3T = 120 K
Z = 2Prism, colorless
F(000) = 2760.30 × 0.25 × 0.20 mm
Data collection top
Bruker APEXII DUO CCD area detector
diffractometer
645 reflections with I > 2σ(I)
phi and ω scansRint = 0.026
Absorption correction: multi-scan
(SADABS; Bruker, 2008)
θmax = 29.0°, θmin = 2.4°
Tmin = 0.701, Tmax = 0.746h = 1313
7859 measured reflectionsk = 1313
671 independent reflectionsl = 1010
Refinement top
Refinement on F2Hydrogen site location: inferred from neighbouring sites
Least-squares matrix: fullH-atom parameters constrained
R[F2 > 2σ(F2)] = 0.041 w = 1/[σ2(Fo2) + (0.0819P)2 + 0.1148P]
where P = (Fo2 + 2Fc2)/3
wR(F2) = 0.108(Δ/σ)max < 0.001
S = 1.03Δρmax = 0.36 e Å3
671 reflectionsΔρmin = 0.22 e Å3
35 parametersAbsolute structure: Refined as an inversion twin
1 restraint
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
O10.00000.13467 (13)0.4774 (2)0.0133 (4)
C10.13482 (17)0.13482 (17)0.5351 (3)0.0120 (5)
H1A0.13840.13840.66090.014*
C20.2745 (2)0.2745 (2)0.4645 (3)0.0149 (4)
H2A0.27160.27160.33870.018*
C30.43522 (17)0.31060 (18)0.5281 (3)0.0198 (4)
H3A0.49380.28520.44620.024*
H3B0.43340.26670.64030.024*
C40.4839 (2)0.4839 (2)0.5319 (4)0.0192 (5)
H4A0.54200.54200.43100.023*
H4B0.53770.53770.63650.023*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
O10.0101 (7)0.0122 (6)0.0170 (7)0.0050 (3)0.0000.0017 (5)
C10.0113 (7)0.0113 (7)0.0135 (9)0.0057 (6)0.0016 (7)0.0016 (7)
C20.0123 (7)0.0123 (7)0.0193 (8)0.0056 (7)0.0002 (7)0.0002 (7)
C30.0118 (6)0.0142 (7)0.0339 (8)0.0068 (6)0.0022 (7)0.0008 (8)
C40.0124 (7)0.0124 (7)0.0304 (9)0.0044 (7)0.0014 (10)0.0014 (10)
Geometric parameters (Å, º) top
O1—C11.4229 (13)C2—H2A1.0000
O1—C1i1.4230 (13)C3—C41.548 (2)
C1—O1ii1.4229 (13)C3—H3A0.9900
C1—C21.505 (3)C3—H3B0.9900
C1—H1A1.0000C4—C3iii1.548 (2)
C2—C31.545 (2)C4—H4A0.9900
C2—C3iii1.545 (2)C4—H4B0.9900
C1—O1—C1i110.23 (18)C2—C3—C488.62 (11)
O1ii—C1—O1110.03 (17)C2—C3—H3A113.9
O1ii—C1—C2108.66 (11)C4—C3—H3A113.9
O1—C1—C2108.66 (11)C2—C3—H3B113.9
O1ii—C1—H1A109.8C4—C3—H3B113.9
O1—C1—H1A109.8H3A—C3—H3B111.1
C2—C1—H1A109.8C3—C4—C3iii88.39 (15)
C1—C2—C3117.97 (13)C3—C4—H4A113.9
C1—C2—C3iii117.97 (13)C3iii—C4—H4A113.9
C3—C2—C3iii88.59 (16)C3—C4—H4B113.9
C1—C2—H2A110.2C3iii—C4—H4B113.9
C3—C2—H2A110.2H4A—C4—H4B111.1
C3iii—C2—H2A110.2
C1i—O1—C1—O1ii58.4 (3)O1—C1—C2—C3iii67.9 (2)
C1i—O1—C1—C2177.20 (10)C1—C2—C3—C4139.29 (18)
O1ii—C1—C2—C367.9 (2)C3iii—C2—C3—C418.09 (19)
O1—C1—C2—C3172.39 (15)C2—C3—C4—C3iii18.05 (19)
O1ii—C1—C2—C3iii172.39 (15)
Symmetry codes: (i) y, xy, z; (ii) x+y, x, z; (iii) y, x, z.
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
C1—H1A···O1iv1.002.523.514 (3)177
Symmetry code: (iv) y, x+y, z+1/2.
 

Acknowledgements

X-ray diffraction data were collected with financial support from the Ministry of Science and Higher Education of the Russian Federation using the equipment of the Center for Mol­ecular composition studies of INEOS RAS.

Funding information

This work was carried out within the State Program of the TIPS RAS.

References

First citationArias-Ugarte, R., Wekesa, F. S. & Findlater, M. (2015). Tetrahedron Lett. 56, 2406–2411.  CAS Google Scholar
First citationBruker (2008). APEX2, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationCarpenter, D. C. & Brockway, L. O. (1936). J. Am. Chem. Soc. 58, 1270–1273.  CrossRef CAS Google Scholar
First citationDeaton, D. N., Gao, E. N., Graham, K. P., Gross, J. W., Miller, A. B. & Strelow, J. M. (2008). Bioorg. Med. Chem. Lett. 18, 732–737.  Web of Science CrossRef PubMed CAS Google Scholar
First citationFunke, C. V. & Cerfontain, H. (1976). J. Chem. Soc. Perkin Trans. 2, pp. 826–827.  CrossRef Web of Science Google Scholar
First citationGroom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171–179.  Web of Science CrossRef IUCr Journals Google Scholar
First citationHay, D. G. & Mackay, M. F. (1980). Acta Cryst. B36, 2367–2371.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationInagaki, F., Sugikubo, K., Oura, Y. & Mukai, C. (2011). Chem. Eur. J. 17, 9062–9065.  Web of Science CrossRef CAS PubMed Google Scholar
First citationKewley, R. (1970). Can. J. Chem. 48, 852–855.  CrossRef CAS Web of Science Google Scholar
First citationRoquitte, B. C. & Walters, W. D. (1962). J. Am. Chem. Soc. 84, 4049–4052.  CrossRef CAS Web of Science Google Scholar
First citationSheldrick, G. M. (2015a). Acta Cryst. A71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015b). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSlobodin, Ya. M. & Blinova, M. V. (1953). Zhur. Obshch. Khim. 23, 1994–1998.  CAS Google Scholar
First citationYusubov, M. S., Gilmkhanova, M. P., Zhdankin, V. V. & Kirschning, A. (2007). Synlett, 4, 563–566.  Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890
Follow Acta Cryst. E
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds