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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Redetermination of the solvent-free crystal structure of L-proline

CROSSMARK_Color_square_no_text.svg

aDepartment für Chemie, Universität zu Köln, Greinstrasse 4, 50939 Köln, Germany
*Correspondence e-mail: mbreugst@uni-koeln.de

Edited by E. V. Boldyreva, Russian Academy of Sciences, Russia (Received 28 February 2018; accepted 2 July 2018; online 10 July 2018)

The title compound, (S)-pyrrolidine-2-carb­oxy­lic acid (C5H9NO2), commonly known as L-proline, crystallized without the inclusion of any solvent or water mol­ecules through the slow diffusion of diethyl ether into a saturated solution of L-proline in ethanol. L-Proline crystallized in its zwitterionic form and the mol­ecules are linked via N—H⋯O hydrogen bonds, resulting in a two-dimensional network. In comparison to the only other publication of a single-crystal structure of L-proline without inclusions [Kayushina & Vainshtein (1965[Kayushina, R. L. & Vainshtein, B. K. (1965). Kristallografiya, 10, 833-844.]). Kristallografiya, 10, 833–844], the R1 value is significantly improved (0.039 versus 0.169) and thus, our data provides higher precision structural information.

1. Chemical context

There are 20 proteinogenic amino acids that form the basis of life. Like most amino acids, L-proline predominantely exists in the zwitterionic form (Boldyreva, 2008[Boldyreva, E. (2008). Models, Mysteries, and Magic of Molecules edited by J. C. A. Boeyens & J. F. Ogilvie, pp. 167-192, Dordrecht: Springer.]; Görbitz, 2015[Görbitz, C. H. (2015). Crystallogr. Rev. 21, 160-212.]). Among those proteinogenic amino acids, L-proline is the only compound featuring a secondary amine that can have a significant influence on the structure of proteins and peptides. For example, L-proline is responsible for the secondary structure of collagen (Hutton et al., 1966[Hutton, J. J. Jr, Tappel, A. L. & Udenfriend, S. (1966). Anal. Biochem. 16, 384-394.]) and often acts as a structural disruptor, which leads to structural changes from helical to coil (Tompa, 2002[Tompa, P. (2002). Trends Biochem. Sci. 27, 527-533.]). Another remarkable influence of the secondary amine can be derived from the hydrogen-bonding pattern in the solid state. Amino acids with primary amino groups commonly form bilayers incorporating two anti­parallel hydrogen-bonded sheets. In contrast, the secondary amino groups in L-proline and its derivatives usually form single-sheet layers, where the amino groups point in the same direction (Görbitz, 2015[Görbitz, C. H. (2015). Crystallogr. Rev. 21, 160-212.]). Similar conclusions were also drawn relying on powder diffraction data (Seijas et al., 2010[Seijas, L. E., Delgado, G. E., Mora, A. J., Fitch, A. N. & Brunelli, M. (2010). Powder Diffr. 25, 235-240.]). Based on the comparison of 40 different amino acids featuring an endocyclic nitro­gen atom, Görbitz concluded that small changes in the mol­ecular composition can cause a significant change in the hydrogen-bonding pattern (Görbitz, 2015[Görbitz, C. H. (2015). Crystallogr. Rev. 21, 160-212.]).

Within the last decade, L-proline has also attracted significant attention in the field of organic chemistry as an organocatalyst. Following earlier reports on the application of L-proline in the Hajos–Parrish–Eder–Sauer–Wiechert reaction (Eder et al., 1971[Eder, U., Sauer, G. & Wiechert, R. (1971). Angew. Chem. Int. Ed. Engl. 10, 496-497.]; Hajos & Parrish, 1974[Hajos, Z. G. & Parrish, D. R. (1974). J. Org. Chem. 39, 1615-1621.]), L-proline was re-discovered as an excellent catalyst for asymmetric aldol reactions (List et al. 2000[List, B., Lerner, R. A. & Barbas, C. F. (2000). J. Am. Chem. Soc. 122, 2395-2396.]; Feng et al., 2015[Feng, Y., Holte, D., Zoller, J., Umemiya, S., Simke, L. R. & Baran, P. S. (2015). J. Am. Chem. Soc. 137, 10160-10163.]). Today, proline and various derivatives are frequently used catalysts that are routinely employed for many different transformations including aldol, Mannich, Diels–Alder or epoxidation reactions (Mukherjee et al., 2007[Mukherjee, S., Yang, J. W., Hoffmann, S. & List, B. (2007). Chem. Rev. 107, 5471-5569.]).

[Scheme 1]

So far, crystal structures with R1 values of less than 0.10 have been published for 19 of the 20 proteinogenic amino acids (Görbitz, 2015[Görbitz, C. H. (2015). Crystallogr. Rev. 21, 160-212.]). However, for L-proline, the only available crystal structure without inclusions dates from 1965 and features a significantly worse R1 value of 0.169 (Kayushina & Vainshtein, 1965[Kayushina, R. L. & Vainshtein, B. K. (1965). Kristallografiya, 10, 833-844.]). To overcome this limitation for the last proteinogenic amio acid, we recently succeeded in determining the crystal structure of L-proline without any inclusions with significantly improved R1 values.

2. Structural commentary

L-Proline crystallized in its zwitterionic form: the oxygen atoms of the carb­oxy­lic acid (O1 and O2) are deprotonated and accordingly, the amine nitro­gen atom N1 is protonated. The pyrrolidine ring within the title compound adopts a slightly bent envelope conformation with the C2 atom out of the plane (Fig. 1[link]). Comparing the obtained values with previously reported crystal structures of enanti­omerically pure L- and D-proline, the racemic compound, as well as the co-crystalized structures, only marginal differences can be observed for the distances N1—C1, N1—C4, and C1—C5 as well as for the binding angles C4—N1—C1 and N1—C1—C5. This indicates that the inclusion of solvents and formation of co-crystals does not influence the structural properties of proline significantly.

[Figure 1]
Figure 1
The mol­ecular structure of the title compound L-proline. Displacement ellipsoids are drawn at the 50% probability level.

3. Supra­molecular features

As a secondary amine, L-proline carries two hydrogen atoms at the nitro­gen atom N1 in its zwitterionic form. These two hydrogen atoms each inter­act with one of the oxygen atoms of the carb­oxy­lic groups (O1 and O2). The different hydrogen-bond parameters between the proline mol­ecules are shown in Table 1[link]. As shown in Fig. 2[link], these hydrogen bonds result in the formation of a single-sheet architecture within the ab plane (also termed sheet L1 in Görbitz, 2015[Görbitz, C. H. (2015). Crystallogr. Rev. 21, 160-212.]). This structure is addionaly stabilized by hydro­phobic inter­actions between the C—H bonds of the pyrrolidine substructure (see Fig. 2[link]). In comparison, the hydrogen-bonding pattern of isoleucin (DAILEU01: Varughese & Srinivasan, 1975[Varughese, K. I. & Srinivasan, R. (1975). J. Cryst. Mol. Struct. 5, 317-328.]) as a typical example of an amino acid with a primary amino group features a double-sheet structure where the hydro­phobic and hydro­philic parts inter­act with each other (Fig. 3[link]). Therefore, the hydrogen-bonding pattern observed for L-proline once again illustrates why proline is considered to be a structural disruptor in proteins. However, as already pointed out above, small structural changes can have a signifcant influence, as the addition of a hy­droxy group in 3-hy­droxy­proline results in the formation of bands in the supra­molecular structure (HOPROL12: Koetzle et al., 1973[Koetzle, T. F., Lehmann, M. S. & Hamilton, W. C. (1973). Acta Cryst. B29, 231-236.]). This again highlights how even small changes such as the addition of a hy­droxy group can change the packing in the crystal structure.

Table 1
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
N1—H1A⋯O2i 0.87 (4) 2.01 (4) 2.759 (3) 144 (3)
N1—H1B⋯O1ii 0.91 (4) 1.82 (4) 2.703 (3) 165 (3)
Symmetry codes: (i) [-x+1, y+{\script{1\over 2}}, -z+{\script{1\over 2}}]; (ii) x+1, y, z.
[Figure 2]
Figure 2
View along the c axis (left) and the a axis (right) showing that L-proline forms layers through hydrogen bonding between the carb­oxy­lic group O1 respectively O2 and amine N1.
[Figure 3]
Figure 3
Hydro­philic and hydro­phobic layers in the crystal structure of isoleucin (DAILEU01: Varughese & Srinivasan, 1975[Varughese, K. I. & Srinivasan, R. (1975). J. Cryst. Mol. Struct. 5, 317-328.]).

4. Database survey

A survey of the Cambridge Structural Database (CSD, Version 5.39, last update Nov. 2017; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) for the L-proline structure resulted in 16 hits. Only one very early entry refers to the single crystal of the pure L-isomer without any inclusions (PROLIN: Kayushina & Vainshtein, 1965[Kayushina, R. L. & Vainshtein, B. K. (1965). Kristallografiya, 10, 833-844.]). However, the determination of this crystal structure was performed in 1965. Nevertheless, Kayushina and Vainshtein could identify the space group as P212121 and determine the cell parameters with a = 5.20 Å, b = 9.02 Å, c = 11.55 Å, which are good, but could be determined with higher precision in this study. Furthermore, the R1 value has now improved substanti­ally to 0.039. Seijas et al. (2010[Seijas, L. E., Delgado, G. E., Mora, A. J., Fitch, A. N. & Brunelli, M. (2010). Powder Diffr. 25, 235-240.]) investigated the powder diffraction data of enanti­opure L-proline and obtained an R1 value of 0.089 with similar structural features. They further compared the four pseudopolymorphs of L-proline, L-proline monohydrate, DL-proline and DL-proline monohydrate and concluded that all show a layered packing, which is stabilized by van der Waals inter­actions (PROLIN01: Seijas et al., 2010[Seijas, L. E., Delgado, G. E., Mora, A. J., Fitch, A. N. & Brunelli, M. (2010). Powder Diffr. 25, 235-240.]).

Besides the single entry for enanti­opure L-proline, one entry refers to enanti­opure L-proline with the inclusion of water (RUWGEV: Janczak & Luger, 1997[Janczak, J. & Luger, P. (1997). Acta Cryst. C53, 1954-1956.]), two entries refer to the racemic compound (QANRUT: Myung et al., 2005[Myung, S., Pink, M., Baik, M.-H. & Clemmer, D. E. (2005). Acta Cryst. C61, o506-o508.]; QANRUT01: Hayashi et al., 2006[Hayashi, Y., Matsuzawa, M., Yamaguchi, J., Yonehara, S., Matsumoto, Y., Shoji, M., Hashizume, D. & Koshino, H. (2006). Angew. Chem. 118, 4709-4713.]) and the racemic product with water (DLPROM01: Padmanabhan et al., 1995[Padmanabhan, S., Suresh, S. & Vijayan, M. (1995). Acta Cryst. C51, 2098-2100.]; DLPROM02: Flaig et al., 2002[Flaig, R., Koritsanszky, T., Dittrich, B., Wagner, A. & Luger, P. (2002). J. Am. Chem. Soc. 124, 3407-3417.]) or chloro­form (WERMIQ: Klussmann et al., 2006[Klussmann, M., White, A. J. P., Armstrong, A. & Blackmond, D. G. (2006). Angew. Chem. Int. Ed. 45, 7985-7989.]). The enanti­opure L-proline was also crystallized with inclusions of p-amino­benzoic acid (CIDBOH: Athimoolam & Natarajan, 2007[Athimoolam, S. & Natarajan, S. (2007). Acta Cryst. C63, o283-o286.]), 1,1-di­cyano-2-(4-hy­droxy­phen­yl)ethene (IHUMAZ: Timofeeva et al., 2003[Timofeeva, T. V., Kuhn, G. H., Nesterov, V. V., Nesterov, V. N., Frazier, D. O., Penn, B. G. & Antipin, M. Y. (2003). Cryst. Growth Des. 3, 383-391.]), S-bi­naphthol (NISVOA: Periasamy et al., 1997[Periasamy, M., Venkatraman, L. & Thomas, K. R. J. (1997). J. Org. Chem. 62, 4302-4306.]; NISVOA01: Hu et al., 2012[Hu, X., Shan, Z. & Chang, Q. (2012). Tetrahedron Asymmetry, 23, 1327-1331.]), p-nitro­phenol (QIRNUC: Sowmya et al., 2013[Sowmya, N. S., Vidyalakshmi, Y., Sampathkrishnan, S., Srinivasan, T. & Velmurugan, D. (2013). Acta Cryst. E69, o1723.]), and thio­urea monohydrate (UFOQEN: Umamaheswari et al., 2012[Umamaheswari, R., Nirmala, S., Sagayaraj, P. & Joseph Arul Pragasam, A. (2012). J. Therm. Anal. Calorim. 110, 891-895.]).

5. Synthesis and crystallization

The crystals were grown from commercially available L-proline (purchased from Carbolution). Crystals suitable for X-ray crystallography were obtained by the slow diffusion of diethyl ether into a saturated solution of L-proline in ethanol. After one night, colourless crystals were obtained and directly investigated via single crystal X-ray analysis. 1H NMR (500 MHz, DMSO-d6) δ = 1.67–1.83 (2 H, m, 3–H), 1.90–2.08 (2 H, m, 2–H), 3.02 (1 H, dt, 2J = 11.2 Hz and 3J = 7.5 Hz, 4–H), 3.22 (1 H, ddd, 2J = 11.2 Hz, 3J = 7.5 Hz, and 5.9 Hz, H–4), 3.65 (1 H, dd, 3J = 8.7 Hz and 6.5 Hz, 1–H). 13C NMR (125 MHz, DMSO-d6) δ = 24.3 (C-3), 29.4 (C-2), 45.7 (C-4), 61.2 (C-1), 169.8 (C-5). [α]D: −85.9° (c 1.0, H2O) (Lit. Monteiro, 1974[Monteiro, H. J. (1974). Synthesis, p. 137.]): −85° ± 2° (c 1.1, H2O), m.p. 486.7–487.2 K (decomposition).

6. Refinement details

Crystal data, data collection and structure refinement details are summarized in Table 2[link]. All H atoms bonded to carbon were placed with idealized geometry and refined using a riding model with C—H = 0.95 Å, Uiso(H) = 1.2 Ueq(C) for CH, C—H = 0.99 Å Uiso(H) = 1.2Ueq(C) for CH2, C—H = 0.98 Å and Uiso(H) = 1.5Ueq(C) for CH3. N-bound H atoms were located in a difference electron map and refined isotropically.

Table 2
Experimental details

Crystal data
Chemical formula C5H9NO2
Mr 115.13
Crystal system, space group Orthorhombic, P212121
Temperature (K) 100
a, b, c (Å) 5.2794 (4), 8.8686 (6), 11.5321 (9)
V3) 539.94 (7)
Z 4
Radiation type Cu Kα
μ (mm−1) 0.92
Crystal size (mm) 0.40 × 0.10 × 0.08
 
Data collection
Diffractometer Bruker D8 Venture
Absorption correction Multi-scan (SADABS; Bruker, 2012[Bruker (2012). APEX3, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.])
Tmin, Tmax 0.553, 0.754
No. of measured, independent and observed [I > 2σ(I)] reflections 4791, 1062, 993
Rint 0.053
(sin θ/λ)max−1) 0.618
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.036, 0.086, 1.11
No. of reflections 1062
No. of parameters 81
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.22, −0.19
Absolute structure Flack x determined using 361 quotients [(I+)−(I)]/[(I+)+(I)] (Parsons et al., 2013[Parsons, S., Flack, H. D. & Wagner, T. (2013). Acta Cryst. B69, 249-259.])
Absolute structure parameter 0.10 (17)
Computer programs: APEX3 and SAINT (Bruker, 2012[Bruker (2012). APEX3, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXT (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2014 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]) and SHELXLE (Hübschle et al., 2011[Hübschle, C. B., Sheldrick, G. M. & Dittrich, B. (2011). J. Appl. Cryst. 44, 1281-1284.]), SCHAKAL99 (Keller & Pierrard, 1999[Keller, E. & Pierrard, J.-S. (1999). SCHAKAL99. University of Freiburg, Germany.]), PLATON (Spek, 2009[Spek, A. L. (2009). Acta Cryst. D65, 148-155.]) and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]).

Supporting information


Computing details top

Data collection: APEX3 (Bruker, 2012); cell refinement: SAINT (Bruker, 2012); data reduction: SAINT (Bruker, 2012); program(s) used to solve structure: SHELXT (Sheldrick, 2015a); program(s) used to refine structure: SHELXL2014 (Sheldrick, 2015b) and SHELXLE (Hübschle et al., 2011); molecular graphics: SCHAKAL99 (Keller & Pierrard, 1999); software used to prepare material for publication: PLATON (Spek, 2009) and publCIF (Westrip, 2010).

(S)-Pyrrolidine-2-carboxylic acid top
Crystal data top
C5H9NO2Dx = 1.416 Mg m3
Mr = 115.13Melting point: 486.9 K
Orthorhombic, P212121Cu Kα radiation, λ = 1.54178 Å
Hall symbol: P 2ac 2abCell parameters from 4791 reflections
a = 5.2794 (4) Åθ = 6.3–72.3°
b = 8.8686 (6) ŵ = 0.92 mm1
c = 11.5321 (9) ÅT = 100 K
V = 539.94 (7) Å3Prism, colourless
Z = 40.40 × 0.10 × 0.08 mm
F(000) = 248
Data collection top
Bruker D8 Venture
diffractometer
993 reflections with I > 2σ(I)
Radiation source: micro focusRint = 0.053
phi / ω scansθmax = 72.3°, θmin = 6.3°
Absorption correction: multi-scan
(SADABS; Bruker, 2012)
h = 66
Tmin = 0.553, Tmax = 0.754k = 1010
4791 measured reflectionsl = 1414
1062 independent reflections
Refinement top
Refinement on F2Hydrogen site location: mixed
Least-squares matrix: fullH atoms treated by a mixture of independent and constrained refinement
R[F2 > 2σ(F2)] = 0.036 w = 1/[σ2(Fo2) + (0.036P)2 + 0.1571P]
where P = (Fo2 + 2Fc2)/3
wR(F2) = 0.086(Δ/σ)max < 0.001
S = 1.11Δρmax = 0.22 e Å3
1062 reflectionsΔρmin = 0.19 e Å3
81 parametersAbsolute structure: Flack x determined using 361 quotients [(I+)-(I-)]/[(I+)+(I-)] (Parsons et al., 2013)
0 restraintsAbsolute structure parameter: 0.10 (17)
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
O10.2943 (3)0.61385 (18)0.31235 (15)0.0182 (4)
O20.2573 (3)0.38601 (19)0.23111 (17)0.0261 (5)
N10.7901 (4)0.5949 (2)0.35050 (17)0.0150 (4)
H1A0.708 (7)0.673 (4)0.326 (3)0.040 (9)*
H1B0.952 (7)0.596 (4)0.325 (3)0.034 (9)*
C10.6604 (4)0.4557 (2)0.3057 (2)0.0134 (5)
H10.74820.41650.23500.016*
C20.6869 (4)0.3449 (2)0.4064 (2)0.0171 (5)
H2A0.85670.29770.40710.020*
H2B0.55630.26500.40240.020*
C30.6479 (5)0.4456 (3)0.5127 (2)0.0186 (5)
H3A0.46630.46850.52460.022*
H3B0.71640.39750.58360.022*
C40.7967 (5)0.5875 (3)0.4816 (2)0.0191 (5)
H4A0.71650.67800.51600.023*
H4B0.97330.58030.51000.023*
C50.3804 (4)0.4883 (3)0.27998 (19)0.0150 (5)
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
O10.0086 (7)0.0153 (8)0.0307 (9)0.0011 (7)0.0002 (7)0.0015 (7)
O20.0135 (8)0.0212 (8)0.0435 (11)0.0007 (8)0.0075 (8)0.0108 (8)
N10.0083 (9)0.0136 (9)0.0230 (10)0.0000 (8)0.0014 (8)0.0008 (8)
C10.0100 (11)0.0126 (10)0.0177 (10)0.0006 (9)0.0005 (8)0.0019 (9)
C20.0167 (12)0.0143 (10)0.0202 (12)0.0003 (9)0.0018 (10)0.0012 (9)
C30.0178 (12)0.0195 (11)0.0186 (11)0.0004 (10)0.0011 (9)0.0015 (9)
C40.0175 (11)0.0196 (11)0.0201 (12)0.0014 (10)0.0013 (10)0.0036 (9)
C50.0115 (10)0.0167 (11)0.0168 (10)0.0006 (9)0.0004 (9)0.0015 (9)
Geometric parameters (Å, º) top
O1—C51.260 (3)C2—C31.531 (3)
O2—C51.250 (3)C2—H2A0.9900
N1—C11.504 (3)C2—H2B0.9900
N1—C41.514 (3)C3—C41.526 (3)
N1—H1A0.87 (4)C3—H3A0.9900
N1—H1B0.91 (4)C3—H3B0.9900
C1—C21.527 (3)C4—H4A0.9900
C1—C51.535 (3)C4—H4B0.9900
C1—H11.0000
C1—N1—C4108.53 (18)H2A—C2—H2B109.1
C1—N1—H1A108 (2)C4—C3—C2102.92 (18)
C4—N1—H1A112 (2)C4—C3—H3A111.2
C1—N1—H1B109 (2)C2—C3—H3A111.2
C4—N1—H1B108 (2)C4—C3—H3B111.2
H1A—N1—H1B111 (3)C2—C3—H3B111.2
N1—C1—C2103.03 (18)H3A—C3—H3B109.1
N1—C1—C5110.50 (18)N1—C4—C3105.00 (18)
C2—C1—C5110.87 (18)N1—C4—H4A110.7
N1—C1—H1110.7C3—C4—H4A110.7
C2—C1—H1110.7N1—C4—H4B110.7
C5—C1—H1110.7C3—C4—H4B110.7
C1—C2—C3102.82 (17)H4A—C4—H4B108.8
C1—C2—H2A111.2O2—C5—O1126.0 (2)
C3—C2—H2A111.2O2—C5—C1116.8 (2)
C1—C2—H2B111.2O1—C5—C1117.18 (19)
C3—C2—H2B111.2
C4—N1—C1—C221.2 (2)C2—C3—C4—N128.2 (2)
C4—N1—C1—C597.3 (2)N1—C1—C5—O2172.9 (2)
N1—C1—C2—C338.5 (2)C2—C1—C5—O273.5 (3)
C5—C1—C2—C379.7 (2)N1—C1—C5—O18.7 (3)
C1—C2—C3—C441.5 (2)C2—C1—C5—O1104.9 (2)
C1—N1—C4—C34.4 (2)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
N1—H1A···O2i0.87 (4)2.01 (4)2.759 (3)144 (3)
N1—H1B···O1ii0.91 (4)1.82 (4)2.703 (3)165 (3)
Symmetry codes: (i) x+1, y+1/2, z+1/2; (ii) x+1, y, z.
 

Acknowledgements

We thank Professor Dr Albrecht Berkessel and his group for support.

Funding information

Financial support from the Fonds der Chemischen Industrie (Liebig-Scholarship to MB) and the University of Cologne within the excellence initiative is gratefully acknowledged.

References

First citationAthimoolam, S. & Natarajan, S. (2007). Acta Cryst. C63, o283–o286.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationBoldyreva, E. (2008). Models, Mysteries, and Magic of Molecules edited by J. C. A. Boeyens & J. F. Ogilvie, pp. 167–192, Dordrecht: Springer.  Google Scholar
First citationBruker (2012). APEX3, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationEder, U., Sauer, G. & Wiechert, R. (1971). Angew. Chem. Int. Ed. Engl. 10, 496–497.  CrossRef CAS Web of Science Google Scholar
First citationFeng, Y., Holte, D., Zoller, J., Umemiya, S., Simke, L. R. & Baran, P. S. (2015). J. Am. Chem. Soc. 137, 10160–10163.  Web of Science CrossRef Google Scholar
First citationFlaig, R., Koritsanszky, T., Dittrich, B., Wagner, A. & Luger, P. (2002). J. Am. Chem. Soc. 124, 3407–3417.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationGörbitz, C. H. (2015). Crystallogr. Rev. 21, 160–212.  Google Scholar
First citationGroom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171–179.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationHajos, Z. G. & Parrish, D. R. (1974). J. Org. Chem. 39, 1615–1621.  CrossRef CAS Web of Science Google Scholar
First citationHayashi, Y., Matsuzawa, M., Yamaguchi, J., Yonehara, S., Matsumoto, Y., Shoji, M., Hashizume, D. & Koshino, H. (2006). Angew. Chem. 118, 4709–4713.  CrossRef Google Scholar
First citationHu, X., Shan, Z. & Chang, Q. (2012). Tetrahedron Asymmetry, 23, 1327–1331.  Web of Science CrossRef Google Scholar
First citationHübschle, C. B., Sheldrick, G. M. & Dittrich, B. (2011). J. Appl. Cryst. 44, 1281–1284.  Web of Science CrossRef IUCr Journals Google Scholar
First citationHutton, J. J. Jr, Tappel, A. L. & Udenfriend, S. (1966). Anal. Biochem. 16, 384–394.  CrossRef Web of Science Google Scholar
First citationJanczak, J. & Luger, P. (1997). Acta Cryst. C53, 1954–1956.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationKayushina, R. L. & Vainshtein, B. K. (1965). Kristallografiya, 10, 833–844.  Google Scholar
First citationKeller, E. & Pierrard, J.-S. (1999). SCHAKAL99. University of Freiburg, Germany.  Google Scholar
First citationKlussmann, M., White, A. J. P., Armstrong, A. & Blackmond, D. G. (2006). Angew. Chem. Int. Ed. 45, 7985–7989.  Web of Science CrossRef Google Scholar
First citationKoetzle, T. F., Lehmann, M. S. & Hamilton, W. C. (1973). Acta Cryst. B29, 231–236.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationList, B., Lerner, R. A. & Barbas, C. F. (2000). J. Am. Chem. Soc. 122, 2395–2396.  Web of Science CrossRef Google Scholar
First citationMonteiro, H. J. (1974). Synthesis, p. 137.  CrossRef Google Scholar
First citationMukherjee, S., Yang, J. W., Hoffmann, S. & List, B. (2007). Chem. Rev. 107, 5471–5569.  Web of Science CrossRef PubMed CAS Google Scholar
First citationMyung, S., Pink, M., Baik, M.-H. & Clemmer, D. E. (2005). Acta Cryst. C61, o506–o508.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationPadmanabhan, S., Suresh, S. & Vijayan, M. (1995). Acta Cryst. C51, 2098–2100.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationParsons, S., Flack, H. D. & Wagner, T. (2013). Acta Cryst. B69, 249–259.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationPeriasamy, M., Venkatraman, L. & Thomas, K. R. J. (1997). J. Org. Chem. 62, 4302–4306.  CSD CrossRef PubMed CAS Web of Science Google Scholar
First citationSeijas, L. E., Delgado, G. E., Mora, A. J., Fitch, A. N. & Brunelli, M. (2010). Powder Diffr. 25, 235–240.  Web of Science CrossRef 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 citationSowmya, N. S., Vidyalakshmi, Y., Sampathkrishnan, S., Srinivasan, T. & Velmurugan, D. (2013). Acta Cryst. E69, o1723.  CrossRef IUCr Journals Google Scholar
First citationSpek, A. L. (2009). Acta Cryst. D65, 148–155.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationTimofeeva, T. V., Kuhn, G. H., Nesterov, V. V., Nesterov, V. N., Frazier, D. O., Penn, B. G. & Antipin, M. Y. (2003). Cryst. Growth Des. 3, 383–391.  Web of Science CSD CrossRef CAS Google Scholar
First citationTompa, P. (2002). Trends Biochem. Sci. 27, 527–533.  Web of Science CrossRef Google Scholar
First citationUmamaheswari, R., Nirmala, S., Sagayaraj, P. & Joseph Arul Pragasam, A. (2012). J. Therm. Anal. Calorim. 110, 891–895.  Web of Science CrossRef Google Scholar
First citationVarughese, K. I. & Srinivasan, R. (1975). J. Cryst. Mol. Struct. 5, 317–328.  CSD CrossRef CAS Web of Science Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals 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