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

Journal logoSTRUCTURAL
CHEMISTRY
ISSN: 2053-2296

The α-D-anomer of 2′-de­­oxy­cyti­dine: crystal structure, nucleo­side conformation and Hirshfeld surface analysis

crossmark logo

aLaboratory of Bioorganic Chemistry and Chemical Biology, Center for Nanotechnology, Heisenbergstrasse 11, 48149 Münster, Germany, bOrganisch-Chemisches Institut, Westfälische Wilhelms-Universität Münster, Corrensstrasse 40, 48149 Münster, Germany, and cLaboratorium für Organische und Bioorganische Chemie, Institut für Chemie neuer Materialien, Universität Osnabrück, Barbarastrasse 7, Osnabrück 49069, Germany
*Correspondence e-mail: frank.seela@uni-osnabrueck.de

Edited by I. Oswald, University of Strathclyde, United Kingdom (Received 24 February 2021; accepted 30 March 2021; online 9 April 2021)

β-2′-De­oxy­ribo­nucleo­sides are the constituents of nucleic acids, whereas their anomeric α-analogues are rarely found in nature. Moreover, not much infor­mation is available on the structural and conformational parameters of α-2′-de­oxy­ribo­nucleo­sides. This study reports on the single-crystal X-ray structure of α-2′-de­oxy­cyti­dine, C9H13N3O4 (1), and the conformational param­eters characterizing 1 were determined. The conformation at the glycosylic bond is anti, with χ = 173.4 (2)°, and the sugar residue adopts an almost symmetrical C2′-endo-C3′-exo twist ([^2_3T]; S-type), with P = 179.7°. Both values lie outside the conformational range usually preferred by α-nucleo­sides. In addition, the amino group at the nucleobase shows partial double-bond character. This is supported by two separated signals for the amino protons in the 1H NMR spectrum, indicating a hindered rotation around the C4—N4 bond and a relatively short C—N bond in the solid state. Crystal packing is controlled by N—H⋯O and O—H⋯O contacts between the nucleobase and sugar moieties. Moreover, two weak C—H⋯N contacts (C1′—H1′ and C3′—H3′A) are observed. A Hirshfeld surface analysis was carried out and the results support the inter­molecular inter­actions observed by the X-ray analysis.

1. Introduction

Nucleosides with an α-configuration at the anomeric carbon are seldom found in nature (Ni et al., 2019[Ni, G., Du, Y., Tang, F., Liu, J., Zhao, H. & Chen, Q. (2019). RSC Adv. 9, 14302-14320.]). α-Nucleosides are not building blocks of naturally occurring DNA or RNA. However, α-nucleo­sides have been isolated as constituents of small mol­ecules in living cells, such as vitamin B12 (Bonnett, 1963[Bonnett, R. (1963). Chem. Rev. 63, 573-605.]) or a nicotinamide adenine dinucleotide (NAD) derivative isolated from Azobacter vinelandii (Suzuki et al., 1965[Suzuki, S., Suzuki, K., Imai, T., Suzuki, N. & Okuda, S. (1965). J. Biol. Chem. 240, PC554-PC556.]). Also, chemical nucleo­side synthesis yields α-nucleo­sides together with the β-anomers in ratios depending on the structures of the starting materials and the experimental conditions. Protocols were developed for the stereoselective synthesis of α-D nucleo­sides or by anomerization of β-D anomers. This topic has been reviewed recently by Ni et al. (2019[Ni, G., Du, Y., Tang, F., Liu, J., Zhao, H. & Chen, Q. (2019). RSC Adv. 9, 14302-14320.]).

α-Nucleosides were also incorporated into oligonucleotides, replacing single β-nucleo­sides (Guo & Seela, 2017[Guo, X. & Seela, F. (2017). Chem. Eur. J. 23, 11776-11779.]), or α-oligo­nucleotides were constructed which are entirely com­posed of α-nucleo­sides (Morvan et al., 1990[Morvan, F., Rayner, B. & Imbach, J.-L. (1990). In Genetic Engineering, Vol. 12, edited by J. K. Setlow. New York: Plenum Press.]). α-Oligonucleotides form duplexes with an anti­parallel orientation, with com­plementary strands also having an α-configuration (Morvan et al., 1987a[Morvan, F., Rayner, B., Imbach, J.-L., Chang, D.-K. & Lown, J. W. (1987a). Nucleic Acids Res. 15, 4241-4255.]), while duplexes with a parallel alignment are formed when the com­plementary strand is a β-oligonucleotide (Morvan et al., 1987b[Morvan, F., Rayner, B., Imbach, J.-L., Lee, M., Hartley, J. A., Chang, D.-K. & Lown, J. W. (1987b). Nucleic Acids Res. 15, 7027-7044.]).

Recently, we reported on the stability and recognition of silver-mediated heterochiral DNA with com­plementary α/β-strands (Chai et al., 2020[Chai, Y., Guo, X., Leonard, P. & Seela, F. (2020). Chem. Eur. J. 26, 13973-13989.]). Also, silver-mediated homochiral duplexes were constructed in which single residues were replaced by α-dC (1) (Scheme 1) (Guo & Seela, 2017[Guo, X. & Seela, F. (2017). Chem. Eur. J. 23, 11776-11779.]). The silver-mediated base pair formed by anomeric α-dC (1) with β-dC (2) shows significantly higher stability than that formed by the silver-mediated β-dC–β-dC pair. Not only the stability of the metal-mediated base pair is higher, but also the metal-free α-dC–β-dC mismatch is more stable.

Conformational studies on α-nucleo­sides revealed distinct differences com­pared to their β-anomeric counterparts (Sundaralingam, 1971[Sundaralingam, M. (1971). J. Am. Chem. Soc. 93, 6644-6647.]; Latha & Yathindra, 1992[Latha, Y. S. & Yathindra, N. (1992). Biopolymers, 32, 249-269.]). In general, the flexibility around the glycosylic linkage, as well as the sugar pucker of α-nucleo­sides, seems to be more restricted than for β-nucleo­sides. The different conformational properties of the α/β-anomers were attributed to the differences in the steric inter­actions between the nucleobase and the sugar moiety (Sundaralingam, 1971[Sundaralingam, M. (1971). J. Am. Chem. Soc. 93, 6644-6647.]; Latha & Yathindra, 1992[Latha, Y. S. & Yathindra, N. (1992). Biopolymers, 32, 249-269.]). However, com­pared to the number of X-ray analyses of β-nu­cleo­sides, studies on α-nucleo­sides are extremely limited (Sundaralingam, 1971[Sundaralingam, M. (1971). J. Am. Chem. Soc. 93, 6644-6647.]; Latha & Yathindra, 1992[Latha, Y. S. & Yathindra, N. (1992). Biopolymers, 32, 249-269.]). Surprisingly, among the canonical α-nucleo­sides, only the solid-state conformations of α-cyti­dine (Post et al., 1977[Post, M. L., Birnbaum, G. I., Huber, C. P. & Shugar, D. (1977). Biochim. Biophys. Acta Nucleic Acids Protein Synth. 479, 133-142.]) and α-2′-de­oxy­thy­midine (3) (Görbitz et al., 2005[Görbitz, C. H., Nelson, W. H. & Sagstuen, E. (2005). Acta Cryst. E61, o1207-o1209.]) have been reported. Moreover, some X-ray studies on modified α-nucleo­side analogues have been reported, e.g. α-5-acetyl-2′-de­oxy­uridine (Hamor et al., 1977[Hamor, T. A., O'Leary, M. K. & Walker, R. T. (1977). Acta Cryst. B33, 1218-1223.]), α-5-aza-7-de­aza-2′-de­oxy­guanosine (Seela et al., 2002[Seela, F., Rosemeyer, H., Melenewski, A., Heithoff, E.-M., Eickmeier, H. & Reuter, H. (2002). Acta Cryst. C58, o142-o144.]), α-5-iodo-2′-de­oxy­cyti­dine (Müller et al., 2019[Müller, S. L., Zhou, X., Leonard, P., Korzhenko, O., Daniliuc, C. & Seela, F. (2019). Chem. Eur. J. 25, 3077-3090.]) and α-5-octa­diynyl-2′-de­oxy­cyti­dine (Zhou et al., 2019[Zhou, X., Müller, S. L., Leonard, P., Daniliuc, C., Chai, Y., Budow-Busse, S. & Seela, F. (2019). J. Mol. Struct. 1190, 37-46.]). Among the α-2′-de­oxy­ribo­nucleo­sides, the solid-state conformations of the α-anomers of 2′-de­oxy­cyti­dine (1), 2′-de­oxy­adenosine and 2′-de­oxy­guanosine are still unknown.

[Scheme 1]

The single-crystal X-ray analysis of α-2′-de­oxy­cyti­dine (1) was performed in order to obtain a deeper insight into the conformational properties of 1 in the solid-state. This is the second report of an α-anomer of a canonical pyrimidine 2′-de­oxy­ribo­nucleo­side besides α-2′-de­oxy­thymidine (3) (Gör­bitz et al., 2005[Görbitz, C. H., Nelson, W. H. & Sagstuen, E. (2005). Acta Cryst. E61, o1207-o1209.]). The results are com­pared to the structure of β-dC (Young & Wilson, 1975[Young, D. W. & Wilson, H. R. (1975). Acta Cryst. B31, 961-965.]). For both 1 and 2, the sugar conformation in solution was determined using a 1H NMR-based method. Moreover, a Hirshfeld surface analysis of 1 was carried out to visualize the packing inter­actions.

2. Experimental

2.1. Synthesis and crystallization of α-dC (1)

α-2′-De­oxy­cyti­dine (1) was synthesized as reported pre­viously (Chai et al., 2019[Chai, Y., Leonard, P., Guo, X. & Seela, F. (2019). Chem. Eur. J. 25, 16639-16651.]). For crystallization, com­pound 1 was dissolved in methanol containing 10% water (10 mg in 1 ml) and was obtained as colourless prisms (m.p. 203–204 °C; Yamaguchi & Saneyoshi, 1984[Yamaguchi, T. & Saneyoshi, M. (1984). Chem. Pharm. Bull. 32, 1441-1450.]) by slow evaporation of the solvent at room temperature. A colourless prism-like speci­men of 1 was used for the X-ray crystallographic analysis.

2.2. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 1[link]. The H atoms on N4, O3′ and O5′ were refined freely.

Table 1
Experimental details

Crystal data
Chemical formula C9H13N3O4
Mr 227.22
Crystal system, space group Orthorhombic, P212121
Temperature (K) 100
a, b, c (Å) 6.8378 (4), 11.4334 (7), 12.7595 (8)
V3) 997.53 (11)
Z 4
Radiation type Cu Kα
μ (mm−1) 1.02
Crystal size (mm) 0.22 × 0.18 × 0.16
 
Data collection
Diffractometer Bruker APEXII Kappa CCD
Absorption correction Multi-scan (SADABS; Bruker, 2014[Bruker (2014). SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.])
Tmin, Tmax 0.75, 0.85
No. of measured, independent and observed [I > 2σ(I)] reflections 11824, 1768, 1719
Rint 0.037
(sin θ/λ)max−1) 0.596
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.025, 0.061, 1.12
No. of reflections 1768
No. of parameters 161
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.12, −0.18
Absolute structure Flack x determined using 684 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 0.04 (9)
Computer programs: SAINT (Bruker, 2015[Bruker (2015). SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXS2014 (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2014 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), APEX3 (Bruker, 2016[Bruker (2016). APEX3. Bruker AXS Inc., Madison, Wisconsin, USA.]) and XP (Bruker, 1998[Bruker (1998). XP - Interactive molecular graphics. Version 5.1. Bruker AXS Inc., Madison, Wisconsin, USA.]).

3. Results and discussion

3.1. Mol­ecular geometry and conformation of α-dC (1)

The three-dimensional (3D) structure of α-dC (1) is shown in Fig. 1[link] and selected geometric parameters are presented in Table 2[link]. The 3D structure of 1 clearly indicates the α-orientation of the nucleobase (Fig. 1[link]), which in addition is supported by the Flack parameter (see Table 1[link]; Parsons et al., 2013[Parsons, S., Flack, H. D. & Wagner, T. (2013). Acta Cryst. B69, 249-259.]). Moreover, according to the synthetic pathway, the anomeric centre at C1′ shows an S-configuration, confirming the α-D anomeric structure of 1.

Table 2
Selected geometric parameters (Å, °)

N1—C1′ 1.499 (2) N4—C4 1.337 (3)
       
C6—N1—C1′ 122.33 (16) O5′—C5′—C4′ 112.26 (16)
C2—N1—C1′ 117.08 (15) C1′—O4′—C4′ 110.96 (15)
       
C4—N3—C2—O2 −178.13 (18) C1′—C2′—C3′—C4′ −33.03 (18)
N4—C4—C5—C6 175.05 (18) C3′—C4′—C5′—O5′ 55.9 (2)
C2—N1—C1′—O4′ 173.39 (16)    
[Figure 1]
Figure 1
Perspective view of the α-D anomer of 2′-de­oxy­cyti­dine (1), showing the atomic numbering scheme. Displacement ellipsoids are drawn at the 50% probability level and H atoms are shown as small spheres of arbitrary size.

The crystal structure of the related canonical β-2′-de­oxy­cyti­dine (2) has been reported previously (Young & Wilson, 1975[Young, D. W. & Wilson, H. R. (1975). Acta Cryst. B31, 961-965.]). The β-anomer 2 shows two conformers (2a and 2b) in the unit cell. As not many single-crystal X-ray analyses of α-2′-de­oxy­ribo­nucleo­sides exist, it was of inter­est to com­pare the geometric parameters of the α/β-anomers of 2′-de­oxy­cyti­dine (1 and 2).

For pyrimidine nucleo­sides, the orientation of the nucleobase with respect to the sugar moiety (syn/anti) is defined by the torsion angle χ (O4′—C1′—N1—C2) (IUPAC–IUB Joint Commission on Biochemical Nomenclature, 1983[IUPAC-IUB Joint Commission on Biochemical Nomenclature (1983). Eur. J. Biochem. 131, 9-15.]). In the anti conformation, atom O2 of the six-membered ring is pointing away from the sugar, while in the syn conformation, O2 is pointing towards the sugar ring (Saenger, 1984[Saenger, W. (1984). In Principles of Nucleic Acid Structure, edited by C. R. Cantor. New York: Springer-Verlag.]). The preferred conformation for canonical pyrimidine β-2′-de­oxy­ribo­nucleo­sides, including β-dC (2a and 2b), is anti (2a: χ = 201.2°; 2b: χ = 222.2°) (Young & Wilson, 1975[Young, D. W. & Wilson, H. R. (1975). Acta Cryst. B31, 961-965.]). In contrast to the broad range of anti conformations adopted by β-nucleo­sides, a rather narrow preferred anti range, together with a preference of χ to adopt lower anti values, has been reported for α-nucleo­sides (Latha & Yathindra, 1992[Latha, Y. S. & Yathindra, N. (1992). Biopolymers, 32, 249-269.]). For instance, α-2′-de­oxy­thymidine (3) adopts a χ value of 124° (Görbitz et al., 2005[Görbitz, C. H., Nelson, W. H. & Sagstuen, E. (2005). Acta Cryst. E61, o1207-o1209.]). However, in case of the title com­pound α-2′-de­oxy­cyti­dine (1), an anti conformation with χ = 173.39 (16)° is observed which is significantly greater. In addition, other solid-state structures of modified pyrimidine α-2′-de­oxy­ribo­nucleo­sides with χ values around 168° have been reported recently, which also fall into this range (Zhou et al., 2019[Zhou, X., Müller, S. L., Leonard, P., Daniliuc, C., Chai, Y., Budow-Busse, S. & Seela, F. (2019). J. Mol. Struct. 1190, 37-46.]; Müller et al., 2019[Müller, S. L., Zhou, X., Leonard, P., Korzhenko, O., Daniliuc, C. & Seela, F. (2019). Chem. Eur. J. 25, 3077-3090.]).

The second conformational parameter of inter­est for nucleo­sides is the sugar puckering mode. The β-2′-de­oxy­ribo­furanosyl moiety shows a preference for two principal sugar conformations, namely C3′-endo (N) and C2′-endo (S) (Altona & Sundaralingam, 1972[Altona, C. & Sundaralingam, M. (1972). J. Am. Chem. Soc. 94, 8205-8212.]; Sundaralingam, 1971[Sundaralingam, M. (1971). J. Am. Chem. Soc. 93, 6644-6647.]). In contrast, studies on α-2′-de­oxy- and ribo­nucleo­sides showed that these anomers prefer mainly C3′-exo, C2′-exo and C4′-endo conformations (Latha & Yathindra, 1992[Latha, Y. S. & Yathindra, N. (1992). Biopolymers, 32, 249-269.]; Sundaralingam, 1971[Sundaralingam, M. (1971). J. Am. Chem. Soc. 93, 6644-6647.]). As can be seen in Fig. 1[link], the sugar moiety of α-dC (1) adopts an almost symmetrical C2′-endo-C3′-exo twist ([^2_3T]; S-type), with a pseudorotational phase angle P = 179.7° and a maximum amplitude τm = 33.0° (Altona & Sundaralingam, 1972[Altona, C. & Sundaralingam, M. (1972). J. Am. Chem. Soc. 94, 8205-8212.]; Saenger, 1984[Saenger, W. (1984). In Principles of Nucleic Acid Structure, edited by C. R. Cantor. New York: Springer-Verlag.]). Thus, the α-2′-de­oxy­ribo­furanosyl moiety of 1 exhibits a C2′-endo conformation which is outside the preferred conformational range of α-2′-de­oxy­ribo­nucleo­sides. Other examples of α-2′-de­oxy­ribo­nucleo­sides with a C2′-endo conformation of the sugar residue include α-5-acetyl-2′-de­oxy­uridine (Hamor et al., 1977[Hamor, T. A., O'Leary, M. K. & Walker, R. T. (1977). Acta Cryst. B33, 1218-1223.]) and the α-anomer of 5-aza-7-de­aza-2′-de­oxy­guanosine (Seela et al., 2002[Seela, F., Rosemeyer, H., Melenewski, A., Heithoff, E.-M., Eickmeier, H. & Reuter, H. (2002). Acta Cryst. C58, o142-o144.]). For com­parison, the conformers of the canonical β-dC (2) exhibit two different conformations, namely C3′-endo (N-type) for conformer 2a and C2′-endo (S-type) for conformer 2b (Young & Wilson, 1975[Young, D. W. & Wilson, H. R. (1975). Acta Cryst. B31, 961-965.]).

The torsion angle γ (O5′—C5′—C4′—C3′) characterizes the orientation of the exocyclic 5′-hy­droxy group relative to the sugar ring (Saenger, 1984[Saenger, W. (1984). In Principles of Nucleic Acid Structure, edited by C. R. Cantor. New York: Springer-Verlag.]). Earlier studies on α-nucleo­sides indicate that the conformational preference about the C4′—C5′ bond is similar to that of β-nucleo­sides (Sundaralingam, 1971[Sundaralingam, M. (1971). J. Am. Chem. Soc. 93, 6644-6647.]). For α-dC (1), γ is 55.9 (2)°, referring to a +sc (gauche, gauche) conformation which is similar to that found for β-dC (2) (56.7 and 62.5°; +sc, gauche, gauche) (Young & Wilson, 1975[Young, D. W. & Wilson, H. R. (1975). Acta Cryst. B31, 961-965.]).

3.2. Hydrogen bonding and mol­ecular packing of α-dC (1)

Fig. 2[link] displays the crystal packing mode and hydrogen-bonding pattern for the crystal of α-dC (1). The corresponding hydrogen-bonding data and symmetry codes are summarized in Table 3[link]. The particular nucleo­side units of 1 are connected by hydrogen bonds between (i) the nucleobases, (ii) nucleobases and sugars, as well as (iii) two sugar moieties. The crystal structure is formed by a repetition of nucleo­side units which are arranged in chains in a zigzag-like manner within the ac plane (Fig. 2[link]). This arrangement is different to that of the canonical β-dC (2). Space-filling models of α-dC (1) and β-dC (2) shown in Figs. 3[link](a) and 3(b), respectively, visualize the different crystal packing modes.

Table 3
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
N4—H4A⋯O3′i 0.86 (3) 2.10 (3) 2.949 (2) 168 (2)
N4—H4B⋯O2ii 0.93 (3) 2.05 (3) 2.937 (2) 159 (2)
C1′—H1⋯N3iii 1.0 2.46 3.317 (3) 144
C3′—H3A⋯N3iv 1.0 2.53 3.524 (3) 170
O3′—H3⋯O5′v 0.94 (3) 1.86 (4) 2.793 (2) 175 (3)
O5′—H5⋯O2iii 0.91 (4) 1.88 (3) 2.716 (2) 152 (3)
Symmetry codes: (i) [x+{\script{1\over 2}}, -y+{\script{3\over 2}}, -z+1]; (ii) [-x+2, y+{\script{1\over 2}}, -z+{\script{1\over 2}}]; (iii) [-x+2], [y-{\script{1\over 2}}, -z+{\script{1\over 2}}]; (iv) [-x+1, y-{\script{1\over 2}}, -z+{\script{1\over 2}}]; (v) [-x+1, y+{\script{1\over 2}}, -z+{\script{1\over 2}}].
[Figure 2]
Figure 2
Crystal packing of α-2′-de­oxy­cyti­dine (1), shown along the ac plane (ball-and-stick model), and with magnifications of designated areas of the crystal packing, showing the hydrogen-bonding pattern.
[Figure 3]
Figure 3
Space-filling models of (a) α-dC (1) and (b) β-dC (2). Schematic view of the inter­molecular hydrogen-bonding inter­actions of two nucleo­sides for (c) α-dC (1) and (d) β-dC (2).

In more detail, two cytosine residues of β-dC (2) (Young & Wilson, 1975[Young, D. W. & Wilson, H. R. (1975). Acta Cryst. B31, 961-965.]) form a mismatch connected by two hydrogen bonds, each between atom N3 and the 4-amino group of the respective second mol­ecule (Fig. 3[link]d). Fig. 3[link](c) shows that the situation is com­pletely different for the α-anomer of 2′-de­oxy­cyti­dine (1). Only one hydrogen bond is formed between the nucleobases (N4—H4B⋯O2ii) (Fig. 2[link], motif II, and Table 3[link]). The chains are further stabilized by a nucleobase-to-sugar contact (O5′—H5⋯O2iii; Fig. 2[link], motif II, and Table 3[link]) and a sugar-to-sugar contact (O3′—H3′⋯O5′v; Fig. 2[link], motif I, and Table 3[link]). In addition, two weak C—H⋯N contacts with C1′—H1′ (motif II) and C3′—H3′A (motif I) as the hydrogen-bond donors and N3 as the acceptor (N3iii and N3iv, respectively) are observed. In most crystal structures of 2′-de­oxy­ribo­nucleo­sides, the C—H groups of the sugar moiety do not participate as hydrogen-bond donors in hydrogen bonding. However, a few examples have been reported, e.g. α-5-iodo-2′-de­oxy­cyti­dine (Müller et al., 2019[Müller, S. L., Zhou, X., Leonard, P., Korzhenko, O., Daniliuc, C. & Seela, F. (2019). Chem. Eur. J. 25, 3077-3090.]) and 7-iodo-5-aza-7-de­aza­guanosine (Kondhare et al., 2020[Kondhare, D., Budow-Busse, S., Daniliuc, C. & Seela, F. (2020). Acta Cryst. C76, 513-523.]). Moreover, two neighbouring chains are connected by a N4—H4A⋯O3′ hydrogen bond, as illustrated in Fig. 2[link] (motif III), thereby generating a network.

3.3. Hirshfeld surface analysis of α-dC (1)

To visualize the inter­molecular inter­actions of α-dC (1) in the solid-state, a Hirshfeld surface analysis was conducted and two-dimensional (2D) fingerprint plots were analysed (Spackman & Jayatilaka, 2009[Spackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19-32.]). The CrystalExplorer program (Version 17; Spackman & Jayatilaka, 2009[Spackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19-32.]; Turner et al., 2017[Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). CrystalExplorer17. University of Western Australia. https://hirshfeldsurface.net.]) was used to carry out the Hirshfeld surface analysis mapped over a dnorm range from −0.5 to 1.5 Å, shape index (−1.0 to 1.0 Å) and curvedness (−4.0 to 0.4 Å), as well as their associated 2D fingerprint plots (Fig. 4[link]). On the dnorm surface of α-dC (1), several red areas (intense red spots) are observed (Figs. 4[link]b and 4[link]c), corresponding to the close contacts of the nucleobase and sugar residue (N—H⋯O and O—H⋯O). These inter­actions are shorter than the sum of the van der Waals radii and show negative dnorm. Small and light-red coloured spots are also found (Fig. 4[link]b) and can be assigned to the weak contacts with C1′—H1′ and C3′—H3′A as hydrogen-bond donors and N3 as acceptor. The results of the Hirshfeld analyses are consistent with the hydrogen-bonding data (Table 3[link]). The shape index (Fig. 4[link]d) indicates ππ stacking inter­actions by the presence of red and blue triangles, and flat surface patches within the curvedness surfaces (Fig. 4[link]e) are characteristic for planar stacking. However, as also indicated by Fig. 2[link], these inter­actions are less pronounced in the crystal structure of α-dC (1).

[Figure 4]
Figure 4
(a) Perspective view of α-dC (1), showing the atomic numbering scheme. The Hirshfeld surface of 1 mapped with (b) dnorm (−0.5 to 1.5, front view), (c) dnorm (−0.5 to 1.5, back view), (d) shape index and (e) curvedness, and (f) the corresponding fingerprint plots. Full inter­actions (left) and the resolved contacts (left, C⋯H/H⋯C; middle left, N⋯H/H⋯N; middle right, O⋯H/H⋯O; right, H⋯H) are shown, together with the percentages of their contribution to the total Hirshfeld surface area of α-anomer 1.

The 2D fingerprint plots provide a visual summary of the inter­molecular contacts in the crystal structure of 1 and can be resolved to particular atom-pair inter­actions and their relative contributions to the Hirshfeld surface, as illustrated in Fig. 4[link](f). Strong inter­actions are found for O⋯H/H⋯O (31.2%) and N⋯H/H⋯N (13.4%), which agrees with the fact that the crystal packing of α-dC (1) is largely controlled by N—H⋯O and O—H⋯O hydrogen bonds (Table 3[link]).

3.4. Conformation of the α- and β-anomers of 2′-de­oxy­cyti­dine in solution

For canonical nucleo­sides with a β-D configuration, numer­ous studies exist describing their crystal structures and conformation in the solid-state and in solution. Compared to the information available for β-anomeric nucleo­sides, reports on α-anomers are limited (Poznański et al., 2001[Poznański, J., Felczak, K., Bretner, M., Kulikowski, T. & Remin, M. (2001). Biochem. Biophys. Res. Commun. 283, 1142-1149.]). Moreover, the conformational change of anomeric nucleo­sides from β to α has an effect on the stability of the DNA double helix (Thibaudeau & Chattopadhyaya, 1997[Thibaudeau, C. & Chattopadhyaya, J. (1997). Nucleoside Nucleotides Nucleic Acids, 16, 523-529.]).

To ascertain the sugar conformation of α-dC (1) in solution, a conformational analysis of the furan­ose puckering of α-dC (1) and, for com­parison, of β-dC (2) was performed. To this end, high resolution (600 MHz) 1H NMR spectra were measured in dimethyl sulfoxide (DMSO) and coupling constants were determined (Table 4[link]). The conformational analysis of the puckering of the 2′-de­oxy­ribo­furanosyl moiety was performed using the PSEUROT program (Version 6.3; Van Wijk et al., 1999[Van Wijk, L., Haasnoot, C. A. G., de Leeuw, F. A. A. M., Huckriede, B. D., Westra Hoekzema, A. J. A. & Altona, C. (1999). PSEUROT 6.3. Leiden Institute of Chemistry, Leiden University, The Netherlands.]). This program calculates the population of N- and S-type conformers on the basis of five 3J(H,H) coupling constants, namely, 3J(H1′,H2′), 3J(H1′,H2′′), 3J(H2′,H3′), 3J(H2′′,H3′) and 3J(H3′,H4′). The coupling con­stants are summarized in Table 4[link] and the spectra are available in the supporting information.

Table 4
1H NMR chemical shifts, proton–proton vicinal and geminal coupling constants, and the conformation of nucleo­side sugar residues in solution

  Chemical shift/ppm
  H-1′ H-2′ H-2′′ H-3′ H-4′ H-5′ H-5′′ 3′-OH 5′-OH NH H5 H6
1 6.04 (dd) 2.50 (m) 1.81 (dt) 4.18 (td) 4.12 (td) 3.38 (m) 3.38 (m) 5.18 (d) 4.82 (t) 6.99 (s) 7.07 (s) 5.69 (d) 7.74 (d)
2 6.15 (dd) 2.10 (ddd) 1.92 (ddd) 4.19 (ddd) 3.76 (td) 3.54 (qdd) 3.54 (qdd) 5.18 (d) 4.95 (t) 7.10 (s) 7.15 (s) 5.71 (d) 7.78 (d)
  Coupling constant/Hz [J(H,H)] Conformation
  1′2′ 1′2′′ 2′2′′ 2′3′ 2′′3′ 3′4′ 4′5′ 4′5′′ 5′5′′ %N %S
1 7.5 2.8 −14.1 5.4 2.3 2.1 4.8 4.8 21 79
2 7.6 6.0 −13.3 6.0 3.2 3.1 4.0 4.0 −11.8 28 72
Measured in DMSO-d6 at 298 K; r.m.s. < 0.4 Hz. H-2′= H-2′β; H-2′′ = H-2′α. For PSEUROT (Van Wijk et al., 1999[Van Wijk, L., Haasnoot, C. A. G., de Leeuw, F. A. A. M., Huckriede, B. D., Westra Hoekzema, A. J. A. & Altona, C. (1999). PSEUROT 6.3. Leiden Institute of Chemistry, Leiden University, The Netherlands.]) calculations, the coupling constants 3J(H1′–H2′), 3J(H1′–H2′′), 3J(H2′–H3′), 3J(H2′′–H3′) and 3J(H3′–H4′) were used

The PSEUROT analysis of α-dC (1) and β-dC (2) revealed that both nucleo­sides prefer an S-type sugar conformation (72 and 79% S-type, respectively) in solution. Accordingly, α-dC (1) adopts the same sugar conformation (S-type) in solution and the solid-state. The S-conformation is also the preferred conformation of the canonical sugar residues as constituents of DNA. In this regard, the sugar residue of α-nucleo­side 1 fits into the DNA backbone (Fig. 5[link]).

[Figure 5]
Figure 5
N and S conformations of α-D and β-D nucleo­sides in solution. `B' corresponds to a nucleobase, with ax indicating axial and eq equatorial.

Moreover, the 1H NMR spectra of α-dC (1) and β-dC (2) show, in both cases, two signals for the amino protons (Table 4[link] and Figs. S1 and S2 in the supporting information). The appearance of two separated resonances for the amino protons indicates a hindered rotation about the C4—N4 bond due to partial double-bond character. Moreover, in the solid-state structure of α-dC (1), the C4—N4 bond is relatively short [1.337 (3) Å; Table 2[link]]. This suggests that the lone electron pair of the amino group is at least partially delocalized into the pyrimidine ring. These observations are consistent with earlier reports of 1-methyl­cytosine also reporting on the partial double-bond character of the amino group (Rossi & Kistenmacher, 1977[Rossi, M. & Kistenmacher, T. J. (1977). Acta Cryst. B33, 3962-3965.]; Fonseca Guerra et al., 2014[Fonseca Guerra, C., Sanz Miguel, P. J., Cebollada, A., Bickelhaupt, F. M. & Lippert, B. (2014). Chem. Eur. J. 20, 9494-9499.]).

4. Conclusion

In this work, the crystal structure of the α-anomeric analogue of 2′-de­oxy­cyti­dine (1) has been studied. α-2′-De­oxy­ribo­nucleo­sides are not widespread in nature as they are not part of canonical DNA. Literature reports on the conformational properties of α-2′-de­oxy­ribo­nucleo­sides are also limited. The single-crystal X-ray analysis of α-2′-de­oxy­cyti­dine revealed conformational properties which are outside the preferred range of α-nucleo­sides. This is rather unexpected as α-dC (1) is a rather `simple' α-nucleo­side without any further modifications at the nucleobase or sugar moiety. The anti conformation [χ = 173.39 (16)°] at the glycosylic bond is shifted to a higher χ value and the sugar moiety shows an almost symmetrical C2′-endo-C3′-exo twist ([^2_3T]; S-type), with P = 179.7°. The 2′-endo conformation is energetically less favoured in α-nucleo­sides com­pared to β-nucleo­sides, where this conformation is the preferred conformation of the DNA constituents. In addition, the C4—N4 bond between the amino group and the nucleobase is relatively short. Together with the appearance of two separated signals for the amino protons in the 1H NMR spectrum, this indicates a hindered rotation around the C4—N4 bond due to partial double-bond character. Within the crystal, the individual nucleo­side units of 1 are arranged in chains in a zigzag-like manner (ac plane). The crystal packing is controlled by N—H⋯O and O—H⋯O contacts between the nucleobase and sugar moieties. Moreover, two weak C—H⋯N contacts (C1′—H1⋯N3iii and C3′—H3′A⋯N3iv) are observed.

Although the flexibility at the glycosylic bond and the sugar conformation are generally more restricted for α-nucleo­sides, α-2′-de­oxy­cyti­dine (1) is an example of an α-nucleo­side with properties found outside the energetically favoured conformational range. This work constitutes a useful contribution to the field of nucleic acid chemistry and expands the state of knowledge on α-nucleo­sides.

Supporting information


Computing details top

Data collection: APEX3 (Bruker, 2016); cell refinement: SAINT (Bruker, 2015); data reduction: SAINT (Bruker, 2015); program(s) used to solve structure: SHELXS2014 (Sheldrick, 2015a); program(s) used to refine structure: SHELXL2014 (Sheldrick, 2015b); molecular graphics: APEX3 (Bruker, 2016); software used to prepare material for publication: APEX3 (Bruker, 2016) and XP (Bruker, 1998).

2'-Deoxycytidine top
Crystal data top
C9H13N3O4Dx = 1.513 Mg m3
Mr = 227.22Cu Kα radiation, λ = 1.54178 Å
Orthorhombic, P212121Cell parameters from 9144 reflections
a = 6.8378 (4) Åθ = 5.2–66.9°
b = 11.4334 (7) ŵ = 1.02 mm1
c = 12.7595 (8) ÅT = 100 K
V = 997.53 (11) Å3Prism, colourless
Z = 40.22 × 0.18 × 0.16 mm
F(000) = 480
Data collection top
Bruker APEXII Kappa CCD
diffractometer
1768 independent reflections
Radiation source: fine-focus sealed tube, fine-focus sealed tube1719 reflections with I > 2σ(I)
Graphite monochromatorRint = 0.037
Detector resolution: 8.3333 pixels mm-1θmax = 66.9°, θmin = 5.2°
φ and ω scansh = 88
Absorption correction: multi-scan
(SADABS; Bruker, 2014)
k = 1312
Tmin = 0.75, Tmax = 0.85l = 1515
11824 measured 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)] = 0.025 w = 1/[σ2(Fo2) + (0.0234P)2 + 0.3118P]
where P = (Fo2 + 2Fc2)/3
wR(F2) = 0.061(Δ/σ)max < 0.001
S = 1.12Δρmax = 0.12 e Å3
1768 reflectionsΔρmin = 0.18 e Å3
161 parametersAbsolute structure: Flack x determined using 684 quotients [(I+)-(I-)]/[(I+)+(I-)] (Parsons et al., 2013)
0 restraintsAbsolute structure parameter: 0.04 (9)
Primary atom site location: structure-invariant direct methods
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.

Refinement. Reflections were merged by SHELXL according to the crystal class for the calculation of statistics and refinement.

_reflns_Friedel_fraction is defined as the number of unique Friedel pairs measured divided by the number that would be possible theoretically, ignoring centric projections and systematic absences.

The hydrogens at N4, O3' and O5' atoms were refined freely.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
N10.8363 (2)0.56682 (14)0.36354 (13)0.0141 (4)
N30.9151 (2)0.75960 (14)0.30990 (13)0.0150 (4)
N40.8602 (3)0.91354 (15)0.42076 (14)0.0179 (4)
H4A0.851 (4)0.941 (2)0.484 (2)0.025 (7)*
H4B0.912 (4)0.960 (2)0.368 (2)0.041 (8)*
O20.9632 (2)0.60359 (12)0.20186 (11)0.0168 (3)
C20.9078 (3)0.64392 (17)0.28828 (15)0.0139 (4)
C40.8596 (3)0.79804 (17)0.40431 (15)0.0143 (4)
C50.8004 (3)0.72036 (16)0.48517 (15)0.0159 (4)
H5A0.77020.7480.55350.019*
C60.7887 (3)0.60552 (17)0.46094 (16)0.0154 (4)
H60.74670.55120.51260.018*
C1'0.8145 (3)0.44122 (16)0.33182 (15)0.0150 (4)
H10.94040.41170.30190.018*
C2'0.6507 (3)0.42608 (17)0.25202 (15)0.0154 (4)
H2A0.67790.35990.20420.018*
H2B0.6320.49810.21010.018*
C3'0.4725 (3)0.40127 (17)0.32044 (16)0.0155 (4)
H3A0.37380.35280.28230.019*
C4'0.5609 (3)0.33435 (17)0.41243 (15)0.0143 (4)
H40.48970.35570.47810.017*
C5'0.5584 (3)0.20248 (17)0.39946 (15)0.0170 (4)
H5B0.62420.1660.46030.02*
H5C0.42110.1750.39850.02*
O3'0.3873 (2)0.50722 (12)0.35979 (11)0.0179 (3)
H30.381 (5)0.559 (3)0.303 (3)0.054 (9)*
O4'0.7623 (2)0.37426 (11)0.42000 (11)0.0155 (3)
O5'0.6537 (2)0.16570 (12)0.30536 (11)0.0187 (3)
H50.786 (5)0.167 (3)0.316 (3)0.050 (9)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
N10.0156 (8)0.0115 (8)0.0152 (8)0.0007 (7)0.0008 (7)0.0004 (7)
N30.0166 (8)0.0129 (8)0.0154 (8)0.0008 (6)0.0002 (7)0.0003 (7)
N40.0262 (10)0.0129 (9)0.0147 (8)0.0017 (7)0.0015 (8)0.0008 (7)
O20.0211 (7)0.0146 (7)0.0149 (7)0.0002 (6)0.0040 (6)0.0016 (6)
C20.0124 (9)0.0144 (10)0.0150 (10)0.0007 (7)0.0003 (8)0.0027 (7)
C40.0131 (9)0.0139 (9)0.0157 (9)0.0001 (8)0.0015 (8)0.0001 (8)
C50.0187 (11)0.0162 (10)0.0129 (9)0.0001 (9)0.0001 (8)0.0009 (8)
C60.0167 (10)0.0162 (9)0.0134 (9)0.0003 (8)0.0008 (8)0.0019 (8)
C1'0.0196 (10)0.0110 (9)0.0145 (10)0.0005 (8)0.0025 (8)0.0008 (7)
C2'0.0206 (11)0.0127 (9)0.0130 (9)0.0001 (8)0.0003 (8)0.0010 (8)
C3'0.0180 (10)0.0132 (9)0.0152 (9)0.0003 (8)0.0022 (8)0.0027 (8)
C4'0.0149 (9)0.0127 (9)0.0154 (9)0.0011 (8)0.0018 (8)0.0018 (8)
C5'0.0191 (10)0.0138 (10)0.0181 (10)0.0007 (8)0.0039 (8)0.0001 (8)
O3'0.0222 (7)0.0142 (7)0.0173 (7)0.0044 (6)0.0008 (6)0.0004 (6)
O4'0.0181 (7)0.0133 (7)0.0151 (7)0.0022 (5)0.0021 (6)0.0032 (5)
O5'0.0191 (8)0.0156 (7)0.0214 (7)0.0009 (6)0.0025 (7)0.0037 (6)
Geometric parameters (Å, º) top
N1—C61.359 (3)C1'—H11.0
N1—C21.392 (3)C2'—C3'1.525 (3)
N1—C1'1.499 (2)C2'—H2A0.99
N3—C41.337 (3)C2'—H2B0.99
N3—C21.352 (3)C3'—O3'1.435 (2)
N4—C41.337 (3)C3'—C4'1.526 (3)
N4—H4A0.86 (3)C3'—H3A1.0
N4—H4B0.93 (3)C4'—O4'1.454 (2)
O2—C21.254 (2)C4'—C5'1.517 (3)
C4—C51.420 (3)C4'—H41.0
C5—C61.351 (3)C5'—O5'1.429 (2)
C5—H5A0.95C5'—H5B0.99
C6—H60.95C5'—H5C0.99
C1'—O4'1.407 (2)O3'—H30.94 (3)
C1'—C2'1.524 (3)O5'—H50.91 (4)
C6—N1—C2120.59 (16)C1'—C2'—H2A111.2
C6—N1—C1'122.33 (16)C3'—C2'—H2A111.2
C2—N1—C1'117.08 (15)C1'—C2'—H2B111.2
C4—N3—C2119.66 (17)C3'—C2'—H2B111.2
C4—N4—H4A120.2 (16)H2A—C2'—H2B109.1
C4—N4—H4B116.9 (17)O3'—C3'—C2'111.57 (16)
H4A—N4—H4B120 (2)O3'—C3'—C4'108.37 (16)
O2—C2—N3121.88 (17)C2'—C3'—C4'102.54 (16)
O2—C2—N1118.67 (17)O3'—C3'—H3A111.3
N3—C2—N1119.45 (17)C2'—C3'—H3A111.3
N4—C4—N3117.71 (18)C4'—C3'—H3A111.3
N4—C4—C5120.30 (18)O4'—C4'—C5'109.27 (16)
N3—C4—C5121.99 (18)O4'—C4'—C3'105.60 (15)
C6—C5—C4117.28 (18)C5'—C4'—C3'114.21 (16)
C6—C5—H5A121.4O4'—C4'—H4109.2
C4—C5—H5A121.4C5'—C4'—H4109.2
C5—C6—N1120.77 (18)C3'—C4'—H4109.2
C5—C6—H6119.6O5'—C5'—C4'112.26 (16)
N1—C6—H6119.6O5'—C5'—H5B109.2
O4'—C1'—N1109.30 (15)C4'—C5'—H5B109.2
O4'—C1'—C2'106.64 (16)O5'—C5'—H5C109.2
N1—C1'—C2'111.26 (16)C4'—C5'—H5C109.2
O4'—C1'—H1109.9H5B—C5'—H5C107.9
N1—C1'—H1109.9C3'—O3'—H3106.4 (19)
C2'—C1'—H1109.9C1'—O4'—C4'110.96 (15)
C1'—C2'—C3'103.04 (15)C5'—O5'—H5109 (2)
C4—N3—C2—O2178.13 (18)C2—N1—C1'—C2'69.1 (2)
C4—N3—C2—N12.1 (3)O4'—C1'—C2'—C3'27.75 (19)
C6—N1—C2—O2175.01 (18)N1—C1'—C2'—C3'91.35 (18)
C1'—N1—C2—O25.2 (3)C1'—C2'—C3'—O3'82.77 (18)
C6—N1—C2—N35.2 (3)C1'—C2'—C3'—C4'33.03 (18)
C1'—N1—C2—N3174.55 (18)O3'—C3'—C4'—O4'90.74 (17)
C2—N3—C4—N4176.89 (18)C2'—C3'—C4'—O4'27.35 (18)
C2—N3—C4—C52.8 (3)O3'—C3'—C4'—C5'149.18 (16)
N4—C4—C5—C6175.05 (18)C2'—C3'—C4'—C5'92.73 (19)
N3—C4—C5—C64.6 (3)O4'—C4'—C5'—O5'62.1 (2)
C4—C5—C6—N11.4 (3)C3'—C4'—C5'—O5'55.9 (2)
C2—N1—C6—C53.3 (3)N1—C1'—O4'—C4'109.54 (17)
C1'—N1—C6—C5176.44 (19)C2'—C1'—O4'—C4'10.83 (19)
C6—N1—C1'—O4'6.8 (3)C5'—C4'—O4'—C1'112.58 (17)
C2—N1—C1'—O4'173.39 (16)C3'—C4'—O4'—C1'10.69 (19)
C6—N1—C1'—C2'110.68 (19)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
N4—H4A···O3i0.86 (3)2.10 (3)2.949 (2)168 (2)
N4—H4B···O2ii0.93 (3)2.05 (3)2.937 (2)159 (2)
C1—H1···N3iii1.02.463.317 (3)144
C3—H3A···N3iv1.02.533.524 (3)170
O3—H3···O5v0.94 (3)1.86 (4)2.793 (2)175 (3)
O5—H5···O2iii0.91 (4)1.88 (3)2.716 (2)152 (3)
Symmetry codes: (i) x+1/2, y+3/2, z+1; (ii) x+2, y+1/2, z+1/2; (iii) x+2, y1/2, z+1/2; (iv) x+1, y1/2, z+1/2; (v) x+1, y+1/2, z+1/2.
1H NMR chemical shifts, proton–proton vicinal and geminal coupling constants and conformation of nucleoside sugar residues in solution top
H-1'H-2'H-2'H-3'H-4'H-5'H-5''3'-OH5'-OHNHH5H6
16.04 (dd)2.50 (m)1.81 (dt)4.18 (td)4.12 (td)3.38 (m)3.38 (m)5.18 (d)4.82 (t)6.99 (s) 7.07 (s)5.69 (d)7.74 (d)
26.15 (dd)2.10 (ddd)1.92 (ddd)4.19 (ddd)3.76 (td)3.54 (qdd)3.54 (qdd)5.18 (d)4.95 (t)7.10 (s) 7.15 (s)5.71 (d)7.78 (d)
1H NMR chemical shifts, proton–proton vicinal and geminal coupling constants and conformation of nucleoside sugar residues in solution top
1'2'1'2''2'2''2'3'2''3'3'4'4'5'4'5''5'5''%N%S
17.52.8-14.15.42.32.14.84.82179
27.66.0-13.36.03.23.14.04.0-11.82872
Measured in DMSO-d6 at 298 K; r.m.s. < 0.4 Hz. H-2'= H-2'β; H-2'' = H-2'α. For PSEUROT calculations the coupling constants 3J(H1'–H2'), 3J(H1'–H2'), 3J(H2'–H3'), 3J(H2'–H3') and 3J(H3'–H4') were used.
 

Acknowledgements

We thank Dr Peter Leonard for critical reading of the manuscript. We would like to thank Professor Dr B. Wünsch, Institut für Pharmazeutische und Medizinische Chemie, Universität Münster, for providing us with 600 MHz NMR spectra. Funding by ChemBiotech, Münster, Germany, is gratefully acknowledged. Open access funding enabled and organized by Projekt DEAL.

References

First citationAltona, C. & Sundaralingam, M. (1972). J. Am. Chem. Soc. 94, 8205–8212.  CrossRef CAS PubMed Web of Science Google Scholar
First citationBonnett, R. (1963). Chem. Rev. 63, 573–605.  CrossRef Web of Science Google Scholar
First citationBruker (1998). XP – Interactive molecular graphics. Version 5.1. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationBruker (2014). SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationBruker (2015). SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationBruker (2016). APEX3. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationChai, Y., Guo, X., Leonard, P. & Seela, F. (2020). Chem. Eur. J. 26, 13973–13989.  Web of Science CrossRef CAS PubMed Google Scholar
First citationChai, Y., Leonard, P., Guo, X. & Seela, F. (2019). Chem. Eur. J. 25, 16639–16651.  Web of Science CrossRef CAS Google Scholar
First citationFonseca Guerra, C., Sanz Miguel, P. J., Cebollada, A., Bickelhaupt, F. M. & Lippert, B. (2014). Chem. Eur. J. 20, 9494–9499.  Web of Science CAS PubMed Google Scholar
First citationGörbitz, C. H., Nelson, W. H. & Sagstuen, E. (2005). Acta Cryst. E61, o1207–o1209.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationGuo, X. & Seela, F. (2017). Chem. Eur. J. 23, 11776–11779.  Web of Science CrossRef CAS PubMed Google Scholar
First citationHamor, T. A., O'Leary, M. K. & Walker, R. T. (1977). Acta Cryst. B33, 1218–1223.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationIUPAC–IUB Joint Commission on Biochemical Nomenclature (1983). Eur. J. Biochem. 131, 9–15.  CrossRef PubMed Google Scholar
First citationKondhare, D., Budow-Busse, S., Daniliuc, C. & Seela, F. (2020). Acta Cryst. C76, 513–523.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationLatha, Y. S. & Yathindra, N. (1992). Biopolymers, 32, 249–269.  CrossRef PubMed CAS Web of Science Google Scholar
First citationMorvan, F., Rayner, B. & Imbach, J.-L. (1990). In Genetic Engineering, Vol. 12, edited by J. K. Setlow. New York: Plenum Press.  Google Scholar
First citationMorvan, F., Rayner, B., Imbach, J.-L., Chang, D.-K. & Lown, J. W. (1987a). Nucleic Acids Res. 15, 4241–4255.  CrossRef CAS PubMed Web of Science Google Scholar
First citationMorvan, F., Rayner, B., Imbach, J.-L., Lee, M., Hartley, J. A., Chang, D.-K. & Lown, J. W. (1987b). Nucleic Acids Res. 15, 7027–7044.  CrossRef CAS PubMed Web of Science Google Scholar
First citationMüller, S. L., Zhou, X., Leonard, P., Korzhenko, O., Daniliuc, C. & Seela, F. (2019). Chem. Eur. J. 25, 3077–3090.  PubMed Google Scholar
First citationNi, G., Du, Y., Tang, F., Liu, J., Zhao, H. & Chen, Q. (2019). RSC Adv. 9, 14302–14320.  Web of Science CrossRef CAS Google Scholar
First citationParsons, S., Flack, H. D. & Wagner, T. (2013). Acta Cryst. B69, 249–259.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationPost, M. L., Birnbaum, G. I., Huber, C. P. & Shugar, D. (1977). Biochim. Biophys. Acta Nucleic Acids Protein Synth. 479, 133–142.  CSD CrossRef CAS Web of Science Google Scholar
First citationPoznański, J., Felczak, K., Bretner, M., Kulikowski, T. & Remin, M. (2001). Biochem. Biophys. Res. Commun. 283, 1142–1149.  Web of Science PubMed Google Scholar
First citationRossi, M. & Kistenmacher, T. J. (1977). Acta Cryst. B33, 3962–3965.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationSaenger, W. (1984). In Principles of Nucleic Acid Structure, edited by C. R. Cantor. New York: Springer-Verlag.  Google Scholar
First citationSeela, F., Rosemeyer, H., Melenewski, A., Heithoff, E.-M., Eickmeier, H. & Reuter, H. (2002). Acta Cryst. C58, o142–o144.  Web of Science CSD CrossRef CAS IUCr Journals 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 citationSpackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19–32.  Web of Science CrossRef CAS Google Scholar
First citationSundaralingam, M. (1971). J. Am. Chem. Soc. 93, 6644–6647.  CrossRef CAS PubMed Web of Science Google Scholar
First citationSuzuki, S., Suzuki, K., Imai, T., Suzuki, N. & Okuda, S. (1965). J. Biol. Chem. 240, PC554–PC556.  CrossRef PubMed CAS Google Scholar
First citationThibaudeau, C. & Chattopadhyaya, J. (1997). Nucleoside Nucleotides Nucleic Acids, 16, 523–529.  CrossRef CAS Web of Science Google Scholar
First citationTurner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). CrystalExplorer17. University of Western Australia. https://hirshfeldsurface.netGoogle Scholar
First citationVan Wijk, L., Haasnoot, C. A. G., de Leeuw, F. A. A. M., Huckriede, B. D., Westra Hoekzema, A. J. A. & Altona, C. (1999). PSEUROT 6.3. Leiden Institute of Chemistry, Leiden University, The Netherlands.  Google Scholar
First citationYamaguchi, T. & Saneyoshi, M. (1984). Chem. Pharm. Bull. 32, 1441–1450.  CrossRef CAS Google Scholar
First citationYoung, D. W. & Wilson, H. R. (1975). Acta Cryst. B31, 961–965.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationZhou, X., Müller, S. L., Leonard, P., Daniliuc, C., Chai, Y., Budow-Busse, S. & Seela, F. (2019). J. Mol. Struct. 1190, 37–46.  Web of Science CSD CrossRef CAS 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 logoSTRUCTURAL
CHEMISTRY
ISSN: 2053-2296
Follow Acta Cryst. C
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds