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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Structure and NMR properties of the dinuclear complex di-μ-azido-κ4N1:N1-bis­­[(azido-κN)(pyridine-2-carboxamide-κ2N1,O)zinc(II)]

CROSSMARK_Color_square_no_text.svg

aInstituto de Ciencias, Benemérita Universidad Autónoma de Puebla, Av. San Claudio y 18 Sur, 72570 Puebla, Pue., Mexico, and bInstituto de Física, Benemérita Universidad Autónoma de Puebla, 72570 Puebla, Pue., Mexico
*Correspondence e-mail: sylvain_bernes@hotmail.com

Edited by C. Schulzke, Universität Greifswald, Germany (Received 27 November 2020; accepted 23 December 2020; online 8 January 2021)

The new diamagnetic complex, [Zn2(N3)4(C6H6N2O)2] or [Zn2(pca)2(μ1,1-N3)2(N3)2] was synthesized using pyridine-2-carboxamide (pca) and azido ligands, and characterized using various techniques: IR spectroscopy and single-crystal X-ray diffraction in the solid state, and nuclear magnetic resonance (NMR) in solution. The mol­ecule is placed on an inversion centre in space group P[\overline{1}]. The pca ligand chelates the metal centre via the pyridine N atom and the carbonyl O atom. One azido ligand bridges the two symmetry-related Zn2+ cations in the end-on coordination mode, while the other independent azido anion occupies the fifth coordination site, as a terminal ligand. The resulting five-coordinate Zn centres have a coordination geometry inter­mediate between trigonal bipyramidal and square pyramidal. The behaviour of the title complex in DMSO solution suggests that it is a suitable NMR probe for similar or isostructural complexes including other transition-metal ions. The diamagnetic nature of the complex is reflected in similar 1H and 13C NMR chemical shifts for the free ligand pca as for the Zn complex.

1. Chemical context

Polynuclear complexes have received the attention of coordination chemists as they are ideal candidates for developing new functional mol­ecular materials. In the design and preparation of such systems, a number of synthetic strategies have been used for propagating new motifs, affording a large number of polynuclear complexes with potential applications (Miller & Drillon, 2002[Miller, J. S. & Drillon, M. (2002). Magnetism: Molecules to Materials. Weinheim: Wiley-VCH.]). Complexes based on Zn2+ ions are of inter­est because of the versatility of this transition metal towards different kinds of chelating ligands, and its ability to bind ligands with different coordination numbers, ranging from two to six (Sakai et al., 2006[Sakai, K., Imakubo, T., Ichikawa, M. & Taniguchi, Y. (2006). Dalton Trans. pp. 881-883.]). Some complexes have been proposed as models for the active sites of zinc-containing enzymes (Parkin, 2000[Parkin, G. (2000). Chem. Commun. pp. 1971-1985.]; Döring et al., 2002[Döring, M., Ciesielski, M., Walter, O. & Görls, H. (2002). Eur. J. Inorg. Chem. pp. 1615-1621.]), while others have been studied for their catalytic properties (Dey et al., 2002[Dey, M., Rao, C. P., Saarenketo, P., Rissanen, K. & Kolehmainen, E. (2002). Eur. J. Inorg. Chem. pp. 2207-2215.]) or for the purpose of producing OLED devices (Sano et al., 2000[Sano, T., Nishio, Y., Hamada, Y., Takahashi, H., Usuki, T. & Shibata, K. (2000). J. Mater. Chem. 10, 157-161.]; Tokito et al., 2000[Tokito, S., Noda, K., Tanaka, H., Taga, Y. & Tsutsui, T. (2000). Synth. Met. 111-112, 393-396.]; Ray et al., 2012[Ray, S., Konar, S., Jana, A., Jana, S., Patra, A., Chatterjee, S., Golen, J. A., Rheingold, A. L., Mandal, S. S. & Kar, S. K. (2012). Polyhedron, 33, 82-89.]).

Upon coordination of a ligand to a metal centre, the ligand properties, such as electrophilic or nucleophilic character, acidity, susceptibility to oxidation or reduction, can be significantly altered, thereby enhancing or inhibiting its reactivity (Konidaris et al., 2012[Konidaris, K. F., Polyzou, C. D., Kostakis, G. E., Tasiopoulos, A. J., Roubeau, O., Teat, S. J., Manessi-Zoupa, E., Powell, A. K. & Perlepes, S. P. (2012). Dalton Trans. 41, 2862-2865.]). Co-ligands are also important for the structure and properties of the complex, especially if they can bridge metal centres. Among them, the azido ligand, N3, has been widely used in the building of mol­ecular magnetic materials with a rich diversity of topologies (Ribas et al., 1999[Ribas, J., Escuer, A., Monfort, M., Vicente, R., Cortés, R., Lezama, L. & Rojo, T. (1999). Coord. Chem. Rev. 193-195, 1027-1068.]; Hong & Chen, 2009[Hong, M.-C. & Chen, L. (2009). Design and Construction of Coordination Polymers. Chichester: John Wiley & Sons.]). The challenging aspect of N3 is its great coordination flexibility, which turns out to be rather a drawback since structures are poorly predictable. However, the correlation between the structures of polynuclear complexes including azido bridges and their magnetic properties is now well understood (Husain et al., 2012[Husain, A., Turnbull, M. M., Nami, S. A. A., Moheman, A. & Siddiqi, K. S. (2012). J. Coord. Chem. 65, 2593-2611.]; Yu et al., 2007[Yu, M.-M., Ni, Z.-H., Zhao, C.-C., Cui, A.-L. & Kou, H.-Z. (2007). Eur. J. Inorg. Chem. pp. 5670-5676.]).

The azido ion can link two or more metal ions in different configurations. The most representative are the end-to-end (EE) mode, in which two terminal N atoms bridge the metals, and the end-on (EO) mode, in which only one terminal N atom is used (Dori & Ziolo, 1973[Dori, Z. & Ziolo, R. F. (1973). Chem. Rev. 73, 247-254.]; Mautner et al., 2013[Mautner, F. A., Louka, F. R., Hofer, J., Spell, M., Lefèvre, A., Guilbeau, A. E. & Massoud, S. S. (2013). Cryst. Growth Des. 13, 4518-4525.]). Based on a survey of the CSD (Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]), the prevalence of the EO mode is much higher than the EE mode, by a factor of about ten. Mixed species having both terminal (i.e. non-bridg­ing) and EE/EO bridging azides are known, but are not so common. Several architectures occur depending on whether EE or EO bridges are present, which can be symmetric or asymmetric, single or multiple, and associated or not with other bridges (Goher et al., 2000[Goher, M. A. S., Cano, J., Journaux, Y., Abu-Youssef, M. A. M., Mautner, F. A., Escuer, A. & Vicente, R. (2000). Chem. Eur. J. 6, 778-784.]; Maji et al., 2001[Maji, T. K., Mukherjee, P. S., Chaudhuri, N. R., Mostafa, G., Mallah, T. & Cano-Boquera, J. (2001). Chem. Commun. pp. 1012-1013.]).

In this context, our group has paid attention to the synthesis of Zn2+ complexes including azido ligands, with the aim of using these diamagnetic compounds as NMR probes for other structurally related or analogous complexes. Herein, we report the mol­ecular structure of a dinuclear complex with bridging and non-bridging azido ligands, synthesized with picolinamide, a pyridine derivative with an amido group, suitable for the chelation of transition metals (Đaković et al., 2008[Đaković, M., Popović, Z., Giester, G. & Rajić-Linarić, M. (2008). Polyhedron, 27, 210-222.]).

[Scheme 1]

2. Structural commentary

The dinuclear complex [Zn2(N3)4(pca)2], where pca is picolinamide (IUPAC name: pyridine-2-carboxamide), crystallizes in the triclinic space group P[\overline{1}], with the mol­ecule placed on the inversion centre (Fig. 1[link]). The central [Zn2N2] core is thus planar by symmetry, with azido ligand N3/N4/N5 bridging the metal centres in the EO configuration. The double bridge is asymmetric, with Zn—N3 bond lengths of 2.057 (3) and 2.218 (3) Å (Table 1[link]). These bond lengths are comparable to those observed in other Zn2+ complexes bearing Schiff bases (Ray et al., 2012[Ray, S., Konar, S., Jana, A., Jana, S., Patra, A., Chatterjee, S., Golen, J. A., Rheingold, A. L., Mandal, S. S. & Kar, S. K. (2012). Polyhedron, 33, 82-89.]; Đaković et al., 2015[Đaković, M., Jaźwiński, J. & Popović, Z. (2015). Acta Chim. Slov. 62, 328-336.]; Sheng et al., 2014[Sheng, G.-H., Cheng, X.-S., You, Z.-L. & Zhu, H.-L. (2014). J. Struct. Chem. 55, 1106-1110.]), and are in agreement with IR spectroscopy data (You et al., 2009[You, Z.-L., Hou, P., Ni, L.-L. & Chen, S. (2009). Inorg. Chem. Commun. 12, 444-446.]; Qian & You, 2011[Qian, H.-Y. & You, Z.-L. (2011). J. Chem. Crystallogr. 41, 1593-1597.]). The pca mol­ecule behaves as a κ2-N,O-chelating ligand, forming a common five-membered metallacycle. This mode of coordination is almost universally found in other complexes including pca as ligand: there are very few occurrences of κ2-N,N-pca ligands reported so far in the CSD. Finally, each Zn centre coordinates one terminal azido ion, N6/N7/N8, with the short distance Zn—N6 = 1.991 (4) Å. Both independent azido ligands are nearly linear, and the bridging azido has a bent coordination with the metal centre. In the dinuclear complex, the Zn⋯Zn separation is 3.2760 (11) Å.

Table 1
Selected geometric parameters (Å, °)

Zn1—N6 1.991 (4) Zn1—O1 2.119 (3)
Zn1—N1 2.057 (3) Zn1—N3i 2.218 (3)
Zn1—N3 2.057 (3)    
       
N6—Zn1—N1 133.42 (14) N3—Zn1—O1 92.16 (12)
N6—Zn1—N3 112.66 (15) N6—Zn1—N3i 94.82 (14)
N1—Zn1—N3 113.87 (13) N1—Zn1—N3i 95.16 (13)
N6—Zn1—O1 98.33 (13) N3—Zn1—N3i 80.02 (12)
N1—Zn1—O1 77.87 (12) O1—Zn1—N3i 166.53 (12)
Symmetry code: (i) [-x+1, -y+1, -z+1].
[Figure 1]
Figure 1
Mol­ecular structure of the title compound, showing 50% probability displacement ellipsoids for non-H atoms. Non-labelled atoms are generated by the symmetry operation 1 − x, 1 − y, 1 − z.

The IR spectrum of the solid shows the stretching modes of coordinated pca ligands (Fig. 2[link]). The band at 1678 cm 1 is assigned to the νC= O vibration, which is shifted towards lower energy because of the C=O bond lengthening upon coord­ination [C6=O1: 1.250 (4) Å]. In contrast, the N—H stretching band of the amide group is not displaced in comparison to the free ligand, indicating that the NH2 group does not coordinate to Zn2+ ions (Konidaris et al., 2012[Konidaris, K. F., Polyzou, C. D., Kostakis, G. E., Tasiopoulos, A. J., Roubeau, O., Teat, S. J., Manessi-Zoupa, E., Powell, A. K. & Perlepes, S. P. (2012). Dalton Trans. 41, 2862-2865.]). The medium intensity band at 1296 cm−1 can be attributed to the νC N vibration in the pyridyl ring. The most useful IR vibrations are those related to azido ligands, which are clearly split over two frequencies, at 2094 and 2065 cm−1 (Fig. 2[link], inset). Based on previous reports in the literature, the former can be assigned to bridging-EO azido ligands and the latter to terminal azido ligands (Đaković et al., 2015[Đaković, M., Jaźwiński, J. & Popović, Z. (2015). Acta Chim. Slov. 62, 328-336.]; Forster & Horrocks, 1966[Forster, D. & Horrocks, W. D. (1966). Inorg. Chem. 5, 1510-1514.]). Similar intensities for these bands are in agreement with the X-ray structure. Finally, Zn—N vibrations give a low-intensity band at 412 cm−1 (Majumder et al., 2006[Majumder, A., Rosair, G. M., Mallick, A., Chattopadhyay, N. & Mitra, S. (2006). Polyhedron, 25, 1753-1762.]).

[Figure 2]
Figure 2
IR spectrum (KBr pellet) of the title complex, with assignment of the main bands. The inset is an expansion of the anti­symmetric stretching vibrations for azide groups.

The resulting dinuclear complex has five-coordinate Zn2+ ions, for which the Addison geometric parameter is τ5 = 0.55, midway between an ideal square-pyramidal (τ5 = 0) and a trigonal–bipyramidal geometry (τ5 = 1; Addison et al., 1984[Addison, A. W., Rao, T. N., Reedijk, J., van Rijn, J. & Verschoor, G. C. (1984). J. Chem. Soc. Dalton Trans. pp. 1349-1356.]). The strain caused by the five-membered metallacycle formed by the pca ligand [bite angle: 77.87 (12)°], together with the geometric restraint imposed by the central [Zn2N2] ring [N3—Zn1—N3i angle: 80.02 (12)°] account for the observed trigonal distortion. Such distortion has been observed in other similar dinuclear five-coordinate Zn2+ complexes bearing both terminal and bridging azido ligands: for nine complexes retrieved from the CSD, the Addison parameter ranges from τ5 = 0.40 (Sun & Wang, 2007[Sun, J.-Y. & Wang, X.-X. (2007). Acta Cryst. E63, m1140-m1141.]) to τ5 = 0.93 (Wang et al., 2004[Wang, L.-Y., Zhang, C.-X., Liao, D.-Z., Jiang, Z.-H. & Yan, S. P. (2004). Chin. J. Struct. Chem. 23, 171-175.]).

Non-covalent inter­molecular inter­actions are present in the crystal structure. Given that the NH2 groups in the pca ligands are not engaged in coordination, they form instead weak inter­molecular N—H⋯N hydrogen bonds with terminal N atoms of azide groups (Table 2[link]). These bonds form a 2D framework parallel to plane (100) in the crystal. The mol­ecules are then arranged in such a way that pyridyl rings are stacked in the [100] direction, with an offset face-to-face arrangement characterized by centroid-to-centroid distances for pyridyl rings of 4.702 (3) and 5.141 (3) Å along a stack (Fig. 3[link]).

Table 2
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
N2—H2A⋯N5ii 0.85 (2) 2.41 (3) 3.184 (5) 151 (4)
N2—H2B⋯N8iii 0.88 (2) 2.13 (2) 2.994 (5) 166 (4)
Symmetry codes: (ii) [-x+1, -y+1, -z]; (iii) [x, y+1, z-1].
[Figure 3]
Figure 3
Part of the crystal structure of the title complex showing the arrangement of N—H⋯N hydrogen bonds. The proximity between π systems is reflected in the inter­molecular C4⋯C4 separations, as measured using Mercury (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]; thin red lines): C4⋯C4i = 3.632 Å, and C4i⋯C4ii = 3.414 Å [symmetry codes: (i) 1 − x, 2 − y, −z; (ii) −1 + x, y, z]. The strongest inter­molecular hydrogen bond (Table 2[link], entry 2) is represented by blue dotted lines.

3. NMR measurements and chemical shift calculations

Using DMSO-d6 solutions of the free ligand pca and the title complex, 1H and 13C-NMR spectra were recorded on a Bruker Avance III 500 MHz spectrometer. Computationally, the geometry for the complex was optimized with the BLYP functional (Becke, 1993[Becke, A. D. (1993). J. Chem. Phys. 98, 5648-5652.]) and the 6-31+G(2d,p) basis to correlate the experimental structural information, time-dependent DFT, and NMR chemical shift estimations. Bond lengths and angles are similar in the DFT-optimized structure and in the X-ray crystal structure, validating the correctness of the calculations (GAUSSIAN09; Frisch et al., 2009[Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Mennucci, B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratchian, H. P., Izmaylov, A. F., Bloino, J., Zheng, G., Sonnenberg, J. L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery, J. A. Jr, Peralta, J. E., Ogliaro, F., Bearpark, M., Heyd, J. J., Brothers, E., Kudin, K. N., Staroverov, V. N., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Rega, N., Millam, J. M., Klene, M., Knox, J. E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E., Yazyev, O., Austin, A. J., Cammi, R., Pomelli, C., Ochterski, J. W., Martin, R. L., Morokuma, K., Zakrzewski, V. G., Voth, G. A., Salvador, P., Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, Ö., Foresman, J. B., Ortiz, J. V., Cioslowski, J. & Fox, D. J. (2009). GAUSSIAN09.. Revision D.01. Gaussian Inc., Wallingford, CT, USA. https://www.gaussian.com]). The shielding scales were converted to chemical shift scales by applying reference shielding of 32.0531 and 178.5112 ppm for 1H and 13C in TMS, respectively.

The 1H and 13C data of pca together with those of the complex are displayed in Table 3[link]. Moreover, 1H and 13C chemical shifts were calculated, allowing the assignment of all signals in the experimental spectra (Figs. 4[link] and 5[link]). The aromatic 1H spin systems are identified assuming doublet-like signals for H1 and H4, and triplet-like signals for H2 and H3. The presence of two NH broad signals with short relaxation times is due to the presence of the N and Zn atoms, which are more electronegative than H. The proton signals are slightly deshielded upon complexation, with the magnitude of deshielding decreasing while the distance from the metal centre increases. As seen in Fig. 4[link], the 3d10 cation does not affect the position of the signals very much. The most affected signals are those corresponding to the amide NH groups, which are shifted by ca 0.2 ppm and broadened upon coordination. This behaviour is probably related to different hydrogen-bonding schemes involving the NH2 group: free pca is strongly stabilized in the solid state by R22(8) ring motifs (Évora et al., 2012[Évora, A. O. L., Castro, R. A. E., Maria, T. M. R., Rosado, M. T. S., Silva, M. R., Canotilho, J. & Eusébio, M. E. S. (2012). CrystEngComm, 14, 8649-8657.]), which are no longer present once the mol­ecule is coordinated to the metal centre. The small influence of the metal centres on NMR properties is confirmed by experimental 13C-NMR chemical shifts, which are almost identical for pca and the title complex (Fig. 5[link]). However, a broadening is observed for the quaternary carbon atom C5, which is located in the close vicinity of the N and Zn sites, resulting in a very short relaxation time.

Table 3
1H-NMR (500 MHz) and 13C-NMR (125 MHz) chemical shifts, δ (ppm), and coupling constants JH—H (Hz), for the ligand pca and the diamagnetic complex [Zn2(N3)4(pca)2], in DMSO-d6

  1H-NMR (experimental) H⋯H coupling 1H-NMR (calculated) 13C-NMR (experimental) 13C-NMR (calculated)
Picolinamide H4: 8.01 d, J = 7.8 H4: 8.18 C4: 122.36 C4: 127.02
  H3: 7.97 td, J = 7.6, 1.7 H3: 7.94 C3: 138.12 C3: 142.08
  H2: 7.57 ddd, J = 7.5, 4.8, 1.3 H2: 7.56 C2: 126.94 C2: 130.65
  H1: 8.61 ddd, J = 4.7 H1: 8.68 C1: 148.92 C1: 153.41
  NH2: 8.11, 7.62 broad s NH2: 5.18, 7.66 C5: 150.66 C5: 155.63
        C6: 166.55 C6: 171.00
           
[Zn2(N3)4(pca)2] H4: 8.09 d, J = 7.7 H4: 8.00 C4: 122.51 C4: 128.35
  H3: 8.03 td, J = 7.6 H3: 8.33 C3: 138.56 C3: 146.74
  H2: 7.64 m H2: 7.92 C2: 127.24 C2: 134.07
  H1: 8.65 d, J = 4.6 H1: 9.08 C1: 148.93 C1: 154.42
  NH2: 8.29, 7.85 broad s NH2: 6.84, 6.13 C5: 149.81 C5: 149.48
        C6: 166.53 C6: 170.84
[Figure 4]
Figure 4
Experimental 1H-NMR spectra of pca (blue) and [Zn2(N3)4(pca)2] (red) in DMSO-d6. Chemical shifts and coupling constants are given in Table 3[link].
[Figure 5]
Figure 5
Experimental 13C-NMR spectra of pca (blue) and [Zn2(N3)4(pca)2] (red) in DMSO-d6. Chemical shifts are given in Table 3[link].

These data corroborate that proton chemical shifts for pca are only marginally affected by coordination to a diamagnetic metal centre as Zn2+. Very different spectra would be expected with paramagnetic centres, such as Mn2+, Co2+, or Cu2+. Most often, NMR spectra are difficult to inter­pret for these complexes, due to their broad and out of tune signals. However, our NMR data do not allow determination of whether the complex survives as a dimeric compound in solution, and whether the hydrogen bonding scheme observed in the crystal structure is retained in solution.

4. Synthesis and crystallization

An aqueous solution of pca (0.122 g, 1.0 mmol in 10 mL) was slowly poured onto an aqueous solution of Zn(SO4)·7H2O (0.287 g, 1.0 mmol in 10 mL) and an aqueous solution of NaN3 (0.130 g, 2.0 mmol in 5 mL). After one week at room temperature, colourless crystals formed in the mixture. Yield: 90%. Melting point: 471 K. The complex is soluble in water, DMSO, DMF and ethanol. IR data (cm−1, KBr pellet): 3394 (amide νN—H), 3302 (amide νsymN—H), 2094, 2065 (νasymm(N3)), 1678 (νC=O), 1570 (νC C), 1296 (νC N). UV–Vis (λmax/nm, H2O, ca 10−5 M): 215 (ππ*), 264 (n → π*).

5. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 4[link]. All C-bound H atoms were placed in calculated positions and refined as riding on their carrier C atoms, while amide H atoms were found in a difference map and refined with free orientation. The geometry of the NH2 group was restrained with distance targets N—H = 0.87 (2) Å, and isotropic displacement parameters for these H atoms were calculated as Uiso(H) = 1.2Ueq(N2).

Table 4
Experimental details

Crystal data
Chemical formula [Zn2(N3)4(C6H6N2O)2]
Mr 543.12
Crystal system, space group Triclinic, P[\overline{1}]
Temperature (K) 295
a, b, c (Å) 6.7689 (8), 8.3283 (10), 9.4835 (11)
α, β, γ (°) 69.942 (9), 75.447 (9), 75.901 (9)
V3) 478.74 (10)
Z 1
Radiation type Ag Kα, λ = 0.56083 Å
μ (mm−1) 1.35
Crystal size (mm) 0.27 × 0.04 × 0.03
 
Data collection
Diffractometer Stoe Stadivari
Absorption correction Multi-scan (X-AREA; Stoe & Cie, 2018[Stoe & Cie (2018). X-AREA and X-RED32. Stoe & Cie, Darmstadt, Germany.])
Tmin, Tmax 0.426, 1.000
No. of measured, independent and observed [I > 2σ(I)] reflections 11210, 2088, 1344
Rint 0.099
(sin θ/λ)max−1) 0.639
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.040, 0.078, 0.80
No. of reflections 2088
No. of parameters 151
No. of restraints 2
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.65, −0.52
Computer programs: X-AREA (Stoe & Cie, 2018[Stoe & Cie (2018). X-AREA and X-RED32. Stoe & Cie, Darmstadt, Germany.]), SHELXT2018/2 (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2018/3 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), ORTEP-3 for Windows (Farrugia, 2012[Farrugia, L. J. (2012). J. Appl. Cryst. 45, 849-854.]), Mercury (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]) and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]).

Supporting information


Computing details top

Data collection: X-AREA (Stoe & Cie, 2018); cell refinement: X-AREA (Stoe & Cie, 2018); data reduction: X-AREA (Stoe & Cie, 2018); program(s) used to solve structure: SHELXT2018/2 (Sheldrick, 2015a); program(s) used to refine structure: SHELXL2018/3 (Sheldrick, 2015b); molecular graphics: ORTEP-3 for Windows (Farrugia, 2012) and Mercury (Macrae et al., 2020); software used to prepare material for publication: publCIF (Westrip, 2010).

Di-µ-azido-κ4N1:N1-bis[(azido-κN)(pyridine-2-carboxamide-κ2N1,O)zinc(II)] top
Crystal data top
[Zn2(N3)4(C6H6N2O)2]F(000) = 272
Mr = 543.12Dx = 1.884 Mg m3
Triclinic, P1Melting point: 471 K
a = 6.7689 (8) ÅAg Kα radiation, λ = 0.56083 Å
b = 8.3283 (10) ÅCell parameters from 6594 reflections
c = 9.4835 (11) Åθ = 2.5–22.2°
α = 69.942 (9)°µ = 1.35 mm1
β = 75.447 (9)°T = 295 K
γ = 75.901 (9)°Needle, colourless
V = 478.74 (10) Å30.27 × 0.04 × 0.03 mm
Z = 1
Data collection top
Stoe Stadivari
diffractometer
2088 independent reflections
Radiation source: Sealed X-ray tube, Axo Astix-f Microfocus source1344 reflections with I > 2σ(I)
Graded multilayer mirror monochromatorRint = 0.099
Detector resolution: 5.81 pixels mm-1θmax = 21.0°, θmin = 2.5°
ω scansh = 88
Absorption correction: multi-scan
(X-AREA; Stoe & Cie, 2018)
k = 1010
Tmin = 0.426, Tmax = 1.000l = 1112
11210 measured reflections
Refinement top
Refinement on F2Primary atom site location: dual
Least-squares matrix: fullSecondary atom site location: difference Fourier map
R[F2 > 2σ(F2)] = 0.040Hydrogen site location: mixed
wR(F2) = 0.078H atoms treated by a mixture of independent and constrained refinement
S = 0.80 w = 1/[σ2(Fo2) + (0.0255P)2]
where P = (Fo2 + 2Fc2)/3
2088 reflections(Δ/σ)max < 0.001
151 parametersΔρmax = 0.65 e Å3
2 restraintsΔρmin = 0.52 e Å3
0 constraints
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Zn10.71118 (8)0.52257 (7)0.37011 (6)0.03385 (16)
O10.7775 (4)0.5513 (4)0.1331 (3)0.0381 (7)
N10.7228 (5)0.7831 (4)0.2758 (4)0.0297 (8)
N20.8242 (6)0.7454 (5)0.1008 (4)0.0412 (9)
H2A0.859 (6)0.663 (4)0.141 (5)0.049*
H2B0.827 (7)0.852 (3)0.164 (4)0.049*
N30.4244 (5)0.4627 (5)0.3951 (4)0.0346 (8)
N40.3234 (5)0.5162 (4)0.2967 (4)0.0378 (9)
N50.2268 (6)0.5678 (5)0.2013 (4)0.0562 (12)
N60.9229 (5)0.3132 (5)0.4351 (4)0.0430 (9)
N70.8669 (5)0.1971 (5)0.5432 (4)0.0390 (9)
N80.8183 (7)0.0821 (5)0.6474 (5)0.0589 (12)
C10.6936 (6)0.8957 (6)0.3551 (5)0.0366 (10)
H10.6782970.8537630.4609230.044*
C20.6855 (7)1.0727 (6)0.2841 (5)0.0419 (11)
H20.6670831.1477610.3415740.050*
C30.7048 (6)1.1356 (6)0.1285 (5)0.0413 (11)
H30.6969261.2540220.0793210.050*
C40.7364 (6)1.0207 (5)0.0448 (5)0.0347 (10)
H40.7507481.0600980.0610390.042*
C50.7458 (6)0.8456 (5)0.1235 (4)0.0290 (9)
C60.7848 (6)0.7035 (5)0.0490 (5)0.0303 (9)
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Zn10.0414 (3)0.0265 (3)0.0289 (3)0.0046 (2)0.0060 (2)0.0036 (2)
O10.0548 (19)0.0256 (16)0.0287 (16)0.0070 (14)0.0027 (13)0.0050 (13)
N10.0309 (18)0.0261 (19)0.0302 (19)0.0045 (15)0.0083 (14)0.0043 (15)
N20.066 (3)0.028 (2)0.027 (2)0.005 (2)0.0079 (18)0.0076 (17)
N30.044 (2)0.043 (2)0.0190 (18)0.0124 (18)0.0077 (16)0.0078 (16)
N40.049 (2)0.024 (2)0.038 (2)0.0150 (17)0.0050 (18)0.0025 (17)
N50.075 (3)0.056 (3)0.043 (2)0.015 (2)0.033 (2)0.003 (2)
N60.043 (2)0.031 (2)0.040 (2)0.0021 (17)0.0037 (17)0.0019 (18)
N70.041 (2)0.034 (2)0.038 (2)0.0014 (18)0.0099 (17)0.0098 (18)
N80.075 (3)0.036 (2)0.049 (3)0.012 (2)0.012 (2)0.010 (2)
C10.046 (3)0.036 (3)0.030 (2)0.004 (2)0.0099 (19)0.012 (2)
C20.051 (3)0.032 (3)0.047 (3)0.009 (2)0.011 (2)0.015 (2)
C30.046 (3)0.029 (2)0.045 (3)0.009 (2)0.006 (2)0.007 (2)
C40.040 (2)0.027 (2)0.032 (2)0.0039 (19)0.0071 (19)0.0024 (19)
C50.027 (2)0.028 (2)0.030 (2)0.0026 (17)0.0039 (17)0.0070 (18)
C60.027 (2)0.031 (2)0.030 (2)0.0026 (18)0.0058 (17)0.0056 (19)
Geometric parameters (Å, º) top
Zn1—N61.991 (4)N4—N51.151 (4)
Zn1—N12.057 (3)N6—N71.197 (5)
Zn1—N32.057 (3)N7—N81.159 (5)
Zn1—O12.119 (3)C1—C21.389 (6)
Zn1—N3i2.218 (3)C1—H10.9300
O1—C61.250 (4)C2—C31.370 (6)
N1—C51.339 (5)C2—H20.9300
N1—C11.342 (5)C3—C41.388 (6)
N2—C61.314 (5)C3—H30.9300
N2—H2A0.852 (19)C4—C51.385 (5)
N2—H2B0.884 (19)C4—H40.9300
N3—N41.193 (4)C5—C61.515 (5)
N6—Zn1—N1133.42 (14)N7—N6—Zn1117.3 (3)
N6—Zn1—N3112.66 (15)N8—N7—N6178.1 (5)
N1—Zn1—N3113.87 (13)N1—C1—C2122.1 (4)
N6—Zn1—O198.33 (13)N1—C1—H1119.0
N1—Zn1—O177.87 (12)C2—C1—H1119.0
N3—Zn1—O192.16 (12)C3—C2—C1119.2 (4)
N6—Zn1—N3i94.82 (14)C3—C2—H2120.4
N1—Zn1—N3i95.16 (13)C1—C2—H2120.4
N3—Zn1—N3i80.02 (12)C2—C3—C4119.3 (4)
O1—Zn1—N3i166.53 (12)C2—C3—H3120.3
C6—O1—Zn1114.5 (3)C4—C3—H3120.3
C5—N1—C1118.1 (4)C5—C4—C3118.1 (4)
C5—N1—Zn1116.5 (3)C5—C4—H4121.0
C1—N1—Zn1125.2 (3)C3—C4—H4121.0
C6—N2—H2A117 (3)N1—C5—C4123.1 (4)
C6—N2—H2B125 (3)N1—C5—C6112.3 (3)
H2A—N2—H2B117 (4)C4—C5—C6124.6 (4)
N4—N3—Zn1123.8 (3)O1—C6—N2122.8 (4)
N4—N3—Zn1i121.5 (3)O1—C6—C5118.5 (4)
Zn1—N3—Zn1i99.98 (12)N2—C6—C5118.7 (4)
N5—N4—N3179.7 (5)
C5—N1—C1—C20.4 (6)C3—C4—C5—N11.1 (6)
Zn1—N1—C1—C2174.2 (3)C3—C4—C5—C6178.3 (4)
N1—C1—C2—C31.0 (6)Zn1—O1—C6—N2179.9 (3)
C1—C2—C3—C41.3 (6)Zn1—O1—C6—C50.3 (4)
C2—C3—C4—C50.4 (6)N1—C5—C6—O14.7 (5)
C1—N1—C5—C41.4 (6)C4—C5—C6—O1175.8 (4)
Zn1—N1—C5—C4173.6 (3)N1—C5—C6—N2175.5 (4)
C1—N1—C5—C6178.0 (3)C4—C5—C6—N24.0 (6)
Zn1—N1—C5—C66.9 (4)
Symmetry code: (i) x+1, y+1, z+1.
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
N2—H2A···N5ii0.85 (2)2.41 (3)3.184 (5)151 (4)
N2—H2B···N8iii0.88 (2)2.13 (2)2.994 (5)166 (4)
C1—H1···N3i0.932.693.243 (5)119
C4—H4···N8iii0.932.643.541 (6)165
Symmetry codes: (i) x+1, y+1, z+1; (ii) x+1, y+1, z; (iii) x, y+1, z1.
1H-NMR (500 MHz) and 13C-NMR (125 MHz) chemical shifts, δ (ppm), and coupling constants JH—H (Hz), for the ligand pca and the diamagnetic complex [Zn2(N3)4(pca)2], in DMSO-d6 top
1H-NMR (experimental)H···H coupling1H-NMR (calculated)13C-NMR (experimental)13C-NMR (calculated)
PicolinamideH4: 8.01d, J = 7.8H4: 8.18C4: 122.36C4: 127.02
H3: 7.97td, J = 7.6, 1.7H3: 7.94C3: 138.12C3: 142.08
H2: 7.57ddd, J = 7.5, 4.8, 1.3H2: 7.56C2: 126.94C2: 130.65
H1: 8.61ddd, J = 4.7H1: 8.68C1: 148.92C1: 153.41
NH2: 8.11, 7.62broad sNH2: 5.18, 7.66C5: 150.66C5: 155.63
C6: 166.55C6: 171.00
[Zn2(N3)4(pca)2]H4: 8.09d, J = 7.7H4: 8.00C4: 122.51C4: 128.35
H3: 8.03td, J = 7.6H3: 8.33C3: 138.56C3: 146.74
H2: 7.64mH2: 7.92C2: 127.24C2: 134.07
H1: 8.65d, J = 4.6H1: 9.08C1: 148.93C1: 154.42
NH2: 8.29, 7.85broad sNH2: 6.84, 6.13C5: 149.81C5: 149.48
C6: 166.53C6: 170.84
 

Acknowledgements

We are thankful to the Laboratorio Nacional de Supercómputo del Sureste de México for computer time.

Funding information

Funding for this research was provided by: Vicerrectoría de Investigación y Estudios de Posgrado, Benemérita Universidad Autónoma de Puebla (award No. 100049155-VIEP-2019); Consejo Nacional de Ciencia y Tecnología (award No. 268178; scholarship No. 784569).

References

First citationAddison, A. W., Rao, T. N., Reedijk, J., van Rijn, J. & Verschoor, G. C. (1984). J. Chem. Soc. Dalton Trans. pp. 1349–1356.  CSD CrossRef Web of Science Google Scholar
First citationBecke, A. D. (1993). J. Chem. Phys. 98, 5648–5652.  CrossRef CAS Web of Science Google Scholar
First citationĐaković, M., Jaźwiński, J. & Popović, Z. (2015). Acta Chim. Slov. 62, 328–336.  Web of Science PubMed Google Scholar
First citationĐaković, M., Popović, Z., Giester, G. & Rajić-Linarić, M. (2008). Polyhedron, 27, 210–222.  Google Scholar
First citationDey, M., Rao, C. P., Saarenketo, P., Rissanen, K. & Kolehmainen, E. (2002). Eur. J. Inorg. Chem. pp. 2207–2215.  CrossRef Google Scholar
First citationDori, Z. & Ziolo, R. F. (1973). Chem. Rev. 73, 247–254.  CrossRef CAS Web of Science Google Scholar
First citationDöring, M., Ciesielski, M., Walter, O. & Görls, H. (2002). Eur. J. Inorg. Chem. pp. 1615–1621.  Google Scholar
First citationÉvora, A. O. L., Castro, R. A. E., Maria, T. M. R., Rosado, M. T. S., Silva, M. R., Canotilho, J. & Eusébio, M. E. S. (2012). CrystEngComm, 14, 8649–8657.  Google Scholar
First citationFarrugia, L. J. (2012). J. Appl. Cryst. 45, 849–854.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationForster, D. & Horrocks, W. D. (1966). Inorg. Chem. 5, 1510–1514.  CrossRef CAS Google Scholar
First citationFrisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Mennucci, B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratchian, H. P., Izmaylov, A. F., Bloino, J., Zheng, G., Sonnenberg, J. L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery, J. A. Jr, Peralta, J. E., Ogliaro, F., Bearpark, M., Heyd, J. J., Brothers, E., Kudin, K. N., Staroverov, V. N., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Rega, N., Millam, J. M., Klene, M., Knox, J. E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E., Yazyev, O., Austin, A. J., Cammi, R., Pomelli, C., Ochterski, J. W., Martin, R. L., Morokuma, K., Zakrzewski, V. G., Voth, G. A., Salvador, P., Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, Ö., Foresman, J. B., Ortiz, J. V., Cioslowski, J. & Fox, D. J. (2009). GAUSSIAN09.. Revision D.01. Gaussian Inc., Wallingford, CT, USA. https://www.gaussian.com  Google Scholar
First citationGoher, M. A. S., Cano, J., Journaux, Y., Abu-Youssef, M. A. M., Mautner, F. A., Escuer, A. & Vicente, R. (2000). Chem. Eur. J. 6, 778–784.  CrossRef PubMed CAS 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 citationHong, M.-C. & Chen, L. (2009). Design and Construction of Coordination Polymers. Chichester: John Wiley & Sons.  Google Scholar
First citationHusain, A., Turnbull, M. M., Nami, S. A. A., Moheman, A. & Siddiqi, K. S. (2012). J. Coord. Chem. 65, 2593–2611.  CSD CrossRef CAS Google Scholar
First citationKonidaris, K. F., Polyzou, C. D., Kostakis, G. E., Tasiopoulos, A. J., Roubeau, O., Teat, S. J., Manessi-Zoupa, E., Powell, A. K. & Perlepes, S. P. (2012). Dalton Trans. 41, 2862–2865.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationMacrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226–235.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMaji, T. K., Mukherjee, P. S., Chaudhuri, N. R., Mostafa, G., Mallah, T. & Cano-Boquera, J. (2001). Chem. Commun. pp. 1012–1013.  CSD CrossRef Google Scholar
First citationMajumder, A., Rosair, G. M., Mallick, A., Chattopadhyay, N. & Mitra, S. (2006). Polyhedron, 25, 1753–1762.  Web of Science CSD CrossRef CAS Google Scholar
First citationMautner, F. A., Louka, F. R., Hofer, J., Spell, M., Lefèvre, A., Guilbeau, A. E. & Massoud, S. S. (2013). Cryst. Growth Des. 13, 4518–4525.  CSD CrossRef CAS Google Scholar
First citationMiller, J. S. & Drillon, M. (2002). Magnetism: Molecules to Materials. Weinheim: Wiley-VCH.  Google Scholar
First citationParkin, G. (2000). Chem. Commun. pp. 1971–1985.  Web of Science CrossRef Google Scholar
First citationQian, H.-Y. & You, Z.-L. (2011). J. Chem. Crystallogr. 41, 1593–1597.  CSD CrossRef CAS Google Scholar
First citationRay, S., Konar, S., Jana, A., Jana, S., Patra, A., Chatterjee, S., Golen, J. A., Rheingold, A. L., Mandal, S. S. & Kar, S. K. (2012). Polyhedron, 33, 82–89.  CSD CrossRef CAS Google Scholar
First citationRibas, J., Escuer, A., Monfort, M., Vicente, R., Cortés, R., Lezama, L. & Rojo, T. (1999). Coord. Chem. Rev. 193–195, 1027–1068.  Web of Science CrossRef CAS Google Scholar
First citationSakai, K., Imakubo, T., Ichikawa, M. & Taniguchi, Y. (2006). Dalton Trans. pp. 881–883.  CSD CrossRef Google Scholar
First citationSano, T., Nishio, Y., Hamada, Y., Takahashi, H., Usuki, T. & Shibata, K. (2000). J. Mater. Chem. 10, 157–161.  CrossRef CAS 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 citationSheng, G.-H., Cheng, X.-S., You, Z.-L. & Zhu, H.-L. (2014). J. Struct. Chem. 55, 1106–1110.  CSD CrossRef CAS Google Scholar
First citationStoe & Cie (2018). X-AREA and X-RED32. Stoe & Cie, Darmstadt, Germany.  Google Scholar
First citationSun, J.-Y. & Wang, X.-X. (2007). Acta Cryst. E63, m1140–m1141.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationTokito, S., Noda, K., Tanaka, H., Taga, Y. & Tsutsui, T. (2000). Synth. Met. 111–112, 393–396.  Web of Science CrossRef CAS Google Scholar
First citationWang, L.-Y., Zhang, C.-X., Liao, D.-Z., Jiang, Z.-H. & Yan, S. P. (2004). Chin. J. Struct. Chem. 23, 171–175.  CAS Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationYou, Z.-L., Hou, P., Ni, L.-L. & Chen, S. (2009). Inorg. Chem. Commun. 12, 444–446.  Web of Science CSD CrossRef CAS Google Scholar
First citationYu, M.-M., Ni, Z.-H., Zhao, C.-C., Cui, A.-L. & Kou, H.-Z. (2007). Eur. J. Inorg. Chem. pp. 5670–5676.  Web of Science CSD CrossRef 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