Next Article in Journal
Zeolite-Like Boron Imidazolate Frameworks (BIFs): Synthesis and Application
Next Article in Special Issue
σ- and π-Hole Interactions
Previous Article in Journal
The Design Strategy for an Aggregation- and Crystallization-Induced Emission-Active Molecule Based on the Introduction of Skeletal Distortion by Boron Complexation with a Tridentate Ligand
Previous Article in Special Issue
Local Vibrational Mode Analysis of π–Hole Interactions between Aryl Donors and Small Molecule Acceptors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

N/N Bridge Type and Substituent Effects on Chemical and Crystallographic Properties of Schiff-Base (Salen/Salphen) Niii Complexes

by
Cynthia S. Novoa-Ramírez
1,†,
Areli Silva-Becerril
1,†,
Fiorella L. Olivera-Venturo
1,2,
Juan Carlos García-Ramos
1,3,
Marcos Flores-Alamo
4 and
Lena Ruiz-Azuara
1,*
1
Department of Inorganic Chemistry, Faculty of Chemistry, Universidad Nacional Autónoma de México, Av. Universidad 3000, Circuito Exterior S/N, Coyoacán, C.P. 04510 Ciudad de México, Mexico
2
Department of Science, Faculty of Science and Philosophy, Av. Honorio Delgado 430, 15102 San Martín de Porres, Peru
3
School of Health Sciences, Universidad Autónoma de Baja California, Campus Ensenada, Ensenada, 21100 Baja California, Mexico
4
USAII, Faculty of Chemistry, Universidad Nacional Autónoma de México, Av. Universidad 3000, Circuito Exterior S/N, Coyoacán, C.P. 04510 Ciudad de México, Mexico
*
Author to whom correspondence should be addressed.
Equal participation in this work.
Crystals 2020, 10(7), 616; https://doi.org/10.3390/cryst10070616
Submission received: 9 June 2020 / Revised: 13 July 2020 / Accepted: 13 July 2020 / Published: 15 July 2020
(This article belongs to the Special Issue σ- and π-Hole Interactions)

Abstract

:
In total, 13 ligands R-salen (N,N’-bis(5-R-salicylidene)ethylenediamine (where R = MeO, Me, OH, H, Cl, Br, NO2) and R-salphen (N,N’-bis(5-R-salicylidene)-1,2-phenylenediamine (where R = MeO, Me, OH, H, Cl, Br) and their 13 nickel complexes NiRsalen and NiRsalphen were synthesized and characterized using IR (infrared) spectroscopy, mass spectrometry, elemental analysis, magnetic susceptibility, NMR (nuclear magnetic resonance), UV-vis (ultraviolet-visible) spectroscopy, cyclic voltammetry, and X-ray crystal diffraction. Previous studies have shown that all complexes have presented a square planar geometry in a solid state and as a solution (DMSO). In electrochemical studies, it was observed that in N/N aliphatic bridge complexes, the NiII underwent two redox reactions, which were quasi-reversible process, and the half-wave potential followed a trend depending on the ligand substituent in the 5,5’-R position. The electron-donor substituent—as -OH, and -CH3 decreased the E1/2 potential—favored the reductor ability of nickel. The crystals of the complexes NiMesalen, NiMeOsalen, NiMeOsalphen, and Nisalphen were obtained. It was shown that the crystal packaging corresponded to monoclinic systems in the first three cases, as well as the triclinic for Nisalphen. The Hirshfeld surface analysis showed that the packaging was favored by H∙∙∙H and C∙∙∙H/H∙∙∙C interactions, and C-H∙∙∙O hydrogen bridges when the substituent was -MeO and π-stacking was added to an aromatic bridge. Replacing the N/N bridge with an aromatic ring decreased distortion in square-planar geometry where the angles O-Ni-N formed a perfect square-planar.

Graphical Abstract

1. Introduction

Salen type ligands, derived from the condensation between salicylaldehyde and a primary diamine, are considered versatile ligands in coordination chemistry because the steric and electronic properties can be modulated by different amine aldehydes precursors. The metal salen complexes have been widely studied in diverse areas of chemistry. The interest lays in its easy synthesis, versatility, and kinetic and thermodynamic stability provided by the chelating capacity of the tetradentate ligand with N2O2 donor atoms. For the aforementioned reasons, such complexes not only play an important role in coordination chemistry, but in various areas such as asymmetric catalysis [1], epoxide formation [2,3,4], olefin hydrogenation [5,6], or in polymerization reactions [7,8]. Moreover, they have been extensively studied in connection with metalloprotein models and, more recently, in bioinorganic chemistry. A complex of NiII, FeII, and CuII-salen has been synthesized as a biomimetic compound for the study of metalloenzyme active sites and their catalytic mechanism [9,10,11,12].
It is well known that the steric and electronic effect plays an important role in the reactivity of M-salen compounds. The incorporation of electron-donor/withdrawing groups on salen skeleton allows redox potential modulation [9,13] and favors certain geometries [14,15] that impact the interaction with the substrates or specific recognition sites. In recent years, attention has focused on the biological properties of Schiff bases and their metal-compounds [16], showing that ligands by themselves can inhibit a carbonic anhydrase enzyme. The efficiency of these enzymes depend on the properties of the bridge N/N [16]. Several metal-salen complexes have presented different biological activities, such as an antibacterial [17,18] and antiproliferative against different tumor lines [19,20,21,22]. One of the proposed mechanisms of action of these complexes with salen-type ligands is the specific interactions with DNA and RNA. Different types of damage can occur depending on the chemical reactivity of the metal complex. MnIII-salen in the presence of an oxidant induces DNA cleavages [23,24], whereas CoIII-salen can cleave DNA under aerobic conditions [25]. CoII-salen and FeIII-salen bind DNA at the minor groove [26], while NiII-salen causes damage to nucleic acids, specifically causing divisions in guanine residues in the hairpin region of a single chain [27]. On the other hand, it has been reported that square coordination compounds with conjugated systems in their coordination spheres present stacking interactions with DNA [28] and G-quadruplexes. The binding affinity and selectivity of interaction with the latter is modulated by changing the substituents on the salen skeleton and modifying the nature of the N/N-bridge for the Schiff base [29,30]. The correct choice of the N/N bridge plays an important role in the geometry of these compounds, as it lengthens the chain and favors octahedral geometry [31] while adding aromatic rings that favor planar-square geometry. Moreover, it increases π-π interactions [32]. These type of complexes also present a square-planar geometry [33] and multiple studies have shown how the reactivity of nickel-salen complexes can modulate by the design and control of the nature of the ligands by the choice of the N/N bridge or its substituents. Therefore, in this work, we study the effect the N/N bridge has on geometry and how the half-wave potential (E1/2) changes when a substitute is added. We report the structure of four Nickel complexes with tetradentate N2O2 ligands, analyzing the intermolecular interactions favored by the substituent and the N/N bridge, which modify the distance of Ni-Ni interaction found between dimers.

2. Materials and Methods

The experiments were carried out in ambient conditions. Nickel acetate tetrahydrate, salicylaldehyde, 2-hydroxi-5-metoxybenzaldehyde, 2-hydroxi-5-metylbenzaldehyde, 2-hydroxi-5-nitrobenzaldehyde, 2-hydroxi-5-clorobenzaldehyde, 2-hydroxi-5-bromorobenzaldehyde, and o-phenylenediamine were acquired from Sigma-Aldrich (Saint Louis, MO, USA), whereas 1,2-diaminoethane was acquired from Merck (Darmstadt, Germany.). The solvents used were acquired from Quimica Alvi (Ciudad de México, Mexico). All reactive materials were used without further purification. Elemental analysis was carried out in the Reach and Industry Support Services Unit (USAII for its Spanish abbreviations), using an EAGER 200 elemental analyzer (EAGER 200 CHNS/method, Ciudad de México, Mexico). IR (infrared) spectra were performed on a Nicolet AVATAR 320 FT-IR (Ciudad de México, Mexico) in an interval of 4000–400 cm−1. The samples incorporated onto a KBr disk in the range of 3500–500. UV-VIS (ultraviolet-visible) spectra were obtained on a Hewlett Packard 845× UV-visible system diode array spectrophotometer in a range from 250 nm to 800 nm in dimethylsulfoxide (DMSO) solutions Sigma-Aldrich (Saint Louis, MO, USA). The 1H-NMR 13C-NMR (nuclear magnetic resonance) (Ciudad de México, Mexico), COSY (correlation spectroscopy), and HSQC (heteronuclear single quantum correlation) (Tables S7 and S8, Figures S21–S39, Supplementary Materials), were performed with the USAII, collected by a VARIAN VNMRS 400 MHz. Chemical shifts were reported in ppm relative to the internal TMS (tetramethylsilane) standard. The solvents used were CDCl3, Acetone-d6, DMF-d7, and DMSO-d6, all of which were acquired from Sigma-Aldrich (Saint Louis, MO, USA). Mass spectrometers were acquired in the USAII (Ciudad de México, Mexico). All of the ligands and NiRsalen were obtained using FAB+ in a LECO PEGASUSIII. NMR and mass spectrometry were not obtained for NiClsalphen and NiBrsalphen due to their low solubility. Cyclic voltammetry was performed using PAR27 potentiostat/galvanostat (Ciudad de México, Mexico) with a conventional three-electrode array. Tetrabutylammonium hexafluorophosphate (Sigma-Aldrich, Saint Louis, MO, USA), served as a support electrolyte and DMSO (99.995, Sigma-Aldrich, Misuri, United State). Voltammogram were referenced with an internal adjustment using ferrocene (99.9%, Sigma-Aldrich, Saint Louis, MO, USA).

2.1. Abbreviation

Crystals 10 00616 i001

2.2. Synthesis of Schiff Base Ligands

Next, 2 mmol of appropriate salicylaldehyde was dissolved in acetonitrile, then 1 mmol of 1,2-diaminoethane or o-phenylenediamine was dissolved in acetonitrile and added slowly. The dissolution was stirred for 20 min. The volume was reduced and a solid precipitated (range of yellows in the case of 1,2-diaminoethane bridge and range of orange for o-phenylenediamine derivatives), which was vacuum filtered and recrystallized in methanol. The ligands were characterized by elemental analysis, FT-IR spectroscopy, NMR (1H and 13C), and mass spectrometry:
MeOsalen C18H20N2O4 analysis (%Calculated (Found)): C, 65.86 (66.07); H, 6.09 (6.02); N, 8.54 (8.84). FT-IR (cm−1): υC=N, 1639; υC-O, 1276. FT-IR values were comparable to those reported [34]. NMR 1H (ppm, CDCl3): 8.30 (s, CH=N), 12.63 (s, CAr-O-H); 13C (ppm): 166 (C=N), 155 (CAr-O). M+ (m/z): 328 (328).
Mesalen C18H20N2O2 analysis (%C(F)): C, 72.97 (73.10); H, 6.76 (6.38); N, 9.46 (10.05). FT-IR (cm−1): υC=N, 1637; υC-O, 1282. FT-IR values were comparable to those reported [35]. NMR 1H (ppm, CDCl3): 8.29 (s, CH=N), 12.95 (s, CAr-O-H); 13C (ppm): 166 (C=N), 159 (CAr-O). M+ (m/z): 296 (296).
OHsalen C16H16N2O4; Analysis (%C(F)): C,63.55 (63.85); H, 5,37 (5.03); N, 9.32 (9.47). FT-IR (cm−1): υC=N, 1640; υC-O, 1258. FT-IR values were slightly lower than those reported [36]. NMR 1H (ppm, Acet-d6): 8.46 (s, CH=N), 12.40 (s, CAr-O-H). Insoluble for NMR and mass spectrometry.
Salen C16H16N2O2 analysis (%C(F)): C,71.69 (71.62); H, 5.64 (6.01); N, 10.79 (10.44). FT-IR (cm−1): υC=N, 1636; υC-O, 1284. FT-IR values were comparable to those reported [37]. 1H NMR (ppm, CDCl3): 8.36 (s, CH=N), 13.19 (s, CAr-O-H); 13C (ppm): 166 (C=N), 161 (CAr-O). M+ (m/z): 268 (268).
Clsalen C16H14N2Cl2O2 analysis (%C(F)): C, 56.96 (57.13); H, 4.15 (3.88); N, 8.30 (8.52). FT-IR (cm−1): υC=N, 1631; υC-O, 1274. FT-IR values were slightly lower than those reported [38]. 1H NMR (ppm, CDCl3): 8.29 (s, CH=N), 13.08 (s, CAr-O-H); 13C (ppm): 165 (C=N), 159 (CAr-O). M+ (m/z): 336 (336).
Brsalen C16H14Br2N2O2 analysis (%C(F)): C, 45.07 (45.18); H, 3.29 (2.80); N, 6.57(7.22). FT-IR (cm−1): υC=N, 1635; υC-O, 1273. FT-IR values were comparable to those reported [34]. 1H NMR (ppm, CDCl3): 8.28 (s, CH=N), 13.10 (s, CAr-O-H); 13C (ppm): 165 (C=N), 160 (CAr-O). M+ (m/z): 426 (426).
NO2salen C16H14N4O6 analysis (%C(F)): C, 53.63 (53.69); H, 3.93 (3.59); N, 15.63 (15.76). FT-IR (cm−1): υC=N, 1647; υC-O, 1326. FT-IR values were comparable to those reported [39]. 1H NMR (ppm, DMSO-d6): 8.77 (s, CH=N). M+ (m/z): 358 (358).
MeOsalphen C22H20N2O4 analysis (%C(F)): C, 70.19 (70.30); H, 5.35 (5.02); N, 7.44 (7.95). FT-IR (cm−1): υC=N, 1616; υC-O, 1275. NMR 1H (ppm, Acet-d6): 8.85 (s, CH=N), 12.46 (s, CAr-O-H); 13C NMR (ppm): 166 (C=N), 155 (CAr-O). M + H+ (m/z): 376 (376). The structure of this ligand was already reported [40].
Mesalphen C22H20N2O2 analysis (%C(F)): C 76.72 (76.83); H, 5.85 (5.63); N, 8.13 (8.68). FT-IR (cm−1): υC=N, 1618; υC-O, 1283. FT-IR values were comparable to those reported [40] NMR 1H (ppm, CD3CN): 8.70 (s, CH=N), 12.86 (s, CAr-O-H); 13C NMR (ppm): 165 (C=N), 159 (CAr-O). M + H+ (m/z): 344 (345).
OHsalphen C20H16N2O4 analysis (%C(F)): C, 68.95 (68.96); H, 4.62 (4.34); N, 8.04 (8.26). FT-IR (cm−1): υC=N, 1614; υC-O, 1277. NMR 1H (ppm, Acet-d6): 8.76 (s, CH=N), 12.30 (s, CAr-O-H); 13C NMR (ppm): 165 (C=N), 155 (CAr-O). M + H+ (m/z): 348 (349).
Salphen C20H16N2O2 analysis (%C(F)): C,75.89 (75.93); H,5.0 (5.1); N,8.91 (8.85). FT-IR (cm−1): υC=N, 1612; υC-O, 1276. IR values were comparable to those reported [41]. M + H+ (m/z): 316 (317). Insoluble for NMR
Clsalphen C20H14N2Cl2O2 analysis (%C(F)): C, 62.45 (62.25); H, 3.66 (3.19); N, 7.27(8.04). FT-IR (cm−1): υC=N, 1614; υC-O, 1273. M + H+ (m/z): 384 (385). Insoluble for NMR. The structure of this ligand has already been reported [42].
Brsalphen C20H14N2Br2O2 analysis (%C(F)): C, 50.66 (50.74); H, 2.97 (2.57); N, 5.90 (6.42). FT-IR (cm−1): υC=N, 1612; υC-O, 1373. FT-IR values were comparable to those reported [18]. NMR 1H (ppm, Acet-d6): 8.91 (s, CH=N), 13.09 (s, CAr-O-H); 13C NMR (ppm): 164 (C=N), 160 (CAr-O). M + H+ (m/z): 474 (474).

2.3. Synthesis of Nickel Complexes

In total, 1 mmol of nickel acetate was dissolved in methanol and the ligand, previously dissolved in methanol/chloroform, was added the dropwise to the nickel acetate solution, in case the ligand NO2salen was dissolved in the DMF (dimethylformamide). The mixture of the reaction was stirred for 15 min and a solid compound precipitated. In the case of the NiRsalen complex, a brown-orange precipitate was obtained. For the NiRsalphen complexes, a red solid was precipitated. The solid compound was vacuum filtered and washed with methanol and chloroform. NiNO2salen was washed with cold DMF [43]. The compounds were characterized by elemental analysis, FT-IR spectroscopy, NMR (1H and 13C), and mass spectrometry:
NiMeOsalen, N,N’-bis(5-metoxisalicylidene)ethylenediamine, nickel(II), NiC18H18N2O4 · H2O analysis (%C(F)): C,53.63 (53.12); H, 5.00 (4.96); N, 6.95 (6.98). FT-IR (cm−1): υC=N, 1626; υC-O, 1328 FT-IR values were slightly lower than those reported [34]. NMR 1H (ppm, DMSO-d6): 7.80 (s, CH=N); 13C (ppm): 160 (C=N), 162 (CAr-O). M + H+ (m/z): 384 (385).
NiMesalen, N,N’-bis(5-metylsalicylidene)ethylenediamine, nickel(II), NiC18H18N2O2 analysis (%C(F)): C,61.19 (61.23); H, 5.10 (5.13); N, 7.93 (7.93). FT-IR (cm−1): υC=N, 1624; υC-O, 1316 FT-IR values were comparable to those reported [35]. NMR 1H (ppm, CDCl3): 7.25 (s, CH=N); 13C (ppm): 161 (C=N), 163 (CAr-O). M + H+ (m/z): 352 (353).
NiOHsalen, N,N’-bis(5-hidroxisalicylidene)ethylenediamine, nickel(II), NiC16H14N2O4 · 2H2O analysis (%C(F)): C, 48.94 (48.49); H, 4.15 (4.61); N, 7.39 (7.12). FT-IR (cm−1): υC=N, 1614; υC-O, 1301. NMR 1H (ppm, DMSO-d6): 7.75 (s, CH=N); 13C (ppm): 159 (C=N), 162 (CAr-O). M + H+ (m/z): 356 (357). The structure of this complex was already reported [44].
Nisalen, N,N’-bis(salicylaldehyde)ethylenediamine, nickel(II), NiC16H14N2O2 analysis (%C(F)): C, 58.77 (59.51); H, 3.89 (3.92); N, 8.57 (8.92). FT-IR (cm−1): υC=N, 1624; υC-O, 1320. NMR 1H (ppm, CDCl3): 7.38 (s, CH=N); 13C (ppm): 162 (C=N), 165 (CAr-O). M + H+ (m/z): 324 (325). The structure of this complex was already reported [33].
NiClsalen, N,N’-bis(5-chlorosalicylidene)ethylenediamine, nickel(II), NiC16H12N2Cl2O2 · H2O analysis (%C(F)): C, 46.22 (46.22); H, 3.42 (2.75); N, 6.80 (6.89). FT-IR (cm−1): υC=N, 1624; υC-O, 1312. FT-IR values were comparable to those reported [45]. NMR 1H (ppm, CDCl3): 7.44 (s, CH=N). M+H+ (m/z): 394 (394).
NiBrsalen, N,N’-bis(5-bromosalicylidene)ethylenediamine, nickel(II) NiC16H12Br2N2O2 · 2H2O analysis (%C(F)): C, 37.04 (36.85); H, 3.10 (3.05); N, 5.39 (5.46). FT-IR (cm−1): υC=N, 1626; υC-O, 1309 FT-IR values are comparable to those reported [34]. NMR 1H (ppm, CDCl3): 7.45 (s, CH=N). M + H+ (m/z): 482 (483).
NiNO2salen, N,N’-bis(5-nitrosalicylidene)ethylenediamine), nickel(II), NiC16H12N4O6 · 1.6H2O analysis (%C(F)): C, 42.98 (43.18); H, 3.11 (3.46); N, 12.74 (12.59). FT-IR (cm−1): υC=N, 1639; υC-O, 1321. FT-IR values were comparable to those reported [39]. NMR 1H (ppm, DMSO-d6): 7.95 (s, CH=N). Insoluble for mass spectrometry.
NiMeOsalphen (N,N’-bis(5-metoxisalicylidene)-1,2-phenylenediamine, nickel(II) NiC22H18N2O4 analysis (%C(F)): C, 61.01 (61.50); H, 4.18 (3.71); N, 6.46 (7.24). FT-IR (cm−1): υC=N, 1616; υC-O, 1213. NMR 1H (ppm, CDCl3): 8.23 (s, CH=N). M + H+ (m/z): 432 (433).
NiMesalphen, N,N’-bis(5-metylsalicylidene)-1,2-phenylenediamine, nickel(II) NiC22H18N2O2, analysis (%C(F)): C, 65.88 (66.30); H, 4.52 (4.02); N, 6.98 (7.31). FT-IR (cm−1): υC=N, 1624; υC-O, 1213. M + H+ (m/z): 400 (401). Insoluble for NRM. The structure of this complex was already reported [35].
NiOHsalphen, N,N’-bis(5-hidroxisalicylidene)-1,2-phenylenediamine, nickel(II) NiC20H14N2O4 · CH3OH analysis (%C(F)): C, 57.70 (57.63); H, 4.15 (3.66); N, 6.40 (6.62). FT-IR (cm−1): υC=N, 1610; υC-O, 1220. NMR 1H (ppm, DMF-7): 8.00 (s, CH=N); 13C (ppm): 162 (C=N), 155 (CAr-O). M+H+ (m/z): 404 (405).
Nisalphen, N,N’-bis(salicylaldehyde) -1,2-phenylenediamine, nickel(II) NiC20H14N2O2 analysis (%C(F)): C, 64.8 (64.4); H, 3.70(3.78); N, 7,32 (7.51). FT-IR (cm−1): υC=N, 1604; υC-O, 1295 FT-IR values were comparable to those reported [41]. M + H+ (m/z):372 (373). Insoluble for NRM.
NiClsalphen, N,N’-bis(5-chlorosalicylidene)-1,2-phenylenediamine, nickel(II), NiC20H12N2Cl2O2; analysis (%C(F)): C, 54.35 (55.59); H, 2.73 (2.28); N, 6.33 (7.16). FT-IR (cm−1): υC=N, 1608; υC-O, 1290. FT-IR values were comparable to those reported [38]. Insoluble for NRM and mass spectrometry.
NiBrsalphen, N,N’-bis(5-bromosalicylidene)-1,2-phenylenediamine, nickel(II), NiC20H12N2Br2O2; analysis (%C(F)): C, 45.25 (45.77); H, 2.27 (2.16); N, 5.27 (5.87). FT-IR (cm−1): υC=N, 1606; υC-O, 1328. FT-IR values were comparable to those reported [18]. Insoluble for NRM and mass spectrometry.

2.4. X-Ray Crystallography

Suitable single crystals for compounds NiMesalen, NiMeOsalen, NiMeOsalphen, and Nisalphen were mounted on a glass fiber. Crystallographic data were collected with an Oxford Diffraction Gemini "A" diffractometer with a CCD area detector, with λMoKα = 0.71073 Å for NiMesalen, NiMeOsalen, NiMeOsalphen, and λCuKα = 1.54184 Å for Nisalphen at 130 K. Unit cell parameters were determined with a set of three runs of 15 frames (1° in ω). The double pass method of scanning was used to exclude any noise [46]. The collected frames were integrated by using an orientation matrix determined from the narrow frame scans. Final cell constants were determined by a global refinement. Collected data were corrected for absorbance by using an analytical numeric absorption correction with a multifaceted crystal model based on expressions upon the Laue symmetry with equivalent reflections [47]. Structure solutions and refinement were carried out with the SHELXS-2014 [48] and SHELXL-2014 [49] packages. WinGX v2018.3 [50] software was used to prepare material for publication. Full-matrix least-squares refinement was carried out by minimizing (Fo2Fc2)2. All non-hydrogen atoms were refined anisotropically. H atoms attached to C atoms were placed in geometrically idealized positions and refined as riding on their parent atoms, with C-H = 0.95 – 0.99 Å and with Uiso(H) = 1.2Ueq(C) for aromatic and methylene groups, and 1.5Ueq(C) for methyl groups. On the other hand, for the compound NiMesalen, the solvent molecules were significantly disordered and could not be modeled properly (i.e., SQUEEZE [51]). Part of the PLATON package of crystallographic software was used to calculate the solvent disorder area and remove contributions to the overall intensity data. The disordered solvents area was centered around the 0.500–0.034 position and showed an estimated total of 60 electrons and a void volume of 180 Å3. Crystallographic data for all complexes are presented in Table 1. The crystallographic data for the structures reported in this paper was deposited with the Cambridge Crystallographic Data Centre as supplementary publication no. CCDC 2006691–2006694. Copies of the data can be obtained free of charge on application to CCDC, 12 Union Road, Cambridge, CB2 1EZ, UK. (fax: (+44) 1223-336-033, e-mail: [email protected]).

2.5. Cyclic Voltammetry

The cyclic voltammetry was carried out with a conventional arrangement of three electrodes: a vitreous carbon working electrode, a platinum counter electrode, and a silver pseud-electrode. The potentials were referenced to the saturated calomel electrode (SCE) with ferrocene as an internal standard (E°Fc+/Fc = +0.46 V vs. SCE). The experiments were collected in 0.001 M DMSO solutions under nitrogen atmosphere. The supporting electrolyte was 0.1 M of tetrabutylammonium hexafluorophosphate.

3. Results and Discussion

3.1. Electronic Spectra

The electronic spectra of the Rsalen ligand showed three absorption maxima. The first was the 255–270 nm region and the second was the 315–350 nm region. These two bands were attributed to π → π* transitions, with a high molar absorptivity coefficient. They also had a third band between 420–430 nm due to n → π* of the group C=N. Electronic spectra of Rsalphen were similar to Rsalen spectra. The difference was that Rsalphen spectra showed a small shoulder next to the 260 nm and Rsalphen ligands had another band in the 270–400 region due to the π → π* transition for the third aromatic ring in N/N bridge. All Rsalphen transitions were shifted to a major wavelength values and had bigger values of molar extinction coefficients, because the higher aromaticity of the ligands favored the delocalization of electron density.
NiRsalen showed four characteristics bands (Figures S9–S20, Supplementary Materials): the first two were in the 260–268 nm and 320–380 nm regions, with a high molar absorptivity coefficient, both due to π → π* transitions of the ligand. The third ws the 405–518 nm due to a ligand-metal charge transfer (LMCT) transition, from the phenolate to M due to 1A1g1Eg transition [52,53]. The last band in the 500–680 nm region was owed to the d-d transition [52,54]. These bands could not be characterized with precision because of the low solubility of the compounds, since they presented a very low molar absorptivity coefficient. These bands were attributed to 1A1g1A2g, which is characteristic for a square planar geometry. These electronic transitions were confirmed measuring the magnetic moment (µeff 0.5, Table S6, Supplementary Materials), meaning that the nickel complexes presented a diamagnetic property, consequences of the eight paired electrons. NiRsalphen had the same trend that their ligands and had one more π → π* transition due to the third aromatic ring and higher molar extinction coefficients (Data Table 2). All maxima shifted to major wavelength values [52,54].
The last band involved with the d-d transition provided an approximation of the intensity of the complex field, since the energy of this electronic transition was associated with 10 Dq. This band could not be observed for complexes with the imine aliphatic bridge since it was masked by high intensity transitions. On the other hand, in the compounds with the aromatic bridge, we observed that those with a substituted electron-withdrawn (-Br, -Cl) had a greater wavelength value, thus decreasing the energy necessary to carry out this transition, especially when compared to the substituted electron-donor (-MeO, -OH) [43].

3.2. X-Ray

From the single crystal X-ray diffraction analysis we found that the compounds NiMesalen and NiMeOsalen of NiII were polymorphs of (2,2’-(ethane-1,2-diylbis((nitrilo)methylylidene))bis(4-methylphenolato))-nickel(II) methanol solvate [35] and dinuclear bis(2,2’-(ethane-1,2-diylbis((nitrilo)methylylidene))bis(4-methoxyphenolato))-di-nickel(II) methanol solvate [35], respectively. Polymorphism was found in the crystalline arrangement, since the compound NiMesalen (Figure 1) crystallized in the monoclinic crystal system with a space group of C2/c, while in literature it was found that methanol solvated compound crystallized in the triclinic crystal system with a space group of P-1. On the other hand, NiMeOsalen crystallized in the monoclinic crystal system with a space group of P21/c, while the previously reported methanol solvated compound was space group P21/n and monoclinic crystal system [35].
Discrete unit NiMesalen contained one central NiII ion and one unit of deprotonated 2,2′-(ethane-1,2-diylbis((nitrilo)methylylidene))bis(4-methylphenolato) tetradentate ligand. Figure 1 shows the tetracoordinated metal center of NiII despites having a N2O2 coordination environment. Selected bond and angle parameters are given in Table 3.
The NiII center had almost a perfect square-planar geometry, which was defined by two N and two O atoms with τ4 of 0.02 and torsion angles O1-Ni1-N1-C8 (169.73(17)°), N1’-Ni1-N1-C7 (173.6(2)°), C2-O1-Ni1-N1 (7.8(2)°), and C2-O1-Ni1-O1’ (171.2(2)°). In fact, the Ni (II) atom was 0.016 Å out of the plane and formed by O1/O1’/N1’/N1. Each discrete molecule coplanar had an rms (root-mean squeare) of 0.030. There were intermolecular interactions of type C-H∙∙∙O, a hydrogen bond, and π∙∙∙π contacts that stabilized the crystal packing (Figure 2). Intermolecular interactions were established as follows: the hydrogen atom (C) carbon donor atom interacted with the (O) oxygen acceptor atom C8-H8A∙∙∙O1 (2.45 Å), thus forming a R12(4) motif along the c axes. In this same crystallographic direction, the interaction of type π∙∙∙π was represented by the centroid Cg4 and six membered ring C1/C6. The intermolecular contacts of the hydrogen bond and π-stacking array formed a bidimensional complex array along the a-b plane.
Figure 3 shows the perspective view of the molecular structure of polymorphic compound NiMeOsalen. The NiMeOsalen discrete unit contained one central NiII ion and one 2,2′-(ethane-1,2-diylbis((nitrilo)methylylidene))bis(4-methoxyphenolato) tetradentate ligand. The NiII ion was tetracoordinated with an N2O2 coordination environment. All bond length and angles corresponded to those reported in the literature [43]. Selected bond and angles parameters are given in Table 3.
The NiII center had a perfect square-planar geometry, defined by two N and two O atoms with rms of 0.004, τ4 of 0.006, and torsion angles O1-Ni1-N1-C8 (176.01(2)°), N2-Ni1-N1-C7 (176.33(2)°), C2-O1-Ni1-N1 (10.03 (2)°), and C2-O1-Ni1-O2 (170.32(2)°). In fact, the NiII atom was 0.005 Å out of the plane and formed by O1/O2/N2/N1. Nonetheless, it was observed that the six-membered rings deviated slightly from the coplanarity, thus finding an angle of 2.60(1)° between the aromatic rings.
Similarly with NiMesalen, in the crystalline arrangement for the compound NiMeOsalen, there were intermolecular interactions for the C-H∙∙∙O hydrogen bonding and intermolecular contacts of type π∙∙∙π (Figure 4). For the no classical hydrogen bond, these interactions were formed between the (C) carbon donor atom and two (O) oxygen acceptor atoms (C8-H8B∙∙∙O1 (2.61 Å) and C8-H8B∙∙∙O1(2.45 Å)), thus forming an R12(4) motif along the b axes. Additionally, there were C17-H17C∙∙∙O4 (2.75 Å) and C18-H18C∙∙∙O3 (2.82 Å), which formed an R22(6) motif along the b-c plane. The weak interaction π∙∙∙π had a distance of 3.95(8) Å between Cg4 and Cg5. Cg4 represent the six membered ring C1/C6 and Cg5 correspond to the C11/C16 ring. Finally, the intermolecular contacts of no classical hydrogen bond and π-stacking formed a tridimensional supramolecular array.
The compound NiMeOsalphen crystallized in a monoclinic crystal system with the space group P21/n. The asymmetric unit consisted of one nickel(II) center and one 2,2′-{1,2-phenylenebis[(azanylylidene)methylylidene]}bis(4-methoxyphenolato) ligand. The ORTEP diagram is shown in Figure 5. The squared plane of NiII center was chelated by two oxygen and two nitrogen atoms that derived from a salen ligand, with Ni-O and Ni-N bond distances ranging from 1.8394 to 1.8643 (13) Å. O-Ni-O, N-Ni-N, and O-Ni-N bond angles of 84.12 to 179.34°. The length distance of Ni-N was, on average, 1.8634(12) Å (Table 3), which was slightly higher than that observed in compounds NiMesalen and NiMeOsalen. Nevertheless, the tetracoordinate NiII in compound NiMeOsalphen had a square plane geometry with a rms of 0.0147 and a τ4 of 0.020. An analysis of the coplanarity shows that there were angles of 3.07 (7) and 5.06 (7)° between the square plane N2O2 at the metal center. The planes formed by the six-membered rings C1/C6 and C11/C16, respectively. Additionally, there was a perfect coplanarity between the square plane N2O2 at the metal center and the ring formed by the C8-C9-C1 /C22 atoms with an angle of 0.52 (7)°.
In the crystal packing, there were C-H∙∙∙O no classic hydrogen bonds and π∙∙∙π intermolecular contacts (Figure 6).
The interactions of type hydrogen bond were observed between C17-H17C∙∙∙O2 (2.43 Å), C18-H18C∙∙∙O3 (2.50 Å), and C21-H21∙∙∙O3 (2.42 Å). These intermolecular contacts formed R22(30) and R22(11) motifs along the a-c plane. Moreover, there were weak π-π interactions with a distance of centroids Cg4-Cg6 (3.84 Å) and Cg5-Cg6 (3.56 Å). Cg4 represented the six membered ring C1/C6, while Cg5 corresponded to the C8-C9-C19/C22. Cg6 was formed by the C11/C16 ring. Finally, all intermolecular contacts formed a tridimensional supramolecular array.
Unlike previously mentioned compounds NiMesalen, NiMeOsalen, and NiMeOsalphen, the single crystal X-ray diffraction analysis revealed that compound Nisalphen crystalized in the triclinic space group P-1. The asymmetric unit of Nisalphen contained two molecules of the nickel coordination compound and three molecules of the chloroform, which was then used as a solvent (Figure 7). Each metal central of NiII ion was tetracoordinated with one unit of a deprotonated salphen ligand with salphen = N,N′-o-phenylenebis(salicylideneimine). Selected bond and angles parameters are given in Table 3.
While investigating the plane formed by three aromatic rings and the square symmetry N2O2 at the metallic center of NiII, we found a coplanarity in each molecule with rms of 0.060 and 0.019 Å for molecules A and B, respectively. Furthermore, these molecules had a parallel arrangement between them, with an angle of 1.4°.
In molecule A, there was a perfect square planar geometry with a τ4 de 0.026 with torsion angles O1A-Ni1A-N1A-C8A (179.72(15)°), N2A-Ni1A-N1A-C7A (177.8(2)°), C2A-O1A-Ni1A-N1A (0.90 (2)°), and C2A-O1A-Ni1A-O2A (179.6(2)°). Similarly, molecule B had a perfect square planar geometry (τ4 de 0.024) with torsion angles O1B-Ni1B-N1B-C8B (179.26(15)°), N2B-Ni1B-N1B-C7B (178.8(2)°), C2B-O1B-Ni1B-N1B (6.0(2)°), and C2B-O1B-Ni1B-O2B (173.7(2)°).
In the crystalline arrangement of compound Nisalphen, a Ni-Ni distance of 3.26 Å was observed. The short distance found between both metal centers was favored by the interaction of the π∙∙∙π. There were electronic densities in the coplanar and parllel A-A and A-B molecules.
This system obtained dinuclear structural arrangements with possible applications in molecular modeling and bioinorganic systems. Additionally, there were intermolecular interactions for C-H∙∙∙O hydrogen bonding. Figure 8 shows the crystalline array with intermolecular contacts.
For the no classical hydrogen bond, interactions were formed between the carbon donor atom of the chloroform molecule solvent and the two oxygen acceptor atoms of the salphen ligand: C41-H41∙∙∙O1A (2.26 Å), C41-H41∙∙∙O2A (2.26 Å), C51-H51∙∙∙O1B (2.29 Å), and C51-H51∙∙∙O2B (2.19 Å). These show a linear, bifurcated, and trifurcated form for two, three, and four centers, respectively, in the intermolecular interaction. Additionally, there were C51∙∙∙H16A (2.99 Å) and C52∙∙∙H14B (3.01 Å) intermolecular contacts.
Despite the small differences on the NiII-donor atoms length, the nature of the N/N bridge and electron-donor/withdrawn character of the substituents in the 5- and 5′-position of the Schiff base play a key role in packing NiII coordination compounds. This can be observed in the Ni-Ni distance found in the different crystal structures obtained here and those previously reported.
Another important factor that influenced crystalline packing was the solvent. The Nisalen solvate reported by Siegler et al. showed a crystalline arrangement; the interactions stabilizing the crystal depended on it. It favored dimers when the Ni-Ni distances were modified according to acetone, 3.16 Å, CHCl3, 3.13 Å (system monoclinic); CHCl3, 3.19 Å (system orthorhombic); CH2Cl2, 3.28 Å; C2H4O2, 3.37 Å; DMF, 3.3901 Å; or the favor 1-D chain, as was the case for the methanol solvate Ni-Ni 3.44 Å, wherein the solvent joined the monomers through C-H∙∙∙O interactions in one direction [33].
Comparing the intermolecular interactions found on Nisalen and Nisalphen, the incorporation of an extra aromatic ring in the ligand structure increased the number on the π∙∙∙π and C-H∙∙∙π interactions. In the two Nisalen structures reported with the same crystalline system (triclinic), the π∙∙∙π interaction found a length of 3.63 [55] and 4.43 Å [33]. Meanwhile, the two C-H∙∙∙π were observed. On the other hand, the Nisalphen structure presented two π∙∙∙π interactions with lengths of 3.89 and 4.55 and four C-H∙∙∙π interactions of 3.22, 3.39, 3.65, and 3.68 Å. The sum of all interactions led to a Ni-Ni distance of 3.26 Å, which was slightly smaller than the length found in both Nisalen with 3.63 and 3.36 Å, respectively.
Substituents also played an important role in crystal packing. NiMeOsalphen presented three π∙∙∙π interactions (i.e., 3.56, 3.65, and 3.84 Å), three C-H∙∙∙ π interactions (3.24, 3.33 and 3.38 Å), and a C-H∙∙∙O interaction with a length of 2.50 Å. For NiMeOsalen, only one π∙∙∙π interaction of 3.95 Å was observed. There were two C-H∙∙∙π interactions with lengths of 3.61 and 3.77 Å and two C-H∙∙∙O interactions with lengths of 2.75 and 2.82 Å. The Ni-Ni distance observed in these examples could be closely related with the C-H∙∙∙O and π∙∙∙π interactions from the methoxy groups and the extra aromatic ring for NiMeOsalphen. The Ni-Ni distance on NiMeOsalphen could be longer than NiMeOsalen due to the π∙∙∙π interaction found between the two dimeric units.
For NiMesalen, the π∙∙∙π interaction was retained but the main contribution for the crystal stabilization relied on the C-H∙∙∙π interaction with distance values of 3.64 and 3.66 Å. These interactions kept the two units close enough to establish a Ni-Ni distance of 3.39 Å. The C-H∙∙∙O interactions elicited by the methoxy groups contributed a shorter Ni-Ni distance for NiMeOsalen (3.18 Å) than NiMesalen (3.39 Å). The same was observed for the compounds NiMeOsalphen and NiMesalphen [35].
In the crystalline structure NiOHsalen, two interactions of π-π were shown. However, the -OH groups in the structure stabilized the crystalline packing mainly by the interaction of the hydrogen bridges for the Osolvent-H∙∙∙Osalen and Osalen-H∙∙∙Osolvent with methanol molecules [44]. The Ni-Ni distance was 3.61 Å, which, when compared to NiMeOsalen (3.18 Å), increased because of the sovlent’s role in the packing. One methanol molecule formed a hydrogen bridge interaction with two neighboring molecules, Osolvent-H∙∙∙Osalen and Osalen-H∙∙∙Osolvent [44]. These solvent interactions also occurred in the Nisalen structure when methanol was the solvate [33].

3.3. Hirshfeld Surface Analysis

Hirshfeld’s surface (HS) analysis provided detailed information regarding intermolecular interactions. A better understanding of the problem may help address the challenge of quantitatively understanding intermolecular contacts using visual information on color and shadow on surfaces [56].
The Crystal Explorer 17 program [57] was used to generate the HS and 2D fingerprint plots of the complexes (i.e., NiMesalen, NiMeOsalen, NiMeOsalphen, and Nisalphen). The dnorm HS was obtained, which combined the normalized distances from the closer atom inside the surface (di) and outside the surface (de) to the HS, showing all contacts of the crystal structure. The red regions indicate the contacts were shorter than the sum of the van der Waals radii of the involved atoms. The blue and white regions indicated that the contacts were longer and closer to the van der Waals limit. Figure 9 shows the HS and all compound interactions.
The dnorm HS of the compounds showed red spots, which indicated close-contacts in the crystal structure, i.e., non-classical hydrogen bonds C-H∙∙∙O and π∙∙∙π, as well as intermolecular interactions between centroids of six-membered rings in phenyl groups. The shape index was a function of HS and very helpful when investigating the π∙∙∙π stacking interaction. The blue and red zone indicated a region with a stacking arrangement. Figure 10 presents the shape index mapped on the compounds’ HS. The blue zone indicated the presence of π∙∙∙π stacking interactions in the crystal structure. The π∙∙∙π interaction in compound Nisalphen stabilized and favored the 3.26 Å distance between the NiII metal centers, due to the presence of molecules A and B in the asymmetric unit of Nisalphen. Figures S1–S8 present the details of the fingerprint plots for each compound. In them, they describe the intermolecular interactions around the HS.
Figure 11 shows the contributions of contacts obtained from the decomposition of the fingerprint plots. The fingerprint plots of NiMesalen, NiMeOsalen, and NiMeOsalphen were similar, indicating that the H∙∙∙H and C∙∙∙H/H∙∙∙C were the most important contributors for crystal packing. H∙∙∙H contacts contributed 64.4% (NiMesalen), 46.4% (NiMeOsalen), and 32.4% (NiMeOsalphen), while C∙∙∙H/H∙∙∙C contacts contributed 16% (NiMesalen), 21.2% (NiMeOsalen), and 20% (NiMeOsalphen). A similar trend was observed in the fingerprint plot for Nisalphen, where the H∙∙∙H and X∙∙∙H/H∙∙∙X contacts had greater contributors for stabilizing interactions, with H∙∙∙H contacts contributing 32.5% and 27.5% in molecules A and B, respectively. The contributions for C∙∙∙H/H∙∙∙C, O∙∙∙H, and C∙∙∙C contacts were approximately of 20%, 5%, and 8% for molecules A and B, while the Cl∙∙∙H/H∙∙∙Cl contact contributed 19.9% and 23.8% in molecules A and B, respectively.

3.4. Cyclic Voltammetry

NickelII-salen compounds have a neutral charge and show low solubility. Adding an extra aromatic ring in the ligand structure (i.e., salphen-type ligands) causes the solubility to decrease even more. When the NiRsalphen solution bubbled with nitrogen, it started to precipitate. Because the low solubility of the compounds, it was only possible to characterize NiRsalen compounds in the electrochemical study.
We performed voltammetry of the ligands (Figure 12b). The salen ligand ran in the direction of the positive potential in an interval of −3.2 to 1.0 V. In an inversion study, reduction signals 3a and 3b were associated with C=N reduction and an irreversible oxidation signal, 4a [58]. Figure 12a shows the Nisalen voltammogram under the same condition, caused by the nickel oxidation process ([NiIIL] → [NiIIIL ] + 1 ē) and 2a and 2b due to nickel reduction process ([NiIIL] + 1 ē → [NiIL]) [14,59]. Signal 3a and 3b was also observed to shift lower potential values. The other NiRsalen complexes showed a similar behavior with the signals that shifted to different potentials due to the substituent in the 5,5′position (Figures S40–S46, Supplementary Materials). In this work, only the processes associated with the reduction and oxidation of nickel were reported. The voltammograms were run in an interval of −2.4 to 1.0 V (Figure 12c,d).
The cathodic and anodic peak current were plotted in the square root function of the sweep speed (ʋ1/2). Only the complexes NiMeOHsalen, NiOHsalen, NiClsalen, and NiBrsalen presented a linear dependence, which means that the oxidation of nickel was a diffusion-controlled process. A coupled reaction was suggested to impact the reversible process, thus confirming that Nisalen and NiMesalen via plotting ipc/ipa vs. logV. The oxidation process for the complexes were irreversible due to the ∆E being too big. The electron transference was a slow process, as is shown in Table 4. The oxidation process involved an EC mechanism and the NiIII-salen complex coordinates solvents, such as DMSO, in their axial position to stabilize the NiIII oxidations for electronic density [14,60].
With regard to the oxidation process, the reductions were a quasi-reversible process and we found that all nickel complex reductions were diffusion-controlled processes, except for NiOHsalen, which presented coupled reactions. In comparation with oxidation reactions, the reduction of NiII was a more quantitative process. ∆E values were close to 59 mV and the ipc/ipa ratio was closer to 1 (Table 4).
For both processes, we found a trend between E1/2 and the effect of the substituent. Correlations were made with the Hammett sigma in the para-position. The metal center’s acidity was influenced by the effect of the substituent. Therefore, the oxidative and reductive capacity of nickel modulated with the correct use of these substituents [13,61,62,63]. Electron-donor substituents shifted the E1/2 to a lower potential value and the electron-withdrawn groups shifted toward a more positive potential value. Thus, an electron-donor group improved the reductive capacity and electron-withdrawn groups improved the oxidative capacity of nickel, as shown in Figure 13.

4. Conclusions

In this study, various Schiff bases and their NiII complexes were synthesized. All the prepared ligand and complexes were analyzed via C, H, and N analyses. They were assigned molecular structures and geometries using information obtained from UV-Vis, magnetic susceptibility, and X-ray crystallography, all of which corresponded to square-planar geometry in the solid state. The Hirshfeld surface analysis was used to study intermolecular interactions. This analysis revealed that the O∙∙∙H, H∙∙∙H and π∙∙∙π contacts were the most significant in the crystal array of the compounds NiMesalen, NiMeOsalen, and NiMeOsalphen, and O∙∙∙H, H∙∙∙H, Cl∙∙∙H and π∙∙∙π contacts in the crystal array of the compound Nisalphen. The no classical hydrogen bonding and π∙∙∙π stacking information conveyed by Hirshfeld surface analysis were consistent with the crystal structure analysis. The substituents and the N/N bridge affected the crystal packing and electronic properties of nickel. According to the structures obtained for Nisalphen, NiMeOsalen, and NiMeOsalphen, it was possible to observe that the addition of an aromatic ring in the N/N bridge increased the number on π∙∙∙π and C-H∙∙∙π interactions and decreased their length. Substituents also played an important role in crystal packing for NiMeOsalen and NiMeOsalphen. i.e., a higher contribution for the O∙∙∙H interaction. Due to this contribution, the length of π∙∙∙π interactions were minor in both complexes. In N/N aliphatic bridge complexes, the substituents also had an important role. The E1/2 depended on the electron-withdrawn or electron-donor nature of the R (R’) substituent, which followed a correlation with the σp of Hammet in such a way that, for the electron-donor substituent, -OH and -CH3 decreased the half-wave potential, instead favoring nickel’s reductor ability.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4352/10/7/616/s1, Table S1. Atomic coordinates (x 104) and equivalent isotropic displacement parameters (Å2 103) for compound NiMesalen. U(eq) is defined as one third of the trace of the orthogonalized Uij tensor. Table S2. Atomic coordinates (×104) and equivalent isotropic displacement parameters (Å2 × 103) for compound NiMeOsalen. U(eq) is defined as one third of the trace of the orthogonalized Uij tensor. Table S3. Atomic coordinates (x 104) and equivalent isotropic displacement parameters (Å2 × 103) for compound NiMeOsalphen. U(eq) is defined as one third of the trace of the orthogonalized Uij tensor. Table S4. Atomic coordinates (×104) and equivalent isotropic displacement parameters (Å2 × 103) for compound Nisalphen. U(eq) is defined as one third of the trace of the orthogonalized Uij tensor. Table S5. Bond lengths [Å] and angles [°] for compound Nisalphen. Table S6. Effective magnetic moment and number of unpaired electrons of NiII complexes. Table S7. 1H-NMR values for the ligands and nickel complexes. Table S8. 13C-NMR values for the ligands and nickel complexes. Figure S1. Normalized contact distance (dnorm, defined in terms of de, di, and the van der Waals radii of the atoms) mapped on the Hirshfeld surface of the compound NiMeOsalen, represented with one surrounding moiety to visualize the intermolecular interaction. Figure S2. Hirshfeld surface with dnorm mapped and fingerprint plots for compound NiMesalen, with C∙∙∙H interaction (first row) and H∙∙∙H, O∙∙∙H interactions (row 2–3). The color ranges from dark blue to red with increasing frequency (relative area of the surface), corresponding to each kind of interaction. Figure S3. Normalized contact distance (dnorm, defined in terms of de, di, and the van der Waals radii of the atoms) mapped on the Hirshfeld surface of the compound NiMeOsalen, represented with one surrounding moiety to visualize the intermolecular interaction. Figure S4. Hirshfeld surface with dnorm mapped and fingerprint plots for NiMeOsalen, with C∙∙∙H interaction (first row) and H∙∙∙H, O∙∙∙H interactions (row 2–3). The color ranges from dark blue to red with increasing frequency (relative area of the surface) corresponding to each kind of interaction. Figure S5. Normalized contact distance (dnorm, defined in terms of de, di, and the van der Waals radii of the atoms) mapped on the Hirshfeld surface of the compound NiMeOsalphen, represented together with one surrounding moiety to visualize the intermolecular interaction. Figure S6. Hirshfeld surface with dnorm mapped and fingerprint plots for compound NiMeOsalphen, with C∙∙∙H interaction (first row) and H∙∙∙H, O∙∙∙H interactions (row 2–3). The color ranges from dark blue to red with increasing frequency (relative area of the surface) corresponding to each kind of interaction. Figure S7. Normalized contact distance (dnorm, defined in terms of de, di, and the van der Waals radii of the atoms) mapped on the Hirshfeld surface of the compound Nisalphen, represented together with one surrounding moiety to visualize the intermolecular interaction. Figure S8. Hirshfeld surface with dnorm mapped and fingerprint plots of the two molecules name A and B in compound Nisalphen for C∙∙∙H interaction (first row) and H∙∙∙H, Cl∙∙∙H interactions (row 2–3). The color ranges from dark blue to red with increasing frequency (relative area of the surface) corresponding to each kind of interaction. Figure S9. UV-vis of NiMeOsalen in DMSO solution. Figure S10. UV-vis NiMesalen in DMSO solution. Figure S11. UV-vis NiOHsalen in DMSO solution. Figure S12. UV-vis of Nisalen in DMSO solution. Figure S13. UV-vis of NiClsalen in DMSO solution. Figure S14. UV-vis of NiBrsalen in DMSO solution. Figure S16. UV-vis NiMeOsalphen in DMSO solution. Figure S17. UV-vis NiOHsalphen in DMSO solution. Figure S18. UV-vis Nisalphen in DMSO solution. Figure S19. UV-vis NiClsalphen in DMSO solution. Figure S20. UV-vis NiBrsalphen in DMSO solution. Figure S21. 1H-NMR of NiMeOsalen in DMSO-d6. Figure S22. 13C-NMR of NiMeOsalen in DMSO-d6. Figure S23. 1H-NMR of NiMesalen in chloroform. Figure S24. 13C-NMR of NiMesalen in chloroform. Figure S25. 1H-NMR of NiOHsalen in DMSO-d6. Figure S26. 13C-NMR of NiOHsalen in DMSO-d6. Figure S27. 1H-NMR of Nisalen in chloroform. Figure S28. 13C-NMR of Nisalen in chloroform. Figure S29. COSY spectrum of Nisalen in chloroform. Figure S30. HSQC spectrum of Nisalen in choloform. Figure S31. 1H-NMR of NaClsalen in chloroform. Figure S32. 1H-NMR of NiBrsalen in chloroform. Figure S33. 1H-NMR of NiNO2salen in DMSO-d6. Figure S34. 1H-NMR of NiOHsalphen in DMF-d7. Figure S35. 13C-NMR of NiOHsalphen in DMF-d7. Figure S36. COSY spectrum of NiOHsalphen in DMF-d7. Figure S37. HSQC spectrum of NiOHsalphen in DMF-d7. Figure S38. 1H-NMR of NiMeOsalphen in CDCl3. Figure S39. COSY spectrum of NiMeOsalphen in CDCl3. Figure S40. Voltammogram of NiMeOsalen 1 mM in DMSO. Figure S41. Voltammogram of NiMesalen 1 mM in DMSO. Figure S42. Voltammogram of NiOHsalen 1 mM in DMSO. Figure S43. Voltammogram of Nisalen 1 mM in DMSO. Figure S44. Voltammogram of NiClsalen 1 mM in DMSO. Figure S45. Voltammogram of NiBrsalen 1 mM in DMSO. Figure S46. Voltammogram of NiNO2salen 1 mM in DMSO.

Author Contributions

Conceptualization, C.S.N.-R. and A.S.-B.; methodology, C.S.N.-R., A.S.-B., and F.L.O.-V.; software, M.F.-A.; formal analysis, C.S.N.-R., A.S.-B., J.C.G.-R., and M.F.-A.; investigation, L.R.-A.; direction and resources, L.R.-A.; data curation, M.F.-A.; writing—original draft preparation, C.S.N.-R., A.S.-B., and M.F.-A.; writing—review and editing, L.R.-A. and J.C.G.-R., supervision,. L.R.-A.; project administration, L.R.-A.; funding acquisition, L.R.-A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

We thank USAII for elemental analysis, mass spectrometry, nuclear magnetic resonance tests and chromatography coupled to mass spectrometry tests; PAPIIT IN230020 for financial support and PAPIIT IN217613 for scholarships (CSNR, and ASB).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ito, Y.N.; Katsuki, T. Asymmetric Catalysis of New Generation Chiral Metallosalen Complexes. Bull. Chem. Soc. Jpn. 1999, 72, 603–619. [Google Scholar] [CrossRef]
  2. Naeimi, H.; Karshenas, A. Highly regioselective conversion of epoxides to β-hydroxy nitriles using metal(II) Schiff base complexes as new catalysts under mild conditions. Polyhedron 2013, 49, 234–238. [Google Scholar] [CrossRef]
  3. Dalton, C.T.; Ryan, K.M.; Langan, I.J.; Coyne, É.J.; Gilheany, D.G. Asymmetric alkene epoxidation with chromium oxo salen complexes: Effect of π-rich and other types of additives. J. Mol. Catal. A Chem. 2002, 187, 179–187. [Google Scholar] [CrossRef]
  4. Yang, H.; Zhang, L.; Su, W.; Yang, Q.; Li, C. Asymmetric ring-opening of epoxides on chiral Co(Salen) catalyst synthesized in SBA-16 through the “ship in a bottle” strategy. J. Catal. 2007, 248, 204–212. [Google Scholar] [CrossRef]
  5. Bordoloi, A.; Amrute, A.P.; Halligudi, S.B. [Ru(salen)(NO)] complex encapsulated in mesoporous SBA-16 as catalyst for hydrogenation of ketones. Catal. Commun. 2008, 10, 45–48. [Google Scholar] [CrossRef]
  6. Chatterjee, D.; Bajaj, H.C.; Das, A.; Bhatt, K. First report on highly efficient alkene hydrogenation catalysed by Ni(salen) complex encapsulated in zeolite. J. Mol. Cat. 1994, 92, L235–L238. [Google Scholar] [CrossRef]
  7. Hamdan, H.; Navijanti, V.; Nur, H.; Muhid, M.N.M. Fe(III)-salen encapsulated Al-MCM-41 as a catalyst in the polymerisation of bisphenol-A. J. Solid State Sci. 2005, 7, 239–244. [Google Scholar] [CrossRef]
  8. Ding, L.; Liang, S.; Zhang, J.; Ding, C.; Chen, Y.; Lü, X. Cu2+-templated self-assembly of an asymmetric Salen-Cu(II) complex and its application in catalytic polymerization of methyl methacrylate (MMA). Inorg. Chem. Commun. 2014, 44, 173–176. [Google Scholar] [CrossRef]
  9. Butsch, K.; Günther, T.; Klein, A.; Stirnat, K.; Berkessel, A.; Neudörfl, J. Redox chemistry of copper complexes with various salen type ligands. Inorganica Chimi. Acta 2013, 394, 237–246. [Google Scholar] [CrossRef]
  10. Pratt, R.C.; Stack, T.D.P. Mechanistic insights from reactions between copper(II)-phenoxyl complexes and substrates with activated C-H bonds. Inorg. Chem. 2005, 44, 2367–2375. [Google Scholar] [CrossRef]
  11. Thomas, F. Ten years of a biomimetic approach to the copper(II) radical site of galactose oxidase. Eur. J. Inorg. Chem. 2007, 2379–2404. [Google Scholar] [CrossRef]
  12. Mahapatra, P.; Drew, M.G.B.; Ghosh, A. Ni(II) Complex of N2O3 Donor Unsymmetrical Ligand and Its Use for the Synthesis of NiII-MnII Complexes of Diverse Nuclearity: Structures, Magnetic Properties, and Catalytic Oxidase Activities. Inorg. Chem. 2018, 57, 8338–8353. [Google Scholar] [CrossRef] [PubMed]
  13. Kianfar, A.H.; Paliz, M.; Roushani, M.; Shamsipur, M. Synthesis, spectroscopy, electrochemistry and thermal study of vanadyl tridentate Schiff base complexes. Spectrochim Acta A 2011, 82, 44–48. [Google Scholar] [CrossRef] [PubMed]
  14. Santos, I.C.; Vilas-boas, M.; Piedade, M.F.M.; Freire, C.; Duarte, M.T.; Castro, B. De Electrochemical and X-ray studies of nickel (II) Schiff base complexes derived from salicylaldehyde. Structural effects of bridge substituents on the stabilisation of the +3 oxidation state. Polyhedron 2000, 19, 655–664. [Google Scholar] [CrossRef]
  15. Shehata, E.E.; Masoud, M.S.; Khalil, E.A.; Abdel-Gaber, A.M. Synthesis, spectral and electrochemical studies of some Schiff base N2O2 complexes. J. Mol. Liq. 2014, 194, 149–158. [Google Scholar] [CrossRef]
  16. Carradori, S.; De Monte, C.; D’Ascenzio, M.; Secci, D.; Celik, G.; Ceruso, M.; Vullo, D.; Scozzafava, A.; Supuran, C.T. Salen and tetrahydrosalen derivatives act as effective inhibitors of the tumor-associated carbonic anhydrase XII—A new scaffold for designing isoform-selective inhibitors. Bioorg. Med. Chem. Lett. 2013, 23, 6759–6763. [Google Scholar] [CrossRef]
  17. Hang, Z.X.; Dong, B.; Wang, X.W. Synthesis, crystal structures, and antibacterial activity of zinc(II) complexes with bis-schiff bases. Synth. React. Inorg. Met. Org. Chem. 2012, 42, 1345–1350. [Google Scholar] [CrossRef]
  18. Fasina, T.M.; Ogundele, O.; Ejiah, F.N.; Dueke-Eze, C.U. Biological Activity of Copper (II), Cobalt (II) and Nickel (II) Complexes of Schiff Base Derived from O-phenylenediamine and 5-bromosalicylaldehyde. Int. J. Biol. Chem. 2012, 6, 24–30. [Google Scholar] [CrossRef] [Green Version]
  19. Woldemariam, G.A.; Mandal, S.S. Iron(III)-salen damages DNA and induces apoptosis in human cell via mitochondrial pathway. J. Inorg. Biochem. 2008, 102, 740–747. [Google Scholar] [CrossRef]
  20. Ansari, K.I.; Grant, J.D.; Kasiri, S.; Woldemariam, G.; Shrestha, B.; Mandal, S.S. Manganese(III)-salens induce tumor selective apoptosis in human cells. J. Inorg. Biochem. 2009, 103, 818–826. [Google Scholar] [CrossRef]
  21. Meshkini, A.; Yazdanparast, R. Chemosensitization of human leukemia K562 cells to taxol by a Vanadium-salen complex. Exp. Mol. Pathol. 2010, 89, 334–342. [Google Scholar] [CrossRef]
  22. Immel, T.A.; Grützke, M.; Batroff, E.; Groth, U.; Huhn, T. Cytotoxic dinuclear titanium-salan complexes: Structural and biological characterization. J. Inorg. Biochem. 2012, 106, 68–75. [Google Scholar] [CrossRef] [PubMed]
  23. Gravert, D.J.; Griffin, J.H. Specific DNA cleavage mediated by manganese complex [Salen Mn(III)]+. J. Org. Chem. 1993, 58, 820–822. [Google Scholar] [CrossRef]
  24. Peng, B.; Zhou, W.H.; Yan, L.; Liu, H.W.; Zhu, L. DNA-binding and cleavage studies of chiral Mn(III) salen complexes. Transit. Metal. Chem. 2009, 34, 231–237. [Google Scholar] [CrossRef]
  25. Bhattacharya, S.; Mandal, S.S. Ambient oxygen activating water soluble Cobalt-Salen complex for DNA cleavage. J. Chem. Soc. Chem. Commun. 1995, 2489–2490. [Google Scholar] [CrossRef]
  26. Ali, A.; Kamra, M.; Bhan, A.; Mandal, S.S.; Bhattacharya, S. New Fe(III) and Co(II) salen complexes with pendant distamycins: Selective targeting of cancer cells by DNA damage and mitochondrial pathways. Dalton Trans. 2016, 45, 9345–9353. [Google Scholar] [CrossRef]
  27. Muller, J.G.; Paikoff, S.J.; Rokita, S.E.; Burrows, C.J. DNA modification promoted by water-soluble nickel(II) salen complexes: A switch to DNA alkylation. J. Inorg. Biochem. 1994, 54, 199–206. [Google Scholar] [CrossRef]
  28. Mariappan, M.; Suenaga, M.; Mukhopadhyay, A.; Maiya, B.G. Synthesis, structure, DNA binding and photonuclease activity of a nickel(II) complex with a N,N′-Bis(salicylidene)-9-(3,4-diaminophenyl) acridine ligand. Inorg. Chim. Acta 2012, 390, 95–104. [Google Scholar] [CrossRef]
  29. Zhou, C.Q.; Liao, T.C.; Li, Z.Q.; Gonzalez-Garcia, J.; Reynolds, M.; Zou, M.; Vilar, R. Dinickel–salphen complexes as binders of human telomeric dimeric G-quadruplexes. Chem. Eur. J. 2017, 23, 4713–4722. [Google Scholar] [CrossRef]
  30. Reed, J.E.; Arnal, A.A.; Neidle, S.; Vilar, R. Stabilization of G-quadruplex DNA and inhibition of telomerase activity by square-planar nickel(II) complexes. J. Am. Chem. Soc. 2006, 128, 5992–5993. [Google Scholar] [CrossRef]
  31. Mukherjee, P.; Biswas, C.; Drew, M.G.B.; Ghosh, A. Structural variations in Ni(II) complexes of salen type di-Schiff base ligands. Polyhedron 2007, 26, 3121–3128. [Google Scholar] [CrossRef]
  32. Ghaemi, A.; Fayyazi, K.; Keyvani, B.; Ng, S.W.; Tiekink, E.R.T. {2,2′-[o-Phenylenebis(nitrilomethanylylidene)]diphenolato- κ4 O,N,N′,O′}nickel(II) monohydrate. Acta Crystallogr. Sect. E Struct. Rep. Online 2011, E67, m1481–m1482. [Google Scholar] [CrossRef] [Green Version]
  33. Siegler, M.A.; Lutz, M. Ni(salen): A system that forms many solvates with interacting Ni atoms. Cryst. Growth Des. 2009, 9, 1194–1200. [Google Scholar] [CrossRef]
  34. Choudhary, A.; Das, B.; Ray, S. Enhanced catalytic activity and magnetization of encapsulated nickel Schiff-base complexes in zeolite-Y: A correlation with the adopted non-planar geometry. Dalton Trans. 2016, 45, 18967–18976. [Google Scholar] [CrossRef]
  35. Xu, Y.; Xue, L.; Wang, Z.G. Synthesis, X-ray crystal structures, and antibacterial activities of Schiff base nickel(II) complexes with similar tetradentate Schiff bases. Russ. J. Coord. Chem. 2017, 43, 314–319. [Google Scholar] [CrossRef]
  36. Lamour, E.; Routier, S.; Bernier, J.L.; Catteau, J.P.; Bailly, C.; Vezin, H. Oxidation of Cu(II) to Cu(III), free radical production, and DNA cleavage by hydroxy-salen-copper complexes. Isomeric effects studied by ESR and electrochemistry. J. Am. Chem. Soc. 1999, 121, 1862–1869. [Google Scholar] [CrossRef]
  37. Aein Jamshid, K.; Asadi, M.; Hossein Kianfar, A. Synthesis, characterization and thermal studies of dinuclear adducts of diorganotin(IV) dichlorides with nickel(II) Schiff-base complexes in chloroform. J. Coord. Chem. 2009, 62, 1187–1198. [Google Scholar] [CrossRef]
  38. Biradar, N.S.; Karajagi, G.V.; Roddabasanagoudar, V.L.; Aminabhavi, T.M. Geometrical transformations around nickel(ii) with silicon(iv) tetrachloride. Synth. React. Inorg. Met.-Org. Chem. 1984, 14, 773–783. [Google Scholar] [CrossRef]
  39. Naeimi, H.; Moradian, M. Efficient synthesis and characterization of some novel nitro-schiff bases and their complexes of nickel(II) and copper(II). J. Chem. 2013, 2013. [Google Scholar] [CrossRef] [Green Version]
  40. Eltayeb, N.E.; Teoh, S.G.; Chantrapromma, S.; Fun, H.K.; Ibrahim, K. 4,4′-Dimethoxy-2,2′-[1,2-phenylene-bis(nitrilomethylidyne)] diphenol. Acta Crystallogr. Sect. E Struct. Rep. Online 2007, 63. [Google Scholar] [CrossRef]
  41. Wang, J.; Bei, F.; Xu, X.; Yang, X.; Wang, X. Crystal structure and characterization of 1, 2-N, N-disallicydene-phenylamineato nickel (II) complex. J. Chem. Crystallogr. 2003, 33, 845–849. [Google Scholar] [CrossRef]
  42. Elerman, Y.; Elmali, A.; Kabak, M.; Aydin, M.; Peder, M. Crystal structure of bis-N,N’-p-chloro-salicylideneamine-1,2.diaaminobenzene. J. Chem. Cryst. 1994, 24, 603–606. [Google Scholar] [CrossRef]
  43. Batley, G.; Graddon, D. Nickel(II) hydrosalicylamide complexes. Aust. J. Chem. 1967, 20, 1749. [Google Scholar] [CrossRef] [Green Version]
  44. Kondo, M.; Nabari, K.; Horiba, T.; Irie, Y.; Shimizu, Y.; Fuwa, Y. Synthesis and crystal structure of [Ni{bis(2,5-dihydroxysalicylidene)ethylenediaminato}]: A hydrogen bonded assembly of Ni (II)–salen complex. Inorga Chem Commun 2003, 6, 154–156. [Google Scholar] [CrossRef]
  45. Asadi, M.; Jamshid, K.A.; Kyanfar, A.H. Synthesis, characterization and equilibrium study of the dinuclear adducts formation between nickel(II) Salen-type complexes with diorganotin(IV) dichlorides in chloroform. Inorganica Chimica Acta 2007, 360, 1725–1730. [Google Scholar] [CrossRef]
  46. Oxford Diffraction Ltd. Available online: https://www.rigaku.com/products/smc/crysalis (accessed on 17 July 2019).
  47. Clark, R.C.; Reid, J.S. The analytical calculation of absorption in multifaceted crystals. Acta Crystallogr. A 1995, 51, 887–897. [Google Scholar] [CrossRef]
  48. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. A Found. Crystallogr. 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  49. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. C 2015, C71, 3–8. [Google Scholar] [CrossRef]
  50. Farrugia, L.J. WinGX and ORTEP for Windows: An update. J. Appl. Crystallogr. 2012, 45, 849–854. [Google Scholar] [CrossRef]
  51. Spek, A.L. PLATON SQUEEZE: A tool for the calculation of the disordered solvent contribution to the calculated structure factors. Acta Crystallogr. Sect. C Struct. Chem. 2015, C71, 9–18. [Google Scholar] [CrossRef] [Green Version]
  52. Garg, B.S.; Nandan Kumar, D. Spectral studies of complexes of nickel(II) with tetradentate schiff bases having N2O2 donor groups. Spectrochim Acta A Mol. Biomol. Spectrosc. 2003, 59, 229–234. [Google Scholar] [CrossRef]
  53. Lever, A.B.P. Inorganic Electronic Spectroscopy, 2nd ed.; Elsevier: New York, NY, USA, 1984. [Google Scholar]
  54. Maki, G. Ligand field theory of Ni (II) complexes. I. Electronic energies and singlet ground-state conditions of Ni (II) complexes of different symmetries. J. Chem. Phys. 1958, 28, 651–662. [Google Scholar] [CrossRef]
  55. Kianfar, A.H.; Dostani, M.; Mahmood, W.A.K. An unprecedented DDQ-nickel(II)Salen complex interaction and X-ray crystal structure of nickel(II)Salen.DDH co-crystal. Polyhedron 2015, 85, 488–492. [Google Scholar] [CrossRef]
  56. Spackman, M.A.; Jayatilaka, D. Hirshfeld surface analysis. Cryst. Eng. Comm. 2009, 11, 19–32. [Google Scholar] [CrossRef]
  57. Wolff, S.K.; Grimwood, D.J.; McKinnon, J.J.; Jayatilaka, D.; Spackman, M.A. Available online: http://hirshfeldsurface.net/CrystalExplorer (accessed on 17 July 2019).
  58. Isse, A.A.; Gennaro, A.; Vianello, E. Electrochemical reduction of Schiff base ligands H2salen and H2salophen. Electrochim. Acta 1997, 42, 2065–2071. [Google Scholar] [CrossRef]
  59. Pooyan, M.; Ghaffari, A.; Behzad, M.; Amiri Rudbari, H.; Bruno, G. Tetradentate N 2O 2 type Nickel(II) Schiff base complexes derived from meso -1,2-diphenyle-1,2-ethylenediamine: Synthesis, characterization, crystal structures, electrochemistry, and catalytic studies. J. Coord. Chem. 2013, 66, 4255–4267. [Google Scholar] [CrossRef]
  60. Freire, C. Spectroscopic characterisation of electrogenerated nickel(III) species. Complexes with N2O2 Schiff-base ligands derived from salicylaldehyde. J. Chem. Soc. Dalton Trans. 1998, 1491–1498. [Google Scholar] [CrossRef]
  61. Jäger, E.G.; Schuhmann, K.; Görls, H. Syntheses, characterization, redox behavior and Lewis acidity of chiral nickel(II) and copper(II) Schiff base complexes. Inorg. Chim. Acta 1997, 255, 295–305. [Google Scholar] [CrossRef]
  62. Zolezzi, S.; Spodine, E.; Decinti, A. Electrochemical studies of copper(II) complexes with Schiff-base ligands. Polyhedron 2002, 21, 55–59. [Google Scholar] [CrossRef]
  63. Kianfar, A.H.; Sobhani, V.; Dostani, M.; Shamsipur, M.; Roushani, M. Synthesis, spectroscopy, electrochemistry and thermal study of vanadyl unsymmetrical Schiff base complexes. Inorg. Chim. Acta 2011, 365, 108–112. [Google Scholar] [CrossRef]
Figure 1. View on the perspective of the compound NiMesalen, with a displacement ellipsoid at a 50% probability level for non-H atoms.
Figure 1. View on the perspective of the compound NiMesalen, with a displacement ellipsoid at a 50% probability level for non-H atoms.
Crystals 10 00616 g001
Figure 2. Crystal array of compound NiMesalen shows the view along the b-axis and with perspective to plane formed by a-c axes emphasizing the π-stacking and R12 (4) motif.
Figure 2. Crystal array of compound NiMesalen shows the view along the b-axis and with perspective to plane formed by a-c axes emphasizing the π-stacking and R12 (4) motif.
Crystals 10 00616 g002
Figure 3. Perspective of the compound NiMeOsalen with a displacement ellipsoid at a 50% probability level for non-H atoms.
Figure 3. Perspective of the compound NiMeOsalen with a displacement ellipsoid at a 50% probability level for non-H atoms.
Crystals 10 00616 g003
Figure 4. Crystal array of compound NiMeOsalen. View along the c axes and with a perspective of a plane formed by the a-b axes, which emphasizes the R12(4) and R22(6) motifs, as well as π-stacking.
Figure 4. Crystal array of compound NiMeOsalen. View along the c axes and with a perspective of a plane formed by the a-b axes, which emphasizes the R12(4) and R22(6) motifs, as well as π-stacking.
Crystals 10 00616 g004
Figure 5. Perspective of the compound NiMeOsalphen with a displacement ellipsoids at a 65% probability level for non-H atoms.
Figure 5. Perspective of the compound NiMeOsalphen with a displacement ellipsoids at a 65% probability level for non-H atoms.
Crystals 10 00616 g005
Figure 6. Crystal array of compound NiMeOsalphen. View is a tridimensional perspective emphasizing R22(30) and R22(11) motifs and π-stacking.
Figure 6. Crystal array of compound NiMeOsalphen. View is a tridimensional perspective emphasizing R22(30) and R22(11) motifs and π-stacking.
Crystals 10 00616 g006
Figure 7. Perspective of the compound Nisalphen with a displacement ellipsoids at a 55% probability level for non-H atoms.
Figure 7. Perspective of the compound Nisalphen with a displacement ellipsoids at a 55% probability level for non-H atoms.
Crystals 10 00616 g007
Figure 8. Crystal array of compound Nisalphen, view a tridimensional perspective emphasizing the π-stacking.
Figure 8. Crystal array of compound Nisalphen, view a tridimensional perspective emphasizing the π-stacking.
Crystals 10 00616 g008
Figure 9. Hirshfeld surface (HS) with dnorm mapped and fingerprinted plots of the compounds NiMesalen (a), NiMeOsalen (b), NiMeOsalphen (c), and Nisalphen (d) for all interactions.
Figure 9. Hirshfeld surface (HS) with dnorm mapped and fingerprinted plots of the compounds NiMesalen (a), NiMeOsalen (b), NiMeOsalphen (c), and Nisalphen (d) for all interactions.
Crystals 10 00616 g009
Figure 10. HS of the compounds NiMesalen (a), NiMeOsalen (b), NiMeOsalphen (c), and Nisalphen (d), mapped with shape index.
Figure 10. HS of the compounds NiMesalen (a), NiMeOsalen (b), NiMeOsalphen (c), and Nisalphen (d), mapped with shape index.
Crystals 10 00616 g010
Figure 11. Contribution of some intermolecular contacts for HS of the compounds NiMesalen, NiMeOsalen, and NiMeOsalphen, as well as for molecules A and B of compound Nisalphen.
Figure 11. Contribution of some intermolecular contacts for HS of the compounds NiMesalen, NiMeOsalen, and NiMeOsalphen, as well as for molecules A and B of compound Nisalphen.
Crystals 10 00616 g011
Figure 12. Voltammogram of Nisalen 1 mM (DMSO). (a) Nisalen voltammogram from −3.2 to 1.0 V; (b) salen voltammogram from −3.2 to 1.0 V; (c) Nisalen voltammogram from negative potential to 2.4 to 1.0 V; (d) Nisalen voltammogram from positive potential to −2.4 to 1.0 V. All the experiments were referenced to the pair Fc+/Fc.
Figure 12. Voltammogram of Nisalen 1 mM (DMSO). (a) Nisalen voltammogram from −3.2 to 1.0 V; (b) salen voltammogram from −3.2 to 1.0 V; (c) Nisalen voltammogram from negative potential to 2.4 to 1.0 V; (d) Nisalen voltammogram from positive potential to −2.4 to 1.0 V. All the experiments were referenced to the pair Fc+/Fc.
Crystals 10 00616 g012
Figure 13. Left: correlation between E1/2 of NiII/NiI and the σ p constant for the NiRsalen complexes. Right: correlation between E1/2 of NiIII/NiII and the σ p constant for the NiRsalen complexes.
Figure 13. Left: correlation between E1/2 of NiII/NiI and the σ p constant for the NiRsalen complexes. Right: correlation between E1/2 of NiIII/NiII and the σ p constant for the NiRsalen complexes.
Crystals 10 00616 g013
Table 1. X-ray diffraction data collection and refinement parameters for the compounds NiMesalen, NiMeOsalen, NiMeOsalphen, and Nisalphen.
Table 1. X-ray diffraction data collection and refinement parameters for the compounds NiMesalen, NiMeOsalen, NiMeOsalphen, and Nisalphen.
CompoundNiMesalenNiMeOsalen NiMeOsalphen Nisalphen
Empirical formula C18 H18 N2 Ni O2C18 H18 N2 Ni O4C22 H18 N2 Ni O4C43 H31 Cl9 N4 Ni2 O4
Formula weight 353.05385.05433.091104.19
Crystal system MonoclinicMonoclinicMonoclinicTriclinic
Space group C 2/cP 21/cP 21/nP -1
Unit cell dimensions
a (Å)a = 21.422(2) a = 16.5157(9)a = 13.5517(4)a = 13.7180(10)
b (Å)b = 13.1534(18) b = 7.2835(3) b = 7.7890(2)b = 14.030(2)
c (Å)c = 6.5394(6)c = 13.7886(7)c = 17.0198(4)c = 14.908(2)
α (°)α = 90.α = 90α = 90.α = 115.285(12)
β (°)β = 96.280(9).β = 104.809(5)β = 100.197(3).β = 116.477(11)
γ (°)γ = 90.γ = 90γ = 90.γ = 92.967(10)
Volume(Å3)1831.6(4)1603.56(14)1768.14(8)2215.8(5)
Z4442
Density (calculated) (mg/m3)1.2801.5951.6271.655
Absorption coefficient (mm−1)1.0681.237 1.880 1.441
F(000)7368008961116
Crystal size(mm3)0.470 × 0.090 × 0.0700.550 × 0.160 × 0.130 0.470 × 0.230 × 0.2000.520 × 0.170 × 0.160
Theta range for data collection3.641 to 29.388°.3.445 to 29.394°.3.854 to 73.599°.3.380 to 29.513°.
Index ranges−27 < = h < = 25, −16 < = k < = 16, −5 < = l < = 9−20 < = h < = 17, −9 < = k < = 10, −18 < = l < = 19−16 < = h < = 16, −7 < = k < = 9, −20 < = l < = 19−18 < = h < = 18, −18 < = k < = 18, −19 < = l < = 16
Reflections collected4413119311143518883
Independent reflections2132 [R(int) = 0.0441]3921 [R(int) = 0.0215]3514 [R(int) = 0.0177]10314 [R(int) = 0.0282]
Completeness to theta = 25.242°99.5 % 99.8 % 100.0 % 99.7 %
Absorption correctionAnalyticalAnalyticalAnalyticalAnalytical
Max. and min. transmission0.929 and 0.7500.854 and 0.6460.725 and 0.5560.827 and 0.682
Refinement methodFull-matrix least-squares on F2Full-matrix least-squares on F2Full-matrix least-squares on F2Full-matrix least-squares on F2
Data/restraints/parameters2132/0/1063921/0/2283514/0/26410314/0/559
Goodness-of-fit on F21.0191.0471.0581.027
Final R indices [I > 2sigma(I)]R1 = 0.0506, wR2 = 0.0944R1 = 0.0247, wR2 = 0.0644R1 = 0.0313, wR2 = 0.0854R1 = 0.0394, wR2 = 0.0796
R indices (all data)R1 = 0.0802, wR2 = 0.1062R1 = 0.0287, wR2 = 0.0671R1 = 0.0332, wR2 = 0.0870R1 = 0.0561, wR2 = 0.0892
Largest diff. peak and hole0.601 and −0.625 e.Å−30.373 and −0.344 e.Å−30.238 and −0.567 e.Å−30.914 and −0.815 e.Å−3
Table 2. Electronic spectral data of the Schiff bases and their complexes.
Table 2. Electronic spectral data of the Schiff bases and their complexes.
Compoundλmax (ε, L mol−1 cm−1) in DMSOCompoundλmax (ε, L mol−1 cm−1) in DMSO
MeOsalen260, (21145), 345(14109)NiMeOsalen258(39143), 330(7681), 431(7096)
Mesalen260(17555), 326(8568), 427(176)NiMesalen260(47643), 334(7975), 417(6478)
OHsalen260(13265), 350(8593)NiOHsalen260(40311), 333(7338), 438(6053)
Salen260(8747), 327(18295), 410(299)NiSalen260(58958), 324(8061), 407(5906)
Clsalen258(14461), 327(7794), 420(552)NiClsalen258(44337), 324(8216), 415(6263)
Brsalen260(16272), 327(7479), 419(653)NiBrsalen258(48523), 326(8863), 414(6419)
NO2salen258(17745), 370(20301), 422(30053)NiNO2salen263(18147), 340(11715), 405(22660)
MeOsalphen276(42299), 348(18694)NiMeOsalphen268(50128), 296(2969), 382(26461), 511(2355)
Mesalphen274(18573), 341(14755), 450(790)NiMesalphen260(55055),292(22231), 380(26821), 487(8851), 679(5.6)
OHsalphen274(20172), 370(13706)NiOHsalphen260(39286), 298(19113), 386(22873), 518(8406)
Salphen269(22051), 332(18444), 448(1418)NiSalphen260(37709, 298(14534), 377(22632), 475(7305)
Clsalphen262(18058), 275(16555), 338(8525), 398(1779)NiClsalphen262(59148), 380(27100), 476(10739), 580(179)
Brsalphen274(18364), 339(2102), 404(3323), 451(3341)NiBrsalphen264(36210), 314(13762), 378(21675) 478(23408), 674(750)
Table 3. Selected bond lengths (Å) and angles (°) for compounds NiMesalen, NiMeOsalen, NiMeOsalphen, and Nisalphen.
Table 3. Selected bond lengths (Å) and angles (°) for compounds NiMesalen, NiMeOsalen, NiMeOsalphen, and Nisalphen.
NiMesalenNiMeOsalenNiMeOsalphen
Bond *LengthsBondLengthsBondLengths
C2—O11.323(3)C2—O11.3125(17)C2—O11.3077(18)
C5—C171.525(4)C5—O31.3770(17)C5—O31.3765(18)
C7—N11.297(3)C7—N11.2892(19)C7—N11.298(2)
C8—N11.469(4)C8—N11.4774(17)C8—C91.397(2)
C8—C8#11.515(5)C8—C91.504(2)C8—N11.4233(19)
O1—Ni11.852(2)C17—O31.4256(18)C17—O31.4219(19)
Ni1—N1#11.844(2)O1—Ni11.8505(10)N1—Ni11.8626(12)
Ni1—N11.844(2)O2—Ni11.8544(10)N2—Ni11.8643(13)
Ni1—N21.8476(12)Ni1—O21.8394(11)
Ni1—N11.8520(12)Ni1—O11.8536(11)
BondAnglesBondAnglesBondAngles
N1#1—Ni1—N186.32(15)N1—C8—C9107.28(11)C9—C8—N1113.44(13)
N1#1—Ni1—O1#194.95(9)N2—C9—C8107.04(11)C8—C9—N2114.08(13)
N1—Ni1—O1#1178.40(10)N2—Ni1—O1179.52(5)C8—N1—Ni1113.37(10)
N1#1—Ni1—O1178.40(10)N2—Ni1—N186.03(5)C9—N2—Ni1113.04(10)
N1—Ni1—O194.95(9)O1—Ni1—N194.40(5)O2—Ni1—O184.12(5)
O1#1—Ni1—O183.79(13)N2—Ni1—O293.77(5)O2—Ni1—N1179.34(5)
O1—Ni1—O285.80(4)O1—Ni1—N195.22(5)
N1—Ni1—O2179.60(5)O2—Ni1—N294.60(5)
C8—N1—Ni1114.60(9)O1—Ni1—N2177.72(5)
C9—N2—Ni1113.22(9)N1—Ni1—N286.05(6)
Nisalphen
Molecule AMolecule B
Bond *LengthsBondLengths
C2A-O1A1.306(3)C2B-O1B1.306(3)
C7A-N1A1.308(3)C7B-N1B1.299(3)
C8A-C9A1.395(3)C8B-C9B1.389(4)
C8A-N1A1.417(3)C8B-N1B1.427(3)
C12A-O2A1.312(3)C12B-O2B1.309(3)
O1A-Ni1A1.8398(16)O1B-Ni1B1.8370(17)
O2A-Ni1A1.8346(18)O2B-Ni1B1.8360(16)
N1A-Ni1A1.857(2)N1B-Ni1B1.8567(19)
N2A-Ni1A1.8564(19)N2B-Ni1B1.856(2)
BondAnglesBondAngles
C9A-C8A-N1A113.8(2)C9B-C8B-N1B113.8(2)
C8A-C9A-N2A113.8(2)C8B-C9B-N2B113.8(2)
C8A-N1A-Ni1A113.13(16)C8B-N1B-Ni1B112.87(16)
C9A-N2A-Ni1A112.97(16)C9B-N2B-Ni1B113.22(16)
O2A-Ni1A-O1A83.05(8)O2B-Ni1B-O1B83.03(7)
O2A-Ni1A-N2A95.17(8)O2B-Ni1B-N2B95.47(8)
O1A-Ni1A-N2A178.13(9)O1B-Ni1B-N2B178.47(8)
O2A-Ni1A-N1A178.42(8)O2B-Ni1B-N1B178.24(9)
O1A-Ni1A-N1A95.46(8)O1B-Ni1B-N1B95.23(8)
N2A-Ni1A-N1A86.32(9)N2B-Ni1B-N1B86.27(9)
* Operators for generating equivalent atoms: −x + 1, y, −z + 1/2#1.
Table 4. Cyclic voltammetric parameter for NiRsalen complexes, referenced to the pair Fc+/Fc.
Table 4. Cyclic voltammetric parameter for NiRsalen complexes, referenced to the pair Fc+/Fc.
Process Ni II L Ni III L   +   1   ē Ni II L   +   1   ē   Ni I L
Compound∆E (mV)ipc/ipaE1/2 (V)∆E (mV)ipc/ipaE1/2 (V)
NiMeOsalen1100.190.191100.50−2.08
NiMesalen2400.310.17770.61−2.13
NiOHsalen2000.120.073920.30−2.17
Nisalen4100.160.16760.72−2.09
NiClsalen3900.160.22730.83−1.98
NiBrsalen4000.230.21600.45−1.96
NiNO2salen2000.410.331200.57−1.72

Share and Cite

MDPI and ACS Style

Novoa-Ramírez, C.S.; Silva-Becerril, A.; Olivera-Venturo, F.L.; García-Ramos, J.C.; Flores-Alamo, M.; Ruiz-Azuara, L. N/N Bridge Type and Substituent Effects on Chemical and Crystallographic Properties of Schiff-Base (Salen/Salphen) Niii Complexes. Crystals 2020, 10, 616. https://doi.org/10.3390/cryst10070616

AMA Style

Novoa-Ramírez CS, Silva-Becerril A, Olivera-Venturo FL, García-Ramos JC, Flores-Alamo M, Ruiz-Azuara L. N/N Bridge Type and Substituent Effects on Chemical and Crystallographic Properties of Schiff-Base (Salen/Salphen) Niii Complexes. Crystals. 2020; 10(7):616. https://doi.org/10.3390/cryst10070616

Chicago/Turabian Style

Novoa-Ramírez, Cynthia S., Areli Silva-Becerril, Fiorella L. Olivera-Venturo, Juan Carlos García-Ramos, Marcos Flores-Alamo, and Lena Ruiz-Azuara. 2020. "N/N Bridge Type and Substituent Effects on Chemical and Crystallographic Properties of Schiff-Base (Salen/Salphen) Niii Complexes" Crystals 10, no. 7: 616. https://doi.org/10.3390/cryst10070616

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop