Home Physical Sciences Synthesis, structure, and cytotoxicity of some triorganotin(iv) complexes of 3-aminobenzoic acid-based Schiff bases
Article Open Access

Synthesis, structure, and cytotoxicity of some triorganotin(iv) complexes of 3-aminobenzoic acid-based Schiff bases

  • Ruili Wang , Jing Zhang , Gaoyu Cui and Laijin Tian EMAIL logo
Published/Copyright: December 31, 2022

Abstract

Six new triorganotin(iv) complexes of 3-aminobenzoic acid-based Schiff bases, 3-(R′-CH═N)C6H4COOSnR3 (1–6) (R′, R = 5-Br-2-HOC6H3, Ph (1); 3,5-Br2-2-HOC6H2, Ph (2); 4-NEt2-2-HOC6H3, Cy (3); 3-OCH3-2-HOC6H3, Cy (4); 2-HOC10H6, Ph (5); 2-HOC10H6, Cy (6)), have been synthesized by the one-pot reaction of equimolar 3-aminobenzoic acid, substituted 2-hydroxybenzaldehyde (or 2-hydroxy-1-naphthaldehyde) and triorganotin(iv) hydroxide, and characterized by elemental analysis, FT-IR, NMR spectroscopy, and X-ray single crystal diffraction. The NMR data (1 J(119Sn–13C) and 119Sn chemical shifts) suggested that these organotin(iv) complexes are all four-coordinated in CDCl3 solution. In the crystalline state, the tin atoms in 1–4 and 6 are four-coordinated and possess a distorted tetrahedral geometry. Complex 5 with crystalline solvents (CH3OH and CHCl3) exhibits a zigzag chain, and the five coordination atoms on the tin atom are arranged in a trigonal bipyramidal geometry in which the carboxylate oxygen atom and the phenolic oxygen atom of the adjacent ligand occupy the axial positions. In all complexes, the 3-(arylmethyleneamino)benzoate ligands are coordinated with tin atoms in monodentate mode. Their cytotoxicity against two human cancer cell lines (A549 and HeLa), UV-Vis, and fluorescence have been determined, and the results reveal that complexes 16 have higher cytotoxicity than cisplatin and may be explored for potential blue luminescent materials.

1 Introduction

Organotin(IV) compounds show a diverse range of applications, and are broadly used in the agriculture and industrial production (Davies et al., 2008). In recent years, there has been an increasing interest in organotin(iv) compounds in medicinal chemistry, and the literature shows that the bioactivity potency of organotin(iv) compounds increase in the following order: R4Sn < RSnL3 < R2SnL2 < R3SnL (L = ligand) (Bantia et al., 2019; Pellerito and Nagy, 2002). Triorganotin(iv) carboxylates (R3SnOOCR′) possess rich structural diversity including discrete, dimeric, chain, and macrocyclic structures and shows good application prospects as insecticides and fungicides, and, particularly, in cancer therapy (Bantia et al., 2019; Basu Baul et al., 2010; Davies et al., 2008; Hadi et al., 2021; Liang et al., 2014; Tiekink, 1991, 1994). The R groups on tin atom and carboxylic acid ligand have important effects on the biological activity of triorganotin(iv) carboxylates (Bantia et al., 2019; Basu Baul et al., 2017a). Therefore, to synthesize organotin(iv) complexes with different carboxylic acid ligands is an alternative good strategy to improve their biological activities. Schiff bases (C═N) prepared from the condensation of aldehyde or ketone and primary amine have some characteristic properties such as great synthesis flexibility and biological properties (Kaur et al., 2021; Nath and Saini, 2011). Introducing Schiff base functional groups into triorganotin(iv) carboxylates is expected to achieve the synergistic effect of their biological activities and obtain organotin(iv) complexes with novel structure and good activity. Some triorganotin(iv) complexes with the Schiff base carboxylate ligands have been synthesized and characterized. Basu Baul and co-workers synthesized a series of triorganotin(IV) complexes of amino acid Schiff bases, and found that these complexes have good biological activities, especially anticancer activity (Basu Baul et al., 2007, 2009, 2010, 2017a, 2017b). In addition, they also synthesized a novel series of triphenyltin(iv) 2- or 4-((arylimino)methyl)benzoates, 2/4-(ArN═CH)C6H4COOSnPh3, by the reactions between triphenyltin(iv) 2/4-formylbenzoate and aromatic amines, which revealed high cytotoxic activities against some human cancer cells (Basu Baul et al., 2018). Other groups (Dias et al., 2015; Goh et al., 1998; Saeed et al., 2017; Tzimopoulos et al., 2010; Yin et al., 2005, 2012; Yu et al., 2022) had also done good work on the triorganotin(iv) complexes. Recently, our team has synthesized the triorganotin(iv) complexes of N-(5-bromosalicylidene)-α-amino acid, 5-(salicylideneamino)salicylic acid, and 3-(salicylideneamino)benzoic acid, which exhibited strong cytotoxic efficacy (Chen et al., 2020; Liu et al., 2019; Yao et al., 2017). Organotin(iv) complexes of the carboxylate ligands containing Schiff base are generally synthesized by three different routes: (i) the reaction of organotin(iv) oxide or hydroxide or chloride with the Schiff base carboxylic acid ligand prepared in advance from amino acid and aldehyde (Basu Baul et al., 2017a, 2017b; Dias et al., 2015; Saeed et al., 2017; Tzimopoulos et al., 2010; Yin et al., 2005, 2012; Yu et al., 2022), (ii) the reaction of aldehyde with organotin(iv) aminocarboxylate first prepared from amino acid and organotin(iv) oxides or hydroxides or chlorides (Basu Baul et al., 2018; Tzimopoulos et al., 2010), and (iii) the one-pot reaction of organotin(iv) oxide or hydroxide, amino acid and aldehyde (Beltran et al., 2003; Chen et al., 2020; Yao et al., 2017). As a continuation of our previous works based on organotin(iv) complexes with Schiff base ligands (Chen et al., 2020; Liu et al., 2019; Yao et al., 2017), herein we report one-step preparation, and crystal structure and property of new triphenyltin(iv) and tricyclohexyltin(iv) 3-(arylmethyleneamino)benzoates (1–6), 3-(R′-CH═N)C6H4COOSnR3 (1–6) (R′, R = 5-Br-2-HOC6H3, Ph (1); 3,5-Br2-2-HOC6H2, Ph (2); 4-NEt2-2-HOC6H3, Cy (3); 3-OCH3-2-HOC6H3, Cy (4); 2-HOC10H6, Ph (5); 2-HOC10H6, Cy (6)) (Scheme 1).

Scheme 1 
               Synthesis of the complexes.
Scheme 1

Synthesis of the complexes.

2 Results and discussion

2.1 Synthesis

Equimolar 3-aminobenzoic acid, substituted 2-hydroxybenzaldehyde (or 2-hydroxy-1-naphthaldehyde), and triorganotin(iv) hydroxide are refluxed in benzene for 6 h by azeotropical removal of water to give the products (1–6) 76% to 88% yield (Scheme 1). In this preparation, the Schiff base was formed in situ, and it is simpler than two-step method mentioned above. The complexes are yellow to orange red crystals that are stable in air and soluble in common organic solvents.

2.2 Spectroscopic characterization

In the complexes, IR spectra show a broad band at ∼3,430 cm−1 assigned to ν(OH) of phenol hydroxyl with intramolecular O–H⋯N hydrogen bond. The strong bands attributable to the asymmetric ν as(COO) and symmetric ν s(COO) of carboxylate occur at 1,618–1,637 cm−1 and 1,324–1,343 cm−1, respectively. The ν(C═N) absorption peak of Schiff base unit appears at about 1,615 cm−1, which overlaps with the ν s(COO) absorption in some cases (Vinayak and Nayek, 2019) (Supplementary material). The difference (Δν(COO)) between the ν as(COO) and ν s(COO) bands can provide useful information concerning the coordination behavior of the carboxylate ligand (Deacon and Phillips, 1980; Yin et al., 2005). Generally, the magnitude of Δν(COO) is above 200 cm−1 for a monodentate coordination and below 200 cm−1 for a bidentate coordination. The values (270–314 cm−1) of Δν(COO) in complexes 16 indicate that the carboxylate ligand is bonded to the tin center in a monodentate mode.

The 1H and 13C chemical shift assignments of complexes 16 are from the multiplicity patterns and resonance intensities, as well as related literature (Chen et al., 2020; Tzimopoulos et al., 2010; Yin et al., 2012; Yu et al., 2022). The 1H NMR integration values were consistent with the structures of 16 in Scheme 1. In the complexes, the phenolic O–H proton is observed in the range of 13.07–15.43 ppm as a broad signal due to the formation of intramolecular O–H⋯N hydrogen bond. The CH═O proton signal (∼10 ppm) in the substituted 2-hydroxybenzaldehyde or 2-hydroxy-1-naphthaldehyde does not appear, and the imine CH═N proton signal is observed in the region of 8.50–9.43 ppm, indicating the formation of Schiff base ligand. The chemical shift value of CH═N proton in 14 (∼8.6 ppm) is obviously smaller than that in 5 and 6 (∼9.4 ppm) due to the different effects of benzene ring and naphthalene ring on CH═N proton. The COOH proton resonance signal of Schiff base carboxylic ligand is not observed, which confirms the deprotonation of carboxylic ligand and bonding to the tin atom.

In the 13C NMR spectra, the resonances of carboxylate (COO) carbon appear in the range of 170.52–172.05 ppm. The chemical shifts of Schiff base imine carbon (CH═N) are observed in the range of 160.95–163.30 ppm for complexes 14 and 169.82–170.15 ppm for complexes 5 and 6. The other carbon atoms of the ligand and SnR3 skeletons display the expected resonance signals. The coupling constant, 1 J(13C–119Sn), is closely related to the coordination number of tin atom, and can provide structural information of organotin(iv) compounds in solution (Holecek et al., 1983, 1986). For four-coordinated triphenyltin(iv) and tricyclohexyltin(iv) compounds, 1 J(119Sn–13C) values lie in the range of 550–650 and 295–360 Hz, respectively (Holecek et al., 1983; Zhang et al., 1990). The 1 J(119Sn–13C) value of 16 is 640, 642, 336, 338, 648, and 336 Hz, respectively, clearly indicating that the tin atoms in these complexes are all four-coordinated in non-coordination solvent CDCl3. This conclusion is further supported by the 119Sn NMR chemical shifts of the complexes. Triphenyltin(iv) complexes 1, 2, and 5 exhibit a single sharp resonance at −102.7, −102.4, and −105.4 ppm, and tricyclohexyltin(iv) complexes 3, 4, and 6 display absorption peaks at 18.6, 19.4, and 17.3 ppm, suggesting that the central tin atom has a four-coordinate environment (Holecek et al., 1983; Zhang et al., 1990).

2.3 Crystal structures of the complexes

The molecular structures of these complexes are shown in Figures 16, respectively. The selected bond lengths and bond angles are listed in Table 1. The coordination environment of the tin atom in the triphenyltin(iv) complexes 1 and 2 is a distorted tetrahedron and four coordination atoms are from three carbon atoms of phenyl groups and one carboxyl oxygen atom of the (E)-3-[(2-hydroxybenzylidene)amino]benzoate ligand (Figures 1 and 2). The Sn–C bond lengths and the C–Sn–C bond angles around the tin atom are in the ranges of 2.117(4)–2.143(3) Å and 107.09(12)–119.13(12)°, respectively, which are similar to those observed in the analogous triphenyltin(iv) derivatives, such as 4-(2-HOC6H4CH═N)C6H4COOSnPh3 (Yin et al., 2005), 4-H2NC6H4COOSnPh3 (Tzimopoulos et al., 2009), and 3-(4-Me2NC6H4CH═N)C6H4COOSnPh3 (Basu Baul et al., 2017a). The monodentate nature of the carboxylate ligand is further supported by the long separation of Sn(1)⋯O(2) (2.758(2) Å for 1 and 2.785(3) Å for 2 and short C(19)–O(2) bond length (1.222(4) Å for 1 and 1.217(5) Å for 2). However, the Sn(1)⋯O(2) interaction leads to the distortion from a regular tetrahedron, which expands the C(1)–Sn(1)–C(7) bond angle (119.13(12)° for 1 and 117.52(15)° for 2) and reduces the O(1)–Sn(1)–C(13) bond angle (96.51(10)° for 1 and 94.77(14)° for 2).

Figure 1 
                  The molecular structure of 1. Ellipsoids are drawn at 30% probability level. Hydrogen atoms are omitted for clarity except H3 atom.
Figure 1

The molecular structure of 1. Ellipsoids are drawn at 30% probability level. Hydrogen atoms are omitted for clarity except H3 atom.

Figure 2 
                  (Top) The molecular structure of 2. Ellipsoids are drawn at 30% probability level. Hydrogen atoms are omitted for clarity except H3 atoms. (Bottom) One-dimensional supramolecular chain formed by the intermolecular C–H⋯O hydrogen bonds.
Figure 2

(Top) The molecular structure of 2. Ellipsoids are drawn at 30% probability level. Hydrogen atoms are omitted for clarity except H3 atoms. (Bottom) One-dimensional supramolecular chain formed by the intermolecular C–H⋯O hydrogen bonds.

Figure 3 
                  (Top) The molecular structure of 3. Ellipsoids are drawn at 30% probability level. Hydrogen atoms are omitted for clarity except H3 atoms. (Bottom) Two-dimensional supramolecular network formed by the intermolecular Sn⋯O interactions and C–H⋯O hydrogen bonds. The Cy groups bonded to Sn are omitted for clarity.
Figure 3

(Top) The molecular structure of 3. Ellipsoids are drawn at 30% probability level. Hydrogen atoms are omitted for clarity except H3 atoms. (Bottom) Two-dimensional supramolecular network formed by the intermolecular Sn⋯O interactions and C–H⋯O hydrogen bonds. The Cy groups bonded to Sn are omitted for clarity.

Figure 4 
                  (Top) The molecular structure of 4. Ellipsoids are drawn at 30% probability level. Hydrogen atoms are omitted for clarity except H3 atoms. (Bottom) One-dimensional supramolecular chain formed by the intermolecular C–H⋯O hydrogen bonds.
Figure 4

(Top) The molecular structure of 4. Ellipsoids are drawn at 30% probability level. Hydrogen atoms are omitted for clarity except H3 atoms. (Bottom) One-dimensional supramolecular chain formed by the intermolecular C–H⋯O hydrogen bonds.

Figure 5 
                  (Top) The molecular structure of 5. Ellipsoids are drawn at 30% probability level. For clarity, hydrogen atoms except H1 and H2 atoms and solvent molecules are omitted. (Bottom) One-dimensional chain formed by the intermolecular phenolic O→Sn coordination. The solvent molecule CHCl3 is omitted for clarity.
Figure 5

(Top) The molecular structure of 5. Ellipsoids are drawn at 30% probability level. For clarity, hydrogen atoms except H1 and H2 atoms and solvent molecules are omitted. (Bottom) One-dimensional chain formed by the intermolecular phenolic O→Sn coordination. The solvent molecule CHCl3 is omitted for clarity.

Table 1

Selected bond lengths (Å) and angles (°) for the complexes

1 2 3 4 [CH3OH·CHCl3]2 6
Sn(1)–C(1) 2.127(3) 2.132(4) 2.155(3) 2.149(4) 2.131(3) 2.162(2)
Sn(1)–C(7) 2.125(3) 2.117(4) 2.165(3) 2.157(4) 2.128(2) 2.152(2)
Sn(1)–C(13) 2.143(3) 2.132(4) 2.176(3) 2.158(4) 2.127(3) 2.158(2)
Sn(1)–O(1) 2.064(2) 2.059(3) 2.095(2) 2.081(3) 2.175(2) 2.0712(15)
Sn(1)–O(3) 2.378(2)
C(19)–O(1) 1.307(4) 1.312(5) 1.298(3) 1.310(4) 1.284(3) 1.302(2)
C(19)–O(2) 1.222(4) 1.217(5) 1.224(3) 1.210(5) 1.229(3) 1.217(3)
O(1)–Sn(1)–C(1) 115.27(11) 108.62(13) 105.48(10) 108.52(13) 87.41(8) 104.67(7)
O(1)–Sn(1)–C(7) 104.90(11) 111.99(14) 102.22(10) 105.55(13) 99.00(8) 104.45(9)
O(1)–Sn(1)–C(13) 96.51(10) 94.77(14) 94.92(11) 98.21(12) 94.72(9) 90.00(9)
C(1)–Sn(1)–C(7) 119.13(12) 117.52(15) 124.60(12) 115.89(14) 119.91(11) 124.89(9)
C(1)–Sn(1)–C(13) 111.32(12) 109.07(15) 112.91(12) 116.64(15) 120.00(11) 109.44(3)
C(7)–Sn(1)–C(13) 107.09(12) 112.49(17) 111.30(13) 109.87(14) 118.85(11) 116.26(11)
O(1)–Sn(1)–O(3) 174.18(7)

The other triphenyltin(iv) complex 5 with crystalline solvents (CH3OH and CHCl3) is a coordination polymer, and displays a one-dimensional zigzag chain with the Sn⋯Sn distances of 12.269(3) and 12.419(3) Å (Figure 3). The asymmetric unit contains two molecules of complex 5 with similar geometric parameters. In the chain, the (E)-3-[(2-hydroxy-1-naphthalenyl)methyleneamino]benzoate ligand bridges two tin atoms by a carboxylate oxygen atom and a phenolic oxygen atom, and each tin atom has a distorted trigonal bipyramidal geometry. Three phenyl groups are in the equatorial positions with the three C–Sn–C bond angles of 118.85(11)–120.00(11)° for Sn(1) and 116.73(11)–122.27(12)° for Sn(2), and the axial positions are occupied by a carboxylate oxygen and the phenolic oxygen of an adjacent ligand with the O–Sn–O bond angle of 174.18(7)° for Sn(1) and 175.07(7)° for Sn(2). The Sn–O(carboxylate) bonds (Sn(1)–O(1) 2.175(2) Å and Sn(2)–O(6) 2.167(2) Å) are shorter than the Sn–O(phenol) bonds (Sn(1)–O(1) 2.378(2) Å and Sn(2)–O(4) 2.391(2) Å), which are consistent with that observed in analogue triphenyltin(iv) 4-[(2-hydroxy-1-naphthyl)methyleneamino]benzoate, 4-(2-HOC10H6CH═N)C6H4COOSnPh3 (Yin et al., 2012). The long distance between the carbonyl oxygen of carboxylate and tin atom (Sn(1)⋯O(2) 3.015(3) Å) indicates that the carbonyl oxygen does not coordinate to tin atom and the carboxylate is monodentate.

Figure 6 
                  The molecular structure of 6. Ellipsoids are drawn at 30% probability level. Hydrogen atoms are omitted for clarity except H3 atoms.
Figure 6

The molecular structure of 6. Ellipsoids are drawn at 30% probability level. Hydrogen atoms are omitted for clarity except H3 atoms.

The three tricyclohexyltin(iv) complexes 3, 4, and 6 crystallize in P21/n, C2/c, and P-1 space groups, respectively, and the coordination environment of tin atom is a distorted tetrahedral geometry in which the four vertices are occupied by three carbon atoms of the cyclohexyl groups and the oxygen atom of the aminobenzoate (Figures 3, 4, and 6). The C–Sn–C bond angles around tin atoms are in the range of 111.30(13)–124.89(9)°, and the cyclohexyl groups bound to the tin atom adopt chair conformations with the Sn–C bond lengths of 2.149(4)–2.176(3) Å. The structural parameters are similar to those found in other reported tricyclohexyltin(iv) substituted-benzoates, such as 3-(2-HOC6H4CH═N)C6H4COOSnCy3 (Chen et al., 2020), 5-(2-HOC6H4CH═N)-2-HOC6H3COOSnCy3 (Liu et al., 2019), and 2-(5-Me-2-HOC6H3N═N)C6H4COOSnCy3 (Willem et al., 1998).

In these complexes, the C═N bond lengths (C(26)–N(1)) of the Schiff base carboxylate ligands are in the range of 1.272(5)–1.314(3) Å, and all imine C═N bonds exhibit E-type configuration. Except for complex 2, the two aromatic rings at both ends of the C═N double bond in 1 and 3–6 are not in the same plane, and the dihedral angles between the two aryl rings are 23.58(9)°, 35.43(6)°, 51.11(3)°, 14.81(5)°, and 29.29(5)°, respectively. In 1–4 and 6, there is the intramolecular O–H⋯N hydrogen bond involving hydroxyl O(3)–H(3) and imino N(1) atom, while in 5, this proton transfers from hydroxyl oxygen O(3) to nitrogen N(1) to form the intramolecular N–H⋯O hydrogen bond (Table 2).

Table 2

H-bonding geometry parameters (Å, °) for the complexes

Complex D–H⋯A D–H (Å) H⋯A (Å) D⋯A (Å) D–H⋯A (°) Code #
1 O(3)–H(3)⋯N(1) 0.85 1.89 2.650(4) 148
2 O(3)–H(3)⋯N(1) 0.85 1.77 2.552(5) 151
C(10)–H(10)⋯O(3)# 0.93 2.58 3.455(7) 157 x, 1 + y, z
3 O(3)–H(3)⋯N(1) 0.85 1.85 2.626(3) 151
C(35)–H(35)⋯O(2)# 0.97 2.54 3.508(5) 173 x, 1 − y, 1 − z
4 O(3)–H(3)⋯N(1) 0.85 1.90 2.645(5) 146
C(21)–H(21)⋯O(4)# 0.93 2.49 3.246(6) 138 x, 2 − y, 1/2 + z
[CH3OH·CHCl3]2 N(1)–H(1)⋯O(4) 0.86 1.91 2.597(3) 136
N(2)–H(2)⋯O(3) 0.86 1.91 2.583(3) 135
O(7)–H(7)⋯O(5) 0.85 1.97 2.805(3) 168
O(8)–H(8)⋯O(2) 0.85 1.91 2.751(3) 173
6 O(3)–H(3)⋯N(1) 0.85 1.80 2.549(2) 146

The crystal structure is stabilized by weak intermolecular interactions. In 2 and 4, the molecules are connected to one-dimensional supramolecular chain with Sn⋯Sn distance of 13.74(3) and 12.47(2) Å, respectively, by the intermolecular C–H⋯O hydrogen bonds (Figures 2 and 4, Table 2). In the crystal of 3, a pair of intermolecular C(35)–H(35)⋯O(2) hydrogen bonds connect two molecules into a centrosymmetric R 2 2(28) macrocycle, and the adjacent R 2 2(28) macrocycles are further linked by a pair of Sn(1)⋯O(3) (3.417(3) Å) interactions to form two-dimensional supramolecular network (Figure 4, Table 2).

2.4 UV-Vis absorption and fluorescence of the complexes

UV-Vis absorption and emission spectra of complexes 16 were recorded in dichloromethane solution (3.0×10−5 M) at room temperature (Figures 7 and 8, Table 3). In the complexes, the absorption bands at ∼235 nm are assigned to the π→π∗ electronic transitions of the aromatic rings, and the bands at 270–459 nm are assigned to the π→π∗ and n→π∗ transitions of the conjugated imine (C═N) in the ligand moiety (Gonzalez-Hernandez et al., 2021). The naphthyls (5 and 6) and substituents bonded to the benzene ring (14) affected the position of maximum absorption wavelength by the electronic delocalization. The complexes displayed similar emission at ∼440 nm (438 nm for 1, 2, and 46, 444 nm for 3) when excited at 310 nm (Figure 8a). When excited at their absorption maxima (Figure 8b), their maximum emission wavelengths did not change essentially, but their fluorescence intensity increased. The complexes may be explored for potential blue luminescent materials.

Figure 7 
                   The UV-Vis spectra of complexes 1–6 in CH2Cl2 (3.0 × 10−5 M).
Figure 7

The UV-Vis spectra of complexes 16 in CH2Cl2 (3.0 × 10−5 M).

Figure 8 
                  Fluorescence spectra of the complexes in CH2Cl2 (3.0 × 10−5 M). (a) At λ
                     ex = 310 nm and (b) at λ
                     ex = 352 nm (1), 356 nm (2), 377 nm (3), 356 nm (4), 378 nm (5), and 376 nm (6).
Figure 8

Fluorescence spectra of the complexes in CH2Cl2 (3.0 × 10−5 M). (a) At λ ex = 310 nm and (b) at λ ex = 352 nm (1), 356 nm (2), 377 nm (3), 356 nm (4), 378 nm (5), and 376 nm (6).

Table 3

UV-Vis absorption data of the complexesa

Complex λ max (nm) (ε max (M−1·cm−1))
1 245 (24,433) 270 (15,733) 306 (10,600) 352 (11,900)
2 233 (48,233) 273 (14,766) 311 (12,500) 356 (11,066)
3 234 (21,600) 270 (8,533) 377 (44,167)
4 234 (29,233) 278 (16,700) 310 (16,200) 356 (7,600)
5 235 (33,366) 319 (8,700) 377 (8,666) 438 (5,000) 459 (4,566)
6 235 (53,266) 319 (13,966) 377 (13,800) 438 (8,666) 457 (8,000)

aMeasured in CH2Cl2 (3.0 × 10−5 M).

2.5 In vitro cytotoxicity of the complexes

The cytotoxic activity of complexes 16 against two human cancer cells A549 and HeLa were evaluated using the MTT assay, and the results are listed in Table 4 (a dose-dependent antiproliferative effect was shown in ESI). The IC50 values of exposure to the complexes after 24 h showed that the activity of the complexes against A549 and HeLa cells was higher than that of the reference drug cisplatin, indicating that the compounds are all potent anti-cancer agents. Their cytotoxic activities were similar to those of the reported triorganotin(iv) complexes of aminobenzoic acid-based Schiff bases, such as triphenyltin [4-(dimethylamino)phenylmethyleneamino]benzoates (4-Me2NC6H4CH═N)C6H4COOSnPh3 (IC50 = 0.88–6.18 μM against HeLa) (Basu Baul et al., 2017a) and triorganotin 5-(salicylideneamino)salicylate, 5-(2-HOC6H4CH═N)-2-HOC6H3COOSnR3 (IC50 = 0.9–6.6 μM against A549 and HeLa) (Liu et al., 2019).

Table 4

IC50 (μM) of complexes recorded over a period of 24 ha

Complex A549 Hela
1 1.89 ± 0.13 1.59 ± 0.06
2 1.39 ± 0.09 1.49 ± 0.11
3 2.49 ± 0.12 1.59 ± 0.11
4 1.79 ± 0.05 0.89 ± 0.03
5 2.59 ± 0.10 1.19 ± 0.06
6 1.19 ± 0.07 0.89 ± 0.06
Cisplatin 21.3 ± 1.7 7.8 ± 0.3

aData represent mean value ± SD.

3 Conclusion

Six new triorganotin(iv) 3-(arylmethyleneamino)benzoates (16) have been synthesized by the one-pot three-component reaction. In chloroform solution, the tin atoms in these complexes have a four coordinated environment. In the crystalline state, the compounds adopt a four- or five-coordination mode. Complex 5 exhibits a zigzag chain in which tin atom has a distorted trans-[C3SnO2] trigonal bipyramidal geometry, and other complexes (1–4 and 6) are discrete molecules, and tin atoms possess distorted [C3SnO] tetrahedral configuration. These complexes display fluorescence with an emission at ∼440 nm in CH2Cl2 solution at room temperature, and may be explored for potential blue luminescent materials. Compounds 16 have potent in vitro cytotoxic activity against A549 and HeLa tumor cell lines (IC50 value of 0.89–2.59 μM), indicating their potential as potent anticancer agents, and can be further studied.

Experimental methods

Materials and physical measurements

The chemicals were of reagent grade and were purchased from Sinopharm Chemical Reagent Co. Ltd (Shanghai, China) and Energy Chemical Reagent Co. Ltd (Shanghai, China). The melting points were measured with a WRS-1A digital melting-point apparatus (Shanghai Precision Scientific Instrument Co. Ltd, Shanghai, China). C, H, and N analyses were performed using a Perkin Elmer 2400 Series II elemental analyzer (Perkin Elmer, Waltham, USA). IR spectra were recorded on a Nicolet Nexus 470 FT-IR spectrophotometer (Thermo Nicolet Corporation, USA) using KBr discs in the range of 4,000–400 cm−1. 1H and 13C NMR spectra were recorded on a Bruker Avance III HD500 NMR spectrometer (Bruker Corporation, Switzerland) with CDCl3 as solvent and Me4Si (TMS) as internal standard, and the chemical shifts are reported in δ units (ppm). 119Sn NMR spectra were recorded in CDCl3 on a Varian Mercury Vx300 spectrometer using Me4Sn external reference (Varian Corporation, USA). The UV-visible spectra were obtained on an Agilent 8453 spectrophotometer (Agilent Technologies Inc., USA). Fluorescence spectra were carried out by using a Hitachi F-4600 fluorescence spectrophotometer (HITACHI, Japan).

Synthesis of the complexes (1–6)

3-Aminobenzoic acid (0.137 g, 1 mmol), aldehyde (1 mmol) (for 5-bromo-2-hydroxybenzaldehyde 0.201 g, for 3,5-dibromo-2-hydroxybenzaldehyde 0.280 g, for 4-diethylamino-2-hydroxybenzaldehyde 0.193 g, for 2-hydroxy-3-methoxybenzaldehyde 0.152 g, for 2-hydroxy-1-naphthaldehyde 0.172 g), and triorganotin(iv) hydroxide (1 mmol) (for triphenyltin(iv) hydroxide 0.367 g, for tricyclohexyltin(iv) hydroxide 0.375 g) were mixed in 30 mL of benzene. The mixtures were refluxed under stirring, and the water formed during the reaction was removed using a Dean-Stark water separator. After the reaction is completed (about 6 h), the solution is cooled to room temperature and then filtered to remove the insoluble matter. The filtrate was evaporated by a rotary evaporator, and the residue was recrystallized from chloroform–methanol (1:3, v/v) or dichloromethane-hexane (1:2, v/v). The physical data of the compounds are as follows (for the NMR assignments, refer to the numbering in Scheme 1).

Triphenyltin(iv) 3-(5-bromo-2-hydroxybenzylideneamino)benzoate (1)

Light yellow product of 0.562 g (84%) was obtained. M.p. 149–150°C. Anal. Calcd for C32H24BrNO3Sn: C 57.44, H 3.62, N 2.09; found C 57.24, H 3.36, N 2.12%. IR (KBr, ν): 3,433 (O–H), 1,625 [(COO)as + C═N], 1,328 [(COO)s] cm−1. 1H NMR (CDCl3, δ): 13.07 (bs, 1H, OH), 8.59 (s, 1H, CH═N), 8.08 (d, 3 J = 7.5 Hz, 1H, H-6 of C6H4), 8.03 (d, 4 J = 1.5 Hz, 1H, H-2 of C6H4), 7.82–7.80 (m, 6H, 3 J(119Sn–1H) = 60 Hz, o-H of Ph), 7.51–7.42 (m, 13H, Ar-H), 6.92 (t, 3 J = 9.0 Hz, 1H, H-3 of C6H3) ppm. 13C NMR (CDCl3, δ): 172.05 (C═O), 162.07 (C═N), 160.22, 148.02, 136.00, 134.44, 132.25, 129.42, 126.13, 122.59, 120.48, 119.36, 110.63 (aromatic carbons), 138.19 (1 J(117/119Sn–13C) = 614/640 Hz, i-C of Ph), 136.95 (2 J(119Sn–13C) = 48 Hz, o-C of Ph), 130.33 (4 J(119Sn–13C) = 13 Hz, p-C of Ph), 129.05 (3 J(119Sn–13C) = 63 Hz, m-C of Ph) ppm. 119Sn NMR (CDCl3) δ: −102.7 ppm.

Triphenyltin(iv) 3-(3,5-dibromo-2-hydroxybenzylideneamino)benzoate (2)

The orange crystals (0.605 g, 81%) are obtained. M.p. 169–171°C. Anal. Calcd for C32H23Br2NO3Sn: C 51.38, H 3.10, N 1.87; found C 51.44, H 3.08, N 1.83%. IR (KBr, ν): 3,432 (O–H), 1,637 [(COO)as], 1,612 (C═N), 1,324 [(COO)s] cm−1. 1H NMR (CDCl3, δ): 14.20 (bs, 1H, OH), 8.68 (s, 1H, CH═N), 8.10 (d, 3 J = 7.5 Hz, 1H, H-6 of C6H4), 8.05 (s, 1H, H-2 of C6H4), 7.82–7.80 (m, 6H, 3 J(119Sn–1H) = 60 Hz, o-H of Ph), 7.75 (d, 4 J = 2.5 Hz, 1H, H-6 of C6H2), 7.52–7.44 (m, 12H, Ar-H) ppm. 13C NMR (CDCl3, δ): 171.87 (C═O), 161.02 (C═N), 157.33, 146.86, 138.39, 133.65, 132.42, 129.92, 129.58, 126.36, 122.33, 120.61, 112.25, 110.42 (aromatic carbons), 138.06 (1 J(117/119Sn–13C) = 614/642 Hz, i-C of Ph), 136.98 (2 J(119Sn–13C) = 49 Hz, o-C of Ph), 130.37 (4 J(119Sn–13C) = 13 Hz, p-C of Ph), 129.07 (3 J(119Sn–13C) = 64 Hz, m-C of Ph) ppm. 119Sn NMR (CDCl3) δ: −102.4 ppm.

Tricyclohexyltin(iv) 3-(4-diethylamino-2-hydroxybenzylideneamino)benzoate (3)

The brown crystal of 3 was obtained in 76% (0.516 g) yield. M.p. 136–138°C. Anal. Calcd for C36H52N2O3Sn: C 63.63, H 7.71, N 4.12; found C 63.44, H 7.68, N 4.16%. IR (KBr, ν): 3,440 (O–H), 1,634 [(COO)as + C═N], 1,329 [(COO)s] cm−1. 1H NMR (CDCl3, δ): 13.61 (vbs, 1H, OH⋯N), 8.50 (s, 1H, CH═N), 7.95 (s, 1H, H-2 of C6H4), 7.90 (d, 3 J = 7.5 Hz, 1H, H-6 of C6H4), 7.41 (t, 3 J = 7.5 Hz, 1H, H-5 of C6H4), 7.37 (d, 3 J = 8.5 Hz, 1H, H-4 of C6H4), 7.16 (d, 3 J = 9.0 Hz, 1H, H-6 of C6H3), 6.25 (dd, 4 J = 2.0 Hz, 3 J = 9.0 Hz, 1H, H-5 of C6H3), 6.19 (d, 4 J = 2.0 Hz, 1H, H-3 of C6H3), 3.40 (q, 3 J = 7.0 Hz, 4H, 2NCH2), 2.03–1.33 (m, 33H, Cy), 1.21 (t, 3 J = 7.0 Hz, 6H, 2CH3) ppm. 13C NMR (CDCl3, δ): 171.06 (COO), 160.95 (C═N), 164.21, 151.93, 148.80, 133.90, 133.54, 128.96, 127.18, 125.11, 121.67, 109.11, 103.87, 97.74 (aromatic carbons), 44.59 (2NCH2), 12.71 (2CH3), 33.97 (1 J(117/119Sn–13C) = 322/336 Hz, α-C), 31.14 (2 J(119Sn–13C) = 15 Hz, β-C), 28.94 (3 J(119Sn–13C) = 65 Hz, γ-C), 26.92 (δ-C) (Cy) ppm. 119Sn NMR (CDCl3) δ: 18.6 ppm.

Tricyclohexyltin(iv) 3-(3-methoxy-2-hydroxybenzylideneamino)benzoate (4)

The reaction gave an orange product with 88% (0.562 g) yield. M.p. 86–88°C. Anal. Calcd for C33H45NO4Sn: C 62.08, H 7.10, N 2.19; found C 62.22, H 6.96, N 2.21%. IR (KBr, ν): 3,433 (O–H), 1,641 [(COO)as], 1,617 (C═N), 1,327 [(COO)s] cm−1. 1H NMR (CDCl3, δ): 13.50 (bs, 1H, OH), 8.71 (s, 1H, CH═N), 8.01 (s, 1H, H-2 of C6H4), 8.00 (d, 3 J = 7.5 Hz, 1H, H-6 of C6H4), 7.47 (t, 3 J = 7.5 Hz, 1H, H-5 of C6H4), 7.43 (d, 3 J = 7.5 Hz, 1H, H-4 of C6H4), 7.03 (d, 3 J = 8.0 Hz, 1H, H-6 of C6H3), 7.00 (d, 3 J = 8.0 Hz, 1H, H-4 of C6H3), 6.89 (t, 3 J = 7.5 Hz, 1H, H-5 of C6H3), 3.94 (s, 3H, OCH3), 1.33–2.04 (m, 33H, Cy) ppm. 13C NMR (CDCl3, δ): 170.72 (COO), 163.30 (C═N), 151.47, 148.50, 148.17, 133.83, 129.18, 128.70, 125.69, 123.93, 121.86, 119.11, 118.65, 114.93 (aromatic carbons), 56.23 (OCH3), 34.04 (1 J(117/119Sn–13C) = 323/338 Hz, α-C), 31.15 (2 J(119Sn–13C) = 14 Hz, β-C), 28.94 (3 J(119Sn–13C) = 64 Hz, γ-C), 26.91 (δ-C) (Cy) ppm. 119Sn NMR (CDCl3) δ: 19.4 ppm.

Triphenyltin(iv) 3-((2-hydroxy-1-naphthalenyl)methyleneamino)benzoate (5)

Orange-yellow crystals, yield 0.512 (80%). M.p. 168–171°C. C36H27NO3Sn: C 67.53, H 4.25, N 2.19; found C 67.59, H 4.24, N 2.15%. IR (KBr, ν): 3,431 (O–H), 1,623 [(COO)as + C═N], 1,353 [(COO)s] cm−1. 1H NMR (CDCl3, δ): 15.33 (bs, 1H, O–H⋯N), 9.39 (s, 1H, CH═N), 8.13 (d, 3 J = 8.5 Hz, 1H, H-6 of C6H4), 8.13 (s, 1H, H-2 of C6H4), 8.07–8.05 (m, 1H, H of C10H6), 7.82 (dd, 4 J = 1.0 Hz, 3 J = 7.0 Hz, 3 J(119Sn–1H) = 64 Hz, 6H, o-H of Ph), 7.80 (d, 3 J = 9.0 Hz, 1H, H of C10H6), 7.71 (d, 3 J = 8.0 Hz, 1H, H of C10H6), 7.54–7.48 (m, 12H, (m,p)-H of Ph + H of C10H6), 7.34 (t, 3 J = 8.0 Hz, 1H, H-5 of C6H4), 7.09 (d, 3 J = 9.0 Hz, 1H, H-4 of C6H4) ppm. 13C NMR (CDCl3, δ): 171.93 (COO), 169.82 (CH═N), 155.44, 145.52, 136.86, 133.13, 132.42, 129.49, 129.33, 128.51, 128.15, 127.33, 125.17, 123.65, 121.92, 121.55, 119.08, 108.91 (aromatic carbons), 138.15 (1 J(117/119Sn–13C) = 620/648 Hz, i-C of Ph), 136.92 (2 J(119Sn–13C) = 50 Hz, o-C of Ph), 130.27 (4 J(119Sn–13C) = 13 Hz, p-C of Ph), 129.00 (3 J(119Sn–13C) = 62 Hz, m-C of Ph) ppm. 119Sn NMR (CDCl3) δ: −105.4 ppm.

Tricyclohexyltin(iv) 3-((2-hydroxy-1-naphthalenyl)methyleneamino)benzoate (6)

Orange-yellow crystals, yield 0.533 g (81%). M.p. 117–119°C. Anal. Calcd for C36H45NO3Sn: C 65.67, H 6.89, N 2.13; found C 65.44, H 6.86, N 2.19%. IR (KBr, ν): 3,435 (O–H), 1,618 [(COO)as + C═N], 1,327 [(COO)s] cm−1. 1H NMR (CDCl3, δ): 15.43 (bs, 1H, OH⋯N), 9.42(s, 1H, CH═N), 8.13 (d, 3 J = 9.0 Hz, 1H, H-6 of C6H4), 8.11 (s, 1H, H-2 of C6H4), 8.00–7.98 (m, 1H, H of C10H6), 7.80 (d, 3 J = 9.0 Hz, 1H, H of C10H6), 7.71 (d, 3 J = 8.0 Hz, 1H H of C10H6), 7.54–7.49 (m, 3H, H of C10H6), 7.34 (t, 3 J = 7.5 Hz, 1H, H-5 of C6H4), 7.09 (d, 3 J = 9.0 Hz, 1H, H-4 of C6H4), 2.06–1.32 (m, 33H, Cy) ppm. 13C NMR (CDCl3, δ): 170.52 (COO), 170.15 (CH═N), 154.97, 145.24, 136.79, 134.09, 133.21, 129.41, 129.34, 128.17, 128.11, 127.32, 124.63, 123.60, 122.11, 120.88, 119.01, 108.88 (aromatic carbons), 34.06 (1 J(117/119Sn–13C) = 322/336 Hz, α-C), 31.16 (2 J(119Sn–13C) = 15 Hz, β-C), 29.83 (3 J(119Sn–13C) = 63 Hz, γ-C), 26.91 (4 J(119Sn–13C) = 7 Hz, δ-C) (Cy) ppm. 119Sn NMR (CDCl3) δ: 17.3 ppm.

X-ray crystallography

The yellow to orange single crystals of 16 were obtained by slow evaporation of CHCl3-MeOH (1:2, v/v) solution, respectively. Compound 5 crystallizes with two molecules of each MeOH and CHCl3 (CH3OH·CHCl3). Diffractions data were collected at 295(2) K on a Bruker D8 Quest CCD fitted with graphite monochromatized Mo-Kα radiation (0.71073 Å). The structure solution and refinement were completed using SHELXS Version 2014/5 (Sheldrick, 2015a) and SHELXL-2018 (Sheldrick, 2015b), respectively. Non-hydrogen atoms were refined with anisotropic displacement parameters. Hydrogen atoms bonded to C were included in their calculated positions and refined in the riding-model, and hydrogen atoms of the hydroxyl and amino groups were freely refined. Crystal data and refinement details are listed in Table 5. Crystallographic data have been deposited with the Cambridge Crystallographic Data Center as supplementary publication numbers CCDC 2167238–2167243.

Table 5

Crystallographic and refinement data for the complexes

Complex 1 2 3 4 [CH3OH·CHCl3]2 6
Formula C32H24BrNO3Sn C32H23Br2NO3Sn C36H52N2O3Sn C33H45NO4Sn C76H64Cl6N2O8Sn2 C36H45NO3Sn
Formula weight 669.12 748.02 679.48 638.39 1,583.37 658.42
Crystal system Monoclinic Triclinic Monoclinic Monoclinic Triclinic Triclinic
Space group C2/c P-1 P21/n C2/c P-1 P-1
a (Å) 52.125(7) 7.6169(15) 18.529(2) 43.516(4) 10.0543(14) 8.9941(13)
b (Å) 6.6015(8) 13.743(3) 9.6937(12) 7.8719(7) 16.823(2) 10.6435(15)
c (Å) 17.029(2) 15.096(3) 20.779(3) 19.3847(19) 22.907(3) 18.484(3)
α (°) 90 100.394(4) 90 90 89.664(4) 77.027(4)
β (°) 108.791(4) 100.809(4) 110.565(4) 109.226(3) 78.352(4) 77.660(4)
γ (°) 90 103.453(4) 90 90 73.409(3) 70.385(4)
Volume (Å3) 5,547.4(12) 1,467.6(5) 3,494.4(8) 6,270.0(10) 3,631.2(9) 1,605.7(4)
Z 8 2 4 8 2 2
D c (g·cm−3) 1.602 1.693 1.292 1.353 1.448 1.362
μ (mm−1) 2.395 3.626 0.766 0.851 0.964 0.831
F(000) 2,656 732 1,424 2,656 1,600 684
θ range (°) 3.1–26.0 1.8–26.0 2.3–26.0 2.1–26.0 2.2–26.0 2.5–26.0
Tot. reflections 32,075 9,466 40,142 34,750 75,403 31,932
Uniq. reflections 5,401 5,737 6,807 6,090 14,172 6,232
R int 0.028 0.023 0.025 0.028 0.017 0.018
GOF on F 2 1.05 1.07 1.05 1.08 1.03 1.05
R 1 indices 0.031 0.041 0.035 0.042 0.032 0.023
wR 2 indices 0.072 0.101 0.102 0.119 0.091 0.064
Δρ min (e·Å−3) −1.346 −0.462 −0.523 −0.874 −0.725 −0.505
Δρ max (e·Å−3) 1.311 1.000 0.874 1.549 1.167 0.683

3.1 In vitro screening

The human tumor cell lines A549 (lung tumor cells) and HeLa (cervix tumor cells) were obtained from the Cell Bank of Type Culture Collection of the Chinese Academy of Sciences, Shanghai, China. The cells were cultured in Dulbecco’s modified Eagle medium supplemented with 10% heat-inactivated new-born calf serum and 1% penicillin–streptomycin solution at 310 K in a humidified 5% CO2 incubator.

After plating 5,000 tumor cells per well in 96-well plates, the cells were pre-incubated in drug-free media at 310 K for 24 h before adding various concentrations of the tin complexes to be tested. In order to prepare stock solutions, each complex was dissolved in DMSO and cisplatin was dissolved in phosphate-buffered saline of pH 7.2. This stock was further diluted using cell culture medium until working concentrations were achieved. The drug exposure period was 24 h. Subsequently, 15 μL of 5 mg·mL−1 MTT solution was added to form purple formazan. Afterwards, 100 μL of DMSO was transferred into each well to dissolve the purple formazan. Each well was triplicated and each experiment repeated at least three times. The dose causing 50% inhibition of cell growth (IC50) was calculated and the data were expressed as the mean values ± standard deviation.

  1. Funding information: Authors state no funding involved.

  2. Author contributions: Ruili Wang: experimental work and writing – original draft; Jing Zhang: experimental work; Daoyu Cui: experimental work; Laijin Tian: methodology and writing – review and editing.

  3. Conflict of interest: Authors state no conflict of interest.

  4. Data availability statement: Crystallographic data (CCDC 2167238–2167243) for this article can be obtained free of charge from The Cambridge Crystallographic Data Center via www.ccdc.cam.ac.uk/data_request/cif.

References

Bantia C.N., Hadjikakou S.K., Sismanoglu T., Hadjiliadis N., Anti-proliferative and antitumor activity of organotin(iv) compounds. An overview of the last decade and future perspectives. J. Inorg. Biochem., 2019, 194, 114–152.10.1016/j.jinorgbio.2019.02.003Search in Google Scholar PubMed

Basu Baul T.S., Basu S., Vos D., Linden A., Amino acetate functionalized Schiff base organotin(iv) complexes as anticancer drugs: synthesis, structural characterization, and in vitro cytotoxicity studies. Investig. N. Drugs, 2009, 27, 419–431.10.1007/s10637-008-9189-1Search in Google Scholar PubMed

Basu Baul T.S., Das P., Rivarola E., Song X., Eng G., Synthesis, spectroscopic characterization and structures of tributyltin(iv) 4-[((E)-1-{2-hydroxy-5-[(E)-2-(aryl)-1-diazenyl]-phenyl}methylidene)amino]benzoates.Toxicity studies on the second larval instar of the anopheles stephensi mosquito larvae. J. Inorg. Organomet. Polym. Mater., 2010, 20, 61–68.10.1007/s10904-009-9308-2Search in Google Scholar

Basu Baul T.S., Dutta D., Duthie A., Prasad R., Rana N.K., Koch B., et al., Triphenyltin benzoates with diazenyl/imino scaffold exhibiting remarkable apoptosis mediated by reactive oxygen species. J. Inorg. Biochem., 2017a, 173, 79–92.10.1016/j.jinorgbio.2017.04.020Search in Google Scholar PubMed

Basu Baul T.S., Kehie P., Duthie A., Guchhai N., Raviprakash N., Mokhamatam R.B., et al., Synthesis, photophysical properties and structures of organotin-Schiff bases utilizing aromatic amino acid from the chiral pool and evaluation of the biological perspective of a triphenyltin compound. J. Inorg. Biochem., 2017b, 168, 76–89.10.1016/j.jinorgbio.2016.12.001Search in Google Scholar PubMed

Basu Baul T.S., Longkumer I., Duthie A., Singh P., Koch B., da Silva M.F.C.G., Triphenylstannyl((arylimino)methyl)benzoates with selective potency that induce G1 and G2/M cell cycle arrest and trigger apoptosis via ROS in human cervical cancer cells. Dalton Trans., 2018, 47, 1993–2008.10.1039/C7DT04037GSearch in Google Scholar PubMed

Basu Baul T.S., Masharing C., Ruisi G., Jirasko R., Holcapek M., Vos D., et al., A. Self-assembly of extended Schiff base amino acetate skeletons, 2-[[(2Z)-(3-hydroxy-1-methyl-2-butenylidene)]amino]phenyl-propionate and 2-[[(E)-1-(2-hydroxyaryl)alkylidene]-aminophenylpropionate] skeletons incorporating organotin(iv) moieties: synthesis, spectroscopic characterization, crystal structures, and in vitro cytotoxic activity. J. Organomet. Chem., 2007, 692, 4849–4862.10.1016/j.jorganchem.2007.06.061Search in Google Scholar

Beltran H.I., Zamudio-Rivera L.S., Mancilla T., Santillan R., Farfan, N., One-step preparation, structural assignment, and X-ray study of 2,2-di-nbutyl-and 2,2-diphenyl-6-aza-1,3-dioxa-2-stannabenzocyclononen-4-ones derived from amino acids. Chem. Eur. J., 2003, 9, 2291–2306.10.1002/chem.200204260Search in Google Scholar PubMed

Chen L., Wang Z., Qiu T., Sun R., Zhao Z., Tian L., et al., Synthesis, structural characterization, and properties of triorganotin(iv) complexes of Schiff base derived from 3-aminobenzoic acid and salicylaldehyde or 2,4-pentanedione. Appl. Organomet. Chem., 2020, 34, e5790.10.1002/aoc.5790Search in Google Scholar

Davies A.G., Gielen M., Pannell K.H., Tiekink E.R.T., Tin chemistry: Fundamentals, frontiers, and applications. John Wiley & Sons, Chichester, U.K., 2008.10.1002/9780470758090Search in Google Scholar

Deacon G.B., Phillips R.J., Relationships between the carbon-oxygen stretching frequencies of carboxylato complexes and the type of carboxylate coordination. Coord. Chem. Rev., 1980, 33, 227–250.10.1016/S0010-8545(00)80455-5Search in Google Scholar

Dias L.C., Lima G.M., Takahashi J.A., Ardisson J.D., New di- and triorganotin(iv) carboxylates derived from a Schiff base: synthesis, characterization and in vitro antimicrobial activities. Appl. Organomet. Chem., 2015, 29, 305–313.10.1002/aoc.3292Search in Google Scholar

Goh N.K., Chu C.K., Khoo L.E., Whalen D., Eng G., Smith F.E., The Synthesis, structural characterization and biocidal properties of some triorganotin esters of N-arylideneamino acids. Appl. Organomet. Chem., 1998, 12, 457–466.10.1002/(SICI)1099-0739(199806)12:6<457::AID-AOC720>3.0.CO;2-ZSearch in Google Scholar

Gonzalez-Hernandez A., Leon-Negrete A., Galvan-Hidalgo J.M., Gomez E., Villamil-Ramos R., Barba V., Fused hexacyclic organotin(iv) compounds derived from 3-[((2-hydroxy- naphthalen-1-yl)methylene)amino]naphthalen-2-ol. J. Mol. Struct., 2021, 1242, 130807.10.1016/j.molstruc.2021.130807Search in Google Scholar

Hadi S., Fenska M.D., Noviany N., Satria H., Simanjuntak W., Naseer M.M., Synthesis and antimalarial activity of some triphenyltin(iv) aminobenzoate compounds against Plasmodium falciparum. Main. Group. Met. Chem., 2021, 44, 256–260.10.1515/mgmc-2021-0028Search in Google Scholar

Holecek J., Nadvornik M., Handlir K., 13C and 119Sn NMR spectra of dibutyltin(iv) compounds. J. Organomet. Chem., 1986, 315, 299–308.10.1016/0022-328X(86)80450-8Search in Google Scholar

Holecek J., Nadvornik M., Handlir K., Lycka A., 13C and 119Sn NMR study of some four- and five-coordinate triphenyltin(iv) compounds. J. Organomet. Chem., 1983, 241, 177–184.10.1016/S0022-328X(00)98505-XSearch in Google Scholar

Kaur M., Kumar S., Yusuf M., Lee J., Brown R.J.C., Kim K.-H., et al., Post-synthetic modification of luminescent metal-organic frameworks using Schiff base complexes for biological and chemical sensing. Coord. Chem. Rev., 2021, 449, 214214.10.1016/j.ccr.2021.214214Search in Google Scholar

Liang J., Du D., Xiao X., Han X., Zhu D., Mei Z., A series of organotin(iv) complexes based on (E)-3-(3-nitrophenyl) acrylic acid: Syntheses, crystal structures and biological activities. Inorg. Chem. Comm., 2014, 40, 133–137.10.1016/j.inoche.2013.12.001Search in Google Scholar

Liu J., Lin Y., Liu M., Wang S., Liu X., Tian L., Synthesis, structural characterization and cytotoxic activity of triorganotin 5-(salicylideneamino)salicylates. Appl. Organometal. Chem., 2019, 33, e4715.10.1002/aoc.4715Search in Google Scholar

Nath M., Saini P.K., Chemistry and applications of organotin(iv) complexes of Schiff bases. Dalton Trans., 2011, 40, 7077–7121.10.1039/c0dt01426eSearch in Google Scholar PubMed

Pellerito L., Nagy L., Organotin(iv) complexes formed with biologically active ligands: equilibrium and structural studies, and some biological aspects. Coord. Chem. Rev., 2002, 224, 111–150.10.1016/S0010-8545(01)00399-XSearch in Google Scholar

Saeed A., Channar P.A., Larik F.A., Jabeen F., Muqadar U., Saeed S., et al., Design, synthesis, molecular docking studies of organotin-drug derivatives as multi-target agents against antibacterial, antifungal, α-amylase, α-glucosidase and butyrylcholinesterase. Inorg. Chim. Acta, 2017, 464, 204–213.10.1016/j.ica.2017.05.036Search in Google Scholar

Sheldrick G.M., SHELXT- Integrated space-group and crystal-structure determination. Acta Crystallogr., 2015a, A71, 3–8.10.1107/S2053273314026370Search in Google Scholar PubMed PubMed Central

Sheldrick G.M., Crystal structure refinement with SHELXL, Acta Crystallogr., 2015b, C71, 3–8.10.1107/S2053229614024218Search in Google Scholar PubMed PubMed Central

Tiekink E.R.T., Structural chemistry of organotin carboxylates: a review of the crystallographic literature. Appl. Organometal. Chem., 1991, 5, 1–23.10.1002/aoc.590050102Search in Google Scholar

Tiekink E.R.T., The rich diversity in tin carboxylate structures. Trends Organomet. Chem., 1994, 1, 71–116.Search in Google Scholar

Tzimopoulos D., Czapik A., Gdaniec M., Bakas T., Isab A.A., Varvogli A.-C., et al., Synthesis and study of triorganostannyl esters of 3-,4- and 3,5-pyridinylimino substituted aminobenzoic acids: Crystal structures of dimorphs of aqua-trimethyltin 3-pyridinyliminobenzoate. J. Mol. Struct. 2010, 965, 56–64.10.1016/j.molstruc.2009.11.038Search in Google Scholar

Tzimopoulos D., Gdaniec M., Bakas T., Akrivos P.D., Structural elucidation for triorganotin derivatives of 3-amino, 4-amino and 3,5-diaminobenzoate. Crystal structures of triphenyltin 4-aminobenzoate and trimethyl and triphenyltin 3,5-diaminobenzoate. J. Coord. Chem., 2009, 62, 1218–1231.10.1080/00958970802521084Search in Google Scholar

Vinayak R., Nayek H.P., Organotin metalloligands for selective sensing of metal ions. N. J. Chem., 2019, 43, 7259–7268.10.1039/C9NJ00944BSearch in Google Scholar

Willem R., Verbruggen I., Gielen M., Biesemans M., Mahieu B., Basu Baul T.S., et al., Correlating mossbauer and solution- and solid-state 117Sn NMR data with X-ray diffraction structural data of triorganotin 2-[(E)-2-(2-hydroxy-5-methylphenyl)-1-diazenyl]benzoates. Organometallics, 1998, 17, 5758–5766.10.1021/om980504uSearch in Google Scholar

Yao Y., Yang M., Zheng X., Tian L., Synthesis, characterization, and cytotoxic activity of triphenyltin complexes of N-(5-bromosalicylidene)-α-amino acids. Main. Group. Met. Chem., 2017, 40, 93–99.10.1515/mgmc-2017-0015Search in Google Scholar

Yin H., Liu H., Hong M., Synthesis, structural characterization and DNA-binding properties of organotin(iv) complexes based on Schiff base ligands derived from 2-hydroxy-1-naphthaldy and 3- or 4-aminobenzoic acid. J. Organomet. Chem., 2012, 713, 11–19.10.1016/j.jorganchem.2012.03.027Search in Google Scholar

Yin H., Wang Q., Xue S., Synthesis and spectroscopic properties of [N-(4-carboxyphenyl)-salicylideneiminato] di- and tri-organotin(iv) complexes and crystal structures of {[Bu2Sn(2-OHC6H4CH = NC6H4COO-4)]2O}2 and Ph3Sn(2-OHC6H4CH = NC6H4COO-4)). J. Organomet. Chem., 2005, 690, 435–440.10.1016/j.jorganchem.2004.09.063Search in Google Scholar

Yu S., Li C., Fan S., Wang J., Liang L., Hong M., Three organotin(iv) Schiff-base carboxylates: Synthesis, structural characterization and in vitro cytotoxicity against cis-platin-resistant cancer cells. J. Mol. Struct., 2022, 1257, 132585.10.1016/j.molstruc.2022.132585Search in Google Scholar

Zhang D., Xie Q., Zheng J., Li J., Li S., 13C and 119Sn NMR study of tricyclohexyltin(iv) carboxylates. Chin. J. Magn. Reson., 1990, 7, 101–108.Search in Google Scholar

Received: 2022-04-20
Revised: 2022-11-17
Accepted: 2022-12-07
Published Online: 2022-12-31

© 2022 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Embedded three spinel ferrite nanoparticles in PES-based nano filtration membranes with enhanced separation properties
  2. Research Articles
  3. Syntheses and crystal structures of ethyltin complexes with ferrocenecarboxylic acid
  4. Ultra-fast and effective ultrasonic synthesis of potassium borate: Santite
  5. Synthesis and structural characterization of new ladder-like organostannoxanes derived from carboxylic acid derivatives: [C5H4N(p-CO2)]2[Bu2Sn]4(μ3-O)2(μ2-OH)2, [Ph2CHCO2]4[Bu2Sn]4(μ3-O)2, and [(p-NH2)-C6H4-CO2]2[Bu2Sn]4(μ3-O)2(μ2-OH)2
  6. HPA-ZSM-5 nanocomposite as high-performance catalyst for the synthesis of indenopyrazolones
  7. Conjugation of tetracycline and penicillin with Sb(v) and Ag(i) against breast cancer cells
  8. Simple preparation and investigation of magnetic nanocomposites: Electrodeposition of polymeric aniline-barium ferrite and aniline-strontium ferrite thin films
  9. Effect of substrate temperature on structural, optical, and photoelectrochemical properties of Tl2S thin films fabricated using AACVD technique
  10. Core–shell structured magnetic MCM-41-type mesoporous silica-supported Cu/Fe: A novel recyclable nanocatalyst for Ullmann-type homocoupling reactions
  11. Synthesis and structural characterization of a novel lead coordination polymer: [Pb(L)(1,3-bdc)]·2H2O
  12. Comparative toxic effect of bulk zinc oxide (ZnO) and ZnO nanoparticles on human red blood cells
  13. In silico ADMET, molecular docking study, and nano Sb2O3-catalyzed microwave-mediated synthesis of new α-aminophosphonates as potential anti-diabetic agents
  14. Synthesis, structure, and cytotoxicity of some triorganotin(iv) complexes of 3-aminobenzoic acid-based Schiff bases
  15. Rapid Communications
  16. Synthesis and crystal structure of one new cadmium coordination polymer constructed by phenanthroline derivate and 1,4-naphthalenedicarboxylic acid
  17. A new cadmium(ii) coordination polymer with 1,4-cyclohexanedicarboxylate acid and phenanthroline derivate: Synthesis and crystal structure
  18. Synthesis and structural characterization of a novel lead dinuclear complex: [Pb(L)(I)(sba)0.5]2
  19. Special Issue: Theoretical and computational aspects of graph-theoretic methods in modern-day chemistry (Guest Editors: Muhammad Imran and Muhammad Javaid)
  20. Computation of edge- and vertex-degree-based topological indices for tetrahedral sheets of clay minerals
  21. Structures devised by the generalizations of two graph operations and their topological descriptors
  22. On topological indices of zinc-based metal organic frameworks
  23. On computation of the reduced reverse degree and neighbourhood degree sum-based topological indices for metal-organic frameworks
  24. An estimation of HOMO–LUMO gap for a class of molecular graphs
  25. On k-regular edge connectivity of chemical graphs
  26. On arithmetic–geometric eigenvalues of graphs
  27. Mostar index of graphs associated to groups
  28. On topological polynomials and indices for metal-organic and cuboctahedral bimetallic networks
  29. Finite vertex-based resolvability of supramolecular chain in dialkyltin
  30. Expressions for Mostar and weighted Mostar invariants in a chemical structure
Downloaded on 14.1.2026 from https://www.degruyterbrill.com/document/doi/10.1515/mgmc-2022-0026/html?lang=en
Scroll to top button