Next Article in Journal
The Coumarin Psoralidin Enhances Anticancer Effect of Tumor Necrosis Factor-Related Apoptosis-Inducing Ligand (TRAIL)
Previous Article in Journal
Two New Diterpenoids from the Buds of Wikstroemia chamaedaphne
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Spectroscopy, Thermal Analysis, Magnetic Properties and Biological Activity Studies of Cu(II) and Co(II) Complexes with Schiff Base Dye Ligands

Chemistry Department, Faculty of Sciences, Arak University, Dr. Beheshti Ave., Arak 38156-88349, Iran
*
Author to whom correspondence should be addressed.
Molecules 2012, 17(6), 6434-6448; https://doi.org/10.3390/molecules17066434
Submission received: 9 April 2012 / Revised: 9 May 2012 / Accepted: 10 May 2012 / Published: 29 May 2012

Abstract

:
Three azo group-containing Schiff base ligands, namely 1-{3-[(3-hydroxy-propylimino)methyl]-4-hydroxyphenylazo}-4-nitrobenzene (2a), 1-{3-[(3-hydroxypropyl-imino)methyl]-4-hydroxyphenylazo}-2-chloro-4-nitrobenzene (2b) and 1-{3-[(3-hydroxy-propylimino)methyl]-4-hydroxyphenylazo}-4-chloro-3-nitrobenzene (2c) were prepared. The ligands were characterized by elemental analysis, FTIR spectroscopy, UV-Vis spectroscopy, 13C- and 1H-NMR spectroscopy and thermogravimetric analysis. Next the corresponding copper(II) and cobalt(II) metal complexes were synthesized and characterized by the physicochemical and spectroscopic methods of elemental analysis, FTIR spectroscopy, UV-Vis spectroscopy, magnetic moment measurements, and thermogravimetric analysis (TGA) and (DSC). The room temperature effective magnetic moments of complexes are 1.45, 1.56, 1.62, 2.16, 2.26 and 2.80 B.M. for complexes 3a, 3b, 3c, 4a 4b, and 4c, respectively, indicating that the complexes are paramagnetic with considerable electronic communication between the two metal centers.

1. Introduction

Azo Schiff base complexes contain both azo and azomethine groups. The azo group possesses excellent donor properties and is important in coordination chemistry [1,2,3], and some azo compounds have been shown to possess good antibacterial activity [4,5]. Schiff bases are well known to have antifungal, antitumor and herbicidal activities [6,7,8,9,10,11]. Salicylaldimine-based ligands have found applications in preparation of metallomesogens, optical metal ion detection and enantioselective catalysis [12,13,14,15]. Azo Schiff bases are commonly synthesized by coupling a diazonium reagent with an aromatic aldehyde to form an azo aldehyde [16,17]. The azomethine group has good donor properties and can form stable complexes with transition metal ions [18,19,20]. The azo and azomethine groups on azo Schiff base ligands are oriented in such a way that coordination of both groups to a metal ion is not possible, thus, preferential coordination of the azomethine group while the azo group is left free and uncoordinated has been observed [21,22,23,24]. Schiff bases derived from salicylaldehydes are known polydentate ligands, coordinating to metals in both their deprotonated and neutral forms [25,26]. Some cobalt and copper complexes exhibit diverse biological properties viz. anti-inflammatory, antibacterial and anticancer, etc. [27].
Many Schiff base complexes show excellent catalytic activity for various reactions at high temperatures (>100 °C) and in the presence of water. Over the past few years, there have been many reports on their applications in homogeneous and heterogeneous catalysis [28,29]. A wide variety of cobalt(II) and copper(II) complexes are known to bind dioxygen more or less reversibly and are therefore are frequently studied as model compounds for natural oxygen carriers and for O2 storage [30]. Some thermochemical work on a limited series of cobalt(II) and copper(II) complexes has been reported by various researchers [31,32]. Based on the aforementioned properties of Schiff bases and azo compounds, we reported herein the synthesis and spectroscopic studies as well as thermal investigation of novel salicylaldimine-based ligands 1-{3-[(3-hydroxypropyl-imino)methyl]-4-hydroxyphenylazo}-4-nitrobenzene (2a), 1-{3-[(3-hydroxypropyl-imino)methyl]-4-hydroxyphenyl-azo}-2-chloro-4-nitrobenzene (2b) and 1-{3-[(3-hydroxypropyl-imino)methyl]-4-hydroxyphenylazo}-4-chloro-3-nitrobenzene (2c). 13C- and 1H-NMR spectra were obtained to determine the structure of the ligands 2ac. The copper(II) and cobalt(II) complexes derived from azo-linked salicylaldimine were also prepared and their structures were confirmed by elemental analysis, FTIR spectroscopy, UV-Vis spectroscopy, thermogravimetric analysis and magnetic moment measurements.

2. Results and Discussion

2.1. Biological Studies

Three ligands and six complexes were screened for their activity against Escherichia coli (Gram negative), Staphylococcus aureus (Gram positive), Micrococcus luteus, Enterobacter cloacae, Klebsiella pneumoniae and antifungal activities against Aspergillus niger and Pencillium. The inhibition zones were measured and compared with standard drugs. The antibacterial and antifungal activities of the new compounds are presented in Table 1. The results indicate that the ligands have no activity against fungi and have some activity against the bacteria. The complexes have no activity against fungi, except for complex 3a. complex 3b has no activity against bacteria, while complexes 3a,c have activity against bacteria. The complexes 4ac have activity against bacteria.
Table 1. Antimicrobial activity studies of ligands and their metal complexes.
Table 1. Antimicrobial activity studies of ligands and their metal complexes.
CompoundE. coliS. aureusM. luteuE. cloacaeK. pneumoniaeA. nigerPencillium
2a4300000
2b3400000
2c5333200
3a0630004
3b0000000
3c0200000
4a5210000
4b0000200
4c0020000

2.2. Infrared Spectra

The infrared spectra of the free ligands and the metal complexes were obtained over a spectral range of 4,000–300 cm−1. A comparison of the infrared spectra of the ligands and the respective complexes reveals the absence of absorption bands associated with the O-H stretching of the phenol groups, indicating the loss of the protons upon complexation, along with formation of a metal-oxygen bond. This finding was further supported by an increase in (C-O) frequency in the spectra of the metal complexes compared to the ligands [33]. The bands at 1,647–1,651 cm−1 were assigned to the stretching vibration of the azomethine group of the ligands 2ac. This band is shifted in the complexes toward lower frequencies. Reduction of the double bond character of the C=N bond, which is caused by the coordination of nitrogen into the metal center, is in agreement with results obtained from similar complexes [13,16,34]. Based on data from earlier reports, we could assign the bands at 408–443 and 524–599 cm−1 to M-N and M-O vibrations, respectively [35].

2.3. Electronic Spectra

The features of the electronic spectra of the cobalt(II) complexes are very similar to each other, and they are typical for tetrahedral high-spin cobalt(II) complexes. The electronic structures of cobalt(II) complexes with different ligands have been presented in the literature [36,37]. Based on the simplest model, three spin-allowed crystal field bands are expected in the spectra of tetrahedral cobalt(II) complexes, i.e.,4A2(F) → 4T2(F), 4A2(F) → 4T1(F), 4A2(F) → 4T1(P). Usually this type of cobalt(II) complex exhibits two bands between 610 nm and 830 nm that can be assigned to the 4A2(F) → 4T2(F) and 4A2(F) → 4T1(F) transitions, respectively. The transition 4A2(F) → 4T1(P) is usually observed as a well-defined shoulder at approximately 550 nm [38]. The electronic spectra of the complexes were obtained from solid samples using the diffuse reflectance technique. The cobalt(II) complexes each show a broad shoulder with three bands (605, 489 and 278 nm for 4a, 622, 508 and 270 nm for 4b, and 528, 438 and 285 nm for 4c). These bands are assigned as spin-allowed crystal field transitions. Absorption bands are observed at 278, 270 and 285 nm for in the spectra of 4a, 4b and 4c, respectively. These bands are attributed to charge transfer from the non-bonding orbitals of the oxygen atoms in the ligand to the cobalt(II) d orbitals. The last absorption band at approximately 262 nm is assigned to the π → π* or n → π* transitions of the ligand [39,40,41]. The spectra of the copper(II) complexes display broad bands at 457 nm for complex 3a, 610 nm for complex 3b and 547 nm for complex 3c due to the ligand field transition for the CuNO3 chromophore [42,43]. The second absorption bands at 360, 384 and 441 nm for compounds 3a, 3b and 3c, respectively, are assigned to charge transfer from the non-bonding orbital of the oxygen atoms to the vacant copper(II) d orbitals [42,44]. The last absorption band observed at approximately 265 nm for each of the complexes is associated with π → π* or n → π* transitions of the ligand [45].

2.4. Magnetic Moment

The magnetic moments of the cobalt(II) and copper(II) complexes were determined by the Evans method [46]. This method is based on the principle that the position of a given proton resonance (i.e., t-butyl alcohol) in the NMR spectrum of a molecule is dependent on the bulk volume magnetic susceptibility of the medium in which the molecule is found. The shift of the proton resonance for an inert substance due to the presence of paramagnetic ions is given by theoretical expression (1):
Molecules 17 06434 i001
In this equation, Δυ is the shift, υo is the applied field, χυ is the volume magnetic susceptibility of the solution containing paramagnetic ions and χυ' is the volume magnetic susceptibility of the reference solution. The effective magnetic moments (μeff) of the compounds in this study were determined at room temperature to be 1.45, 1.56, 1.62, 2.16, 2.26 and 2.80 B.M. for 3a, 3b, 3c, 4a, 4b and 4c, respectively. These values were calculated using the spin only equation, μeff = 2.82[T χcorr − Nα]1/2, where T is the absolute temperature, χcorr is the molar magnetic susceptibility corrected for paramagnetism of the constituent atoms, and Nα is the temperature independent pararmagnetism. The magnetic moment of the cobalt(II) complexes at room temperature were found to be 2.16 to 2.80 B.M. range per cobalt(II) ion, which appears to be low for a d7 high spin configuration, indicating that there must be a strong spin-spin interaction through the bridging ligands [47]. The values of μeff for the copper(II) complexes at room temperature were found to be in the range of 1.45–1.64 B.M. per copper(II) ion. These data are low for a d9 configuration. The subnormal magnetic moment was ascribed to a significant electronic coupling between the two copper(II) centers, postulated to occur through the bridging atoms (through a superexchange pathway) [48,49].

2.5. 13C- and 1H-NMR Spectra of the Ligands

1H- and 13C-NMR spectra of the ligands were recorded in DMSO-d6. In the 1H-NMR spectra of the ligands, the singlet resonances in the 14–13.21 and 4.7–4.79 ppm regions can be attributed to the protons of the phenolic and alcoholic OH groups, respectively. All three ligands exhibit singlet signals in the range of 8.2–8.76 ppm, which were attributed to the azomethine group protons. In addition, the multiple signals in the range of approximately 3.7–1.83 ppm were attributed to aliphatic protons. The signals of 13C-NMR of the ligands 2a, 2b and 2c, were appeared in the range of 32.8–58.2 and 115.3–166.7 ppm are assigned to aliphatic and aromatic carbon, respectively. The signal in the range of 176–178.7 ppm, can be assigned to C=N carbon. All 1H- and 13C-NMR of the ligands are attached as the supplementary materials.

2.6. Thermal Analysis

The TGA curves (Figure 1 and Figure 2) indicate that the ligands 2a, 2b and 2c begin to decompose at 277, 260 and 268 °C, respectively. Comparison of the decomposition temperature of the ligands shows that 2a and 2c decompose at higher temperatures than 2b. The TGA curve for 2a displays three stages of mass loss within the temperature range of 220–925 °C. The first stage is at 220–320 °C, and exhibits a mass loss of 26.13%, corresponding to the loss of C4H6N2O (calc. 26.92%). The second stage occurs at 320–790 °C, with a mass loss of 21.11%, corresponding to the loss of C6H4 (calc. 20.08%). The third stage of decomposition occurs at the temperature range 790–925 °C, with a mass loss of 8.7%, corresponding to the loss of NO (calc. 8.2%).
Figure 1. Thermogravimetric curves of ligands 2ac.
Figure 1. Thermogravimetric curves of ligands 2ac.
Molecules 17 06434 g001
Figure 2. Thermogravimetric curves for copper(II) complexes 3ac.
Figure 2. Thermogravimetric curves for copper(II) complexes 3ac.
Molecules 17 06434 g002
Complex 2b shows similar degradation behavior and displays three stages of mass loss within the temperature range of 230–953 °C. The first stage is at 230–300 °C, with a mass loss of 24.28%, corresponding to the loss of C5H10N2O (calc. 25.19%). The second stage occurs at 300–780 °C, with a mass loss of 21.07%, corresponding to the loss of C6H4O (calc. 20.33%). The third stage of decomposition occurs at the temperature range 780–953 °C, with a mass loss of 27.32%, corresponding to the loss of C6H3NO2 (calc. 26.74%). The TG curve for 2c exhibits to two stages of mass loss within the temperature range of 243–953 °C. The first stage occurs at 243–329 °C, with a mass loss of 26.09%, corresponding to the loss of C5H8NO (calc. 27.03%). The second stage of decomposition occurs at 329–953 °C and is roughly assigned to the loss of C6H5O with a mass loss 25.43% (calc. 25.66%). The TG curves indicate that the copper complexes 3a, 3b and 3c begin decomposition at 283, 299 and 308 °C, respectively. A comparison of the thermogravimetric curves for the ligands and the copper complexes shows that the copper complexes are more thermally stable than the azo dye ligands. The order of thermal stability is found to be 3c > 3b > 3a > 2a > 2c > 2b. The organic portion of the copper complexes may decompose in three or four steps. The decompositions of the copper complexes all ended with formation of CuO [50,51]. The thermogravimetric analysis data for the ligands and the copper(II) complexes are presented in Table 2. Metal complexes derived from salicylideneamine are among the best known classes of metallomesogens. The liquid crystalline behavior of all the cobalt(II) complexes was studied by thermal analysis (DSC). The phase transitions and thermodynamic data for the complexes were summarized in Table 3. Surprisingly, none of the cobalt(II) complexes displayed any mesomorphic properties. The lack of liquid crystallinity is generally believed to be attributed to weaker interaction between the cobalt complexes.
Table 2. Thermogravimetric analysis data for the ligands and copper(II) complexes.
Table 2. Thermogravimetric analysis data for the ligands and copper(II) complexes.
Compound Temperature range (°C)TG weight loss % calc./foundAssignments
2a220–32026.92 (26.13)C4H6N2O
320–79020.08 (21.11)C6H4
790–9258.2 (8.7)NO
2b230–30025.19 (24.28)C5H10N2O
300–78020.33 (21.07)C6H4O
780–95326.74 (27.32)C6H3NO2
2c243–32927.03 (26.09)C5H8NO
329–95325.66 (25.43)C6H5O
3a230–33531.96 (31.04)C12H8N4O4
335–65019.51 (19.33)C11H6N2
650–9519.3 (9.1)CuO
3b210–29512.14 (11.22)C6H3N2
295–71023.17 (22.78)C9H7NO2Cl
710–9539.3 (9.7)CuO
3c165–2305.15 (4.59)C3H7N
230–31030.86 (31.16)C12H6N4O4Cl2
310–75020.27 (19.95)C12H6N3O2
750–9537.19 (8.07)CuO
Table 3. Transition temperatures and enthalpy changes for the cobalt(II) complexes.
Table 3. Transition temperatures and enthalpy changes for the cobalt(II) complexes.
CompoundTransitionT/°CΔH/J g−1
4aCr-I (dec)293−147
Cr1-Cr21681.61
4bCr2-Cr31895.27
Cr3-I (dec)248−141
4cCr1-Cr21911.2
Cr2-I (dec)230−3.54
Cr = crystal, I = Isotropic liquid, dec = decomposition.

3. Experimental

3.1. Physical Measurements

C, H and N composition determinations were undertaken using an Elemental Analysis System GmbH Vario EL II. Cobalt and copper determinations were performed using a Perkin-Elmer 2380 Atomic Absorption Spectrophotometer. Electronic spectra of complexes were recorded on a Perkin-Elmer Lambda 900 spectrophotometer using the diffuse reflectance technique; MgO was used as a reference. Electronic spectra of ligands in DMSO were recorded on a Perkin-Elmer Lambda 15 instrument. FTIR spectra of the compounds as KBr disks were obtained in the 4,000–300 cm−1 range with a Galaxy series FTIR 5000 spectrophotometer. The spectra were calibrated using the polystyrene bands at 3,028, 1,601 and 1,208 cm−1. The room temperature magnetic moment of each metal complex was measured according to the Evans method [46]. 13C- and 1H-NMR spectroscopy was recorded using a Bruker 300 MHz spectrometer. Thermal analyses were performed using a Perkin-Elmer Thermogravimetric Analyzer TG/DTA 6300 under a N2 gas flow (20 mL min−1) at ambient pressure. A heating rate of 10 °C min−1 was chosen. In the cases where the TG curve indicated the possibility of stable intermediates, a heating rate of 5 °C min−1 or 1 °C min−1 was applied.

3.2. Materials

Chemicals: All chemicals and solvents were of reagent grade quality and were purchased from Merck Chemical Company and used as received without further purification, except for vaccum dring over P2O5.

3.3. Synthesis of the Starting Materials

1-(3-Formyl-4-hydroxyphenylazo)-4-nitrobenzene (1a). Azo dye 1a was synthesized according to the well-known published procedure [52]. A suspension of 4-nitroaniline (5.52 g, 40 mmol) in hydrochloric acid (36 mL) and water (16 mL) was heated to 70 °C until complete dissolution. The clear solution was poured into ice water and was diazotized below 5 °C with sodium nitrite (2.8 g, 40 mmol) dissolved in water (10 mL). The cold diazonium solution was added over the course of 30 min at 0 °C to a solution of salicylaldehyde (4.26 mL, 40 mmol) in water (75 mL) containing sodium hydroxide (1.6 g) and sodium carbonate (14.8 g). During the addition process, the solution was vigorously stirred. The product was collected by vacuum filtration and washed with NaCl solution (100 mL, 10%). Coupling of the diazonium reagent to the salicylaldehyde occurred at the position para to the hydroxyl group. The diazo compound was recrystallized several times from ethyl alcohol. C13H9N3O4, yellow solid, yield: 90%, m.p.: 185–186 °C. FTIR (KBr, cm−1): 3,101 (-OH group), 1,664 (-CHO group), 1,479 (N=N), 1,346 (NO2 group), 1,286 (C-O) cm−1. UV-Vis: λmax = 397, 547 nm. 1H-NMR (DMSO): δ = 11.49 (1H, s), 10.08 (1H, s), 8.4 (2H, t, J = 7.14 Hz), 8.3 (1H, d, J = 2.34 Hz), 8.22 (1H, q, J = 2.37 Hz), 8.01 (2H, t, J = 7.08 Hz), 7.1 (1H, d, J = 9 Hz) ppm.
1-(3-Formyl-4-hydroxyphenylazo)-2-chloro-4-nitrobenzene (1b). Compound 1b was synthesized from 2-chloro-4-nitroaniline (7.9 g, 45 mmol) following the procedure given above for 1a, using the same molar ratio of the reagents. The purity of the compound was evaluated using thin layer chromatography. C13H8N3O4Cl, yellow solid, yield: 85%, m.p.: 127–128 °C. FTIR (KBr, cm−1): 3,431 (-OH group), 1,660 (-CHO group), 1,481 (N=N), 1,346 (NO2 group), 1,286 (C-O) cm−1. UV-Vis: λmax = 380, 566 nm. 1H-NMR (DMSO): δ = 10.35 (1H, s), 8.5 (1H, s), 7.1 (1H, d, J = 9.28 Hz), 7.8 (1H, d, J = 8.78 Hz), 8.1 (1H, d, J = 8.76 Hz), 8.23 (1H, s), 8.3 (2H, d, J = 8.9 Hz) ppm.
1-(3-Formyl-4-hydroxyphenylazo)-4-chloro-3-nitrobenzene (1c). Compound 1c was synthesized from 4-chloro-3-nitroaniline (7.9 g or 45 mmol) following the procedure given above for 1a, using the same molar ratio of the reagents. The purity of the compound was evaluated using thin layer chromatography. C13H8N3O4Cl, yellow solid, yield: 78%, m.p.: 243–244 °C. FTIR (KBr, cm−1): 3,428 (-OH group), 1,658 (-CHO group), 1,473 (N=N), 1,550 (NO2 group), 1,284 (C-O) cm−1. UV-Vis: λmax = 355, 490 nm. 1H-NMR (DMSO): δ = 11.8 (1H, s), 10.36 (1H, s), 8.4 (1H, s), 8.21 (1H, s), 8.1 (2H, d, J = 9.5 Hz), 7.9 (1H, d, J = 8.3 Hz), 7.2 (1H, d, J = 8.4 Hz) ppm.

3.4. Synthesis of the Schiff Base Ligands

Compounds 2ac were prepared using a method previously reported in the literature [53]. For each ligand, a mixture of 0.02 mol of 3-amino-1-propanol and 0.02 mol of the corresponding azo dye 1ac was dissolved in absolute ethanol (80 mL) with a few drops of glacial acetic acid as a catalyst. The resulting mixture was allowed to stir under reflux for 2–3 h. The product was vacuum-filtered and washed with a small amount of hot ethanol. The products were soluble in solvents such as DMF and DMSO.
1-{3-[(3-Hydroxypropylimino)methyl]-4-hydroxyphenylazo}-4-nitrobenzene (2a). Yellow solid, yield: 80%, m.p.: 168–170 °C. Anal. Calc. for C16H16N4O4. 2H2O, C: 52.7, H: 5.5, N: 15.3. Found: C: 52.8, H: 6.4, N: 14.8%. FTIR (KBr, cm−1): 3,100–3,252 (-OH group), 3,051 (C-H, aromatic), 2,922, 2,841 (C-H, aliphatic), 1,647 (C=N), 1,608 (C=C, aromatic), 1,421 (N=N), 1,273 (C-O, phenolic) cm−1. UV-Vis: λmax = 269 (9,296), 399 (12,933), 462 (10,766) nm. 1H-NMR (DMSO): δ = 13.21 (1H, s), 8.72 (1H, s), 8.36 (2H, s), 8.09 (1H, s), 7.92 (3H, s), 6.67 (1H, s), 4.73 (1H, s), 3.72(1H), 3.5 (1H), 1.83 (1H) ppm. 13C-{1H}NMR (DMSO) δ = 32.9, 50.5, 58.3, 115.6, 123, 123.5, 125.4, 126.6, 136.4, 141.8, 147.6, 156.3, 166.9, 176.5 ppm.
1-{3-[(3-Hydroxypropylimino)methyl]-4-hydroxyphenylazo}-2-chloro-4-nitrobenzene (2b). Light brown solid, yield: 71%, m.p.: 187–190 °C. Anal. Calc. for C16H15N4O4Cl. 5H2O, C: 42.4, H: 5.4, N: 12.3. Found: C: 42, H: 5.2, N: 11.83%. FTIR (KBr, cm−1): 3,100–3,271 (-OH group), 3,057 (C-H, aromatic), 2,958, 2,897 (C-H, aliphatic), 1,649 (C=N), 1,610 (C=C, aromatic), 1,450 (N=N), 1,271 (C-O, phenolic) cm−1. UV-Vis: λmax = 268 (825), 412 (12,030), 530 (16,645) nm. 1H-NMR (DMSO): δ = 13.83 (1H, s); 8.76 (1H, s); 8.4 (1H, s), 8.26 (1H, s), 8.12 (1H, s), 7.94 (1H, s), 7.78 (1H, s), 6.67 (1H, s), 4.79 (1H, s), 3.7 (1H), 3.5 (1H), 1.83 (1H) ppm. 13C-{1H}NMR (DMSO) δ = 32.8, 49.8, 58.2, 115.3, 118.6, 123.8, 124.5, 126.1, 126.7, 132.8, 138.8, 141.9, 147.4, 152.6, 167.2, 178.7 ppm.
1-{3-[(3-Hydroxypropylimino)methyl]-4-hydroxyphenylazo}-4-chloro-3-nitrobenzene (2c). Light brown solid, yield: 46%, m.p.: 121–125 °C. Anal. Calc. for C16H15N4O4Cl. C: 52.9, H: 4.1, N: 15.4. Found: C: 52.4, H: 4.4, N: 15.1%. FTIR (KBr, cm−1): 3,300 (-OH group), 3,090 (C-H, aromatic), 2,922, 2,877 (C-H, aliphatic), 1,651 (C=N), 1,612 (C=C, aromatic), 1,462 (N=N), 1,275 (C-O, phenolic) cm−1. UV-Vis: λmax = 264 (15,463), 417 (19,260), 465 (18,703) nm. 1H-NMR (DMSO): δ = 14 (1H, s), 8.3 (1H, s), 8.1 (1H, d), 8 (2H, t), 7.9 (2H, d), 6.71 (1H, d), 4.7 (1H, s), 3.7 (1H), 3.5 (1H), 1.8 (1H) ppm. 13C-{1H}NMR (DMSO) δ = 33, 50.7, 58.3, 115.7, 118, 123.2, 125.5, 126.5, 127.4, 133, 135.8, 141.4, 148.6, 151.6, 166.7, 176 ppm.

3.5. Synthesis of the Copper(II) Complexes

The copper complexes were prepared following a previously published procedure [54]. For each complex, an alcoholic solution containing 10 mmol of 3-amino-1-propanol was added to a solution of the corresponding azo dye 1a–c (10 mmol) in boiling methanol (50 mL). Then, sodium acetate (10 mmol) dissolved in water (10 mL) was added to the reaction. Finally, boiling methanol solution (50 mL) containing Cu(CH3COO)2•H2O (10 mmol) was added. The stirring mixture was refluxed for 2 h and then the product was vacuum-filtered and washed with a small amount of ethanol.
Complex [(1-{3-[(3-Hydroxypropylimino)methyl]-4-hydroxyphenylazo}-4-nitrobenzenato)2(H2O)4Cu2] (3a). Dark orange solid, yield: 73%, Anal. Calc. for C32H28N8O8Cu2. 4 H2O, C: 45.1, H: 4.2, N: 13.15, Cu: 14.91. Found: C: 45.6, H: 4.3, N: 13.13, Cu: 14.7%. FTIR (KBr, cm−1): 1,626 (C=N), 1,473 (N=N), 1,334 (C-O, phenolic), 422 (Cu-N), 599 (Cu-O) cm−1. UV-Vis: λmax = 278, 360 and 457 nm.
Complex [(1-{3-[(3-Hydroxypropylimino)methyl]-4-hydroxyphenylazo}-2-chloro-4-nitrobenzenato)2Cu2] (3b). Brownish red solid, yield: 65%, Anal. Calc. for C32H26N8O8Cu2Cl2, C: 45.3, H: 3.0, N: 13.2, Cu: 14.97. Found: C: 45.16, H: 1.991, N: 12.52, Cu: 15.4%. FTIR (KBr, cm−1): 1,624 (C=N), 1,475 (N=N), 1,334 (C-O, phenolic), 426 (Cu-N), 563 (Cu-O) cm−1. UV-Vis: λmax = 271, 472 and 620 nm.
Complex [(1-{3-[(3-Hydroxypropylimino)methyl]-4-hydroxyphenylazo}-4-chloro-3-nitrobenzenato)2(CHCl3)2(H2O)2Cu2] (3c). Light brown solid, yield: 42%, Anal. Calc. for C32H26N8O8Cu2Cl2. 2CHCl3•H2O, C: 36.9, H: 2.7, N: 10.1, Cu: 11.48. Found: C: 36.4, H: 2.4, N: 10.02, Cu: 12.4%. FTIR (KBr, cm−1): 1,620 (C=N), 1,471 (N=N), 1,336 (C-O, phenolic), 443 (Cu-N), 524 (Cu-O) cm−1. UV-Vis: λmax = 265, 441 and 546 nm.

3.6. Synthesis of the Cobalt(II) Complexes

The cobalt complexes were prepared according to the procedure described above for the copper complexes using Co(CH3COO)2•H2O instead of Cu(CH3COO)2•H2O.
Complex [(1-{3-[(3-Hydroxypropylimino)methyl]-4-hydroxyphenylazo}-4-nitrobenzenato)2(C2H5OH)2Co2] (4a). Brownish red solid, yield: 49%, Anal. Calc. for C32H28N8O8Co2. 2CH3CH2OH, C: 50.13, H: 4.63, N: 12.98, Co: 13.68. Found: C: 50.30, H: 5.83, N: 13.87, Co: 14.34%. FTIR (KBr, cm−1): 1,635 (C=N), 1,479 (N=N), 1,336 (C-O, phenolic), 408 (Co-N), 576 (Co-O) cm−1. UV-Vis: λmax = 256, 278, 489 and 605 nm.
Complex [(1-{3-[(3-Hydroxypropylimino)methyl]-4-hydroxyphenylazo}-2-chloro-4-nitrobenzenato)2(C2H5OH)Co2] (4b). Brown solid, yield: 65%, Anal. Calc. for C32H26N8O8Co2Cl2. CH3CH2OH, C: 46.12, H: 3.61, N: 12.65 Co: 13.32. Found: C: 46.93, H: 3.298, N: 13.17, Co: 12.72%. FTIR (KBr, cm−1): 1,641 (C=N), 1,479 (N=N), 1,344 (C-O, phenolic), 414 (Co-N), 584 (Co-O) cm−1. UV-Vis: λmax = 257, 270, 508 and 622 nm.
Complex [(1-{3-[(3-Hydroxypropylimino)methyl]-4-hydroxyphenylazo}-4-chloro-3-nitrobenzenato)2(CHCl3)2(H2O)2Cu2] (4c). Yellow solid, yield: 34%, Anal. Calc. for C32H26N8O8Co2Cl2. 2CHCl3•H2O, C: 37.24, H: 2.73, N: 10.21, Co: 10.76. Found: C: 37.3, H: 3.071, N: 10.14, Co: 11.2%. FTIR (KBr, cm−1): 1,602 (C=N), 1,475 (N=N), 1,330 (C-O, phenolic), 443 (Co-N), 592 (Co-O) cm−1. UV-Vis: λmax = 259, 285, 438 and 528 nm.
The various synthetic reactions are summarized below (Scheme 1).
Scheme 1. Preparation of ligands and complexes.
Scheme 1. Preparation of ligands and complexes.
Molecules 17 06434 g003

3.7. Biological Studies

The antimicrobial activities of the compounds in this study were tested by the disc diffusion method [55]. Mueller Hinton Agar (MHA, Oxoid, 15 mL) and Sabouraud Dextrose Agar (SDA, sterilized in a flask and cooled to 45–50 °C) were homogenously distributed onto sterilized Petri dishes. The microorganisms were individually introduced on the surface of the agar plates. Blank sterile discs, measuring 6.4 mm in diameter, were soaked in a known concentration of the test compounds and then implanted on the surface of the agar plates. A blank disc was soaked in the solvent (DMSO) and implanted as a negative control on each plate along with the standard drugs. The plates were incubated at 37 °C (24 h) and 27 °C (48 h) for bacterial and fungal strains, respectively.

4. Conclusions

Three Schiff base ligands containing the azo group, 1-{3-[(3-hydroxypropylimino) methyl]-4-hydroxyphenylazo}-4-nitrobenzene, 1-{3-[(3-hydroxypropylimino)methyl]-4-hydroxyphenylazo}-2-chloro-4-nitrobenzene and 1-{3-[(3-hydroxypropylimino)methyl]-4-hydroxyphenylazo}-4-chloro-3-nitrobenzene were synthesized. The ligands and their complexes with copper(II) and cobalt(II) ions were characterized by elemental analysis, FTIR spectroscopy, UV-Vis spectroscopy, 1H-NMR spectroscopy and thermogravimetric analysis. The room temperature effective magnetic moments of complexes are 1.45, 1.56, 1.62, 2.16, 2.26 and 2.80 B.M. for comlexes 3a, 3b, 3c, 4a 4b, and 4c, respectively. These values indicate that the dimetallic complexes are paramagnetic with considerable electronic communication between the two metal centers. The cobalt(II) and copper(II) complexes are coordinated by the N azomethine and O phenoxyl atoms of the ligands. In addition, the Schiff base ligands and their metal complexes were evaluated for their in vitro antibacterial activity using the disc diffusion method.

Acknowledgments

The authors would like to thank the Research Council of Arak University for financial support of this research.

References and Notes

  1. Wang, Y.; Liu, Y.; Luo, J.; Qi, H.; Li, X.; Nin, M.; Liu, M.; Shi, D.; Zhu, W.; Cao, Y. Metallomesogens based on platinum(II) complexes: Synthesis, luminescence and polarized emission. Dalton Trans. 2011, 40, 5046–5051. [Google Scholar]
  2. Emeleus, L.C.; Cupertino, D.C.; Harris, S.G.; Owens, S.; Parsons, S.; Swart, R.W.; Tasker, P.A.; White, D.J. Diazopyrazolones as weak solvent extractants for copper from ammonia leach solutions. Dalton Trans. 2001, 2001, 1239–1245. [Google Scholar]
  3. Oforka, N.C.; Mkpenie, V.N. A new method of synthesis of azo Schiff base ligands with azo and azomethine donors: Synthesis of N-4-methoxy-benzylidene-2-(3-hydroxyphenylazo)-5-hydroxy-aniline and its nickel(II) complex. Chin. J. Chem. 2007, 25, 869–871. [Google Scholar] [CrossRef]
  4. Hudson, S.A.; Maitlis, P.M. Calamitic metallomesogens: Metal-containing liquid crystals with rodlike shapes. Chem. Rev. 1993, 93, 861–885. [Google Scholar] [CrossRef]
  5. Halve, A.; Goyal, A. Synthesis and crystal structure of 2-(2,3,4-trimethoxy-6-methylbenzylideneamino) phenol. Orient. J. Chem. 2001, 12, 87–88. [Google Scholar]
  6. Pandeya, S.N.; Sriram, D.; Nath, G.; de Clercq, E. Synthesis and pharmacological evaluations of some novel isatin derivatives for antimicrobial activity. Il Farmaco 1999, 54, 624–628. [Google Scholar] [CrossRef]
  7. Lashanizadegan, M.; Jamshidbeigi, M. Synthesis of N,N'-bis[4-(benzeneazo)salicylaldehyde]4-methyl-1,2-phenylenediamine and its transition metal complexes. Synth. React. Inorg. Metal-Org. Nano-Metal Chem. 2012, 42, 507–512. [Google Scholar]
  8. Hodnett, E.M.; Dunn, W.J. Synthesis of novel azo Schiff base bis [5-(4-methoxyphenylazo)-2-hydroxy-3-methoxy benzaldehyde]-1,2-phenylene diimine. J. Med. Chem. 1970, 13, 768–770. [Google Scholar] [CrossRef]
  9. Desai, S.B.; Desai, P.B.; Desai, K.R. Synthesis and spectroscopic studies of new Schiff bases. Hetrocycl. Commun. 2001, 7, 83–90. [Google Scholar] [CrossRef]
  10. Mitu, L.; Ilis, M.; Raman, N.; Imran, M.; Ravichandran, S. Transition metal complexes of isonicotinoyl-hydrazone-4-diphenylaminobenzaldehyde: Synthesis, characterization and antimicrobial studies. E J. Chem. 2012, 9, 365–372. [Google Scholar] [CrossRef]
  11. Samadhiya, S.; Halve, A. 1-(4-{[(E)-(4-Diethylamino-2-hydroxyphenyl)methylene]amino}phenyl)ethanone. Orient. J. Chem. 2001, 17, 119–122. [Google Scholar]
  12. Karakuş, O.O.; Deligöz, H. Synthesis and characterization of three novel azocalix[4]arene Schiff base derivatives and their selective copper extraction. J. Iran. Chem. Soc. 2012, 9, 93–100. [Google Scholar] [CrossRef]
  13. Khandar, A.A.; Rezvani, Z. Preparation and thermal properties of the bis [5-((4-heptyloxyphenyl)azo)-N-(4-alkoxyphenyl)-salicylaldiminato]copper(II) complex homologues. Polyhedron 1999, 18, 129–133. [Google Scholar] [CrossRef]
  14. Reddinger, J.L.; Reynolds, J.R. tunable redox and optical properties using transition metal-complexed polythiophenes. Macromolecules 1997, 30, 673–675. [Google Scholar] [CrossRef]
  15. Angelino, M.D.; Laibinis, P.E. Synthesis and characterization of a polymer-supported salen ligand for enantioselective epoxidation. Macromolecules 1998, 31, 7581–7587. [Google Scholar] [CrossRef]
  16. Rezvani, Z.; Ahar, L.; Nejati, K.; Seyedahmadian, S.M. Synthesis, characterization and liquid crystalline properties of new bis[5-((4-alkoxyphenylazo)-N-(alkyl)-salicylaldeminato]coppe(II) and Nickel(II) complexes homologues. Acta Chim. Slov. 2004, 51, 675–686. [Google Scholar]
  17. Cakir, S.; Bicer, E.; Odabasoghu, M.; Albayrak, C.J. A new method of synthesis of azo schiff base ligands with azo and azomethine donors: Synthesis of N-4-methoxy-benzylidene-2-(3-hydroxyphenylazo)-5-hydroxy-aniline and its Nickel(II) complex. Braz. Chem. Soc. 2005, 16, 711–715. [Google Scholar] [CrossRef]
  18. Bhattacharyya, P.; Parr, J.; Ross, A.T. First synthesis of a unique dilead Schiff base complex. Dalton Trans. 1998, 1998, 3149–3150. [Google Scholar]
  19. Nazir, H.; Sapmaz Akben, N.; Bürke Ateş, M.; Sözeri, H.; Ercan, I.; Atakol, O.; Ercan, F. Synthesis, crystal structure and magnetic behaviourof a mononuclear Fe(III)—Schiff base metal complex. Z. Kristallogr. 2006, 221, 276–289. [Google Scholar] [CrossRef]
  20. Das, N.N.; Dash, A.C. Synthesis, characterization and electrochemistry of a binuclear copper(II) complex of schiff base derived from 2-aminomethyl-benzimidazole and salicylaldehyde. Polyhedron 1995, 14, 1221–1227. [Google Scholar] [CrossRef]
  21. Khandar, A.; Nejati, K. Synthesis and characterization of cyano-bridged polynuclear [Cu(dmpn)2]3[Co(CN)6]2•12H2O and trinuclear [Cu(dmpn)2]2[Co(CN)6]ClO4•3H2O complexes. Polyhedron 2000, 19, 607–613. [Google Scholar] [CrossRef]
  22. Khandar, A.A.; Rezvani, Z.; Nejati, K.; Yanovsky, I.; Martines, J.M. Synthesis, characterization and liquid crystalline properties of new bis[5-((4-alkoxyphenylazo)-N-(alkyl)-salicylaldeminato] Nickel(II) complexes. Acta Chim. Slov. 2002, 49, 733–741. [Google Scholar]
  23. Nejati, L.; Rezvani, Z. Syntheses, characterization and mesomorphic properties of new bis(alkoxyphenylazo)-substituted N,N′ salicylidene diiminato Ni(II), Cu(II) and VO(IV) complexes. New J. Chem. 2003, 27, 1665–1669. [Google Scholar] [CrossRef]
  24. Khandar, A.A.; Nejati, K.; Rezvani, Z. Syntheses, characterization and study of the use of Cobalt (II) schiff-base complexes as catalysts for the oxidation of styrene by molecular oxygen. Molecules 2005, 10, 302–311. [Google Scholar] [CrossRef]
  25. Nejati, K.; Rezvani, Z.; Massoumi, B. The synthesis, characterization, thermal and optical properties of copper, nickel, and vanadyl complexes derived from azo dyes. Dyes Pigm. 2007, 75, 653–657. [Google Scholar] [CrossRef]
  26. Arjmand, F.; Muddassir, M. Design and synthesis of heterobimetallic topoisomerase I and II inhibitor complexes: In vitro DNA binding, interaction with 5'-GMP and 5'-TMP and cleavage studies. J. Photochem. Photobiol. B Biol. 2010, 101, 37–46. [Google Scholar] [CrossRef]
  27. Arjmand, F.; Mohani, B.; Ahmad, S. Synthesis, antibacterial, antifungal activity and interaction of CT-DNA with a new benzimidazole derived Cu(II) complex. Eur. J. Med. Chem. 2005, 40, 1103–1110. [Google Scholar] [CrossRef]
  28. Naeimi, H.; Safari, J.; Heidarnezhad, A. Synthesis of Schiff base ligands derived from condensation of salicylaldehyde derivatives and synthetic diamine. Dyes Pigm. 2007, 73, 251–253. [Google Scholar] [CrossRef]
  29. Lippard, S.J.; Berg, J.M. Principles of Bioinorganic Chemistry; University Science Books: Mill Valley, CA, USA, 1994; pp. 1–160. [Google Scholar]
  30. Tumer, M.; Ekinci, D.; Tumer, F.; Bulut, A. Synthesis, characterization and properties of some divalent metal(II) complexes: Their electrochemical, catalytic, thermal and antimicrobial activity studies. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2007, 67, 916–929. [Google Scholar] [CrossRef]
  31. Allan, J.R.; Milburn, G.H.W.; Richmond, F.; Wilson, A.S.; Gerrared, D.L.; Bimie, J. Thermal, spectral and magnetic studies on the chloro compounds of some first-row transition metals with triethylenediamine. Thermochim. Acta 1990, 170, 147–154. [Google Scholar] [CrossRef]
  32. Thornton, D.A.; Verhoeven, P.F.A. Thermogravimetric analysis of cobalt(II) halide complexes with a series of substituted pyridine ligands. Thermochim. Acta 1987, 113, 161–169. [Google Scholar] [CrossRef]
  33. Liu, S.; Jiang, P.; Song, G.; Liu, R.; Zhu, H. Synthesis and optical properties of a series of thermally stable diphenylanthrazolines. Dyes Pigm. 2009, 81, 218–223. [Google Scholar] [CrossRef]
  34. Erdem, E.; Sari, E.Y.; Kilinçarslan, R.; Kabay, N. Synthesis and characterization of azo-linked Schiff bases and their nickel(II), copper(II), and zinc(II) complexes. Transit. Metal Chem. 2009, 34, 167–174. [Google Scholar] [CrossRef]
  35. Canpolat, E.; Kaya, M. Studies on mononuclear chelates derived from substituted Schiff Base ligands: Synthesis and characterization of a new 5-methoxysalicyliden-p-aminoacetophenoneoxime and its complexes with Co(II), Ni(II), Cu(II), and Zn(II). Russ. J. Coord. Chem. 2005, 31, 790–794. [Google Scholar] [CrossRef]
  36. Shishkin, V.N.; Kudrik, E.V.; Shaposhnikov, G.P. Synthesis and some properties of transition metal complexes with octa(sulfophenyl)tetrapyrazinoporphyrazine. Russ. J. Coord. Chem. 2005, 31, 516–520. [Google Scholar] [CrossRef]
  37. Kurzak, K.; Kuniarska-Biernacka, I.; Freire, C. Solution properties and solvatochromism of bis(N-2-pyridyl-salicylaldiminato)cobalt(II). Struct. Chem. 2010, 21, 377–383. [Google Scholar] [CrossRef]
  38. Lever, A.B.P. Inorganic Electronic Spectroscopy, 2nd ed; Elsevier: Amsterdam, The Netherlands, 1982; pp. 250–340. [Google Scholar]
  39. Conarmond, J.; plunere, P.; Lehn, J.M.; Agnus, Y.; Louis, R.; Kahn, O.; Badarous, M. Dinuclear copper (II) cryptates of macrocyclic ligands: Synthesis, crystal structure, and magnetic properties. Mechanism of the exchange interaction through bridging azido ligands. J. Am. Chem. Soc. 1982, 104, 6330–6340. [Google Scholar]
  40. Nassirinia, N.; Sadeghi, N.; Amani, S. Two new alkoxo-bridged dinuclear cooper(II) complexes with 3-aminopyrazidine-2-carboxylic acidas ligand. J. Coord. Chem. 2008, 61, 796–801. [Google Scholar] [CrossRef]
  41. Rodríguez, A.; Sousa-Pedrares, A.; García-Vázquez, J.A.; Romero Antonio Sousa, J. Electrochemical synthesis and characterization of zinc(II) complexes with pyrimidine-2-thionato ligands and their adducts with N,N donors. Polyhedron 2009, 28, 2240–2248. [Google Scholar] [CrossRef]
  42. Amani, S.; van Albada, G.A.; Reedijk, J. Synthesis, spectroscopic and magnetic properties of methoxo-bridged copper(II) complexes with with 2-amino-4-methylpyridine as the ligand. Transit. Metal Chem. 1999, 24, 104–107. [Google Scholar] [CrossRef]
  43. Sino, C.; Holt, S.A. Polarized spectrum of dichlorobis (triphenylphosphine)cobalt (II). Inorg. Chem. 1968, 7, 2655–2657. [Google Scholar] [CrossRef]
  44. Kurzak, K.; Kuniarska-Biernacka, I.; Kurzak, B. Spectrochemical properties of cobalt (II) complex with bidentate schiff base in various solvents. J. Solution Chem. 1999, 28, 133–151. [Google Scholar] [CrossRef]
  45. Kurzak, K.; Kuniarska-Biernacka, I.; Bartecki, B.; Kurzak, K. UV-Vis-NIR spectroscopy and colour of bis(N-phenylsalicylaldiminato)cobalt(II) in a variety of solvents. Polyhedron 2003, 22, 997–1007. [Google Scholar] [CrossRef]
  46. Evans, D.F. The determination of paramagnetic susceptibility of substances in solution by nuclear magnetic resonance. J. Chem. Soc. 1959, 1959, 2003–2005. [Google Scholar] [CrossRef]
  47. Arab Ahmadi, R.; Safari, N.; Khavasi, H.R.; Amani, S. Four new Co(II) complexes with 2-amino-4-methyylpyridine, 2-amino-3-methylpyridine or 2-amino-5-chloropyridine: Synthesis, spectroscopy, magnetic properties and crystal structure. J. Coord. Chem. 2011, 64, 2056–2065. [Google Scholar] [CrossRef]
  48. Hasanvand, F.; Hoseinzadeh, A.; Zolgharnein, J.; Amani, S. Synthesis and characterization of two acetato-bridged dinuclear copper(II) complexes with 4-bromo-2-((4 or 6-methylpyridine-2-ylimino)methyl)phenol as ligand. J. Coord. Chem. 2010, 63, 346–352. [Google Scholar] [CrossRef]
  49. Hasanvand, F.; Nasrollahi, N.; Vajed, A.; Amani, S. Synthesis, spectroscopy, and characterization of four alkoxo-bridged dinuclear copper(II) complexes containing 2-amino-4-methylpyridine or 2-amino-4-cyanopyridineas the ligands. Malays. J. Chem. 2010, 12, 27–32. [Google Scholar]
  50. Materazz, S.; Kurdzielb, K.; Tentolinia, U.; Bacalonia, A.; Aquilia, S. Thermal stability and decomposition mechanism of 1-allylimidazole coordination compounds: A TG-FTIR study of Co(II), Ni(II) and Cu(II) hexacoordinate complexes. Thermochim. Acta 2002, 395, 133–137. [Google Scholar] [CrossRef]
  51. Kenssey, G.; Werner, P.E.; Wadsten, T.; Liptay, G. Pyridine-type complexes of transition-metal halides X. Structural studies on the dimethylpyridine chloro-, bromo- and iodo-complexes of cobalt(II). Acta Chim. Scand. 1999, 33, 168–17. [Google Scholar]
  52. Botros, R. Azomethine dyes derived from an o-hydroxy aromatic aldehyde and a 2-aminopyridine. US Patent 4051119, 1 September 1977. [Google Scholar]
  53. Rezvani, Z.; Ghanea, M.A.; Nejati, K.; Abodollahzadeh Baghaei, A. Syntheses and mesomorphic properties of new oxygen-bridged dicopper complex homologous derived from azo-containing salicylaldimine Schiff base ligands. Polyhedron 2009, 28, 2913–2918. [Google Scholar]
  54. Ku, S.-M.; Wu, C.-Y.; Lai, C.K. Bimetallomesogens: Formation of calamitic or columnar mesophases by μ-exogenous-bridged groups in N,N'-(propan-2-ol)-bis(4-alkoxysalicylaldimine)copper(II) complexes. Dalton Trans. 2000, 2000, 3491–3492. [Google Scholar]
  55. Ispir, E. The synthesis, characterization, electrochemical character, catalytic and antimicrobial activity of novel, azo-containing Schiff bases and their metal complexes. Dyes Pigm. 2009, 82, 13–19. [Google Scholar] [CrossRef]
  • Sample Availability: Samples of compounds 1a-c, 2a-c, 3a-c, 4a-c are available from the authors.

Share and Cite

MDPI and ACS Style

Ahmadi, R.A.; Amani, S. Synthesis, Spectroscopy, Thermal Analysis, Magnetic Properties and Biological Activity Studies of Cu(II) and Co(II) Complexes with Schiff Base Dye Ligands. Molecules 2012, 17, 6434-6448. https://doi.org/10.3390/molecules17066434

AMA Style

Ahmadi RA, Amani S. Synthesis, Spectroscopy, Thermal Analysis, Magnetic Properties and Biological Activity Studies of Cu(II) and Co(II) Complexes with Schiff Base Dye Ligands. Molecules. 2012; 17(6):6434-6448. https://doi.org/10.3390/molecules17066434

Chicago/Turabian Style

Ahmadi, Raziyeh Arab, and Saeid Amani. 2012. "Synthesis, Spectroscopy, Thermal Analysis, Magnetic Properties and Biological Activity Studies of Cu(II) and Co(II) Complexes with Schiff Base Dye Ligands" Molecules 17, no. 6: 6434-6448. https://doi.org/10.3390/molecules17066434

Article Metrics

Back to TopTop