Translate this page into:
Copper based azo dye catalysts for phenoxazinone synthase mimicking efficiency: Structure characterization and bioactivity evaluation
⁎Corresponding authors at: Chemistry Department, Faculty of Applied Science, Umm Al-Qura University, Makkah, Saudi Arabia. haelghamry@uqu.edu.sa (Hoda A. El-Ghamry), s44286910@st.uqu.edu.sa (Azah A. Alkurbi)
-
Received: ,
Accepted: ,
This article was originally published by Elsevier and was migrated to Scientific Scholar after the change of Publisher.
Peer review under responsibility of King Saud University.
Abstract
The ligands, 4-(2-Hydroxy-naphthalen-1-ylazo)-N-thiazol-2-yl-benzenesulfonamide (H2TNBS) and N-(3,4-Dimethyl-isoxazol-5-yl)-4-(2-hydroxy-naphthalen-1-ylazo)-benzenesulfonamide (H2INBS), synthesized in the current investigation have been characterized and used for synthesizing divalent copper complexes by their reaction with a number of Cu(II) salts. Spectral and analytical methods have been applied for structures’ investigation. Morphology of the synthesized compounds have been investigated using TEM technique which joint with the results of X-ray powder diffraction spectroscopy confirmed the precipitation of both ligands and their complexes in the nanometric scale. Formation of the synthesized complexes in 1:1 or 2:1 (M:L) ratio was asserted by analytical results. Ultraviolet–visible spectra and magnetic moment were used to demonstrate the geometry around the Cu centers to be 4 coordinated square planar. The compounds under interest have been screened against selected microorganisms including Gram-positive and Gram-negative bacteria, unicellular and multicellular fungus showing, in most compounds, enhancement of activity upon chelation. The cell lines A-549 (human lung cancer cell line) and Panc-1 (human pancreatic cancer cell line) have been chosen to check the antitumor efficiency of the synthesized compounds; Vinblastine was used as standard. Finally, the Cu(II) chelates were investigated toward mimicking the protein phenoxazinone synthase using o- aminophenol (OAP) as substrate and DMF is the solvent. The results presented extremely high activity for the chloro complex 4 and nitro complex 6 with TOF numbers from 390.48 and 467.01 h−1, respectively. The least activity afforded by the acetato complexes 2 and 5.
Keywords
Copper chelates
Different anions
Sulfonamides
Biological efficiency
Catalytic activity
1 Introduction
Since the discovery of the antibacterial efficiency of the first obtained sulfonamide compound by Gerhard Domagek (Bentley, 2009), non-countable number of synthetic drugs based on the class of compounds known as sulfa drugs have been utilized (Gawrońska et al., 2022). In addition to the well-recognized antibacterial efficacy (Kamal et al., 2007) of such class of compounds, they also reported to have characteristics as anticancer (Scozzafava et al., 2003), diuretic (Supuran, 2008), hypoglycemic (Kumasaka et al., 2003), anti-inflammatory (Weber et al., 2004), and carbonic anhydrase inhibitory (Gokcen et al., 2016) characteristics. They are also capable of inhibiting the metabolic in folic acid and then limit the cell proliferation (Feng et al., 2022; Mizdal et al., 2018) as analogous to p-aminobenzoic acid and so sulfonamides are used to treat nocardiosis and acute urinary tract contagions trigger by a number of microorganisms such as Escherichia coli, and Proteus mirabilis. The biological significance of sulfonamides is largely concerned with their 4‐aminophenyl sulfonylamide core structure, that can endure alternative substituting groups at sulfonamide N, sulfanilamide’s hydrogen bond or as an aromatic heterocyclic ring in another sulfa drugs (Gaffer, 2019). Among the various sulfonamide derivatives, sulfisoxazole and sulfathiazole were established. Sulfanilamide is the source used for the syntheses of sulfisoxazole; which is a well-known antibacterial drug (Uhlemann et al., 2018; Gladys et al., 2003). Sulfisoxazole, with a weak acidic character, has been applied to treat acute otitis, chronic otitis prevention, and acute urinary tract infections (Althagafi et al., 2019).
Additionally, alteration of the sulfonamides’ structures can lead to motivation and expansion of their biological characteristics (Wan et al., 2021; Guo et al., 2019; Xu et al., 2020). Among the ways which are used to improve the sulfonamides’ structures is their implementation for the synthesis of azo dyes.
For a variety of biological uses, azo dyestuffs have attracted great attention. They were categorized as xenobiotic substances with biodegradation resistant character. They were employed, among other things, as biological stains, textile and food colorants, and acid/base indicators. In the treatment with Prontosil, the first sulphonamide‐based azo drug, that was used in defense against streptococcal infections in mice, the azo bond has been exploited in pharmaceutics to shield medications from unfavorable responses (Khattab et al., 2016; Islam et al., 2010). Heterocyclic azo dyes have given a share to the evolution of the branch of coordination chemistry. Different transition metal elements in low oxidation states are normalized using heterocyclic azo dye derivatives. These substances stand out for their biological effectiveness, analytical uses and catalytic performance (Islam et al., 2010; Saad et al., 2019; Khan et al., 2022). Azo dyes are frequently utilized in many industries, including those of textiles, leather, plastics, paper, liquid crystal displays in addition to laser & inkjet printers (Benkhaya et al., 2020). Due to their value as models for biological systems, it has been discovered that sulfa containing medicines incorporating azo groups are of significant interest in the medical profession (Saad et al., 2017; Mohamed and Mabrouk, 2006; Awad et al., 2010). More interestingly, and due to the accessibility to a large number of potential donor atoms, the fascinating structural variety, the flexibility, and the tendency to ligation as neutral or in patterned form, sulfonamide and its azo analogues offer a wide range of uses as potential bonds with numerous metal ions (Islam et al., 2010; Saad et al., 2019; Khan et al., 2022).
What is appearing to be more interesting is the ability of sulfonamides and their azo derivatives to form complexes with numerous numbers of metal ions with promising biological activity. Thus, these derivatives can be used to design and create novel drugs which exhibit, in most cases, enhanced activity when compared with the unbound compounds (Saad et al., 2019; Saad et al., 2017; Samper et al., 2017). So, the principles of selecting the metal ion as well as the coordination ligands are of a great importance.
Copper, as element number 3 to being the most common transition metal that exists in living system, is a crucial whit metal found in both plants and animals and enables the biological iron uptake. It is a part of many enzymes’ structures (Lee et al., 2021). In addition to its relevance as a trace metal, copper plays a vital role in the biochemistry of humans as component in many externally administered substances while it can connect to albumin and alternative proteins in addition to its ability to bind to a number of ligands, where these complexes can interact with biomolecules, notably proteins and nucleic acids (Iakovidis et al., 2011). Copper-containing metalloenzymes in biological systems perform as dioxygen and substrate activators for a number of bio-functions, including the effectiveness of phenoxazinone synthase which is a naturally occurring enzyme found in the bacteria Streptomyces antibiotics and catalyzes the oxidative coupling of o-aminophenol (OAP) derivatives into phenoxazinone chromophore (Phz). In practice, the latter is used to treat tumor types such as Wilm's tumor and other types of tumers by inhibiting the synthesis of RNA from DNA through DNA intercalation (Sarkar et al., 2017). Therefore, the phenoxazinone synthase is one of the enzymes whose structure is frequently imitated. The fundamental prerequisite for a metal complex to demonstrate prospective phenoxazinone synthase-like efficacy is the existence of labile sites on the metal centers.
As a result of all the previous finding, the current study primarily focuses on preparation and structure investigation two newly designed azo ligands and six of their copper complexes by the reaction of these azo ligands which are named 4-(2-Hydroxy-naphthalen-1-ylazo)-N-thiazol-2-yl-benzenesulfonamide (H2TNBS) and N-(3,4-Dimethyl-isoxazol-5-yl)-4-(2-hydroxy-naphthalen-1-ylazo)-benzenesulfonamide (H2INBS) with three different copper salts. The structure and spectral characteristic of the produced compounds have been studied extensively. The produced complexes' antibacterial, anticancer, and phenoxazinone synthase-like activity towards oxidation of o-aminophenol (OAP) to 2-aminophenoxazine-3-one (APX) have been assigned and described.
2 Experimental
2.1 Chemicals and instruments
All of the reagents employed in this investigation, which are analytical grade and employed as they received, were provided by Sigma-Aldrich company. Details about the tools and methods used for structure assertion and implementation is provided in part S1(supplementary file).
2.2 Preparation of H2TNBS (Z) and H2INBS (G) ligands
The current azo ligands named 4-(2-hydroxy-naphthalen-1-ylazo)-N-thiazol-2-yl-benzenesulfonamide (H2TNBS) and N-(3,4-dimethyl-isoxazol-5-yl)-4-(2-hydroxy-naphthalen-1-ylazo)-benzenesulfonamide (H2INBS) have been prepared in a similar way to literature method (Khedr et al., 2019; Khedr et al., 2022). In this method, 2.55 gm (0.01 mol) of sulfathiazole and 2.67 gm (0.01 mol) of sulfafurazole were separately dissolved in 25 ml of distilled H2O which acidified by the addition of 2.5 ml (12 M, 0.03 mol) of concentrated HCl. These solutions were then kept at temperature below zero before the addition of sodium nitrite (0.69 g and 0.01 mol) solution. The process of adding of sodium nitrite solution should by done very lowly to enable the formation and stabilization of diazonium chloride. These solutions were added in the last step to aqueous solutions containing 1.44 g of 2-naphthol (0.01 mol) and 0.4 g (0.01 mol) of NaOH. The formed colored products were kept with stirring in ice bath for 2 h. The coloured products was then filtered off, rinsed by distilled water, and air-dried. Structures of the two ligands H2TNBS and H2INBS are given in Fig. 1.Structure of H2TNBS and H2INBS ligands.
2.3 Synthesis of copper (II) chelates
For synthesis of Cu complexes, 1 mmol of H2TNBS and H2INBS ligands were dissolved in 40 ml MeOH solvent followed by the addition of just drops of (C2H5)3N base. To this solution, 0.002 mol of the Cu salts; CuCl2·2H2O, Cu(CH3COO)2·H2O or Cu(NO3)2, solubilized in 20 ml of hot MeOH was slowly added. Precipitation took place once all the amount of the metal salt solution has been added. The obtained coloured precipitates were isolated from solutions by filtration, washed by hot methanol solution and left to dry using vacuum desiccator.
2.4 Antimicrobial evaluation
Antimicrobial tendency of the target compounds were evaluated contra the strains Staphylococcus aureus (S. aureus), Bacillus cereus (B. cereus), Escherichia coli (E. coli), Salmonella typhi (S. typhi), Aspergillus flavus (A. flavus), and Candida albicans (C. albicans). The applied procedures in detailed are shown as supporting data (part S2).
2.5 Antitumor evaluation
The compound under interest in the current study were also examined for their antitumor activity contra the human lung cancer cells (A-549) and Pancreatic carcinoma (Panc-1) following the method that reported in literature (Mosmann, 1983; Riyadh et al., 2015). Full description of the methods applied for such test are given as supporting data (Part S3).
2.6 Catalytic oxidation of o-aminophenol (OAP)
The synthesized complexes (1–6) have been applied to investigate their catalytic performance as phenoxazinone synthase using o-aminophenol as the substrate. the initial rate method was utilized to calculate the reaction rate (Oloyede et al., 2019) while using the reported value of 18300 M−1 cm−1 as the molar absorptivity of the oxidation product (i.e. 2-aminobenzoxazine-3-one) to calculate the initial rate (Gasque et al., 2018; Podder and Mandal, 2020). Details about the instrument, concentrations and work strategy have been depicted in section S4 in the supplementary file.
3 Results and discussion
3.1 Ligands’ structure
3.1.1 Mass spectra
Electron impact (EI) mass spectra of the two ligands H2TNBS (Z) and H2INBS (G) are given in Figs S1 & S2. As shown in the figures, the molecular ion peaks are apparent at m/z = 410.49 and 422.14 for H2TNBS and H2INBS, successively, that were found to be consistence with the theoretical molecular weights of the ligands’ suggested structures (calculated molecular weights are 410.47 and 422.46 for H2TNBS and H2INBS, respectively).
3.1.2 1H NMR
1H NMR spectra of H2TNBS and H2INBS assigned in deuterated DMSO solvent and in existence of tetramethylsilane (TMS) internal standard are illustrated in Fig. 2 and S3, respectively.
1H NMR spectrum of Structure of H2TNBS ligand.
The spectrum of H2TNBS shows singlet at 15.77 ppm that assigned to the phenolic OH proton. the two singlet signals at, 12.8 and 9.71 ppm are assigned to sulfonamide OH and NH protons, respectively, and hence supports the coexistence of the ligand H2TNBS in the two tautomeric structures I and II (Scheme 1) (Kanaani et al., 2016). The peaks located within the range 8.41–6.38 ppm were attributed to the aromatic protons whose integration assuring the existence of the two tautomeric structures I and II (Scheme 1).The two tautomeric structures (I and II) of the ligand H2TNBS.
For the ligand H2INBS, the singlet appearing at 9.9 ppm assigned to NH proton. The multiple signals appearing within the range 7.59–6.35 ppm have been integrated for the aromatic protons whereas the two singlet signals assigned to the 2 methyl protons appeared at 2.04 and 1.65 ppm (El-Ghamry et al., 2022).
3.2 Complexes structures
3.2.1 Stoichiometry and general composition
on the basis of elements content data and molar conductance values, the emperical formulae of the generated chelates and the two synthesized ligands are displayed in Table 1 together with information of formula weights, color, and melting temperatures. The values of elements’ percent, which also completely consistent with the data indicated in Table 1, validated the separated chelate structures. Accordingly, these Cu chelates were assured to be produced in a 2: 1 (metal: ligand) ratio for complexes 1–3 and 1:1 for complexes 4–6. From molar conductivity values (measured using 10-4 M in DMSO) which found to be inside the range 8.63–35.6 Ω-1cm2mol−1 range supported non-electrolytic complexes formation (Geary, 1971; Fawzy et al., 2020). It was also shown that the complexes maintained their stability despite prolonged subjection to air. Solubility tests demonstrated the ease solubility of all compounds in DMF and MDSO however, in other polar and non-polar solvents, they provided either poor solubility or total insolubility. The yields were found to be 75 %–80 %, Λm is the molar conductance (Ω−1 cm2 mol−1).
No.
Molecular formula
(Empirical formulae)
M.p. (oC)
(M. Wt.)
Colour
(Λm)
Microanalysis
Found (Calc.) %
%C
%H
%N
%M
H2TNBS
H2TNBS
(C19H14N4O3S2)235
(410.47)
Orange
(-)55.78 (55.60)
3.62
(3.44)13.71
(13.65)–
1
[Cu2(HTNBS)(H2O)Cl3]·0·.5H2O
(C19H16Cl3Cu2N4O4.5S2)Over 300 (669.94)
Reddish brown (9.19)
34.18
(34.06)2.66
(2.41)8.51
(8.36)18.78
(18.97)
2
[Cu2(HTNBS)(OAc)2(H2O)2]·0.5CH3OH
(C23.5H24Cu2N4O9.5S2)Over 300
(705.69)
Brown
(8.63)40.22
(40.00)3.67
(3.43)7.83
(7.94)18.32
(18.01)
3
[Cu2(HTNBS)(NO3)2]·H2O
(C19H14Cu2N6O10S2)Over 300 (677.57)
Deep brown
(20.00)33.74
(33.68)2.19
(2.08)12.61
(12.40)18.66
(18.76)
H2INBS
H2INBS
C21H18N4O4S166
(422.46)
Faint brown
(-)59.91
(59.70)4.47
(4.29)13.43
(13.26)–
4
[Cu(HINBS)(H2O)Cl]·0·.5H2O
(C21H20ClCuN4O5.5S)Over 300 (547.47)
Reddish black (12.60)
47.22
(46.07)3.84
(3.68)10.42
(10.23)11.85
(11.61)
5
[Cu(HINBS)(OAc)(H2O)]
(C23H22CuN4O7S)Over 300 (562.05)
Deep brown
(7.42)49.30
(49.15)3.84
(3.95)9.86
(9.97)11.54
(11.31)
6
[Cu(HINBS)(NO3)(H2O)]·H2O
(C21H21CuN5O9S)Over 300 (583.03)
Faint brown
(35.6)
43.21
(43.26)3.84
(3.63)12.24
(12.01)11.16
(10.90)
3.2.2 EI- mass spectra
Since mass spectrometry offers appropriate spectral analysis to establish the molecular weights and, in turn, the molecular formulae of the synthetic compounds, it has been used to evaluate the acquired Cu(II) complexes, 1–6. The mass spectra of complexes 1–6 are shown in Figs. 3, 4, and S4–S7. These figures show that the molecular ion peaks for complexes 1–6 are observed at m/z = 657.98, 689.25, 659.19, 538.12, 562.75, and 564.43, respectively. These molecular ion peaks are consistent with the calculated molecular weights of 660.93, 689.66, 659.56, 538.64, 562.05, and 565.02 (excluding lattice H2O molecules) for complexes 1–6, successively.Mass spectrum of complex 5.
Mass spectrum of complex 6.
3.2.3 Infrared spectra
The coordinated function groups in the ligand with the metal ions can be determined from the IR spectra of both of them upon comparison with each other, making FTIR spectral analysis an important tool for characterizing the structure of metal complexes. Distinctive function groups are shown in Table 2 along with their vibration wavenumbers.
Comp.
ν(OH)
ν(NH)
ν(C = N) ring
ν(N = N)
ν(SO2)
ν(C-O)
ν(M−O)
ν(M−N)
H2TNBS
3426
3146
1568
1499
1326, 1127
1208
–
–
1
3416
3114
1549
1461
1374, 1139
1201
558
453
2
3382
–
1543
1463
–
1195
561
449
3
3379
–
1547
1452
–
1180
557
450
H2INBS
3368
3235
1552
1452
1347, 1164
1207
–
–
4
3381
3191
1550
1442
1346, 1162
1185
577
481
5
3423
3205
1553
1474
1344, 1164
1185
575
426
6
3406
3217
1550
1474
1346, 1162
1182
527
461
This comparison inferred that the two spectral bands occurring at 1499 and 1208 cm−1 in the spectrum of H2TNBS (Z) and at 1452 and 1207 cm−1 in the spectrum of H2INBS (G) which assigned to the vibration wavenumber of N = N and C-O groups have been shifted to higher or lower values in the spectra of all complexes (1–6) by at least 7 cm−1 and hence ensuring their incorporation in connection to the copper ion centers in metal complexes (Khedr et al., 2019; Khedr et al., 2022). The azomethine groups’ stretching wavenumbers of thiazole and isoxazole rings appeared in the ligands’ H2TNBS (Z) and H2INBS (G) spectra at 1568 and 1552 cm−1, respectively. The first band has been shifted to lower position in the spectra of complexes 1–3 compared with H2TNBS (Z) ligand whereas the second band appeared in complexes’ 4–6 spectra in a position very close to its position in the spectrum of ligand H2INBS (G) assuring that the C = N shared in complex formation in complexes 1–3 but in contrast this band kept free in the spectra of complexes 4–6. Similarly, the great shift in the position of the of asymmetric and symmetrical vibrations of SO2 group in the spectra of complex 1 confirm the involvement of its oxygen atom in complex formation while in complexes 4–6, such group has been confirmed to remain free due to its appearance almost at the same wavenumber as in the ligand H2INBS spectra. Vanishing of SO2 bands in spectra of compounds 2 and 3 is explained by coordination of sulfonamide O to the Cu center in the deprotonated enolic form. The non-ligands bands appearing between 558 and 527 cm−1 and between 481 and 426 cm−1 have been assigned to stretching wavenumbers of Cu–O and Cu–N bonds, successively.
In addition, analysis of the IR spectra allows a precise assignment of the coordinated anion's method of binding to the metal center. The two non-ligand bands that are shown in the spectra of the acetate-containing complexes 2 and 5 have appeared at 1593 and 1341 cm−1 for complex 2 and at 1588 and 1351 cm−1 for complex 5 and attributed to the v(COO−)asy and v(COO−)sym, respectively, of CH3COO groups,. The difference between v(COO−)asy and v(COO−)sym has been calculated and equals 252 and 237 cm−1, matching with the Δν value reported for monodentate acetate groups (Olar et al., 2013).
For 3 and 6 (containing NO3 coordinating groups), the two bands occurring at 1427 & 1297 cm−1 in complex 3 spectrum and at 1402 & 1216 cm−1 for complex 6 with Δν differences of 130 and 186 cm−1, successively, assured the bi- and unidentate ligation fashion of nitrato groups in complex 3 and 6, successively (Takroni et al., 2020).
3.2.4 Magnetic moment
Cu(II) complexes 1–3 were discovered to have magnetic moment values that fell within the range of 2.38–2.56B.M. for the two Cu(II) centers (Table 3). Such values are quite close to the calculated one corresponding to bimetallic chelates with two Cu(II) centers that are not in interaction (i.e. 2.45B.M.) (Li and Jin, 2009).
Complex
Wavelength
(nm)
Geometry
µeff
(B.M.)
1
282, 315 374, 416, 493, 738
Square planar
2.41*
2
277,309 360, 395, 503, 733
Square planar
2.56*
3
277,311 343, 401, 503, 735
Square planar
2.38*
4
251, 310, 392, 498, 710
Square planar
1.81
5
369, 348, 390, 406, 495, 707
Square planar
1.77
6
258, 315,370 392, 495, 701
Square planar
1.79
For complexes 4–6, the magnetic moment values were measured to be 1.77, 1.79 and 1.81B.M., respectively, which are almost within the acceptable spin-allowed limit for a single electron (i.e. 1.73B.M.) in mononuclear Cu(II) chelates (Takroni et al., 2020).
3.2.5 Electronic spectra
The most important study for identifying geometry of the ligand atoms surrounding the core metal ion in metal chelates is the electronic absorption spectra (UV–Vis). Accordingly, the nujol-Mull method was used to measure the UV–Vis spectra of compounds 1–6 over the wavelength range of 200–800 nm.
In all complexes’ spectra, Figs. 5, S8 and Table 3, the bands appearing within the ranges 251–282, 209–348 and 360–392 nm with medium to high absorbance are characteristic for π → π* transitions of benzene ring, π → π* and n → π* transitions in C = N & N = N groups, respectively (Yarkandi et al., 2017). The broad bands apparent in 493–503 nm range assigned to LMCT (Cadranel et al., 2017).UV–Vis spectra of compounds 5–8 using nujol mull technique.
The low and broad intensity bands apparent in the spectra of six complexes in the range 707–738, has been assigned to 2B1g→2A1g transition familiar for square planar architectures (Uysal and Kurşunlu, 2011).
3.2.6 Thermogravimetric analysis (TGA)
The thermogravimetric (TGA) analytical tool is one of the useful instruments that is frequently used to gain understanding of the structure and stability of inorganic compounds. The full analysis of the TG thermograms of complexes 1–6 along with the derivative thermogravimetric data (DTG) the is shown in Figs. 6 & S9 and interpreted in Table 4.TG/DTG thermograms of complexes 1–3.
Complex
(Mol. Wt.)
Temp. range (°C)
Mass loss %
DTG (°C)
Assignment
Calcd
Found
1
[Cu2(HTNBS)(H2O)Cl3]·0·.5H2O
(669.94)
25–84
1.34
1.67
60
-Loss of 0.5 lattice H2O.
84–180
2.68
2.49
164
-Loss of 1 coordinated H2O.
180–483
37.66
37.65
345, 410
-Loss of 1.5Cl2 + C3H3N2OS2.
483–800
Not assigned
26.37
668
- Not assigned.
2
[Cu2(HTNBS)(OAc)2(H2O)2]
·0.5CH3OH
(705.69)
25–72
2.27
2.26
54
-Loss of 0.5 lattice CH3OH.
72–185
5.10
5.05
169
-Loss of 2 coordinated H2O.
185–305
8.36
8.09
259
-Loss of 1CH3COO group.
305–436
23.11
22.81
385
-Loss of 1CH3COO group + C6H4N2.
436–800
Not assigned
32.86
–
- Not assigned.
3
[Cu2(HTNBS)(NO3)2]·H2O
(677.57)
25–97
2.65
2.67
48
-Loss of lattice H2O.
97–304
6.79
6.85
162, 204
-Loss of NO2 group.
304–449
30.72
30.61
371
-Loss of NO2 group + C3H2N2O2S2.
449–800
Not assigned
42.28
550
- Not assigned.
4[Cu(HINBS)
(H2O)Cl]·0·.5H2O
(547.47)
25–92
1.64
1.91
69
-Loss of 0.5 lattice H2O.
92–234
17.26
17.81
139, 195 213
-Loss of 1 coordinated H2O + 0.5Cl2 + C2H3N.
234–302
10.05
10.11
283
-Loss of C3H3O.
302–436
14.44
14.19
334
-Loss of SO2NH.
436–800
19.01
19.92
654
-Loss of C6H4N2.
5[Cu(HINBS)
(OAc)(H2O)]
(562.05)
25–150
3.20
3.15
86
-Loss of 1 coordinated H2O.
150–403
50.05
50.28
281
-Loss of CH3COO + C10H6 + C7H6N2O.
403–800
Not assigned
22.72
458
- Not assigned.
6[Cu(HINBS)
(NO3)(H2O)]·H2O
(583.03)25–133
6.17
6.43
110
-Loss of 1 lattice and 1 coordinated H2O.
133–294
19.03
19.33
261
-Loss of C5H7N2O moiety.
294–715
39.48
39.65
358
- Loss of NO3 group + C7H4N2O2S moiety.
Compounds 1–6 underwent disintegration in three decomposition ranges as for complexes 5 and 6, four decomposition ranges as in the complexes 1 and 3 or five decomposition ranges as in the complexes 2 and 4.
A closer look at the results given in Table 4 revealed that only lattice or lattice and coordinated water molecules are lost during the first stage of disintegration, which began at ambient temperature and continued up to 150 °C. The next step of decomposition occurred between 72 and 403 °C, inside which coordinated water or anions get lost in some compounds. The subsequent steps of breakdown are attributed to the progressive breakdown of the organic ligand molecules included in the complexes’ structures. Table 4 also lists and illustrates the DTG temperatures for each stage. The following Scheme (Scheme 2) also indicate the thermal decomposition of complexes 1 and 2, as example:Decomposition stages of complexes 1 and 2.
3.2.7 X-ray diffraction investigations
In light of current advancements in material science technology and knowledge, X-ray diffraction (XRD) is considered as one of the most significant and frequently employed material characterization technique. It is commonly applied to gather structural information on chemical compounds' microcrystalline structures in the solid state (Ali et al., 2022). As a result, the diffraction patterns of organic ligands H2INBS & H2TNBS and inspected metal complexes (1–6) were implemented within 10° <2θ > 80° scattering angle range (Figs. 7 & S10). The ligands' diffraction patterns allocations are drastically distinct from those of the equivalent metal complexes, which supports complex formation and shows that there are no contaminants or beginning reactants that are smeared (Khedr et al., 2022). The acquired x-ray diffraction patterns denoted that ligand H2TNBS displayed high crystallinity with diffraction patterns totally distinct from those of complexes 1–3 which exhibited good to weak crystallinity. Also, high crystallinity afforded for ligand H2INBS, good crystallinity of complex 4 and weak crystallinity of complex 5, with completely different diffraction patterns, whereas complex 6 is fully amorphous. Depending upon the Deby-Scherrer equation calculations (Sanad et al., 2022), the average crystallite sizes have been calculated:
XRD spectra of H2INBS complex 4.
where, τ is the mean size (crystallite size) of the ordered (crystalline) domains, which may be smaller or adequate to the grain size, λ is the X-ray wavelength, K mean a dimensionless shape factor and its value amounts to unity (0.9), but varies with the actual shape of the crystallite, β refers to line broadening at half the maximum intensity (FWHM), after subtraction of the instrumental line broadening.
Applying this equation to the obtained spectra, the ligand H2TNBS, complex 1, complex 2, and complex 3 exhibited 20.7, 19.0, 5.5 and 16.3 average crystallite sizes, respectively, and ligand H2INBS, complex 4, and complex 5 found to be 33.7, 25.3 and 8.0 nm, respectively. The irregular arrangement of the materials’ particles through their precipitation may be indicated by the complexes' amorphous nature. This amorphous state will predominate if the complexation process is rapid and/or the produced chemical is cooled quickly. Additionally, many of the structural elements of the complex cannot be accommodated in crystalline form. The amorphous compounds with probable small sizes are usually investigated by TEM technique to examine if they are having particles sizes in the nano-metric range. These nanometer-scale compounds usually receive a lot of interest due to their improved functional properties and wide range of potential technological applications, including optics, microelectronics, catalysis, bio- and chemical sensors (Gaber et al., 2017).
3.2.8 Investigation of TEM images
Transmittance electron microscopy (TEM) is one of the most common techniques to study a wide range of fascinating nano-sized materials. It's a method that's usually employed to look at the shape and dimensions of solid materials. Therefore, the particle sizes and crystallinities of the produced complexes 1–6 were further investigated by TEM technique. Images captured by the transmittance electron microscope in bright field for complexes 1–6 are shown in Fig. 8 & S11-S14 demonstrating that the samples are made up of teeny, different-sized nanoparticles. These photos make it abundantly notable that the complexes powders are made up of nanoscale particles with spherical-like shape (either regular or not). Additionally, it was determined that the particle size of metal complexes ranged between 15.9 and 45.6 nm for example complex 1 and 27.89–39.37 nm for 2 while complex 3′s was 18.19–34.12 nm, complex 4′s was 9.48–23.25 nm, complex 5′s was 12.13–18.45 nm and complex 6′s was 13.80–20.63 nm.Compounds 2 and 6 TEM images.
Accordingly, in comparison to their bulk analogues, the biological effectiveness of these innovative synthetic compounds is suspected to be stronger at these nanometric scales. There is a lot of interest that might be sparked by the novel functional properties of the nanosized chelates and the wide range of potential uses (Saad et al., 2019).
As a result, and in light of all the investigations previously deliberated, the structures of compounds 1–6 are depicted in Figs. 9 and 10.Concluded structures for complexes 1–3.
Concluded structures for complexes 4–6.
3.3 Applications
3.3.1 Antimicrobial investigation
As described in the experimental part, the examined two ligands, H2TNBS and H2INBS, and their complexes (1–6) were subjected to antimicrobial screening applying a well diffusion method (Jahangirian et al., 2013) (Table 5). All inspected compounds displayed very strong and promising against S. aureus within inhibition zones ranged from 15 to 26 mm which are higher than the inhibition zone of the used standard Gram-positive ant-bacterial agent Amoxycillin-clavulinic (10 mm) by one and half to two and half. Also, the inhibition zones of examined ligands and complexes towards B. cereus varied from 12 to 28 mm, which exceeds the inhibition zone of standard Gram-negative antibacterial agent (12 mm) by two and half times of value. Complexes 1, and 5 exhibited inhibition zones equal 9 equivalent to that of the standard bactericide (8 mm) against E. coli, whereas complexes 2 showed 19 mm, more than the double value of the common standard towards the same organism. respectively. Complex 4, displayed antibacterial activity (16 mm) double that of the applied standard (8 nm) against S. typhi. At the same time, ligand H2TNBS and complex 3 showed 12 and 9 mm inhibition zones towards it. Only ligand H2INBS showed excellent antifungal activity against the multicellular fungus A. flavus, with 26 nm inhibition zone which is higher than double the value of applied standard fungicide (11 mm). Ligands H2INBS and H2TNBS exhibited inhibition zones 16 and 12 mm against C. albicans. Upon complexation the inhibition zones of complexes 3 and 5 increased compared with the corresponding ligands and arrived 16and 18 mm. These values are equivalent or greater than the standard Amphotericin-p (13 mm) towards the same unicellular fungus. The selected species demonstrate diversity of microorganisms, including Gram-positive and Gram-negative bacteria, unicellular and multicellular fungus (Khedr et al., 2022). In most instances, the increased antibacterial effect during chelate formation can be ascribed to metal ions existence, with more hyper sense to microbial cells than the chelating ligands (Saad et al., 2017). Moreover, the greater action of the metal chelates may be caused by the increased lipophilicity during complexation. The coordinated metal ions may cause detrimental interactions between cellular components, or they may stop cellular enzyme activity. Obvious variation in the efficacy of the alternative compounds under study contra distinct organisms are brought on either by the impermeability of the cells of microorganisms or variations in the ribosomes in microbs’ cells (Singh et al., 2007). Additionally, Overton's notion and Tweedy's chelation hypothesis offer the greatest explanations for this improved efficiency upon complex development (Saad et al., 2019). According to the available information, the studied ligands and their complexes hold a great deal of promise for becoming broad-spectrum, novel, highly effective, fungicides and bactericides.
Compound
Staphylococcus Aureus
Bacillus cereus
Escherichia Coli
Salmonella typhi
Aspergillus flavus
Candida albicans
H2TNBS
15
15
–
12
–
12
1
26
22
9
–
–
6
2
25
28
19
–
–
8
3
24
20
–
9
–
16
H2INBS
16
12
–
–
26
16
4
17
24
–
16
–
9
5
22
20
9
2
–
7
6
21
24
–
–
–
18
Amoxycillin-clavulinic (AMC30)
10
12
8
8
–
–
Amphotericin-p
–
–
–
–
11
13
3.3.2 Performance as antitumor agents
The vast number of cytotoxic chemicals that are currently available includes a significant number of metal-based complexes. Platinum (II) complexes, such as cisplatin, oxaliplatin, and carboplatin, which are particularly target genomic DNA are frequently utilized in clinical settings to treat a variety of malignancies. In about 50% of chemotherapy-treated cancer patients, either alone or in combination, a platinum medication is administered. Despite playing a key role in cancer chemotherapy, platinum medicines have significant limitations, including a narrow range of antitumor activity, systemic dose-related toxicity, and a propensity to induce drug resistance, which frequently results in treatment failure (Gamberi and Hanif, 2022). The exploration of nonplatinum metal-based medications as an efficient substitute has sparked a great interest in response to these observations (Saad et al., 2018). In this regard, we seek to evaluate the antitumor efficacy of azo ligands H2TNBS & H2INBS and their chelates (1–6) against human Lung carcinoma (A-549) and Pancreatic carcinoma (Panc-1) and cells (Table 6 and Fig. 11). Untreated cells used as a control and vinblastine sulphate (VS) as a reference drug which had IC50 values of 4.68 and 24.6 and µg/ml contra pancreatic and lung cancer cell lines, respectively. The inspected compound concentration that caused 50% decrease in cell reproduction are referred IC50. IC50 is defined as the concentration that, under usual conditions, inhibits cell growth by 50%. The average of three separate experimental repetitions is averaged to produce each final value (Khedr et al., 2019). The obtained data made obvious that inspected compounds exhibited an increasing propensity to colonies lung cancer cell lines; H2TNBS < H2INBS < 6 < 4 < 5 < 1 < 3 < 2. The examined compounds reduced the viability of cancer cells, which was significantly influenced by their structures, mainly with the type of anions attached to the complex and geometrical arrangement around the core metal ion (Gaber et al., 2018). With IC50 values of 10.41, 4.99, 7.44, 19.62, 12.65 and 29.54 μg ml−1 for complexes 1–5, respectively, demonstrated strong and very encouraging anticancer actions toward A-549 cells, overcoming the standard vinblastine sulphate having IC50 of 24.60 μg ml−1. Complex 6 displayed very good anticancer efficiency with IC50 equals 29.54 μg ml−1, substantially equal to VS against the same tumor cells. Furthermore, an increase in anticancer effectiveness and inhibition of cell viability against Panc-1 were seen for all drugs under consideration in increasing order; H2TNBS < H2INBS < 6 < 4 < 5 < 1 < 3 < 2. With IC50 values of 2.57 and 3.98 g μg ml−1, respectively, complexes 2 and 3 demonstrated strong and extremely promising anticancer actions toward Panc-1 cells, exceeding the standard drug vinblastine sulphate with IC50 equal 4.68 μg ml−1. Also, complexes 5 and 1 displayed excellent and encouraging Panc-1 antitumor efficiency within 9.20 and 6.54 μg ml−1, IC50 value whereas complex 1 showed acceptable IC50 (14.23 μg ml−1). Depending on these findings, these compounds may undoubtedly be regarded as efficient and potential anticancer therapies for the examined tumor cells.
Compound
IC50 (µg/ml)
A-549
(Human lung carcinoma)
PANC-1
(Pancreatic carcinoma)
H2TNBS
122.5 ± 6.78
78.52 ± 4.78
1
10.41 ± 0.97
6.54 ± 0.65
2
4.99 ± 0.36
2.57 ± 0.17
3
7.44 ± 0.62
3.96 ± 0.23
H2INBS
48.82 ± 3.98
26.59 ± 2.09
4
19.62 ± 0.98
14.23 ± 1.21
5
12.65 ± 0.67
9.20 ± 0.98
6
29.54 ± 0.91
20.09 ± 0.65
Vinblastine Sulfate
24.6 ± 0.65
4.68 ± 0.65
The in-vitro anticancer evaluation test as IC50 (μg/ml) of H2TNBS, H2INBS and their metal chelates (1–6) against A-549 and PANC-1 cancer cells.
3.3.3 Effectiveness toward phenoxazinone synthase and kinetic studies
Recording of APX absorption peak that produced consequent to the oxidation of OAP by the power of Cu(II) chelates; 1–6, exhibited a raise of the main peak APX with time. This band emerged in Cu(II) complexes’ spectra around 435 nm, and consequently supported their catalytic effect for phenoxazinone synthase like activity. Figs. 12, 13 & S15-S18 show the time dependent spectra of complexes 1–6 with concentration of 3x10-5 M with addition of 10-2 M of OAP in DMF medium under aerobic condition. For these catalytic reactions, once OAP was added to the complexes’ solutions, time started, and the compounds’ spectra were recorded each 5 min up to 70 min. When the same experiments were run without catalysts, no observed increase in APX peak was seen (blank test). The observed rate constant (kobs) and the initial rate of reaction (Vo) have been assessed using the initial rate method (Table 7) (ε(APX) = 18300 M−1cm−1) (Podder and Mandal, 2020). .The increase in APX spectral peak at ≈ 435 nm upon addition of 10-2 M of OAP to 3x10-5 M of catalyst 4, the reaction medium is DMF.
The increase in APX spectral peak at ≈ 435 nm upon addition of 10-2 M of OAP to 3x10-5 M of catalyst 6, the reaction medium is DMF.
Comp. code
Kobs (min−1)
V○
(M min−1)
Vmax
(M min−1)
KM (M)
kcat (h−1)
1
0.00715
3.91x10-7
3.04x10−5
0.744
60.8
2
0.0034
1.86x10-7
2.17x10−6
0.091
4.34
3
0.0047
2.60x10−6
1.607x10−5
0.593
32.15
4
0.01451
7.93x10-7
1.95x10−4
2.41
390.48
5
0.00298
1.63x10-7
1.14x10−5
0.692
22.82
6
0.0391
2.14 x10-6
2.33x10−4
0.973
467.01
The kinetics of the reactions were studied in order to assign the applied Cu(II) complexes' level of catalytic performance in comparison to one another and to other known comparable compounds. In these reactions, 3 x10-5 M solutions of each catalyst was treated with series of substrate concentrations (with concentrations in the range 1–8 × 10-3 M). The substrate concentration 10 times, at least, higher than the concentration of the catalyst to make sure that pseudo-first order kinetics is applied. For each catalytic solution, the absorption peak of APX, which is formed as the product of substrate oxidation, was taken once the reaction started and each 3 min intervals until the time of the reaction reached 21 min. The recorded absorbances were used to assign the initial rate for each catalytic solution. Plotting the substrate concentrations ([S]) versus initial rate (Vo) inferred that the majority of the Cu complexes under investigation provided rate saturation kinetics (Figs. 14, 15, and S19–S22). The outcomes clearly demonstrate the pre-equilibrium production of the complex-substrate intermediate, and the irreversible substrate oxidation is the rate-determining stages for catalytic cycle, which eventually developed APX, as demonstrated by the equation:Michaelis–Menten plot (left) and Lineweaver–Burk plot (right) of complex 5 for catalytic oxidation of OAP.
Michaelis–Menten plot (left) and Lineweaver–Burk plot (right) of complex 6 for catalytic oxidation of OAP.
The applied catalysts afforded saturation kinetics, so Michaelis-Menten equation is the most significant equation for treatment of such reactions. The following equation illustrates the connection between [substrate] (i.e [S]) and the rate of the enzyme reaction (V) using this kinetic-like model:
Vmax is the greatest rate which the system can achieve when [substrate] reaches saturation levels. The Michaelis constant, or KM, refers to the substrate’s concentration at which the reaction rate is half of Vmax. By linearizing the Michaelis-Menten equation, a double reciprocal Lineweaver-Burk plot is created, which is applied to examine different kinetic parameters like Vmax and KM. The formula serves as an example of the Lineweaver-Burk Equation.
Fulfillment of the previous equation to the kinetic process of the examined catalysts (Figs. 14 & 15), result in the calculation of the kinetic parameters Vmax, KM and the turnover number (kcat, h−1) which are collected in Table 7.
Table 7 shows that H2INBS chelates (compounds 4–6) had higher activity than H2TNBS chelates (complexes 1–3), indicating that the ligand structure is one of the factors influencing the activity of the complexes.
On comparing the activity of the synthesized catalysts; 1–6, relatively according to each other, the following remarks can be summarized:
For complexes 1–3, the catalytic activity of these complexes that deduced from TOF values indicated complex 1 to exhibit the uppermost efficacy with kcat of 60.8 h−1 followed by compound 3 having kcat value of 32.15 h−1 and finally the least activity afforded by compound 2 with kcat of 4.34 h−1. On the other hand, for H2INBS chelates; 4–6, the obtained results showed an extreme activity for compound 6 having 467.01 h−1 kcat value and the next is compound 4 having kcat equals 390.48 h−1 and the least activity observed for complex 5 having kcat equals of 22.82 h−1.
It is observed that APX generation through oxidation of OAP by the catalytic effect of copper chelates can only take place in existence of molecular dioxygen since there is hardly any peak of APX appears without the availability of O2 in a similar behavior of reported work (Sohtun et al., 2021; Mandal et al., 2020). The suggested mechanism is shown in Scheme S1 (Sarkar et al., 2017).
Comparing the activity of the metal chelates with each other can lead to conclude that the metal center oxidation number and its geometry. The distance between metals in bimetallic center and the structure of organic ligand are also important factors that contributes to catalysts’ activity. Comparing the activity of the metal complexes incorporating the same ligand (i.e. compounds 1–3 with each other and compounds 4–6 with each other) inferred that the key factor controlling the activity is the type of incorporated anion sine all the other factors among each group of compounds are the same. Cloro complexes, 1 & 3, and nitro complexes, 3 & 6, provided the highest activity whereas the other complexes, which are only weakly active, contain CH3COO– anions in their structure (i.e., complexes 2 and 5). Depending on the approachability of the labile sites in the first coordination sphere, which is better in case of chloro and nitro groups, it is possible to explain why Cl- or NO3– complexes have greater activity than CH3COO– complexes (Panja et al., 2018). The comparable efficacy may also be elucidated based on the steric impact brought on by the CH3COO– group's bulkiness when compared to Cl- or NO3–.
4 Conclusion
H2TNBS and H2INBS azo dye ligands which synthesized by coupling of sulfathiazole and sulfafurazole diazonium salts and 2-hydroxy-1-naphthaldehyde were obtained and identified applying spectral and analytical ways. Both ligands were used for synthesizing Cu(II) chelates through their reaction with copper salts of different anions (CuCl2·2H2O, Cu(CH3COO)2·H2O and Cu(NO3)2). The structures of the obtained Cu(II) complexes have been assured to be all square planar in geometry with 1:1 (M:L) composition for H2INBS chelates and 2:1 (M:L) composition for H2TNBS chelates. IR. Morphology of the synthesized compounds have been investigated using TEM technique which joint with the results of X-ray powder diffraction spectroscopy confirmed the precipitation of both ligands and their complexes in the nanometric scale. The compounds under interest have been tested for their activity against different bacterial and fungal strains showing, in most cases, enhancement of activity upon chelation. The antitumor activity of Cu(II) chelates and their ligands has been also examined against the cell lines A-549 cells (human Lung cancer cell line) and Panc-1 (Pancreatic carcinoma); standard drug vinblastin has been used for comparison. The compounds showed the order of activity to be H2TNBS < H2INBS < 6 < 4 < 5 < 1 < 3 < 2 for A-549 cells and H2TNBS < H2INBS < 6 < 4 < 5 < 1 < 3 < 2 for Panc-1 cells. Additionally, the synthesized complexes were estimated for their effectiveness as catalysts for oxidative coupling of 2-aminophenol to 2-aminophenoxazine-3-one. The examined extremely high activity for the chloro complex 4 and nitro complex 6 with TOF numbers of 390.48 and 467.01 h−1, respectively. The rest of compounds afforded TOF numbers ranging between 4.34 and 60.8 h−1.
Acknowledgements
The authors would like to thank the Deanship of Scientific Research at Umm Al-Qura University for supporting this work by Grant Code: (22UQU4281810DSR01).
Declaration of Competing Interest
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
References
- X-Ray Diffraction Techniques for Mineral Characterization: A Review for Engineers of the Fundamentals. Applications, and Research Directions, Minerals.. 2022;12:205.
- [Google Scholar]
- Novel Synthesized Benzesulfonamide Nanosized Complexes; Spectral Characterization, Molecular Docking, Molecular Modeling and Analytical Application. J. Inorg. Organomet. Polym. Mater.. 2019;29:876-892.
- [CrossRef] [Google Scholar]
- Awad, I.M.A., Osman, A.H., Aly, A.A.M. 2010. Heterocyclo-Substituted Sulfa Drugs : Part XII . Mercapto-S-Azo-Benzothiazol Dyes and Their Metal Complexes, Taylor & Francis Inc. 37–41.
- Classifications, properties, recent synthesis and applications of azo dyes. Heliyon. 2020;6:e03271.
- [Google Scholar]
- Different roads to discovery; Prontosil (hence sulfa drugs) and penicillin (hence β-lactams) J. Ind. Microbiol. Biotechnol.. 2009;36:775-786.
- [Google Scholar]
- Trapping intermediate MLCT states in low-symmetry Ru(bpy) complexes. Chem. Sci.. 2017;8:17434-18142.
- [Google Scholar]
- A El‐Ghamry, H., Takroni, K. M., AL‐Rashidi, D. O., Alfear, E. S., Alsaedi, R. A. 2022.Design, spectral, thermal decomposition, antimicrobial, docking simulation and DNA binding tendency of sulfisoxazole azo dye derivative and its metal chelates with Mn2+, Fe2+, Co2+, Ni2+, Cu2+, Zn2+ and Cd2+, Applied organometallic Chemistry, 36, e6813.
- Investigation of the inhibition efficiencies of novel synthesized cobalt complexes of 1,3,4-thiadiazolethiosemicarbazone derivatives for the acidic corrosion of carbon steel. J. Mol. Struct.. 2020;1203:127447
- [Google Scholar]
- Sulfonamides repress cell division in the root apical meristem by inhibiting folates synthesis. J. Hazard. Mater. Adv.. 2022;5:100045
- [Google Scholar]
- Characterization and thermal studies of nano- synthesized Mn(II), Co(II), Ni(II) and Cu(II) complexes with adipohydrazone ligand as new promising antimicrobial and antitumor agents. Appl. Organomet. Chem.. 2017;31
- [Google Scholar]
- Nano-synthesis, characterization, modeling and molecular docking analysis of Mn (II), Co (II), Cr (III) and Cu (II) complexes with azo pyrazolone ligand as new favorable antimicrobial and antitumor agents. Appl. Organomet. Chem.. 2018;32(12):e4606.
- [Google Scholar]
- Gasque, L., Mendieta, A., Ferrer-Sueta, G. 2020. Comment on “mixed azido/phenoxido bridged trinuclear Cu(ii) complexes of Mannich bases: Synthesis, structures, magnetic properties and catalytic oxidase activities”,: Dalton Trans., 2018, 47, 9385-9399 and “tri- And hexa-nuclear NiII-MnII complexes of a N2O2 donor unsymmetrical ligand: Synthesis, structures, magnetic properties and catalytic oxidase activities”, Dalton Trans., 47 (2018) 13957-13971, Dalt. Trans. 49 (2020) 3365–3368.
- Sulfonamides with hydroxyphenyl moiety: Synthesis, structure, physicochemical properties, and ability to form complexes with Rh(III) ion. Polyhedron.. 2022;221:115865
- [Google Scholar]
- The use of conductivity measurements in organic solvents for the characterisation of coordination compounds. Coord. Chem. Rev.. 1971;7:81-122.
- [Google Scholar]
- The effect of pH and triethanolamine on sulfisoxazole complexation with hydroxypropyl-β-cyclodextrin. Eur. J. Pharm. Sci.. 2003;20:285-293.
- [Google Scholar]
- A class of sulfonamides as carbonic anhydrase I and II inhibitors. J. Enzyme Inhib. Med. Chem.. 2016;31:180-188.
- [Google Scholar]
- Design, synthesis and evaluation of novel (S)-tryptamine derivatives containing an allyl group and an aryl sulfonamide unit as anticancer agents. Bioorganic Med. Chem. Lett.. 2019;29(9):1133-1137.
- [Google Scholar]
- Copper and Its Complexes in Medicine: A Biochemical Approach. Mol. Biol. Int.. 2011;2011:1-13.
- [Google Scholar]
- Use of a New Polymer Anchored Cu(II) Azo Complex Catalyst for the Efficient Liquid Phase Oxidation Reactions. J. Inorg. Organomet. Polym.. 2010;20:87-96.
- [Google Scholar]
- Well Diffusion Method for Evaluation of Antibacterial Activity of copper phenyl fatty hydroxamate synthesized from canola and palm kernel oils. J. Nanomater. Biostructures.. 2013;8:1263-1270.
- [Google Scholar]
- Synthesis, structure analysis, and antibacterial activity of some novel 10-substituted 2-(4-piperidyl/phenyl)-5,5-dioxo[1,2,4]triazolo[1,5-b][1,2,4]benzothiadiazine derivatives, Bioorganic Med. Chem. Lett.. 2007;17:5400-5405.
- [Google Scholar]
- Tautomeric stability, molecular structure, NBO, electronic and NMR analyses of salicylideneimino-ethylimino-pentan-2-one. J. Mol. Struct.. 2016;1112:87-96.
- [Google Scholar]
- Synthesis, characterization and applications of schiff base chemosensor for determination of Cu2+ ions. J. Saudi Chem. Soc.. 2022;26:101503
- [Google Scholar]
- Microwave-Assisted Synthesis of Arylazoaminopyrazoles as Disperse Dyes for Textile Printing, Zeitschrift Fur Anorg. Und Allg. Chemie.. 2016;642:766-772.
- [Google Scholar]
- Novel series of nanosized mono- and homobi-nuclear metal complexes of sulfathiazole azo dye ligand: Synthesis, characterization, DNA-binding affinity, and anticancer activity. Inorg. Chem. Commun.. 2019;108:107496
- [Google Scholar]
- Khedr, A.M., El–Ghamry, H.A., El-Sayed, Y.S., 2022. Nano-synthesis, solid state structural characterization, antimicrobial and anticancer assessment of new sulfafurazole azo dye-based metal complexes for further pharmacological applications. Appl. Organomet. Chem. 36, e6548.
- Nano-synthesis approach, elaborated spectral, biological activity and in silico assessment of novel nano-metal complexes based on sulfamerazine azo dye. J. Mol. Liq.. 2022;352:118737
- [Google Scholar]
- Analysis of the oral hypoglycemic agent, glibenclamide, in a health food. Yakugaku Zasshi.. 2003;123:1049-1054.
- [Google Scholar]
- Imaging Cu2+ binding to charged phospholipid membranes by high-throughput second harmonic wide-field microscopy. Chem. Phys.. 2021;155:184704
- [Google Scholar]
- Synthesis, crystal structure, and magnetism of a binuclear Cu(II) complex with asingle end-to-end bridged azido ligand. J. Coord. Chem.. 2009;62:2610-2615.
- [Google Scholar]
- Structure and Synthesis of Copper Based Schiff base and Reduced Schiff base Complex: A Combined Experimental and Theoretical Investigation of Biomimetic Catalytic Activity. Dalton Trans.. 2020;49:15461-15472.
- [Google Scholar]
- The antibacterial and anti-biofilm activity of gold-complexed sulfonamides against methicillin-resistant Staphylococcus aureus. Microb. Pathog.. 2018;123:440-448.
- [Google Scholar]
- Voltammetric, potentiality and thermodynamic studies of some rhodanine sulfa drugs azo compounds in aqueous solutions. 2006;3:155-164.
- Rapid colorimetric assay for cellular growth and survival: application to proliferation and cytotoxicity assays. J. Immunol. Methods. 1983;65:55-63.
- [Google Scholar]
- Studies on thermal behavior of some antibacterial copper (II) complex compounds with a dibiguanide derivative ligand. J. Therm. Anal. Calorim.. 2013;111:1189-1195.
- [Google Scholar]
- The necessity of free and uncrowded coordination environments in biomimetic complex models: oxidative coupling by mixed-ligand cobalt(ii) complexes of diazene-disulfonamide. New J. Chem.. 2019;43:18322-18330.
- [Google Scholar]
- Influence of anions and solvents on distinct coordination chemistry of cobalt and effect of coordination spheres on the biomimetic oxidation of o-aminophenols. Mol. Catal.. 2018;449:49-61.
- [Google Scholar]
- Aerobic oxidation of 2-aminophenol catalysed by a series of mononuclear copper(ii) complexes: phenoxazinone synthase-like activity and mechanistic study. New J. Chem.. 2020;44:12793-12805.
- [Google Scholar]
- ChemInform Abstract: Synthesis and Anticancer Activities of Thiazoles, 1,3-Thiazines, and Thiazolidine Using Chitosan-Grafted-Poly(vinylpyridine) as Basic Catalyst. ChemInform.. 2015;46
- [Google Scholar]
- Density functional theory/B3LYP study of nanometric 4-(2,4-dihydroxy-5-formylphen-1-ylazo)-N-(4-methylpyrimidin-2-yl)benzenesulfonamide complexes: Quantitative structure–activity relationship, docking, spectral and biological investigations. Appl. Organomet. Chem.. 2017;31:1-15.
- [Google Scholar]
- Elaborated spectral, modeling, QSAR, docking, thermal, antimicrobial and anticancer activity studies for new nanosized metal ion complexes derived from sulfamerazine azodye. J. Therm. Anal. Calorim.. 2018;131:1249-1267.
- [Google Scholar]
- Nano-synthesis, Biological Efficiency and DNA Binding Affinity of New Homo-binuclear Metal Complexes with Sulfa Azo Dye Based Ligand for Further Pharmaceutical Applications. J. Inorg. Organomet. Polym. Mater.. 2019;29:1337-1348.
- [Google Scholar]
- Synthesis, structural characterization and DNA binding affinity of new bioactive nano-sized transition metal complexes with sulfathiazole azo dye for therapeutic applications. Appl. Organomet. Chem.. 2019;33:1-14.
- [Google Scholar]
- Anticancer activity of hydroxy- and sulfonamide-azobenzene platinum(II) complexes in cisplatin-resistant ovarian cancer cells. J Inorg Biochem.. 2017;174:102-110.
- [Google Scholar]
- Hetero-valent cations- doped zinc stannate nanoparticles for optoelectronic and dielectric applications. Mater. Chem. Phys.. 2022;291:126700
- [Google Scholar]
- Two new manganese(III) complexes with salicylaldimine Schiff bases: Synthesis, structure, self-assembly and phenoxazinone synthase mimicking activity. Inorganica Chim. Acta.. 2017;457:19-28.
- [Google Scholar]
- Synthesis, characterization and biological activity of complexes of 2-hydroxy-3,5-dimethylacetophenoneoxime (HDMAOX) with copper(II), cobalt(II), nickel(II) and palladium(II), Spectrochim. Acta – Part A Mol. Biomol. Spectrosc.. 2007;68:63-73.
- [Google Scholar]
- Copper(II) complexes of tripodal ligand scaffold (N3O) as functional models for phenoxazinone synthase. J. Inorg. Biochem.. 2021;216:111313
- [Google Scholar]
- Diuretics: From Classical Carbonic Anhydrase Inhibitors to Novel Applications of the Sulfonamides. Curr. Pharm. Des.. 2008;14:641-648.
- [Google Scholar]
- Synthesis, structure elucidation, DNA binding and molecular docking studies of novel copper (II) complexes of two 1,3,4-thiadiazolethiosemicarbazone derivatives. Appl. Organomet. Chem.. 2020;34:1-16. e5860
- [Google Scholar]
- Site-specific binding of a water molecule to the sulfa drugs sulfamethoxazole and sulfisoxazole: a laser-desorption isomer-specific UV and IR study. Phys. Chem. Chem. Phys.. 2018;20:6891-6904.
- [Google Scholar]
- The Synthesis and Characterization of Star Shaped Metal Complexes of Triazine Cored Schiff Bases: Their Thermal Decompositions and Magnetic Moment Values. J. Inorg. Organomet. Polym. Mater.. 2011;21:291-296.
- [Google Scholar]
- Sulfonamide derivatives as potential anti-cancer agents and their SARs elucidation. Eur. J. Med. Chem.. 2021;226:113837
- [Google Scholar]
- Unexpected Nanomolar Inhibition of Carbonic Anhydrase by COX-2-Selective Celecoxib: New Pharmacological Opportunities Due to Related Binding Site Recognition. J. Med. Chem.. 2004;47:550-557.
- [Google Scholar]
- Development and evaluation of a Luminex xTAG assay for sulfonamide resistance genes in Escherichia coli and Salmonella isolates. Mol. Cell. Probes.. 2020;49:101476
- [Google Scholar]
- Synthesis, spectroscopic and DNA binding ability of CoII, NiII, CuII and ZnII complexes of Schiff base ligand (E)-1-(((1H-benzo[d]imidazol-2-yl)methylimino)methyl)naphthalen-2-ol. X-ray crystal structure determination of cobalt (II) complex. Mater. Sci. Eng.. 2017;75:1059-1067.
- [Google Scholar]
Appendix A
Supplementary data
Supplementary data to this article can be found online at https://doi.org/10.1016/j.arabjc.2023.104916.
Appendix A
Supplementary data
The following are the Supplementary data to this article:Supplementary data 1
Supplementary data 1