Next Article in Journal
Reproducing Kernel Hilbert Space Approach to Multiresponse Smoothing Spline Regression Function
Previous Article in Journal
Distributed Energy-Efficient Assembly Scheduling Problem with Transportation Capacity
Previous Article in Special Issue
[3 + 2] Cycloadditions in Asymmetric Synthesis of Spirooxindole Hybrids Linked to Triazole and Ferrocene Units: X-ray Crystal Structure and MEDT Study of the Reaction Mechanism
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of a New Ag(I)-Azine Complex via Ag(I)-Mediated Hydrolysis of 2-(((1-(Pyridin-2-yl)ethylidene)hydrazineylidene) Methyl)phenol with AgClO4; X-ray Crystal Structure and Biological Studies

by
Mezna Saleh Altowyan
1,
Saied M. Soliman
2,*,
Dhuha Al-Wahaib
3,
Assem Barakat
4,
Ali Eldissouky Ali
2,* and
Hemmat A. Elbadawy
2,*
1
Department of Chemistry, College of Science, Princess Nourah bint Abdulrahμan University, P.O. Box 84428, Riyadh 11671, Saudi Arabia
2
Chemistry Department, Faculty of Science, Alexandria University, P.O. Box 426, Alexandria 21321, Egypt
3
Chemistry Department, Faculty of Science, Kuwait University, Kuwait City 13060, Kuwait
4
Department of Chemistry, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Symmetry 2022, 14(11), 2226; https://doi.org/10.3390/sym14112226
Submission received: 16 September 2022 / Revised: 9 October 2022 / Accepted: 18 October 2022 / Published: 22 October 2022
(This article belongs to the Special Issue Symmetry/Asymmetry Applied in Chemical Synthesis)

Abstract

:
A new Ag(I)-azine complex of the formula [Ag(La)]2(ClO4)2.1/2(Lb) was synthesized and characterized using elemental analysis, FTIR, XPS and single-crystal X-ray diffraction. The [Ag(La)]2(ClO4)2.1/2(Lb) complex was obtained via Ag(I)-mediated hydrolysis of 2-(((1-(pyridin-2-yl)ethylidene)hydrazineylidene)methyl)phenol (L) with AgClO4 leading to the formation of the azines (1E,2E)-1,2-bis(1-(pyridin-2-yl)ethylidene)hydrazine (La) and 2,2’-((1E,1’E)-hydrazine-1,2-diylidene bis(methanylylidene)) diphenol (Lb). The former underwent complexation with AgClO4, while the latter was co-crystallized with the resulting Ag(I) complex, leading to the complex formula [Ag(La)]2(ClO4)2.1/2(Lb). In this complex, the azine La is acting as a bridged bidentate chelate connecting the two Ag(I) sites, leading to a tetra-coordinated Ag(I) with a distorted square planar coordination environment. Hirshfeld analysis indicated the importance of the H…H (38.4%), O…H (17.9%), H…C (13.2%), C…C (9.0%) and N…H (8.9%) interactions in the molecular packing. The antimicrobial, antioxidant and cytotoxic activities of the [Ag(La)]2(ClO4)2.1/2(Lb) complex were examined and compared with those of [Ag(La)]2(ClO4)2. The former was found to have lower bioactivity than the latter, which shed light on the lowering of biological actions as a consequence of the incorporation of the azine Lb within the Ag(I) complex. In other words, lowering the %Ag decreases the biological actions. The [Ag(La)]2(ClO4)2.1/2(Lb) complex (ΜIC = 39 μg/mL) has higher activity against the fungus A. fumigatus than the control Ketoconazole (ΜIC =156 μg/mL). This complex also has good cytotoxic activity against colon carcinoma and weak antioxidant activity.

1. Introduction

The construction of supramolecular inorganic–organic systems using self-assembly of small building blocks, including metal ions or metal salts and functional ligands, has attracted the interest of researchers [1,2,3]. These small building blocks are arranged in a specific and well-organized pattern via different types of interactions, including coordination, hydrogen bonding, and π–π interactions [4,5,6]. In the area of coordination chemistry, such methodology has been used not only to build many interesting structures with fascinating molecular and supramolecular characteristics [7], but also to produce functional materials which have interesting properties for applications in different areas, especially in medicine [8].
Silver is a coinage metal, and its different forms (metallic, ionic or coordinated) are well acknowledged to have interesting antimicrobial properties [9,10,11,12,13,14,15,16,17,18,19]. In this regard, researchers searching for new materials that could be used as metal-based drugs are interested in silver(I) coordination compounds due to their interesting biological characteristics. The drug silver sulfadiazine is a well-known example of the success of Ag(I) coordination compounds in the treatment of microbial infections [20,21,22]. Ag(I) complexes are also promising antimicrobial materials for use against drug-resistant microbes [13,23,24]. In addition, Ag(I) coordination compounds have attractive antitumor characteristics [25,26,27,28]. Silver complexes have potential applications [29,30,31,32,33] and are used as catalysts in organic reactions [34,35,36,37]. In this regard, organic transformations of some oxamohydrazides and hydrazones to azines in the presence of Ag(I) were presented by our research group [38,39,40]. Herein, we present a new example of hydrazone to azine conversion in the presence of AgClO4 (Figure 1). The reaction of 1-(pyridin-2-yl)ethylidene)hydrazono)methyl)phenol (L) with AgClO4 proceeded with hydrolysis leading to the formation of the corresponding Ag(I)-azine complex (Scheme 1). The structure of the new Ag(I) complex was confirmed using different spectral analyses, including FTIR, XPS and single-crystal X-ray diffraction analysis. In addition, the antimicrobial and cytotoxicity evaluations are presented.

2. Materials and Methods

2.1. Instrumentations

Instrumental and X-ray structure measurement details [41,42] are depicted in Supplementary data. The crystal data of the Ag(I)-azine complex are depicted in Table 1.

2.2. Synthesis of [Ag(La)]2(ClO4)2.1/2(Lb)

First, the hydrazone (L) was prepared according to the method described in the literature [43,44]. Then, 0.1 μmol silver(I) perchlorate in distilled water was added to an ethanolic solution of L (0.1 μmol), followed by the addition of 5 mL ΜeCN to dissolve the resulting yellow precipitate. The resulting mixture was filtered, and the clear filtrate was left at room temperature for slow evaporation for a week. Then, the resulting yellow crystals were picked from solution by filtration. Anal. Calc. for C35H34Ag2Cl2N9O9 (72%): C, 41.57; H, 3.39; N, 12.46; Ag, 21.33%. Found: 41.34; H, 3.27; N, 12.29; Ag, 21.22%. IR (KBr, cm−1): 3432, 1621, 1578, 1148 (sh), 1133 (sh), 1086.

2.3. Hirshfeld Calculations

The Hirshfeld topology analyses were performed using the program crystal explorer 17.5 [45].

2.4. Biological Studies

The biological activity of the Ag(I) complex and salt were determined using the methods described in the Supplementary data [46,47,48,49].

3. Results and Discussion

3.1. Synthesis and Characterizations

The reaction of AgClO4 with the hydrazone L did not proceed to the formation of the Ag(I)-hydrazone complex. Instead, the hydrazone L was converted into the corresponding azines La and Lb following a similar mechanism to that reported by Saied et al. [32]. After the new azines formed, the Ag(I)-azine complex, [Ag(La)]2(ClO4)2.1/2(Lb), was obtained from the reaction mixture as crystalline yellow product (Scheme 1). The synthesized Ag(I) complex was found to be soluble in ΜeCN, DΜSO and DΜF, but was not soluble in water, EtOH or ΜeOH. The single-crystal X-ray structure of the resulting complex was found to agree with the elemental analysis data, FTIR and XPS spectral analyses. Significant changes in the FTIR spectrum of the Ag(I) complex were detected when compared with that for the hydrazone L. The broad band detected at 1086 cm−1 with two shoulders at 1148 and 1133 cm−1 is good evidence of the perchlorate anion. In addition, the sharp band detected at 1621 cm−1 and the broad band at 3432 cm−1 are assigned to the C=N and O-H stretching vibrations of the coordinated azine La and the co-crystallized ligand Lb, respectively. None of these bands were detected in the free hydrazone ligand L. Additionally, the X-ray photoelectron spectra (XPS) is an important and powerful physicochemical technique for elemental and chemical structure predictions. Table 2 summarizes the elemental composition of the hybrid complex, [Ag(La)]2(ClO4)2.1/2Lb, while Figure 2 is a collective figure showing the XPS spectra of the Ag(I) complex. However, a doublet of binding energy peaks corresponding to 3d5/2 and 3d3/2 were detected at 368.8 eV and 374.8 eV, respectively, with a spin-orbital splitting (Δ) = 6.0 eV characteristic of the monovalent Ag(I) [40,50,51]. The presence of the perchlorate ion was evidenced by the appearance of binding energy peaks corresponding to Cl2p(3/2 and ½) at 208.04 and 209.63 eV, respectively, and Δ = 1.6 eV. The perchlorate and hydroxyl oxygen atoms showed peaks matching to O1s with B.E. = 532.58 and 533.26 eV. Moreover, the complex contains sp3 and sp2 hybridized carbon orbitals in different environments (pyridine ring, azine -C=N, -CH3 and C-H), and thus peaks corresponding to C1s appeared at 288.44, 286.69, 285.42 and 284.66 eV. The nitrogen atoms included in the hybrid complex showed peaks at 399.37, 400.1 and 401.87 eV, where the nitrogen atoms are either coordinated to Ag(I) ions or are covalently bonded to carbon or nitrogen via π or σ bonding.

3.2. X-ray Structure Description of [Ag(La)]2(ClO4)2.1/2(Lb) Complex

The structure of the [Ag(La)]2(ClO4)2.1/2(Lb) complex revealed, without a doubt, the Ag(I)-mediated hydrolysis of the Schiff base hydrazone ligand (L). In this case, the Ag(I)-azine based complex with the co-crystallized Lb molecule is crystallized in the lower symmetry monoclinic crystal system compared to the [Ag(La)]2(ClO4)2 without the co-crystallized Lb [40]. In the former, the unit cell parameters are a = 15.6034(11), b = 14.9707(12), c = 17.3993(12) Å and β = 110.554(3) while Z = 4. In this case, the asymmetric unit comprised one of the dinuclear formula [Ag(La)]2(ClO4)2 and half of the azine (Lb) as a co-crystallized molecule (Figure 3). In the crystal structure of the [Ag(La)]2(ClO4)2.1/2(Lb) complex, the two hydrolysis products La and Lb were trapped, which confirms the Ag(I)-mediated hydrolysis of L to the corresponding azines. The Ag(I) is found to be coordinated with the two molecules of the azine La which act as a bridged bidentate chelate connecting the two Ag(I) sites. The second azine Lb is found to be freely uncoordinated with Ag(I) and co-crystallized with the [Ag(La)]2(ClO4)2 molecule. In addition, the azine Lb was found to be stabilized by a strong intramolecular O-H…N hydrogen bond, which could be the main reason for this fragment being freely uncoordinated. The Ag-N(azine) bonds are slightly longer than the Ag-N(pyridine) ones (Table 3). The shortest Ag-N(azine) and Ag-N(pyridine) bonds are Ag1-N7 (2.383(5) Å) and Ag1-N2 (2.358(5) Å), respectively. The largest angles between the trans Ag-N bonds are the N5-Ag2-N10 (157.77(16)) and N2-Ag1-N7 (155.48(17)). The bite angles of the azine ligand La are in the range of 70.22(17) (N11-Ag2-N10) to 70.82(15) (N5-Ag2-N4). Hence, the Ag(I) is tetra-coordinated and has a distorted square planar coordination environment with almost equidistant diagonals, where the two diagonals are deviated significantly from perpendicularity as a consequence of the geometric preferences of the donor atoms in the bidentate chelate. The perchlorate anions are found to be freely uncoordinated, as indicated by the significantly long Ag…O distances (Ag1…O6: 3.163, Ag2…O7: 3.165 Å, Ag2…O9: 3.197 Å). These results agreed very well with the X-ray structure of the [Ag(La)]2(ClO4)2 complex [40].
The packing scheme of the [Ag(La)]2(ClO4)2.1/2(Lb) complex is shown in Figure 4. It is clear that the perchlorate anions play an important role in connecting the [Ag(La)]2(ClO4)2 complex units with each other, and also connect the co-crystallized azine (Lb) molecules with each other, both along the c-direction. Hence, the packing could be described as alternative hydrogen bonded inorganic and organic layers along the a-direction (Table 4). As can be seen from the view along the ac-plane, the inorganic and organic layers are nearly perpendicular to each other and are interconnected by C-H…O interactions with the perchlorate anion along the a-direction.

3.3. Hirshfeld Analysis

It is well known that the molecules in crystals are packed in order to maximize the crystal stability via complicated sets of intermolecular interactions. Hirshfeld analysis is considered to be a powerful tool for predicting all intermolecular contacts in the crystal. The Hirshfeld surfaces of [Ag(La)]2(ClO4)2.1/2(Lb) complex are shown in Figure 5. The Hirshfeld surfaces were drawn for each of the cationic complex unit [Ag(La)]2 and the co-crystallized organic azine molecule Lb.
The percentages of all intermolecular contacts contributing to the molecular packing of the cationic complex are shown in Figure 6. The most dominant contacts are the H…H (38.4%) and O…H (17.9%) interactions. Other contacts, such as the H…C (13.2%), C…C (9.0%) and N…H (8.9%), also significantly contributed to the molecular packing. As can be seen from Figure 5, only the Ag…O, O…H and H…H contacts appeared as red spots in the dnorm map. The Ag1…O6 (3.163 Å), Ag2…O7 (3.165 Å) and Ag2…O9 (3.197 Å) contacts are shorter than the vdWs radii sum of Ag and O atoms, indicating strong interactions with the perchlorate ion. These distances are significantly longer compared to the covalent radii of the respective atoms, which revealed an ionic perchlorate ion. A list of the other short contacts is depicted in Table 5.
For the co-crystallized organic azine molecule Lb, the major contacts are the H…H (40.0%), O…H (32.2%) and H…C (22.8%), while the minor contributing contacts are the N…H (4.0%) and O…O (0.3%). Among these interactions, the O…H contacts are the most significant (Table 5). The significance of the Ag…O, O…H and H…H interactions are further indicated by their appearance as sharp spikes in the fingerprint plots (Figure 7).

3.4. Biological Studies

3.4.1. Antimicrobial Activity

In the light of the promising antimicrobial activity of silver(I) complexes as antibacterial and antifungal agents, the bioactivity of the [Ag(La)]2(ClO4)2.1/2(Lb) complex was examined against Gram-positive and Gram-negative bacteria, as well as selected harmful fungi [46] (Table 6). The results were compared with the [Ag(La)]2(ClO4)2 complex [34]. The studied Ag(I) complex has interesting antibacterial and antifungal properties against the studied microbes. The sizes of the inhibition zones are in the range of 8–9 mm and 10–11 mm against Gram-positive and Gram-negative bacteria, respectively. In both cases, the bioactivity is found to be less than that of the antibacterial control Gentamycin (24–30 mm). On the other hand, the Ag(I) complex has higher antifungal activity against A. fumigatus than the antifungal control Ketoconazole. The sizes of the inhibition zones are 26 and 17 mm, respectively. In contrast, the antifungal activity of the [Ag(La)]2(ClO4)2.1/2(Lb) complex against C. albicans (16 mm) is lower than Ketoconazole (20 mm). Generally, the antimicrobial activities of the [Ag(La)]2(ClO4)2 were found to be higher than those for the [Ag(La)]2(ClO4)2.1/2(Lb) complex against all microbes. The inhibition zone diameters for the [Ag(La)]2(ClO4)2 complex are 33, 20, 21, 11, 13 and 23 mm against A. fumigatus, C. albicans, S. aureus, B. subtilis, E. coli and P. vulgaris, respectively.
Additionally, we tested the antimicrobial actions of the free Ag(I) ion on the studied microbes under the same experimental conditions. Interestingly, the results were found to be completely different (Table 6). While the [Ag(La)]2(ClO4)2.1/2(Lb) complex and its analogue [Ag(La)]2(ClO4)2 are active against the fungus A. fumigatus, the free Ag(I) is found to be inactive against this microbe. Both complexes also have better antifungal activity against C. albicans than the free Ag(I) ion. In contrast, the [Ag(La)]2(ClO4)2.1/2(Lb) complex has lower antibacterial activity than the free Ag(I) ion against all bacterial strains. On the other hand, the [Ag(La)]2(ClO4)2 complex has better activity against S. aureus and P. vulgaris compared to the free Ag(I). The opposite is true for B. subtilis and E. coli microbes. Such diverse biological actions indicate that the bioactivity of the studied Ag(I) complexes is related not only to the free Ag(I) ion but also depends on the different complex species in solution.
In addition, the Μinimum Inhibitory Concentrations (ΜIC) were found to be the best against A. fumigatus (39 μg/mL) as compared to Ketoconazole (156 μg/mL). It is worth noting that all ΜIC values are higher in this complex as compared to the [Ag(La)]2(ClO4)2 complex when tested against the same microbes. It is clear that the antibacterial and antifungal actions of the [Ag(La)]2(ClO4)2.1/2(Lb) complex are diluted by the presence of the co-crystallized azine Lb, and hence have lower silver content compared to the [Ag(La)]2(ClO4)2 complex. These results shed light on the importance of Ag(I) in antimicrobial activity.

3.4.2. Anticancer and Antioxidant Activities

The cytotoxicity of the Ag(I) complex was examined against colon carcinoma (HCT-116 cell line) [49]. The results of the cytotoxicity experiments are depicted in Table 7. It is clear that the examined Ag(I) complex has good cytotoxic activity against the tested cell line. The %cell viability reaches only 1.87 ± 0.39 at 500 μg/mL. The corresponding value of the [Ag(La)]2(ClO4)2 is 1.46 ± 0.38 μg/mL [39]. In addition, the IC50 values were found to be 25.52 ± 1.28 and 19.72 ± 0.94 µg/mL, respectively. The results indicate that the [Ag(La)]2(ClO4)2 complex has higher cytotoxicity against the colon carcinoma (HCT-116 cell line) than the [Ag(La)]2(ClO4)2.1/2(Lb) complex. In comparison with doxorubicin and cis-platin as positive controls, the IC50 values were determined to be 0.49 ± 0.07 and 5.35 ± 0.49 µg/mL under the same experimental conditions. Hence, the results of the Ag(I) complexes are relatively good compared to these controls.
In addition, the DPPH method [47] was used to determine the antioxidant activity of the [Ag(La)]2(ClO4)2.1/2(Lb) complex (Table 8). The %DPPH scavenging at 1280 μg/mL are 83.02 and 88.62% for the [Ag(La)]2(ClO4)2.1/2(Lb) and [Ag(La)]2(ClO4)2 complexes, respectively. In addition, the IC50 values are 222.64 ± 6.28 and 121.82 ± 4.23 µg/mL, respectively. Hence, the antioxidant activity of the latter is almost twice that of the former. With respect to the control, ascorbic acid, the IC50 value is 12.3 ± 0.51 µg/mL, indicating weak antioxidant activity of the studied Ag(I) complex.

4. Conclusions

The reaction of 1-(pyridin-2-yl)ethylidene)hydrazono)methyl)phenol (L) with AgClO4 afforded the Ag(I)-azine complex [Ag(La)]2(ClO4)2.1/2(Lb). It comprised a tetra-coordinated Ag(I) with (1E,2E)-1,2-bis(1-(pyridin-2-yl)ethylidene)hydrazine (La) as bridged bidentate chelate. The structure comprised half a molecule of 2,2’-((1E,1’E)-hydrazine-1,2-diylidenebis(methanylylidene)) diphenol (Lb) per one [Ag(La)]2(ClO4)2 formula. The packing comprised alternative hydrogen bonded inorganic and organic layers via C-H…O interactions with the perchlorate ions along the a-direction. The presence of the azines La and Lb confirmed the Ag(I)-mediated hydrolysis of the hydrazone L to the corresponding azines. Based on Hirshfeld analysis, the H…H (38.4%), O…H (17.9%), H…C (13.2%), C…C (9.0%) and N…H (8.9%) interactions contributed significantly to the molecular packing. The [Ag(La)]2(ClO4)2.1/2(Lb) complex has lower antimicrobial, antioxidant and cytotoxic activities than the [Ag(La)]2(ClO4)2 complex. Hence, the co-crystallized azine Lb suppress the bioactivity of the studied complex compared to the [Ag(La)]2(ClO4)2 complex. The best antimicrobial action of [Ag(La)]2(ClO4)2.1/2(Lb) complex is found to be against A. fumigatus (ΜIC = 39 μg/mL).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/sym14112226/s1, Figure S1: FTIR spectra of [Ag(La)]2(ClO4)2.1/2(Lb) complex. Biological activity methods, Materials and instrumental details

Author Contributions

Conceptualization, S.M.S., A.E.A. and H.A.E.; methodology, H.A.E. and S.M.S.; software, S.M.S. and H.A.E.; formal analysis, S.M.S., H.A.E., D.A.-W. and A.B.; investigation, H.A.E. and S.M.S.; resources, H.A.E., A.E.A., A.B. and M.S.A.; writing—original draft preparation, H.A.E., S.M.S., A.E.A., A.B. and D.A.-W.; writing—review and editing, H.A.E., A.E.A., A.B., D.A.-W. and S.M.S.; supervision, S.M.S., A.E.A. and H.A.E. All authors have read and agreed to the published version of the manuscript.

Funding

Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2022R86), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2022R86), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shi, Y.; Lu, Z.; Zheng, L.; Cao, Q.-E. Silver-Driven Coordination Self-Assembly of Tetraphenylethene Stereoisomer: Construct Charming Topologies and Their Mechanochromic Behaviors. Inorg. Chem. 2020, 59, 6508–6517. [Google Scholar] [CrossRef] [PubMed]
  2. Lehn, J.M. Supramolecular chemistry—Scope and perspectives molecules, supermolecules, and molecular devices (Nobel Lecture). Angew. Chem. Int. Ed. Engl. 1988, 27, 89–112. [Google Scholar] [CrossRef]
  3. Garrett, T.M.; Koert, U.; Lehn, J.-M.; Rigault, A.; Meyer, D.; Fischer, J. Self-assembly of silver (I) helicates. J. Chem. Soc. Chem. Commun. 1990, 7, 557–558. [Google Scholar] [CrossRef]
  4. Hung-Low, F.; Klausmeyer, K.K. Silver coordination complexes of 2-(diphenylphosphinomethyl) pyridine and their bipyridine derivatives. Inorg. Chim. Acta 2008, 361, 1298–1310. [Google Scholar] [CrossRef]
  5. Pyykkö, P. Strong closed-shell interactions in inorganic chemistry. Chem. Rev. 1997, 97, 597–636. [Google Scholar] [CrossRef] [PubMed]
  6. Janiak, C. A critical account on π–π stacking in metal complexes with aromatic nitrogen-containing ligands. J. Chem. Soc. Dalton Trans. 2000, 3885–3896. [Google Scholar] [CrossRef]
  7. Patroniak, V.; Lehn, J.-M.; Kubicki, M.; Ciesielski, A.; Wałęsa, M. Chameleonic ligand in self-assembly and synthesis of polymeric manganese(II), and grid-type copper(I) and silver(I) complexes. Polyhedron 2006, 25, 2643–2649. [Google Scholar] [CrossRef]
  8. Nachon, F.; Brazzolotto, X.; Dias, J.; Courageux, C.; Drożdż, W.; Cao, X.-Y.; Stefankiewicz, A.R.; Lehn, J.-M. Grid-type Quaternary Metallosupramolecular Compounds Inhibit Human Cholinesterases through Dynamic Multivalent Interactions. ChemBioChem 2022. [Google Scholar] [CrossRef]
  9. Fromm, K.M. Silver coordination compounds with antimicrobial properties. Appl. Organomet. Chem. 2013, 27, 683–687. [Google Scholar] [CrossRef]
  10. Yaqoob, A.A.; Umar, K.; Ibrahim, M.N.M. Silver nanoparticles: Various methods of synthesis, size affecting factors and their potential applications—A review. Appl. Nanosci. 2020, 10, 1369–1378. [Google Scholar] [CrossRef]
  11. Medici, S.; Peana, M.; Nurchi, V.M.; Zoroddu, M.A. Medical uses of silver: History, myths, and scientific evidence. J. Med. Chem. 2019, 62, 5923–5943. [Google Scholar] [CrossRef] [PubMed]
  12. Liang, X.; Luan, S.; Yin, Z.; He, M.; He, C.; Yin, L.; Zou, Y.; Yuan, Z.; Li, L.; Song, X. Recent advances in the medical use of silver complex. Eur. J. Med. Chem. 2018, 157, 62–80. [Google Scholar] [CrossRef] [PubMed]
  13. Sierra, M.A.; Casarrubios, L.; de la Torre, M.C. Bio-Organometallic Derivatives of Antibacterial Drugs. Chem. A Eur. J. 2019, 25, 7232–7242. [Google Scholar] [CrossRef] [PubMed]
  14. Aguzzi, C.; Sandri, G.; Bonferoni, C.; Cerezo, P.; Rossi, S.; Ferrari, F.; Caramella, C.; Viseras, C. Solid state characterisation of silver sulfadiazine loaded on montmorillonite/chitosan nanocomposite for wound healing. Colloids Surf. B Biointerfaces 2014, 113, 152–157. [Google Scholar] [CrossRef] [PubMed]
  15. Batarseh, K.I. Anomaly and correlation of killing in the therapeutic properties of silver (I) chelation with glutamic and tartaric acids. J. Antimicrob. Chemother. 2004, 54, 546–548. [Google Scholar] [CrossRef]
  16. Chastre, J. Preventing ventilator-associated pneumonia: Could silver-coated endotracheal tubes be the answer? JAMA 2008, 300, 842–844. [Google Scholar] [CrossRef]
  17. Mahmodiyeh, B.; Kamali, A.; Zarinfar, N.; Joushani, M.M. The Effect of Silver-Coated Endotracheal Tube on the Incidence of Ventilator-Induced Pneumonia in Intubated Patients Admitted to the Intensive Care Unit (ICU). Syst. Rev. Pharm. 2021, 12, 2499–2503. [Google Scholar]
  18. Baxter, P.N.; Lehn, J.M.; Fischer, J.; Youinou, M.T. Self-Assembly and Structure of a 3× 3 Inorganic Grid from Nine Silver Ions and Six Ligand Components. Angew. Chem. Int. Ed. Engl. 1994, 33, 2284–2287. [Google Scholar] [CrossRef]
  19. Marquis, A.; Kintzinger, J.P.; Graff, R.; Baxter, P.N.; Lehn, J.M. Mechanistic Features, Cooperativity, and Robustness in the Self-Assembly of Multicomponent Silver (i) Grid-Type Metalloarchitectures. Angew. Chem. 2002, 114, 2884–2888. [Google Scholar] [CrossRef]
  20. Nimia, H.H.; Carvalho, V.F.; Isaac, C.; Souza, F.A.; Gemperli, R.; Paggiaro, A.O. Comparative study of Silver Sulfadiazine with other materials for healing and infection prevention in burns: A systematic review and meta-analysis. Burns 2019, 45, 282–292. [Google Scholar] [CrossRef]
  21. Ahmadian, S.; Ghorbani, M.; Mahmoodzadeh, F. Silver sulfadiazine-loaded electrospun ethyl cellulose/polylactic acid/collagen nanofibrous mats with antibacterial properties for wound healing. Int. J. Biol. Macromol. 2020, 162, 1555–1565. [Google Scholar] [CrossRef] [PubMed]
  22. Aziz, Z.; Hassan, B.A.R. The effects of honey compared to silver sulfadiazine for the treatment of burns: A systematic review of randomized controlled trials. Burns 2017, 43, 50–57. [Google Scholar] [CrossRef] [PubMed]
  23. Ueda, Y.; Miyazaki, M.; Mashima, K.; Takagi, S.; Hara, S.; Kamimura, H.; Jimi, S. The effects of silver sulfadiazine on methicillin-resistant staphylococcus aureus biofilms. Microorganisms 2020, 8, 1551. [Google Scholar] [CrossRef]
  24. Möhler, J.S.; Sim, W.; Blaskovich, M.A.; Cooper, M.A.; Ziora, Z.M. Silver bullets: A new lustre on an old antimicrobial agent. Biotechnol. Adv. 2018, 36, 1391–1411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Banti, C.N.; Hadjikakou, S.K. Anti-proliferative and anti-tumor activity of silver (I) compounds. Metallomics 2013, 5, 569–596. [Google Scholar] [CrossRef] [PubMed]
  26. Bharathi, S.; Mahendiran, D.; Kumar, R.S.; Choi, H.J.; Gajendiran, M.; Kim, K.; Rahiman, A.K. Silver (I) metallodrugs of thiosemicarbazones and naproxen: Biocompatibility, in vitro anti-proliferative activity and in silico interaction studies with EGFR, VEGFR2 and LOX receptors. Toxicol. Res. 2020, 9, 28–44. [Google Scholar] [CrossRef]
  27. Hossain, M.S.; Zakaria, C.M.; Kudrat-E-Zahan, M. Metal complexes as potential antimicrobial agent: A review. Am. J. Heterocycl. Chem. 2018, 4, 1. [Google Scholar] [CrossRef] [Green Version]
  28. Chen, X.; Yang, Q.; Chen, J.; Zhang, P.; Huang, Q.; Zhang, X.; Yang, L.; Xu, D.; Zhao, C.; Wang, X. Inhibition of proteasomal deubiquitinase by silver complex induces apoptosis in non-small cell lung cancer cells. Cell. Physiol. Biochem. 2018, 49, 780–797. [Google Scholar] [CrossRef]
  29. Batten, S.R.; Robson, R. Interpenetrating nets: Ordered, periodic entanglement. Angew. Chem. Int. Ed. 1998, 37, 1460–1494. [Google Scholar] [CrossRef]
  30. Venkataraman, D.; Gardner, G.B.; Lee, S.; Moore, J.S. Zeolite-like behavior of a coordination network. J. Am. Chem. Soc. 1995, 117, 11600–11601. [Google Scholar] [CrossRef]
  31. Kitagawa, S.; Kitaura, R.; Noro, S.i. Functional porous coordination polymers. Angew. Chem. Int. Ed. 2004, 43, 2334–2375. [Google Scholar] [CrossRef] [PubMed]
  32. Kitagawa, S.; Noro, S.-i.; Nakamura, T. Pore surface engineering of microporous coordination polymers. Chem. Commun. 2006, 701–707. [Google Scholar] [CrossRef]
  33. Feazell, R.P.; Carson, C.E.; Klausmeyer, K.K. Variability in the structures of luminescent [2-(aminomethyl) pyridine] silver (i) complexes: Effect of ligand ratio, anion, hydrogen bonding, and π-stacking. Eur. J. Inorg. Chem. 2005, 3287–3297. [Google Scholar] [CrossRef]
  34. Wen, C.; Yin, A.; Dai, W.-L. Recent advances in silver-based heterogeneous catalysts for green chemistry processes. Appl. Catal. B Environ. 2014, 160, 730–741. [Google Scholar] [CrossRef]
  35. Hussain, S.; Aneggi, E.; Goi, D. Catalytic activity of metals in heterogeneous Fenton-like oxidation of wastewater contaminants: A review. Environ. Chem. Lett. 2021, 19, 2405–2424. [Google Scholar] [CrossRef]
  36. Hussain-Khil, N.; Ghorbani-Choghamarani, A.; Mohammadi, M. A new silver coordination polymer based on 4, 6-diamino-2-pyrimidinethiol: Synthesis, characterization and catalytic application in asymmetric Hantzsch synthesis of polyhydroquinolines. Sci. Rep. 2021, 11, 1–15. [Google Scholar]
  37. Lo, V.K.-Y.; Chan, A.O.-Y.; Che, C.-M. Gold and silver catalysis: From organic transformation to bioconjugation. Org. Biomol. Chem. 2015, 13, 6667–6680. [Google Scholar] [CrossRef]
  38. Soliman, S.M.; Albering, J.H.; Barakat, A. Unexpected formation of polymeric silver (I) complexes of azine-type ligand via self-assembly of Ag-salts with isatin oxamohydrazide. R. Soc. Open Sci. 2018, 5, 180434. [Google Scholar] [CrossRef] [Green Version]
  39. Soliman, S.M.; Barakat, A. Self-assembly of azine-based hydrolysis of pyridine and isatin oxamohydrazides with AgNO3; synthesis and structural studies of a novel four coordinated Ag (I)-azine 2D coordination polymer. Inorg. Chim. Acta 2019, 490, 227–234. [Google Scholar] [CrossRef]
  40. Elbadawy, H.A.; Khalil, S.M.; Al-Wahaib, D.; Barakat, A.; Soliman, S.M.; Eldissouky, A. Ag (I)-mediated hydrolysis of hydrazone to azine: Synthesis, X-ray structure, and biological investigations of two new Ag (I)-azine complexes. Appl. Organomet. Chem. 2022, 36, e6757. [Google Scholar] [CrossRef]
  41. Sheldrick, G.M. SHELXT: Integrating space group determination and structure solution. Acta Crystallogr. Sect. A Found. Adv. 2014, 70, C1437. [Google Scholar] [CrossRef]
  42. Sheldrick, G.M. SHELXT–Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  43. El-Faham, A.; Soliman, S.M.; Ghabbour, H.A.; Elnakady, Y.A.; Mohaya, T.A.; Siddiqui, M.R.; Albericio, F. Ultrasonic promoted synthesis of novel s-triazine-Schiff base derivatives; molecular structure, spectroscopic studies and their preliminary anti-proliferative activities. J. Mol. Struct. 2016, 1125, 121–135. [Google Scholar] [CrossRef]
  44. Banerjee, S.; Wu, B.; Lassahn, P.-G.; Janiak, C.; Ghosh, A. Synthesis, structure and bonding of cadmium (II) thiocyanate systems featuring nitrogen based ligands of different denticity. Inorg. Chim. Acta 2005, 358, 535–544. [Google Scholar] [CrossRef]
  45. Spackman, P.R.; Turner, M.J.; McKinnon, J.J.; Wolff, S.K.; Grimwood, D.J.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer: A program for Hirshfeld surface analysis, visualization and quantitative analysis of molecular crystals. J. Appl. Crystallogr. 2021, 54, 1006–1011. [Google Scholar] [CrossRef]
  46. Clsi, C. M100-S25: Performance standards for antimicrobial susceptibility testing. In Twenty-Fifth Informational Supplement; Clinical and Laboratory Standards Institute: Wayne, PA, USA, 2012. [Google Scholar]
  47. Yen, G.C.; Duh, P.D. Scavenging effect of methanolic extracts of peanut hulls on free-radical and active-oxygen species. J. Agric. Food Chem. 1994, 42, 629–632. [Google Scholar] [CrossRef]
  48. Cockerill, F.; Clinical and Laboratory Standards Institute. Performance Standards for Antimicrobial Susceptibility Testing: Twenty-First informational Supplement; Clinical and Laboratory Standards Institute: Wayne, PA, USA, 2012; pp. M100–M121. [Google Scholar]
  49. Mosmann, T. Rapid colorimetric assay for cellular growth and survival: Application to proliferation and cytotoxicity assays. J. Immunol. Methods 1983, 65, 55–63. [Google Scholar] [CrossRef]
  50. Ferraria, A.M.; Carapeto, A.P.; do Rego, A.M.B. X-ray photoelectron spectroscopy: Silver salts revisited. Vacuum 2012, 86, 1988–1991. [Google Scholar] [CrossRef]
  51. Volkov, I.L.; Smirnova, A.; Makarova, A.A.; Reveguk, Z.V.; Ramazanov, R.R.; Usachov, D.Y.; Adamchuk, V.K.; Kononov, A.I. DNA with ionic, atomic, and clustered silver: An XPS study. J. Phys. Chem. B 2017, 121, 2400–2406. [Google Scholar] [CrossRef]
Figure 1. Structure of the organic ligands.
Figure 1. Structure of the organic ligands.
Symmetry 14 02226 g001
Scheme 1. Synthesis of the Ag(I)-azine complex.
Scheme 1. Synthesis of the Ag(I)-azine complex.
Symmetry 14 02226 sch001
Figure 2. XPS analysis showing full scan spectra (a) and core levels Ag3d (b), C1s (c), N1S (d), Cl2p (e) and O1s (f) spectra.
Figure 2. XPS analysis showing full scan spectra (a) and core levels Ag3d (b), C1s (c), N1S (d), Cl2p (e) and O1s (f) spectra.
Symmetry 14 02226 g002
Figure 3. Structure of [Ag(La)]2(ClO4)2.1/2(Lb). For better clarity, in this illustration, the relative position of the [Ag(La)]2(ClO4)2 and Lb units are not the same as in the actual crystal.
Figure 3. Structure of [Ag(La)]2(ClO4)2.1/2(Lb). For better clarity, in this illustration, the relative position of the [Ag(La)]2(ClO4)2 and Lb units are not the same as in the actual crystal.
Symmetry 14 02226 g003
Figure 4. Packing scheme of [Ag(La)]2(ClO4)2.1/2(Lb) complex units along the bc and ab planes.
Figure 4. Packing scheme of [Ag(La)]2(ClO4)2.1/2(Lb) complex units along the bc and ab planes.
Symmetry 14 02226 g004
Figure 5. Hirshfeld surfaces of [Ag(La)]2(ClO4)2.1/2(Lb) complex. A, B and C refer to the Ag…O, O…H and H…H contacts.
Figure 5. Hirshfeld surfaces of [Ag(La)]2(ClO4)2.1/2(Lb) complex. A, B and C refer to the Ag…O, O…H and H…H contacts.
Symmetry 14 02226 g005
Figure 6. Percentages of all contacts in the cationic complex unit [Ag(La)]2.
Figure 6. Percentages of all contacts in the cationic complex unit [Ag(La)]2.
Symmetry 14 02226 g006
Figure 7. The decomposed fingerprint plots in [Ag(La)]2(ClO4)2.1/2(Lb) complex.
Figure 7. The decomposed fingerprint plots in [Ag(La)]2(ClO4)2.1/2(Lb) complex.
Symmetry 14 02226 g007
Table 1. Crystal data of [Ag(La)]2(ClO4)2.1/2(Lb) complex.
Table 1. Crystal data of [Ag(La)]2(ClO4)2.1/2(Lb) complex.
CCDC2207700
Empirical formulaC35H34Ag2Cl2N9O9
Fw1011.35 g/moL
Temp (K)296(2)
λ (Å)1.54178
Cryst systΜonoclinic
Space groupP21/c
a (Å)15.6034(11)
b (Å)14.9707(12)
c (Å)17.3993(12)
β (deg)110.554(3)
V3)3805.6(5)
Z4
ρcalc (Μg/m3)1.765
μ (Μo Kα) (mm−1)10.126
No. reflns.45,309
Unique reflns.6708 [R(int) = 0.0597]
Completeness to θ = 33.3399.50%
GOOF (F2)1.054
Final R indices [I > 2sigma(I)]R1 = 0.0624, wR2 = 0.1938
R indices (all data)R1 = 0.0733, wR2 = 0.2099
Table 2. Binding energies and atomic % of [Ag(La)]2(ClO4)2.1/2Lb components.
Table 2. Binding energies and atomic % of [Ag(La)]2(ClO4)2.1/2Lb components.
NamePeak BEFWHM eVAtomic %
C1s284.661.0220
C1s A285.421.2417.72
C1s B286.691.099.03
C1s C288.440.941.02
O1s532.581.222.48
O1s A533.261.2214.3
N1s399.370.70.36
N1s A400.11.261.78
N1s B401.871.232.3
Ag3d, 5/2368.81.131.6
Ag3d, 3/2374.81.121.08
Cl2p, 3/2208.0415.31
Cl2p, 1/2209.631.163.01
Table 3. The important geometric parameters (Å and °) in [Ag(La)]2(ClO4)2.1/2(Lb) complex.
Table 3. The important geometric parameters (Å and °) in [Ag(La)]2(ClO4)2.1/2(Lb) complex.
BondDistanceBondDistance
Ag1-N22.347(4)Ag2-N112.349(5)
Ag1-N82.356(5)Ag2-N52.358(5)
Ag1-N72.368(5)Ag2-N102.371(5)
Ag1-N12.383(5)Ag2-N42.373(5)
BondsAngleBondsAngle
N2-Ag1-N8107.17(16)N11-Ag2-N5106.50(16)
N2-Ag1-N7155.48(17)N11-Ag2-N1070.22(17)
N8-Ag1-N770.63(17)N5-Ag2-N10157.77(16)
N2-Ag1-N170.35(16)N11-Ag2-N4153.97(17)
N8-Ag1-N1155.55(17)N5-Ag2-N470.82(15)
N7-Ag1-N1121.94(16)N10-Ag2-N4122.02(16)
Table 4. Hydrogen bond parameters (Å, °) in [Ag(La)]2(ClO4)2.1/2(Lb) complex.
Table 4. Hydrogen bond parameters (Å, °) in [Ag(La)]2(ClO4)2.1/2(Lb) complex.
D-H...Ad(D-H)d(H...A)d(D...A)<(DHA)Symm. Code
O1-H1...N130.821.992.692(7)144
O1-H1..O20.822.452.892(11)115
C5-H5...O80.932.393.112(15)1341-x,1-y,2-z
C13-H13A...O40.962.533.290(12)136
C13-H13B...O30.962.533.345(12)1422-x,1-y,2-z
C19-H19...O70.932.553.288(13)1371-x,1-y,2-z
C25-H25...O90.932.483.264(16)1421-x,1-y,2-z
C31-H31...O50.932.443.323(10)1582-x,-y,2-z
C34-H34...O50.932.433.314(9)158x,1/2-y,-1/2+z
C35-H35...O20.932.63.481(11)1582-x,-y,2-z
O1-H1...N130.821.992.692(7)144
O1-H1...O20.822.452.892(11)115
Table 5. All short contacts in [Ag(La)]2(ClO4)2.1/2(Lb) complex.
Table 5. All short contacts in [Ag(La)]2(ClO4)2.1/2(Lb) complex.
ContactDistanceContactDistance
Ag1…O63.163O4…H132.444
Ag2…O73.165O2…H12.387
Ag2…O93.197O2…H352.458
O6…H112.569O2…H352.53
O9…H252.364O5…H312.302
O8…H52.287O5…H342.29
O7…H192.439H5…H28C2.067
O5…H14A2.541H25…H13C2.021
O3…H13B2.437
Table 6. Inhibition zone diameters (mm) and ΜIC values (μg/mL) for [Ag(La)]2(ClO4)2.1/2(Lb) complex a.
Table 6. Inhibition zone diameters (mm) and ΜIC values (μg/mL) for [Ag(La)]2(ClO4)2.1/2(Lb) complex a.
MicrobeAg(I) ComplexFree Ag(I) bControl
A. fumigatus26 (39)- (-)17(156) c
C. albicans16 (1250)10 (2500)20(312) c
S. aureus9 (5000)14 (1250)24(10) d
B. subtilis8 (5000)13 (1250)26(5) d
E. coli10 (2500)15 (625)30(5) d
P. vulgaris11 (2500)20 (625)25(5) d
a Values in parntheses for ΜIC; bAgNO3; c Ketoconazole; d Gentamycin.
Table 7. The cytotoxic activity of [Ag(La)]2(ClO4)2.1/2(Lb) complex against colon carcinoma (HCT-116 cell line).
Table 7. The cytotoxic activity of [Ag(La)]2(ClO4)2.1/2(Lb) complex against colon carcinoma (HCT-116 cell line).
Sample Conc. (µg/mL)Viability %Inhibitory %S.D. (±)
5001.8798.130.39
2504.6595.350.67
1259.8390.171.05
62.521.7478.262.08
31.2538.7261.281.56
15.669.4030.62.13
7.885.0614.941.58
3.992.387.620.64
297.652.350.31
199.230.770.49
010000
Table 8. Antioxidant activity using DPPH scavenging assay for [Ag(La)]2(ClO4)2.1/2(Lb) complex.
Table 8. Antioxidant activity using DPPH scavenging assay for [Ag(La)]2(ClO4)2.1/2(Lb) complex.
Sample Conc. (µg/mL)DPPH Scavenging %S.D. (±)
128083.020.94
64071.841.28
32062.952.03
16041.681.94
8029.360.48
4018.610.75
2010.280.42
105.790.13
53.470.25
2.51.950.31
000
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Altowyan, M.S.; Soliman, S.M.; Al-Wahaib, D.; Barakat, A.; Ali, A.E.; Elbadawy, H.A. Synthesis of a New Ag(I)-Azine Complex via Ag(I)-Mediated Hydrolysis of 2-(((1-(Pyridin-2-yl)ethylidene)hydrazineylidene) Methyl)phenol with AgClO4; X-ray Crystal Structure and Biological Studies. Symmetry 2022, 14, 2226. https://doi.org/10.3390/sym14112226

AMA Style

Altowyan MS, Soliman SM, Al-Wahaib D, Barakat A, Ali AE, Elbadawy HA. Synthesis of a New Ag(I)-Azine Complex via Ag(I)-Mediated Hydrolysis of 2-(((1-(Pyridin-2-yl)ethylidene)hydrazineylidene) Methyl)phenol with AgClO4; X-ray Crystal Structure and Biological Studies. Symmetry. 2022; 14(11):2226. https://doi.org/10.3390/sym14112226

Chicago/Turabian Style

Altowyan, Mezna Saleh, Saied M. Soliman, Dhuha Al-Wahaib, Assem Barakat, Ali Eldissouky Ali, and Hemmat A. Elbadawy. 2022. "Synthesis of a New Ag(I)-Azine Complex via Ag(I)-Mediated Hydrolysis of 2-(((1-(Pyridin-2-yl)ethylidene)hydrazineylidene) Methyl)phenol with AgClO4; X-ray Crystal Structure and Biological Studies" Symmetry 14, no. 11: 2226. https://doi.org/10.3390/sym14112226

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

Article Metrics

Back to TopTop