Next Article in Journal
Dereplication and Quantification of Major Compounds of Convolvulus arvensis L. Extracts and Assessment of Their Effect on LPS-Activated J774 Macrophages
Next Article in Special Issue
Structural Investigation of DHICA Eumelanin Using Density Functional Theory and Classical Molecular Dynamics Simulations
Previous Article in Journal
A Molecular-Wide and Electron Density-Based Approach in Exploring Chemical Reactivity and Explicit Dimethyl Sulfoxide (DMSO) Solvent Molecule Effects in the Proline Catalyzed Aldol Reaction
Previous Article in Special Issue
Site Density Functional Theory and Structural Bioinformatics Analysis of the SARS-CoV Spike Protein and hACE2 Complex
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Characterization, Thermal Analysis, DFT, and Cytotoxicity of Palladium Complexes with Nitrogen-Donor Ligands

1
Department of Chemistry, Faculty of Science, Cairo University, Giza 12613, Egypt
2
Department of Chemistry, College of Science, University of Anbar, Ramadi 30001, Iraq
3
Department of Chemistry and Chemical Technologies, University of Calabria, 87036 Arcavacata di Rende, CS, Italy
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(3), 964; https://doi.org/10.3390/molecules27030964
Submission received: 27 December 2021 / Revised: 21 January 2022 / Accepted: 27 January 2022 / Published: 31 January 2022
(This article belongs to the Special Issue Density Functional Theory in the Age of Chemical Intelligence)

Abstract

:
Three new palladium complexes ([Pd(DABA)Cl2], [Pd(CPDA)Cl2], and [Pd(HZPY)Cl2]) bearing dinitrogen ligands (DABA: 3,4-diaminobenzoic acid; CPDA: 4-chloro–o-phenylenediamine; HZPY: 2-hydraziniopyridine) were synthesized, characterized, and tested against breast cancer (MCF-7), prostate carcinoma cell line (PC3) and liver carcinoma cell line (HEPG2). [Pd(DABA)Cl2] complex exhibited the highest inhibition percentage, lying between 68–71%. The hydrolysis mechanism of each palladium complex, the key step preceding the binding to the biological target, as well as their photophysical properties were explored by means of DFT and TDDFT computations. Results indicate a faster hydrolysis process for the Pd(DABA)Cl2 complex. The computed activation energies for the first and second hydrolysis processes suggest that all the compounds could reach DNA in their monohydrated form.

1. Introduction

Coordination chemistry provides a massive number of metal complexes that can be used in different biological applications, including therapeutic and diagnostic medicine. [1,2,3,4,5,6,7,8]. Cisplatin (cis-dichlorodiammineplatinum(II), cis-Pt(Cl2(NH3)2) represents one of the great success stories in the field of cancer chemotherapy for the treatment of testicular tumors whose biological activity arises from its ability to covalently bind DNA, inhibit transcription and replication, and ultimately cause cell death or apoptosis [9,10]. For over three decades, continuous efforts have been made in order to overcome the drawbacks of cisplatin, which include nephrotoxicity, neurotoxicity, ototoxicity, and myelosuppression, with a primary focus on the development of new derivatives with improved pharmacological properties and fewer side effects [11,12]. Palladium was chosen as an alternative for platinum since they share similar chemical behaviors and the possibility to form square planar complexes. Pd(II) complexes are kinetically and thermodynamically stable and are 105 more reactive than analogues Pt(II) complexes. Nevertheless they display lower antitumor activities, probably due to their high reactivity that does not preserve the complex structure until it reach the DNA targets [13,14,15,16,17]. Pd(II) complexes based on nitrogen ligands attracted increasing attention due to their beneficial pharmacological properties [18]. Changing, modifying, and synthesizing novel/existing ligands can be potential ways of improving the biological activities and reducing drug resistance [18]. Tuning and improving the ligands’ design could also enhance the stability of the Pd(II) complexes through the use of chelating agents such as bidentate ligands [18,19,20,21]. Pd(II) complexes based on nitrogen-based ligands such as pyridine, quinolines, pyrazoles, and 1,10-phenanthroline have shown potential antitumor activities. These ligands stabilize and control the palladium-based species in the biological systems rather than reducing the cis-trans isomerism [22]. In the present article, we present a new series of Pd(II) complexes with potential cytotoxic activities. The palladium complexes were prepared using three bidentate ligands, namely, 3,4-diaminobenzoic acid (DABA), 4-chloro–o-phenylenediamine (CPDA), and 2-hydraziniopyridine hydrochloride (hzpy) and were spectroscopically and thermally characterized (Figure 1). The cytotoxic activities of the palladium complexes were studied against cancer cell lines. Moreover, to give insights on the hydrolysis reaction mechanism and kinetics, DFT studies are also herein presented together with TDDFT exploration of the UV–Vis spectra of each compound, providing characterizations of each transition band.

2. Results

2.1. Characterization of Palladium Complexes

2.1.1. Spectral Data

[Pd(DABA)Cl2] Complex

The IR spectrum of [Pd(DABA)Cl2] is shown in Figure 2.
The stretching vibrations of the amino groups of DABA are found at 3328 and 3209 cm−1 in the spectrum of the ligand (Figure S2a). These bands converge in one broad band shifted to a lower frequency (3213 cm−1) in the spectrum of the complex, suggesting the participation of the NH2 group in coordination. The involvement of the amino groups of the DABA ligand is also supported by the observation of the deformation vibrations of NH2, which are found at 1242 cm−1 (ρt NH2), 1172 cm−1 (ρwNH2), and 771 cm−1 (ρrNH2) (Table 1).
The band at 1624 cm−1 due to υC=O in the spectrum of the ligand (Figure S2a), appears shifted to higher frequencies in the complex spectrum (1720 cm−1). The shift of the υC=O of the carboxylic acid group to a higher wave number can be attributed to the liberation of the C=O from the intermolecular hydrogen bonding. The spectrum of the complex also shows an additional band at 428 cm−1 assigned to the M–N bond [23,24]. The mass spectral data is summarized in Table 2.
The mass spectrum of the [Pd(DABA)Cl2] complex (M. wt. = 329.48) gives the parent peak at m/z = 328 (M+-H), a peak at m/z = 193 (DABA-Cl-4H+), in addition to a peak at 141 assigned for (PdCl). The palladium isotopes also appeared at 110, 109, 108, 107, and 106. The magnetic susceptibility measurement of [Pd(DABA)Cl2] complex confirmed the diamagnetic nature of the complex, suggesting that the complex adopts a square planar structure and the Pd(II) (d8) center has the configuration of eg4 a1g2 b2g2 also confirmed by spectral and magnetic data, which prove that the DABA ligand acts as a bidentate ligand via the two amino groups.

[Pd(CPDA)Cl2] Complex

The IR spectrum of [Pd(CPDA)Cl2] is shown in Figure 3.
The stretching vibrations of the amino groups of CPDA are found at 3402 and 3313 cm−1 (Figure S2b), while they appear as a strong peak shifted to lower frequencies (3213 and 3143 cm−1) in the spectrum of the complex, indicating chelation through the NH2 group. The deformation vibrations were found at higher frequencies, specifically 1265 cm−1 (ρt NH2), 1145 cm−1 (ρwNH2), and 779 cm−1 (ρrNH2), confirming the involvement of the NH2 in the coordination to the Pd(II) center (Table 1). The spectrum of the complex shows an additional band at 428 cm−1, assigned to the M–N bond [23,24]. The mass spectral data is summarized in Table 2. The mass spectrum of [Pd(CPDA)Cl2] complex (M. wt. = 319.91) gives a parent peak (M+) at m/z = 320 and two main peaks at m/z = 180, assigned for (CPDA-Cl2-4H+), and at 142, assigned for PdCl. The palladium isotopes appeared at 106, 107, 108, 109, and 110. The magnetic susceptibility measurement of the [Pd(CPDA)Cl2] complex confirms the diamagnetic nature of the complex and corroborates, together with the spectral and magnetic data, the square planar structure for [Pd(CPDA)Cl2], with a Pd(II) having a d8 configuration (eg4 a1g2 b2g2) and with CPDA acting as a bidentate ligand via the two amino groups.

[Pd(hzpy)Cl2] Complex

The IR spectrum of [Pd(hzpy)Cl2] is shown in Figure 4. The stretching of the amino groups of hzpy found at 3305 and 3259 cm−1 in the spectrum of hzpy (Figure S2c) result in shifts to lower frequencies (3274 and 3159 cm−1) in the spectrum of the complex, confirming the complex formation upon the coordination of the amino group and pyridine nitrogen to the Pd ion. Another indicator of the coordination of the two amino groups to the Pd(II) ion is the shift of the deformation vibrations to higher frequencies at 1284 cm−1 (ρt NH2), 1188 cm−1 (ρwNH2), and 763 cm−1 (ρrNH2) (Table 1). The spectrum of the complex also shows an additional band at 466 cm−1 assigned to the M–N bond [23,24]. The mass spectral data is summarized in Table 2. The mass spectrum of [Pd(CPDA)Cl2] complex (M. wt. = 286.46) shows the parent peak (M+) at m/z = 286, in addition to other major peaks at m/z = 109 assigned to the hzpy ligand. An additional peak, belonging to (Cl2), appears at m/z = 71. The palladium isotopes appear at m/z = 109 and 110. The magnetic susceptibility measurement of the [Pd(hzpy)Cl2] complex confirm the diamagnetic nature of the complex, suggesting a square planar geometry with the d8 configuration eg4 a1g2 b2g2 for the Pd(II) center. Spectral and magnetic data support such geometry for the [Pd(hzpy)Cl2] complex, with the hzpy ligand acting as a bidentate ligand via the NH2 group and nitrogen of the pyridine ring.

2.1.2. 1HNMR Spectra

The 1HNMR spectra prove that the three palladium complexes are diamagnetic. For the [Pd(DABA)Cl2] complex, the aromatic protons of the DABA ligand are found at 7.69–8.33 ppm as multiplet, while the NH2 groups appear at 3.49 ppm. The aromatic protons, in the case of the [Pd(CPDA)Cl2] complex, appear at (7.69–8.33, m, 3H), and the NH2 groups at 4.79 ppm as a singlet peak. For the [Pd(hzpy)Cl2] complex, the aromatic protons of the hzpy ligand appear as multiplet (7.69–8.33, 3H), and the NH and NH2 group at 6.69 also appear as multiplet. There was a downfield shift in the peaks’ positions upon the formation of the three palladium complexes, confirming that the chelation process occurs via the NH2 groups [25].

2.2. Thermal Studies

The thermal stabilities of the palladium complexes were investigated through TGA and DTA graphs. The temperature ranges at which thermal decomposition was observed, along with the corresponding mass losses, are given in Table 3. Table 4 gives the decomposition temperature ranges, the DTG peak temperature, the correlation coefficients of the Arrhenius plots, as well as the thermodynamic parameters.

2.2.1. Thermal Analysis of the [Pd(DABA)Cl2] Complex

The thermogravimetric plot (Figure 5) shows that [Pd(DABA)Cl2] decomposes in two steps, as reported in Scheme 1. The weight losses were observed in the range 477–1201 K. The first decomposition peak (477–624 K) associated with the loss of (59.8%, calc. 57.1) is consistent with the elimination of DABA and 1/2 Cl2 species (m/z = 192). The second decomposition peak (708–1201 K), with a mass loss of (9.0%, calc. 10.7), is due to the loss of 1/2 Cl2 species, leaving Pd (31.2%, calc. 32.5%) as the metallic residue (m/z = 107.8).

2.2.2. Thermal analysis of the [Pd(CPDA)Cl2] Complex

The thermogravimetric plot of [Pd(CPDA)Cl2] (Figure 6) shows two decomposition steps, as reported in Scheme 2. The weight losses were observed in the temperature range 550–682 K. The first decomposition peak (550–682 K), with a mass loss of (56.4%, calc. 57.6%) is due to the loss of CPDA and 1/2 Cl2 species (m/z = 180). The second decomposition step (1057–1224 K), accompanied by the loss of (11.9%, calc. 11.09), could be attributed to the loss of 1/2 Cl2 species, leaving Pd metal (31.7%, calc. 33.06% as the metallic residue (m/z = 106.5).

2.2.3. Thermal analysis of the [Pd(hzpy)Cl2] Complex

Even the thermogravimetric plot of [Pd(hzpy)Cl2] (Figure 7) shows two decomposition steps (483–617 K), as reported in Scheme 3. The first decomposition peak (483–617 K) is linked to the loss of (49.4%, calc. 50.5%) and is consistent with the elimination of hzpy (m/z = 109). The second decomposition peak (620–1179 K) is connected to the loss of (22.36%, calc. 23.8%), and could be attributed to the loss of Cl2, leaving Pd metal (39.2%, calc. 38.24%) as the metallic residue (m/z = 110).
The thermodynamic parameters of the decomposition were calculated using Horowitz–Metzger and Coats–Redfern equations [26,27,28,29,30]. The correlation coefficients of the Arrhenius plots were found in the range from 0.62–0.98, indicating good fitness of the linear function. The three complexes are thermally stable with overall activation energy of 107, 573, and 580 kJ mol–1 for [Pd(DABA)Cl2], [Pd(CPDA)Cl2], and [Pd(hzpy)Cl2], respectively.

2.3. Theoretical Studies

The optimized structures of the investigated compounds are displayed in Figure 8. The structural parameters of main interest, such as bond lengths and angles surrounding the metal center, are listed in Table S1, following the atom numbering defined therein.
An inspection of the computed parameters reveals that the Pd–N and Pd–Cl bond lengths are within the regular range reported in the literature [31], and the Pd metal center adopts a quite regular square planar conformation, with angles ranging between 90° and 95° between Pd and the chloride ions, and slightly smaller between Pd and the nitrogens of the bidentate ligands (81°–83°) (See Table S1).
Frontier molecular orbitals explorations (HOMO−1, HOMO, LUMO, LUMO+1) allow a better characterization of the ground-state electronic structures of these complexes. The energy diagram of the frontier MOs, with the indication of the HOMO–LUMO energy gaps and molecular orbitals plots, are reported in Figure 9. The main orbital composition (%) for each frontier orbital is reported in Figure S3.
A significant contribution of the metal characterizes the HOMO-1 orbitals of all the investigated compounds, being the Pd–d contribution comprised by 41 and 48%. The involvement of chloride ligands is also recognizable from the plots. The HOMO orbital is clearly metal-based for Pd(DABA)Cl2 and Pd(CPDA)Cl2 complexes, with an average Pd–d contribution of 46%. Interestingly, in the case of bidentate hzpy ligand, HOMO is clearly N–N ligand-based, with a consequent increase in energy of the orbital, and a consequent decrease of the HOMO–LUMO gap, as can be observed in Figure 8. Indeed, the LUMO orbital is predominantly based on the N–N ligands, with a degree of Pd–d mixing character similar for all the investigated compounds, and no changes in energy emerge with the choosing of the bidentate ligand, remaining quite constant for all the Pd compounds considered. Concerning the LUMO+1, an analysis of the orbital composition reveals that it is almost exclusively based on the bidentate ligand, and the metal contribution has almost vanished.
The inspection of the computed UV–Vis spectra of the investigated compounds (Figure S4) reveals the most intense absorption bands below 300 nm, which can be characterized as N–N ligand-centered for Pd complexes with DABA and CPDA, and assigned to a combination of mixed LC and LMCT characters for the Pd(hzpy)Cl2 complex, as shown by NTOs reported in Figure S5.
The experimental peaks recorded above 300 nm are quite well reproduced by computations. In particular, weak transitions at 329 nm, 331 nm, and 334 nm are found in the computed spectra for Pd(DABA)Cl2, Pd(CPDA)Cl2, and Pd(hzpy)Cl2, respectively (vs 329 nm, 331 nm, and 334 nm, experimentally determined, Figure S2), and they have a significant LMCT contribution, as suggested by the NTOs analysis reported in Figure 10.
The leaving electron comes from π orbital, whose largest contribution comes from the chloride ligands to a delocalized orbital over the metal center and the ligands. The very weak absorptions around 450 nm can be assigned to metal d–d and LLCT transitions.
As the analogous Pt(II) compounds, these new complexes display molecular electrostatic potentials (MEP, Figure 11) with the main negative region located on the chloride ions, which make them good leaving groups.
Indeed, it is well known that the mechanism of action of cisplatin-like compounds starts with the activation, by hydrolysis, of the metal complexes generating reactive aquated species able to subsequently interact with DNA, ultimately leading to apoptosis. The chloride ligands’ replacement occurs once the complex enters the cell nucleus, due to the lower chlorine concentration compared with water. The exploration of the chloride–aqua substitution reactions is, therefore, essential to provide useful insights on the kinetics of the process. DFT studies have proven to be accurate in providing both mechanistic details as well as potential energy surfaces for Pt(II) and even Pd(II) compounds [32,33,34,35,36,37,38,39,40,41].
Aiming at providing information on this process, even for our new Pd(II) complexes, the hydrolysis reaction was explored at the DFT-level of theory. The obtained potential energy surface (PES) for all the complexes, together with snapshots of the relevant distances involved in the transition states, are reported in Figure 12. All the stationary points located along the PES are depicted in Figure S6, and their Cartesian coordinates are provided in Table S2. Inspection of the mechanism confirms that the reaction paths follow a second-order nucleophilic substitution (SN2) reaction, with each step of the reaction proceeding via an associative mechanism through a penta-coordinated trigonal bipyramid transition state, in analogy with the established hydrolysis paths of similar compounds.
For each considered complex, the reaction starts with the catalytic water molecule located in the Pd(II) second-coordination sphere (R). The two extra water molecules assist the process, creating a hydrogen-bonding network involving –NH2 or –Cl groups of DABA, CPDA, or hzpy ligands. The first transition state (TS1) is characterized by the incoming water molecule at distances comprised between 2.39 Å and 2.47 Å, and leaving the chloride at an average distance of 2.76 Å from the metal center (Figure 12). The latter arranges in a trigonal bipyramidal geometry with water molecules and restores its square-planar geometry in the first aqua-complex (INT1). A second water molecule approaches the Pd(II) center in the second associative transition state (TS2), which represents the rate-determining step (RDS) for all the investigated complexes, ultimately leading to the diaqua complexes (P). From our data emerges that the faster hydrolysis occurs for the Pd(DABA)Cl2 complex, with the latter requiring an activation energy to be overcome equal to 15.3 kcal/mol. The first chloride–water substitution requires slightly higher energies for the other two considered complexes; however, this is still very feasible. As previously mentioned, the activation barrier for the second hydrolysis process appears higher, compared to the first one, for all the complexes. Even in the second step, the Pd(DABA)Cl2 complex shows the lowest activation barrier, while the rate-determining step is considerably higher in energy for compounds bearing the hzpy bidentate ligand, being the difference between the two significant barriers. Accordingly, results lead us to suggest that our Pd(II) complexes could mainly interact with DNA in their monohydrated form. The small differences between the two computed activation barriers for Pd(DABA)Cl2 and Pd(CPDA)Cl2, however, also suggest that the formation of diaqua complexes is still a plausible mechanism of action. On the contrary, the high RDS found for Pd(hzpy)Cl2 lead us to exclude for it the possibility of reaching the biological target while fully hydrated.

2.4. The Antitumor Activities of the Pd Complexes

The antitumor activities of the three palladium complexes were studied in vitro versus three cell lines: breast cancer (MCF-7), prostate carcinoma cell line (PC3) and liver carcinoma cell line (HEPG2). The inhibition and survival percentages are tabulated in (Table 5).
The [Pd(DABA)Cl2] complex exhibited the highest inhibition percentage, lying between 68–71%, as reported in Table 5—much closer to that of the reference Vinblastine sulfate, whose measured inhibition % lies between 82–88%. Although smaller, [Pd(DABA)Cl2] shows promising activity that deserves further investigation. On the contrary, compared to the other two complexes, [Pd(hzpy)Cl2] showed the weakest performance, especially against PC3 and MCF7 cell lines (Table 5). The weak activities of the [Pd(hzpy)Cl2] complex could be linked with the higher activation energy found to undergo full hydrolysis, whose possibility cannot be excluded, instead, for the other two compounds, as suggested by the DFT mechanistic studies and free energy profiles. The inhibition values do not have a distinct pattern among the three complexes to enable the structure–activity correlation. Further study may be also conducted in the future to study the cytotoxicity of samples against normal cells.

3. Materials and Methods

3.1. Materials

All the chemicals and solvents used in the current work were of high purity and were used without further purification. Sodium tetrachloro palladate, 4-chloro–o-phenylenediamine (CPDA), 3,4-diaminobenzoic acid (DABA) and 2-hydraziniopyridine hydrochloride (hzpy) were purchased from Sigma–Aldrich (Saint Louis, MO, USA).

3.2. Physical Measurements

The IR spectra of the three complexes were recorded via FT-IR–460 plus (JASCO, Tokyo, Japan). The absorption spectra were recorded using an Optizen spectrophotometer using quartz cell (K Lab, Gunpo-si, Korea). A Varian-Oxford Mercury VX-300 NMR (300 MHz) spectrometer (Agilent, Santa Clara, CA, USA) was used for recording the 1H NMR spectra in DMSO-d6. Fast atom bombardment mass spectroscopy (FAB-MS) measurements of the Pd(II) complexes were performed via a JMS-AX 500 spectrometer (JEOL, Tokyo, Japan). Thermal analyses of the palladium complexes were performed by using a thermo gravimetric analyzer TGA-50H (Shimadzu, Kyoto, Japan) under a continuous flow of nitrogen (20 mL/min) and heating rate of 10 °C/min, over starting from ambient temperature up to 1000 °C.

3.3. Computational Details

All the computations herein presented were performed at DFT- and TDDFT- [42] levels of theory, as implemented in the GAUSSIAN 16 program package. [43] Ground-state optimizations were carried out in water without constraints, by means of the B3LYP exchange–correlation functional (XC) in conjunction with the 6–31+G(d,p) basis and the quasirelativistic Stuttgart–Dresden pseudopotential [44] to describe Pd metal centers. The integral equation formalism polarizable continuum model (IEFPCM) [45,46], setting a dielectric constant equal to ε = 80, was used to simulate the effect of the environment in conjunction with the CPCM polarizable conductor calculation model [47]. Absorption spectra were obtained in water for the singlet ground-state equilibrium structures, using the same basis set and XC functional as for the optimizations. To better describe the kinetic of the hydrolysis process of the Pd(II) compounds, in addition to the implicit solvent, two extra explicit water molecules beyond the catalytic one were added in the system during the exploration of the hydrolysis mechanism. The resulting model was then optimized without constraints during the search of the intermediates and the transition states located along the hydrolysis free energy profiles. Vibrational frequency analysis was carried out at the same level of theory to confirm the nature of all the stationary points. Imaginary frequency values are reported in the SI for each transition state found. The use of implicit solvent and explicit water molecules allow for the proper solvation of the leaving chloride, better accounting for the hydrogen-bonding network during the substitution process. Such a model was previously adopted for the study of similar reactions [36,37,38,39,40,41]. The total free energy in solution, including all non-electrostatic terms (solute–solvent dispersion interaction energy, solute–solvent repulsion interaction energy, and solute cavitation energy), were obtained through single-point calculations with the larger basis set 6–31++G(2df,2pd), and were used to draw the relative free energy profiles of the hydrolysis mechanism. The used protocol allows a direct comparison with results previously obtained for analogous Pt(II) compounds [36,37,38,39,40,41].

3.4. Antitumor Activity

The cytotoxicity testing was performed by the assessing of the cellular growth and survival by using a rapid colorimetric assay method at the Tissue Culture Unit at the Regional Centre for Mycology and Biotechnology (RCMB), Al-Azhar University, Cairo, Egypt [48,49]. Three different cancer cell lines were used to investigate the antitumor activities of the palladium complexes, as well as of the reference Vinblastine sulfate, including human breast cancer (MCF-7), prostate carcinoma cell line (PC3), and liver carcinoma cell line (HEPG2). All cell lines were obtained from VACSERA Tissue Culture Unit. The cells were propagated in Dulbecco’s modified Eagle’s medium (DMEM) or RPMI-1640, depending on the type of cell line, supplemented with 10% heat-inactivated fetal bovine serum, 1% L-glutamine, HEPES buffer, and 50 µg/mL gentamycin. All cells were maintained at 37 °C in a humidified atmosphere with 5% CO2, and were sub-cultured two times a week during experimentation [48,49]. The concentration of the complexes used in the screening was 100 mg/well. Vinblastine sulfate (100 mg/well) was used by RCMB as a reference, and was treated the same way as the Pd(II) complexes.

3.5. Synthesis of the Palladium Complexes

3.5.1. Synthesis of [Pd(DABA)Cl2] Complex

The [Pd(DABA)Cl2] complex was prepared by mixing Na2PdCl4 (0.294 g, 1.0 mmol) with DABA ligand (0.152 g, 1.0 mmol) in 20 mL of ethanol. The mixture was adjusted to pH 3.5 and stirred for 4 h. The solid of each complex was filtered off and washed thoroughly with ethanol, followed by diethyl ether; the solid complex was dried under vacuum and subjected to analysis. Yield: (0.237 g, 72%). Found: C, 25.35; H, 2.23; N, 7.56. Anal. Calc. for C7H8N2O2PdCl2: C, 25.52; H, 2.45; N, 8.50. FT-IR (KBr, cm−1) υ: 3213 (s), 1720 (s), 1272 (m), 1141 (s), 771 (s), and 428 (m). 1H-NMR (δ, ppm) (DMSO-d6): aromatic protons of DABA (7.69–8.33, 3H), NH2 groups (3.49 ppm, s, 2H). MS (FAB-MS): m/z = 328 (M+-H). UV–Vis: 315, 330, 340, 380, 690, and 715 nm. UV–Vis spectra are reported in Figure S1.

3.5.2. Synthesis of [Pd(CPDA)Cl2] Complex

The [Pd(CPDA)Cl2] complex was prepared by mixing Na2PdCl4 (0.294 g, 1.0 mmol) with CPDA ligand (0.142 g, 1.0 mmol) in 20 mL of ethanol. The mixture was adjusted to pH 3.5 and stirred for 4 h. The solid of each complex was filtered off and washed thoroughly with ethanol, followed by diethyl ether; the solid complex was dried under vacuum and subjected to analysis. Yield: 0.236 g (74%). Found: C, 21.89; H, 2.54; N, 7.98. Anal. Calc. for C6H7N2PdCl3: C, 22.53; H, 2.21; N, 8.76. FT-IR (KBr, cm−1) υ: 3213 (s), 3243 (s), 1265 (m), 1145 (s), 779 (s), and 428 (m). 1H-NMR (δ, ppm) (DMSO-d6): aromatic protons of CPDA ligand (7.69–8.33, 3H), NH2 groups (4.79 ppm, s, 2H). MS (FAB-MS): m/z = 320 (M+). UV–Vis: 315, 365, 380, 560, 685, and 715 nm.

3.5.3. Synthesis of [Pd(hzpy)Cl2] Complex

The [Pd(hzpy)Cl2] complex was prepared by mixing Na2PdCl4 (0.294 g, 1.0 mmol) with hzpy ligand (0.182 g, 1.0 mmol) in 20 mL of ethanol. The mixture was adjusted to pH 3.5 and stirred for 4 h. The solid of each complex was filtered off and washed thoroughly with ethanol, followed by diethyl ether; the solid complex was dried under vacuum and subjected to analysis. Yield: (0.21 g, 73%). Found: C, 20.32; H, 2.67; N, 13.58. Anal. Calc. for C5H7N3PdCl2: C, 20.96; H, 2.46; N, 14.67. FT-IR (KBr, cm−1) υ: 3274 (s), 3159 (s), 1284 (m), 1188 (s), 763 (s), and 466 (m). 1H-NMR (δ, ppm) (DMSO-d6): aromatic protons of hzpy ligand (7.69–8.33, 3H), NH and NH2 group (6.69, m, 3H). MS (FAB-MS): m/z = 286 (M+). UV–Vis: 310, 370, and 515 nm.

4. Conclusions

A joint experimental and theoretical study was carried out on three novel palladium complexes with potential activities as anticancer agents. The compounds were synthesized, spectroscopically and thermally characterized, their hydrolysis mechanism as well as their photophysical properties were explored by means of DFT and TDDFT calculations, and their cytotoxic effects were tested against breast cancer (MCF-7), prostate carcinoma cell line (PC3) and liver carcinoma cell line (HEPG2). The [Pd(DABA)Cl2] complex exhibited the highest inhibition percentage, lying between 68–71%. In addition to elucidating the steps of the SN2 mechanism employed by these complexes to hydrolyze, the outcomes of the DFT exploration also indicate a faster chloride–water substitution in the case of the [Pd(DABA)Cl2] complex, that could be linked with the superior performances of that compound. For all the complexes, the rate determining step is represented by the second hydrolysis process. In any case, while the significant higher activation energy found for the second chloride–water substitution mechanism computed for Pd(hzpy)Cl2 leads to exclude the possibility for such complex to undergo full hydrolysis, the same possibility appears more plausible for the other two compounds for which the difference between the two activation energies is smaller and for which the formation of diaqua complexes could be still a feasible process.

Supplementary Materials

The following are available online, Figure S1: UV–Vis spectra; Figure S2: Spectral data of the N,N diamino ligands; Figure S3: HOMO-1, HOMO, LUMO, LUMO+1 MOs plots with the indication of the main orbital composition (%) and IR spectra; Figure S4: Computed Spectra in DMF at the B3LYP/6-31+G(d,p)/SDD-level of theory, for all the Pd complexes; Figure S5: Occupied (NTOo) and virtual (NTOv) Natural Transition Orbitals of the main electronic transitions quoted as A1–A4, as reported in Figure S4; Figure S6: Optimized structures for the first and second hydrolysis processes; Table S1: Selected bond lengths and angles; Table S2: Cartesian coordinates of all the intermediates and transition states located along the hydrolysis free energy profiles.

Author Contributions

S.R.M., M.A.A. and F.A.A. were in charge of investigation and data curation; M.A.A. also carried out preliminary DFT calculations; M.E.A. supervised and refined the computational part and explored the hydrolysis pathways along with conceptualization and writing; A.A.S. supervised the experimental part along with conceptualization and writing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

M.E.A. acknowledges the CINECA award, under the ISCRA initiative, for the availability of high-performance computing resources and support.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Sun, Y.-G.; Sun, D.; Yu, W.; Zhu, M.-C.; Ding, F.; Liu, Y.-N.; Gao, E.-J.; Wang, S.-J.; Xiong, G.; Dragutan, I.; et al. Synthesis, characterization, interaction with DNA and cytotoxicity of Pd(ii) and Pt(ii) complexes containing pyridine carboxylic acid ligands. Dalton Trans. 2013, 42, 3957–3967. [Google Scholar] [CrossRef] [PubMed]
  2. Jain, M.L.; Bruice, P.Y.; Szabó, I.E.; Bruice, T.C. Incorporation of Positively Charged Linkages into DNA and RNA Backbones: A Novel Strategy for Antigene and Antisense Agents. Chem. Rev. 2012, 112, 1284–1309. [Google Scholar] [CrossRef] [PubMed]
  3. Eryazici, I.; Moorefield, C.N.; Newkome, G.R. Square-Planar Pd(II), Pt(II), and Au(III) Terpyridine Complexes: Their Syntheses, Physical Properties, Supramolecular Constructs, and Biomedical Activities. Chem. Rev. 2008, 108, 1834–1895. [Google Scholar] [CrossRef] [PubMed]
  4. Sigel, H. Metal Ions in Biological Systems: Volume 42: Metal Complexes in Tumor Diagnosis and as Anticancer Agents, 1st ed.; CRC Press: Boca Raton, FL, USA, 2004; pp. 143–177. [Google Scholar]
  5. Soliman, A.A.; Amin, M.A.; Sayed, A.M.; Abou-Hussein, A.A.A.; Linert, W. Cobalt and copper complexes with formamidine ligands: Synthesis, crystal X-ray study, DFT calculations and cytotoxicity. Polyhedron 2019, 161, 213–221. [Google Scholar] [CrossRef]
  6. Roque, J.A., III; Barrett, P.C.; Cole, H.D.; Lifshits, L.M.; Bradner, E.; Shi, G.; Von Dohlen, D.; Kim, S.; Russo, N.; Deep, G.; et al. Os(II) Oligothienyl Complexes as a Hypoxia-Active Photosensitizer Class for Photodynamic Therapy. Inorg. Chem. 2020, 59, 16341–16360. [Google Scholar] [CrossRef] [PubMed]
  7. Alberto, M.E.; Russo, N.; Adamo, C. Synergistic Effects of Metals in a Promising RuIIPtII Assembly for a Combined Anticancer Approach: Theoretical Exploration of the Photophysical Properties. Chem. Eur. J. 2016, 22, 9162–9168. [Google Scholar] [CrossRef] [Green Version]
  8. Roque, J.A., III; Barrett, P.C.; Cole, H.D.; Lifshits, L.M.; Shi, G.; Monro, S.; Von Dohlen, D.; Kim, S.; Russo, N.; Deep, G.; et al. Breaking the Barrier: An Osmium Photosensitizer with Unprecedented Hypoxic Phototoxicity for Real World Photodynamic Therapy. Chem. Sci. 2020, 11, 9784–9806. [Google Scholar] [CrossRef]
  9. Rosenberg, B.; Van Camp, L.; Krigas, T. Inhibition of Cell Division in Escherichia coli by Electrolysis Products from a Platinum Electrode. Nature 1965, 205, 698–699. [Google Scholar] [CrossRef]
  10. Jamieson, E.R.; Lippard, S.J. Structure, Recognition, and Processing of Cisplatin−DNA Adducts. Chem. Rev. 1999, 99, 2467–2498. [Google Scholar] [CrossRef]
  11. Johnstone, T.C.; Suntharalingam, K.; Lippard, S.J. The Next Generation of Platinum Drugs: Targeted Pt(II) Agents, Nanoparticle Delivery, and Pt(IV) Prodrugs. Chem. Rev. 2016, 116, 3436–3486. [Google Scholar] [CrossRef] [Green Version]
  12. Bruijnincx, P.C.; Sadler, P.J. New trends for metal complexes with anticancer activity. Curr. Opin. Chem. Biol. 2008, 12, 197–206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Carneiro, T.J.; Martins, A.S.; Marques, M.P.M.; Gil, A.M. Metabolic Aspects of Palladium(II) Potential Anti-Cancer Drugs. Front. Oncol. 2020, 10, 2218. [Google Scholar] [CrossRef] [PubMed]
  14. Abu-Surrah, A.S.; Kettunen, M. Platinum group antitumor chemistry: Design and development of new anticancer drugs complementary to cisplatin. Curr. Med. Chem. 2006, 13, 1337–1357. [Google Scholar] [CrossRef] [PubMed]
  15. Abu-Safieh, K.A.; Abu-Surrah, A.S.; Tabba, H.D.; AlMasri, H.A.; Bawadi, R.M.; Boudjelal, F.M.; Tahtamouni, L.H. Novel Palladium(II) and Platinum(II) Complexes with a Fluoropiperazinyl Based Ligand Exhibiting High Cytotoxicity and Anticancer Activity In Vitro. J. Chem. 2016, 2016, 7508724. [Google Scholar] [CrossRef] [Green Version]
  16. Lüköová, A.; Drweesh, E.A.; Volarevic, V.; Miloradovic, D.; Simovic Markovic, B.; Smolková, R.; Samoľová, E.; Kuchár, J.; Vilková, M.; Potočňák, I. Low-dimensional compounds containing bioactive ligands. Part XIII: Square planar anti-cancer Pd(II) complexes with halogenderivatives of 8-quinolinol and dimethylamine. Polyhedron 2020, 184, 114535. [Google Scholar] [CrossRef]
  17. Omondi, R.O.; Bellam, R.; Ojwach, S.O.; Jaganyi, D.; Fatokun, A.A. Palladium(II) complexes of tridentate bis(benzazole) ligands: Structural, substitution kinetics, DNA interactions and cytotoxicity studies. J. Inorg. Biochem. 2020, 210, 111156. [Google Scholar] [CrossRef] [PubMed]
  18. Małecki, J.G.; Maroń, A. Spectroscopic, structure and DFT studies of palladium(II) complexes with pyridine-type ligands. Transit. Met. Chem. 2011, 36, 297–305. [Google Scholar] [CrossRef] [Green Version]
  19. Najajreh, Y.; Perez, J.M.; Navarro-Ranninger, C.; Gibson, D. Novel Soluble Cationic trans-Diaminedichloroplatinum(II) Complexes that Are Active against Cisplatin Resistant Ovarian Cancer Cell Lines. J. Med. Chem. 2002, 45, 5189–5195. [Google Scholar] [CrossRef]
  20. Garoufis, A.; Hadjikakou, S.K.; Hadjiliadis, N. 46Pd The Use of Palladium Complexes in Medicine. In Metallotherapeutic Drugs and Metal-Based Diagnostic Agents; Wiley & Sons: Chichester, UK, 2005; pp. 399–419. [Google Scholar]
  21. Shoukry, A.; Rau, T.; Shoukry, M.; van Eldik, R. Kinetics and mechanisms of the ligand substitution reactions of bis(amine)(cyclobutane-1,1-dicarboxylato)palladium(II). J. Chem. Soc. Dalton Trans. 1998, 18, 3105–3112. [Google Scholar] [CrossRef]
  22. González, M.L.; Tercero, J.M.; Matilla, A.; Niclós-Gutiérrez, J.; Fernández, M.T.; López, M.C.; Alonso, C.; González, S. cis-Dichloro(α,ω-diamino carboxylate ethyl ester)palladium(II) as Palladium(II) versus Platinum(II) Model Anticancer Drugs:  Synthesis, Solution Equilibria of Their Aqua, Hydroxo, and/or Chloro Species, and in Vitro/in Vivo DNA-Binding Properties. Inorg. Chem. 1997, 36, 1806–1812. [Google Scholar] [CrossRef]
  23. Soliman, A.A.; Khattab, M.M.; Linert, W. Kinetic and characterization studies of iron(II) and iron(III) complex formation reactions with hydrazinopyridine. J. Coord. Chem. 2008, 61, 2017–2031. [Google Scholar] [CrossRef]
  24. Nakamoto, K. Applications in Inorganic Chemistry. In Infrared and Raman Spectra of Inorganic and Coordination Compounds; John Wiley & Sons: Hoboken, NJ, USA, 2009; pp. 149–354. [Google Scholar]
  25. Samota, M.K.; Seth, G. Synthesis, characterization, and antimicrobial activi25 of palladium(II) and platinum(II) complexes with 2-substituted benzoxazole ligands. Heteroat. Chem. 2010, 21, 44–50. [Google Scholar]
  26. Horowitz, H.H.; Metzger, G. A New Analysis of Thermogravimetric Traces. Anal. Chem. 1963, 35, 1464–1468. [Google Scholar] [CrossRef]
  27. Nath, M.; Arora, P. Spectral and Thermal Studies of Cobalt(II), Nickel(II) and Copper(II) Complexes of Schiff Bases Obtained from o-Hydroxyacetophenone and Amino Acids. Synth. React. Inorg. Met.-Org. Chem. 1993, 23, 1523–1546. [Google Scholar] [CrossRef]
  28. Coats, A.W.; Redfern, J.P. Kinetic Parameters from Thermogravimetric Data. Nature 1964, 201, 68–69. [Google Scholar] [CrossRef]
  29. Soliman, A.A.; El-Medani, S.M.; Ali, O.A.M. Thermalstudy of chromium and molybdenum complexes with some nitrogen and nitrogen-oxygendonors ligands. J. Therm. Anal. Calorim. 2006, 83, 385–392. [Google Scholar] [CrossRef]
  30. Soliman, A.A.; Taha, A.; Linert, W. Spectral and thermal study on the adduct formation between square planar nickel(II) chelates and some bidentate ligands. Spectrochim. Acta Part A 2006, 64, 1058–1064. [Google Scholar] [CrossRef]
  31. Dikmen, G.; Hür, D. Palladium (II) complex: Synthesis, spectroscopic studies and DFT calculations. Chem. Phys. Lett. 2019, 716, 49–60. [Google Scholar] [CrossRef]
  32. Alberto, M.E.; Butera, V.; Russo, N. Which One among the Pt-Containing Anticancer Drugs More Easily Forms Monoadducts with G and A DNA Bases? A Comparative Study among Oxaliplatin, Nedaplatin, and Carboplatin. Inorg. Chem. 2011, 50, 6965–6971. [Google Scholar] [CrossRef]
  33. Alberto, M.E.; Cosentino, C.; Russo, N. Hydrolysis mechanism of anticancer Pd(II) complexes with coumarin derivatives: A theoretical investigation. Struct. Chem. 2012, 23, 831–839. [Google Scholar] [CrossRef]
  34. Zhang, Y.; Guo, Z.; You, X.-Z. Hydrolysis Theory for Cisplatin and Its Analogues Based on Density Functional Studies. J. Am. Chem. Soc. 2001, 123, 9378–9387. [Google Scholar] [CrossRef] [PubMed]
  35. Ponte, F.; Alberto, M.E.; De Simone, B.C.; Russo, N.; Sicilia, E. Photophysical Exploration of Dual-Approach PtII–BODIPY Conjugates: Theoretical Insights. Inorg. Chem. 2019, 58, 9882–9889. [Google Scholar] [CrossRef] [PubMed]
  36. Raber, J.; Zhu, C.; Eriksson, L.A. Activation of anti-cancer drug cisplatin—Is the activated complex fully aquated? Mol. Phys. 2004, 102, 2537–2544. [Google Scholar] [CrossRef]
  37. Pavelka, M.; Lucas, M.F.A.; Russo, N. On the Hydrolysis Mechanism of the Second-Generation Anticancer Drug Carboplatin. Chem. Eur. J. 2007, 13, 10108–10116. [Google Scholar] [CrossRef]
  38. Alberto, M.E.; Lucas, M.F.A.; Pavelka, M.; Russo, N. The Second-Generation Anticancer Drug Nedaplatin: A Theoretical Investigation on the Hydrolysis Mechanism. J. Phys. Chem. B 2009, 113, 14473–14479. [Google Scholar] [CrossRef]
  39. Lucas, M.F.A.; Pavelka, M.; Alberto, M.E.; Russo, N. Neutral and Acidic Hydrolysis Reactions of the Third Generation Anticancer Drug Oxaliplatin. J. Phys. Chem. B 2009, 113, 831–838. [Google Scholar] [CrossRef]
  40. Alberto, M.E.; Adamo, C. Synergistic Effects in PtII–Porphyrinoid Dyes as Candidates for a Dual-Action Anticancer Therapy: A Theoretical Exploration. Chem. Eur. J. 2017, 23, 15124–15132. [Google Scholar] [CrossRef] [Green Version]
  41. Alberto, M.E.; Mazzone, G.; Regina, C.; Russo, N.; Sicilia, E. Two-components RuII-Porphirin dyes as promising assemblies for a combined antitumor effect: Theoretical elucidation of the photophysical properties. Dalton Trans. 2020, 49, 12653–12661. [Google Scholar] [CrossRef]
  42. Casida, M.E. Time-Dependent Density Functional Response Theory of Molecular Systems: Theory, Computational Methods, and Functionals. In Theoretical and Computational Chemistry; Seminario, J.M., Ed.; Elsevier: Amsterdam, The Netherlands; New York, NY, USA, 1996; pp. 155–192. [Google Scholar]
  43. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016.
  44. Andrae, D.; Häußermann, U.; Dolg, M.; Stoll, H.; Preuß, H. Energy-adjustedab initio pseudopotentials for the second and third row transition elements. Theor. Chim. Acta 1990, 77, 123–141. [Google Scholar] [CrossRef]
  45. Cossi, M.; Barone, V. Solvent effect on vertical electronic transitions by the polarizable continuum model. J. Chem. Phys. 2000, 112, 2427–2435. [Google Scholar] [CrossRef]
  46. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef]
  47. Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation model. J. Comput. Chem. 2003, 24, 669–681. [Google Scholar] [CrossRef] [PubMed]
  48. 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]
  49. Riyadh, S.M.; Gomha, S.M.; Mahmmoud, E.A.; Elaasser, M.M. ChemInform Abstract: Synthesis and Anticancer Activities of Thiazoles, 1,3-Thiazines, and Thiazolidine Using Chitosan-Grafted-Poly(vinylpyridine) as Basic Catalyst. Heterocycles 2015, 46, 1227–1243. [Google Scholar] [CrossRef]
Figure 1. Dinitrogen ligands DABA, CPDA, and hzpy used in the preparation of palladium complexes.
Figure 1. Dinitrogen ligands DABA, CPDA, and hzpy used in the preparation of palladium complexes.
Molecules 27 00964 g001
Figure 2. Experimental IR spectrum of the [Pd(DABA)Cl2] complex.
Figure 2. Experimental IR spectrum of the [Pd(DABA)Cl2] complex.
Molecules 27 00964 g002
Figure 3. Experimental IR spectrum of the [Pd(CPDA)Cl2] complex.
Figure 3. Experimental IR spectrum of the [Pd(CPDA)Cl2] complex.
Molecules 27 00964 g003
Figure 4. Experimental IR spectrum of the [Pd(hzpy)Cl2] complex.
Figure 4. Experimental IR spectrum of the [Pd(hzpy)Cl2] complex.
Molecules 27 00964 g004
Figure 5. Thermogram of the [Pd(DABA)Cl2] complex (TG: black; DTG: blue).
Figure 5. Thermogram of the [Pd(DABA)Cl2] complex (TG: black; DTG: blue).
Molecules 27 00964 g005
Scheme 1. Thermal decomposition of the [Pd(DABA)Cl2] complex.
Scheme 1. Thermal decomposition of the [Pd(DABA)Cl2] complex.
Molecules 27 00964 sch001
Figure 6. Thermogram of the [Pd(CPDA)Cl2] complex (TG: black; DTG: blue).
Figure 6. Thermogram of the [Pd(CPDA)Cl2] complex (TG: black; DTG: blue).
Molecules 27 00964 g006
Scheme 2. Thermal decomposition of the [Pd(CPDA)Cl2] complex.
Scheme 2. Thermal decomposition of the [Pd(CPDA)Cl2] complex.
Molecules 27 00964 sch002
Figure 7. Thermogram of the [Pd(hzpy)Cl2] complex (TG: black; DTG: blue).
Figure 7. Thermogram of the [Pd(hzpy)Cl2] complex (TG: black; DTG: blue).
Molecules 27 00964 g007
Scheme 3. Thermal decomposition of the [Pd(hzpy)Cl2] complex.
Scheme 3. Thermal decomposition of the [Pd(hzpy)Cl2] complex.
Molecules 27 00964 sch003
Figure 8. Optimized structures of the investigated complexes at the B3LYP/6-31+G(d,p)/SDD-level of theory, in water environment.
Figure 8. Optimized structures of the investigated complexes at the B3LYP/6-31+G(d,p)/SDD-level of theory, in water environment.
Molecules 27 00964 g008
Figure 9. Energy diagram (eV) of the HOMO-1 (H-1), HOMO (H), LUMO (L), and LUMO+1 (L+1) orbitals for all the investigated compounds, also reporting the H–L energy gap (eV) and graphical molecular orbital plots.
Figure 9. Energy diagram (eV) of the HOMO-1 (H-1), HOMO (H), LUMO (L), and LUMO+1 (L+1) orbitals for all the investigated compounds, also reporting the H–L energy gap (eV) and graphical molecular orbital plots.
Molecules 27 00964 g009
Figure 10. Occupied (NTOo) and virtual (NTOv) Natural Transition Orbitals characterizing the electronic transitions at (a) above 300 nm and (b) above 400 nm for Pd complexes.
Figure 10. Occupied (NTOo) and virtual (NTOv) Natural Transition Orbitals characterizing the electronic transitions at (a) above 300 nm and (b) above 400 nm for Pd complexes.
Molecules 27 00964 g010
Figure 11. Molecular electrostatic potentials mapping (MEP) for (a) for Pd(DABA)Cl2, (b) Pd(CPCA)Cl2, (c) Pd(HZPY)Cl2, complexes.
Figure 11. Molecular electrostatic potentials mapping (MEP) for (a) for Pd(DABA)Cl2, (b) Pd(CPCA)Cl2, (c) Pd(HZPY)Cl2, complexes.
Molecules 27 00964 g011
Figure 12. Potential energy profiles for the first and second hydrolysis reaction of the investigated Pd(II) complexes, with snapshots of the first and second transition states (TS1 and TS2), computed in water at the B3LYP/6-31++G(2df,2pd)/SDD-level of theory. The reactants, first intermediate and product of the reaction are referred as R, INT1 and P in the energetic profile, respectively. More details can be found in SI section.
Figure 12. Potential energy profiles for the first and second hydrolysis reaction of the investigated Pd(II) complexes, with snapshots of the first and second transition states (TS1 and TS2), computed in water at the B3LYP/6-31++G(2df,2pd)/SDD-level of theory. The reactants, first intermediate and product of the reaction are referred as R, INT1 and P in the energetic profile, respectively. More details can be found in SI section.
Molecules 27 00964 g012
Table 1. Experimental vibrational frequencies of the Pd complexes.
Table 1. Experimental vibrational frequencies of the Pd complexes.
ComplexObsAssignmentComplexObsAssignmentComplexObsAssignment
[Pd(DABA)Cl2]3213υ(NH2)[Pd(CPDA)Cl2]3213υ(NH2)[Pd(hzpy)Cl2]3274υ(NH2)
1720υ(C=O) 3143υ(NH2) 3159υ(NH2)
1242(ρt NH2) 1265(ρt NH2) 1284(ρt NH2)
1172(ρwNH2) 1145(ρwNH2) 1188(ρwNH2)
771(ρrNH2) 779(ρrNH2) 763(ρrNH2)
428υ(M-N) 428υ(M-N) 466υ(M-N)
Table 2. Important mass peaks of palladium complexes.
Table 2. Important mass peaks of palladium complexes.
ComplexMolar Massm/z Values
[Pd(DABA)Cl2]329.48328, 192, 150, 147, 109, 107, 106
[Pd(CPDA)Cl2]319.91320, 180, 141,110, 109, 106
[Pd(hzpy)Cl2]286.46286, 142, 109, 110, 109, 106, 104
Table 3. Thermo-analytical data of palladium complexes.
Table 3. Thermo-analytical data of palladium complexes.
ComplexesTG Range (K)DTAmax
(K)
Mass Loss Found (calc. %)Assignment of the
Removed Species
Metallic Residue Found (calc.%)
[Pd(DABA)Cl2]477–64263159.8, (57.1)C7H8N2O2, 1/2Cl2Pd
708–12017519.0, (10.7)1/2Cl231.2, (32.5)
[Pd(CPDA)Cl2]550–68257255.21, (55.6)(C6H7N2Cl), 1/2Cl2Pd
1057–1224105612.74, (11.76)1/2 Cl233.06, (33.28)
[Pd(hzpy)Cl2]483–61848851.2, (53.9)C5H7N3+1/2Cl2Pd
620–1179114110.6, (13.2)Cl38.2, (38.4)
Table 4. The thermodynamic parameters of the thermal decomposition of the palladium complex.
Table 4. The thermodynamic parameters of the thermal decomposition of the palladium complex.
ComplexDecomposition Temperature (K)∆E/
KJ mol−1
R2∆S/
J K−1 mol−1
∆H/
KJ mol−1
∆G/
KJ mol−1
[Pd(DABA)Cl2]477–642
708–1201
55
62
0.74
0.62
−170
−228
−51
−52
140
313
107−398103453
[Pd(CPDA)Cl2]550–682
1057–1224
217
356
0.98
0.98
76
38.5
211
346
162
303
573114.5557465
[Pd(hzpy)Cl2]483–617
620–1179
43
537
0.86
0.98
−177
599
39
531
128
153
580422570281
Table 5. Cytotoxicity activity of the palladium complexes.
Table 5. Cytotoxicity activity of the palladium complexes.
ComplexesPC3MCF7HEPG2
Inhibition%Inhibition%Inhibition%
[Pd(DABA)]Cl2717068
[Pd(CPDA)Cl2]565553
[Pd(hzpy)Cl2]444962
Vinblastine Sulfate828188
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Majeed, S.R.; Amin, M.A.; Attaby, F.A.; Alberto, M.E.; Soliman, A.A. Synthesis, Characterization, Thermal Analysis, DFT, and Cytotoxicity of Palladium Complexes with Nitrogen-Donor Ligands. Molecules 2022, 27, 964. https://doi.org/10.3390/molecules27030964

AMA Style

Majeed SR, Amin MA, Attaby FA, Alberto ME, Soliman AA. Synthesis, Characterization, Thermal Analysis, DFT, and Cytotoxicity of Palladium Complexes with Nitrogen-Donor Ligands. Molecules. 2022; 27(3):964. https://doi.org/10.3390/molecules27030964

Chicago/Turabian Style

Majeed, Sattar R., Mina A. Amin, Fawzy A. Attaby, Marta E. Alberto, and Ahmed A. Soliman. 2022. "Synthesis, Characterization, Thermal Analysis, DFT, and Cytotoxicity of Palladium Complexes with Nitrogen-Donor Ligands" Molecules 27, no. 3: 964. https://doi.org/10.3390/molecules27030964

Article Metrics

Back to TopTop