Open Access Article
Yokari Godínez-Loyolaa,
Jesús Gracia-Mora
a,
Iván D. Rojas-Montoya
a,
Luis Felipe Hernández-Ayala
a,
Miguel Reinaa,
Luis Antonio Ortiz-Fradeb,
Luisa Alondra Rascón-Valenzuelac,
Ramón Enrique Robles-Zepedac,
Virginia Gómez-Vidalesd,
María Josefa Bernad-Bernad*a and
Lena Ruiz-Azuara
*a
aFacultad de Química, Universidad Nacional Autónoma de México, Av. Universidad 3000, Circuito Exterior S/N, CU, Ciudad de México, C.P. 04510, Mexico. E-mail: bernadf@comunidad.unam.mx; lenar701@gmail.com
bCentro de Investigación y Desarrollo Tecnológico en Electroquímica, Sanfandila, Querétaro, Mexico
cDepartamento de Ciencias Químico-Biológicas, Universidad de Sonora, Boulevard Luis Encinas y Rosales S/N, Hermosillo, Sonora C.P. 83000, Mexico
dInstituto de Química, Universidad Nacional Autónoma de México, Av. Universidad 3000, Circuito Exterior S/N, CU, Ciudad de México, C.P. 04510, Mexico
First published on 3rd August 2022
Seven new Casiopeinas® were synthesized and properly characterized. These novel compounds have a general formula [Cu(N–N)(Indo)]NO3, where Indo is deprotonated indomethacin and N–N is either bipyridine or phenanthroline with some methyl-substituted derivatives, belonging to the third generation of Casiopeinas®. Spectroscopic characterization suggests a square-based pyramid geometry and voltammetry experiments indicate that the redox potential is strongly dependent on the N–N ligand. All the presented compounds show high cytotoxic efficiency, and most of them exhibit higher efficacy compared to the well-known cisplatin drug and acetylacetonate analogs of the first generation. Computational calculations show that antiproliferative behavior can be directly related to the volume of the molecules. Besides, a chitosan (CS)–polyacrylamide (PNIPAAm) nanogel was synthesized and characterized to examine the encapsulation and release properties of the [Cu(4,7-dimethyl-1,10-phenanthroline)(Indo)]NO3 compound. The results show good encapsulation performance in acidic conditions and a higher kinetic drug release in acidic media than at neutral pH. This result can be described by the Peppas–Sahlin model and indicates a release mechanism predominantly by Fick diffusion.
To overcome the drug resistance and the side effects several approaches have been developed, one of them is the drug design based on the essential trace metals such as copper. Copper is a micronutrient with a fundamental role in several biological processes such as mitochondrial metabolism19 and cellular protection against oxidant species.20 Although copper is essential in healthy cells, an elevated concentration favors angiogenesis, tumor growth and metastasis.21,22 This difference in copper response between the tumoral and normal cell environments is the reason for the evolution of copper complexes as anticancer agents. These complexes show high cytotoxicity and they have proved anticancer performance. Recently, copper complexes have been synthesized with several donor atoms (N, O, S and P principally) ligand-based. Some of them as disulfiram, elesclomol, and thiosemicarbazones based ligands are currently in clinical trials.23–25 These complexes are multitarget and multifunctional agents, and their anticancer performance involves several mechanisms. Copper chelators diminish their endogenous concentration and difficult the growth of malignant cells and angiogenesis. Other systems interact with proteosomes and inhibit their function.26 For instance, a Cu-esclemol complex can interfere in the ferredoxin function and produce cuproptosis, a cell death that occurs independently of apoptosis pathways.27
In the search for new anticancer drugs with fewer side effects, Casiopeinas®, mixed chelate copper complexes, have been synthesized, reported, and patented.28,29 Casiopeinas® are well-known copper compounds with proved potent anticancer activity; their general formulae is [Cu(N–N)(L–L)]n+ (n = 1, 2), where N–N = 1,10-phenanthroline or 2,2′-bipyridine and L–L = secondary ligand being different bidentate chelate, one of them (CasIII-ia: [Cu(44′dmbipy)(acac)]+) is now being tested in clinical trials. The second-generation possess a neutral L–L ligand as ethylenediamine and substituted benzimidazoles and 1,2-dianilines,30,31 and the third generation is characterized by the presence of molecules with O–O donor atoms as secondary ligands with proven biological activity as curcumin and indomethacin.32 The main difference between first and third generation of Casiopeinas® is that the secondary ligand for first generation is any monocharged N–O or O–O donor atoms molecule, and for third generation is a molecule with the same donor atoms, also monocharged, but these molecules present by themselves a proved biological activity. The insertion of this type of secondary ligand may modulate and increase the anticancer activity and the anti-inflammatory activity that would enhance the antitumor one. However, the main mechanism of action that have been obtained for the first generation would not be different for the third generation because the main weak oxidant property and ROS generation and apoptosis induction, through the reduction of CuII to CuI is supposed to be the same. The mechanisms of action have been studied in more than 20 compounds, founding the ROS generation towards induction of apoptosis, interaction with DNA, and their nuclease action.33–35 These compounds represent a viable, attractive, and accesible alternative for cancer treatment, including lung, cervix, and breast cancer. Additionally, Casiopeinas® show low toxicity against normal cells, which suggests high selectivity.36 It has been also demonstrated that Casiopeinas® possess a multitarget cytotoxicity mechanism that converges in the cell apoptosis induction.37–39 In addition, in silico studies suggest that all the three Casiopeinas® generations can act as antiviral agents by the inhibition of the main protease of SARS-CoV-2.40
Indomethacin is a nonsteroidal anti-inflammatory drug (NSAID) that inhibits cyclooxygenases (COX).41 In cancer disease, COX-2, an isoform of COX enzymes, is expressed in numerous solid tumors and their neovasculature.42 Thus, the use of NSAIDs as COX-2 inhibitors emerge as a strategy to limit tumoral growth. In fact, in vitro, and in vivo models show a proliferation rate decrease, an increment in the apoptosis, and an attenuation of metastasis.43 Copper–indomethacin complexes also have been investigated as anticancer agents in mammary cells.44 These complexes show cytotoxic activity in micromolar order and a COX-2 inhibitors with a better performance than indomethacin.45 From above, indomethacin complexes represent good candidates to possess relevant antiproliferative properties.
Although the investigation of therapeutic systems has been a prolific research field in the last decades, drug delivery on specific target cells to treat cancer has remained an elusive and complex topic.46 One interesting strategy to overcome this obstacle is drug encapsulation with polymers or nanoparticles as vehicles.47–51 Encapsulation involves the modification of different physicochemical and biochemical features including solubility, stability, and Burst release, resulting in the prevention of drug degradation, increased therapeutic efficacy and decreased side effects.52–56 On that matter, hydrogels present an available and versatile alternative to pharmaceutical applications, especially since they do not dissolve in physiological temperature and pH.57–59 Hydrogels are generally made of low-cost materials, possess tunable three-dimension structures, remarkable mechanic properties, high water content, and great biocompatibility.60 Hydrogels made of chitosan, a biopolymer of acetylated/deacetylated N-acetyl-β-D-glucosamine, have received great attention for their inherent biological properties and considerable swelling in an aqueous medium.61,62 This is a valuable property since these systems can respond to media changes counting pH and once cross-linked to thermo-responsive polymers, temperature,63,64 allowing their focalized use for drug release. These features are suitable for anticancer drugs since tumor microenvironments are characterized by acid pH and increased temperature.65
Herein, the synthesis, characterization, DFT studies, and antiproliferative activity of novel third Casiopeinas® generation are presented. These compounds have general formula [Cu(N–N)(Indo)]NO3, where Indo is the deprotonated form of indomethacin and N–N are different methyl-substituted bipyridines and phenanthrolines. In addition, the synthesis and characterization of a nanopolymer formed by chitosan (CS), poly-N-isopropylacrylamide (pNIPAAm), and N,N′-methylenebis(acrylamide) (MBA) is also presented. Finally, the studies of [Cu(4,7-dimethyl-1,10-phenanthroline)(Indo)]NO3 nanoencapsulation efficiency and kinetics release are discussed.
To calculate redox potentials, we followed previous reports that have proven to be useful for transition metal complexes,77–79 and in which geometries of ferrocene (Cp2Fe) and ferrocenium ([Cp2Fe]+) in the eclipsed conformation (D5h) were taken into account and optimized in the gas phase. Gas phase and solvation ΔG energies were calculated by the same level of theory and using SMD continuum solvation model80 and DMSO to simulate the same environment of the electrochemical experiments. To obtain gas phase and solvation free energies of all species, a thermochemical analysis at 298.15 K and 1 atm were performed following the next cycle:
Redox potential was determined through the free energy changes and according to the one-electron exchange of Nernst equation:
| ΔG0 redoxsolv = ΔG0 redoxgas + ΔG0 reds − ΔG0 oxs | (1) |
| ΔG0 redoxsolv = −FE0calc | (2) |
O) stretching vibration suggests that Indo is properly coordinated to the Cu atom by the carboxylate groups. Additionally, all compounds' spectra show an absorption band around 3400 cm−1, corresponding to the ν(OH) vibration, probably caused by the presence of coordinated water on axial positions, and 1385 cm−1 associated with the NO3− as counterion.84 Mass spectra (Fig. S2†) of the complexes show the molecular ion [M+] peaks in the range of 577–741 m/z ratio and they are in accordance with the proposed molecular formulae. In addition, conductivity evaluations in a range from 82.4 to 105.9 ohm−1 cm2 mol−1 indicate 1
:
1 electrolyte behavior, in agreement with FTIR. Moreover, μeff values range from 1.66 to 2.22 BM, which is expected for mononuclear Cu(II) compounds with d9 configuration (S = 1/2), allowing EPR characterization.85
The EPR spectra (Table S1 and Fig. S3†) showed g// < 2.3, which indicates Cu–L bonds have dominant covalent character, this is explained by the σ-donor capacity of diimines and the electronic inductive effect caused by methyl groups.86 According to Peishach and Blumberg, A// and g// are useful to determinate the nature of the atoms bonded to copper, the data behavior shows copper centers have mixed copper–nitrogen and copper–oxygen bonds (CuN2O2 center).87 Dependence g// > g⊥ > ge purpose that the unpaired electron is in the dx2–y2 orbital, suggesting copper centers with axial symmetry D4h (square planar or octahedral with distortion) and confirming mixed copper–nitrogen and copper–oxygen bonds. Consistent with Aiso and giso values, the geometry adopted by all compounds is square-based pyramid C4v.86,88 The small f values (f = g///A//) showed by the seven copper compounds demonstrate their low distortion.89
Furthermore, for the systems under examination, λmax transitions (at 230 and 290 nm and ε ≈ 10
000 M−1 cm−1) correspond to transitions from π bonding orbitals to π antibonding orbitals of diimine aromatic rings. The metal–ligand charge transfer from d orbitals with t62ge3g configuration to empties π antibonding orbitals are found around 310 nm (ε ≈ 5000 M−1 cm−1).85,90 Finally, 2B1 → 2E and 2B1 → 2B2 transitions are observed at 650 nm (ε ≈ 35 M−1 cm−1) and 765 nm (ε ≈ 20 M−1 cm−1), respectively.91,92
Solution stability of the compounds was studied at different conditions time: 0–55 h, T = 22–40 °C and pH = 3.4–8.4 (37 °C) by UV-Vis techniques. As an example, the spectra of [Cu(4,7-dimethyl-1,10-phenanthroline)(Indo)]NO3 are presented in Fig. S4–S6.† Regarding this, the metal complexes were stable at the conditions tested, neither the dissociation of the ligands nor the formation of unknown species was detected.
Cyclic voltammetry was carried out to use half-wave potentials values (E1/2 for the CuII/CuI process) as possible descriptors related to the biological activity. Fig. 1 shows the voltammogram of [Cu(4,4′-dimethyl-2,2′-bipyridine)(Indo)]NO3 acquired with a glassy carbon electrode from open circuit potential to cathodic direction. Neither 1,10-phenanthroline and 2,2′-bipyridine nor HIndo processes interfere with copper signals.93 A peak was assigned to the copper reduction process at approximately −0.78 V/Fc+–Fc, and its respective oxidative (CuI → CuII) process was detected around at −0.55 V/Fc+–Fc. Even though some other signals can be found in the systems analyzed, for this study, only CuII/CuI redox values were considered. In general, the systems under study present a quasi-reversible behavior and since cathodic peak currents are not proportional to v1/2, non-diffusion-controlled processes are implied. Half-wave potentials are presented in Table 1 for all the investigated complexes.
![]() | ||
| Fig. 1 Cyclic voltammogram of 0.001 mol L−1 [Cu(4,4′-dimethyl-2,2′-bipyridine)(Indo)]NO3, 0.1 mol L−1 TBAPF6 in DMSO, v = 0.1 V s−1. | ||
| Compounds | E1/2 CuI/CuII (V/Fc+–Fc) |
|---|---|
| [Cu(1,10-phenanthroline)(Indo)]NO3 | −0.66 |
| [Cu(4,7-dimethyl-1,10-phenanthroline)(Indo)]NO3 | −0.83 |
| [Cu(5,6-dimethyl-1,10-phenanthroline)(Indo)]NO3 | −0.63 |
| [Cu(3,4,7,8-tetramethyl-1,10-phenanthroline)(Indo)]NO3 | −0.65 |
| [Cu(2,2′-bipyridine)(Indo)]NO3 | −0.65 |
| [Cu(4,4′-dimethyl-2,2′-bipyridine)(Indo)]NO3 | −0.72 |
| [Cu(5,5′-dimethyl-2,2′-bipyridine)(Indo)]NO3 | −0.76 |
Table 1 presents the redox potential values for copper(II) mixed compounds. It has been already suggested that this redox potential could play a key role in modulating the biological activity for some metal-containing systems, including copper.29 For the complexes under study, E1/2 for CuI/CuII redox pair are in a range between −0.63 and −0.83 V/Fc+–Fc values. In general, higher redox potential values are found in complexes containing 1,10-phenanthroline ligands. This can be explained by the extra aromatic ring compared to the 2,2′-bipyridine systems. The enhancement of the aromaticity is related to the electron delocalization capability. In those systems, this feature can explain the electron density decreasing around the copper atom through π back-bonding, thus, facilitating its reduction. Besides, methyl substituted systems can produce an electron density increased nearby the metal center by σ donation, without reducing the π back-bonding diimines capabilities as observed for [Cu(3,4,7,8-tetramethyl-1,10-phenanthroline)(Indo)]NO3.
IC50 = 1.10). Also, IC50 for Cu(NO3)2 is shown to truly assess the importance of copper complexes compared to Cu(II) in an aqueous solution (IC50 > 100.00 μmol L−1). The obtained values present a short-range since few structural modifications are present, i.e. the only differences are the number of aromatic rings and the presence in some compounds of methyl groups in different positions. This is related to the electron delocalization efficiency and the electron donating and withdrawing capability of each system. In this regard, IC50 values vary from 0.67 to 25.20 μmol L−1. According to the established compounds groups, substituted 1,10-phenanthroline systems showed higher antiproliferative activity compared to substituted 2,2′-bipyridine compounds. In each complex group, the less active system corresponds to those that are not substituted, i.e. [Cu(1,10-phenanthroline)(Indo)]NO3 and [Cu(2,2′-bipyridine)(Indo)]NO3 with IC50 values of 2.3 and 25.2 μmol L−1, respectively. This trend is in good agreement with previous Casiopeinas® reports29 and excluding [Cu(2,2′-bipyridine)(Indo)]NO3, all the analyzed systems show a lower IC50 value compared to cisplatin. In this regard, indomethacin derivatives show higher antiproliferative efficiency than the acetylacetonate derivatives.29 The trend in the primary donor is the same in both class of compounds, thus indicating that the difference arises from the secondary ligand. The inherent antiproliferative properties of the indomethacin can enhance the cytotoxic behavior of the copper compound.
| Compounds | IC50 (μmol L−1) | Log (IC50) | IC50 acaca (μmol L−1) |
|---|---|---|---|
| a Ref. 29, N.R.: not reported. | |||
| Cisplatin | 12.5 ± 0.70 | 1.10 | — |
| [Cu(1,10-phenanthroline)(Indo)]NO3 | 2.30 ± 0.02 | 0.36 | 10.7 ± 0.9 |
| [Cu(4,7-dimethyl-1,10-phenanthroline)(Indo)]NO3 | 0.72 ± 0.10 | −0.14 | 1.4 ± 0.1 |
| [Cu(5,6-dimethyl-1,10-phenanthroline)(Indo)]NO3 | 0.67 ± 0.02 | −0.17 | 3.4 ± 0.5 |
| [Cu(3,4,7,8-tetramethyl-1,10-phenanthroline)(Indo)]NO3 | 1.00 ± 0.03 | 0.00 | 1.9 ± 0.2 |
| [Cu(2,2′-bipyridine)(Indo)]NO3 | 25.2 ± 1.07 | 1.40 | 42.0 ± 3.1 |
| [Cu(4,4′-dimethyl-2,2′-bipyridine)(Indo)]NO3 | 7.87 ± 0.40 | 0.90 | 18.2 ± 2.7 |
| [Cu(5,5′-dimethyl-2,2′-bipyridine)(Indo)]NO3 | 2.87 ± 1.02 | 0.46 | N.R. |
| Cu(NO3)2 | >100 | >2 | — |
From all the studied complexes, [Cu(5,6-dimethyl-1,10-phenanthroline)(Indo)]NO3 and [Cu(4,7-dimethyl-1,10-phenanthroline)(Indo)]NO3 are the most effective (IC50 = 0.67 and 0.72 μmol L−1, respectively). Those systems are characterized by having three aromatic rings and –CH3 groups in different positions, which enables them to promote a better electron delocalization and to increment electron density on the copper metal atom. Usually, copper containing compounds might offer some advantages. They have high stability in physiological conditions, could be safer in comparison to other metal complexes, and they can produce several cell death mechanisms including cuproptosis or apoptosis.27,34 Hence, six of seven synthesized compounds belonging to the third Casiopeinas® generation with indomethacin might be suitable and considered as promising candidates for further studies as metallodrugs against cancer. Furthermore, those trends are in concordance with previous inquiries29,33,34 in which copper complexes are assumed to react with some important endogen reducers molecules such as glutathione and nicotinamide adenine dinucleotide (NADH). These reactions can generate Cu(I) species, which in turn can react with molecular oxygen (O2), hydrogen peroxide (H2O2), and superoxide radical
producing reactive oxygen species, thus causing oxidative damage and mitochondrial dysfunction.34 Finally, for the subsequent analysis, we have selected as the drug model for hydrogel design, [Cu(4,7-dimethyl-1,10-phenanthroline)(Indo)]NO3 compound.
The CuII complexes present a square planar arrangement where the indomethacin and diimines are both bidentate O–O (C–O distance average 1.97 Å) and N–N (Cu–N distance average 1.97 Å) coordinated, respectively. The reduced CuI species has a trigonal planar geometry, the diimine remains N–N bonded (Cu–N distance average 1.92 and 1.99 Å) and the indomethacin is bonded only for a single O atom (Cu–O average distance 1.90 Å). The remaining oxygen atom remains far from the Cu atom, but still, a weak intermolecular interaction could be proceeding (Cu–O average distance 3.02 Å).
By QSAR studies, our research group demonstrates that the cytotoxic activity of the first Casiopeinas® generation can be described by the redox potential.29,94,95 The redox potential has been used commonly in structure–activity studies for several metal containing compounds with biological activity.94 Although calculated redox potential values, E0calc, are in good agreement with experimental ones (R2 = 0.98, Fig. S8†), in this case; this parameter does not provide useful information regarding the antiproliferative effect. However, when the molar volume, V, is employed, some interesting relations can be made. In Fig. 2, we present those correlations for the two involved compound families. On the one hand, the 1,10-phenanthroline group, and on the other hand the 2,2′-bipyridine molecules. As can be seen in this Fig. 2, only the molar volume was considered, and molecules are well separated according to the group they belong to (black circles for 2,2′-bipyridine molecules and red squares for 1,10-phenanthroline compounds, with R2 = 0.9998, and R2 = 0.9117, respectively). The most important result from this figure, is that for all complexes under study, the larger the volume, the better the antiproliferative effect. This trend can be observed regardless of the molecules group. In particular, the steric effect is much more pronounced on the cytotoxic efficiency for bipyridine derivatives compared to phenanthroline ones. As previously observed, the less active complexes correspond to the non-methylated substituted complexes. Contrarily, comparing methyl substitutions, the one in position 5 enhances the biological activity. Interestingly, the 5,6-dimethyl-1,10-phenanthroline derivative shows the best cytotoxic behavior. This complex owns the most oxidizing power and the biggest molar volume of all the analyzed systems. In this regard, it has been proved that this ligand enhances the capability of Casiopeinas® of diffusion across the cell membrane due to the minimized electronic and structural repulsion provided by the methyl substitution in the 5,6-positions compared to others.93 Accordingly, the change in the biological activity can be attributed to the interactions that facilitate their passage through the membrane, that is, its biodistribution.
The characterization of the nanogel was made by FT-IR spectroscopy. The signals in 1648 (νs C
O), 1553 (νas N–H), and 1261 cm−1 (νs C–N) confirm the presence of the amide group.97 Broadband around 1066 cm−1 is according to the pyranosic ring.98 Additionally, the bending of methylene of the methylene group (νsc C–H) that produces a signal in 1469 cm−1 can be observed.98 Finally, the peak in 1027 cm−1 is attributed to the ether groups (νs C–O–C).98
To emulate biological conditions and to evaluate the behavior of the nanogel in those situations, various characterization parameters were obtained at different temperatures and pH values. Table 3 contains the obtained hydrodynamic diameter, zeta potential (Zp), and polydispersity index (PDI).
| pH | T (°C) | Hydrodynamic diameter (nm) | Zp (mV) | PDI |
|---|---|---|---|---|
| 5.0 | 25 | 690.27 ± 7.10 | 22.97 ± 0.55 | 0.530 |
| 37 | 654.13 ± 32.07 | 23.50 ± 0.35 | 0.512 | |
| 7.4 | 25 | 108.92 ± 12.85 | −6.81 ± 0.45 | 0.573 |
| 37 | 100.89 ± 8.33 | −5.81 ± 1.90 | 0.540 |
PDI values indicate that, in the studied conditions, the nanogel is not a monodisperse system. The temperature does not have a great effect on the hydrodynamic diameter, nevertheless, the particles swell when pH values diminish; these results are explained by the polymers ratio minish in the hydrogel. The swelling in acidic conditions can be explained since at low pH values, the lateral chains of CS are protonated, and thus, repulsion is increased and so is the hydrodynamic diameter. In the same way, the acidic pH generates a positive potential difference while the neutral polymer (pH = 7.4) has a negative Zp value. Anyway, in both values of investigated pH, the CS–NIPAAm polymer is electrically stable. The difference in the revised indexes suggests that the polymer responds to pH and temperature changes.99
The changes in the differential calorimetry screening (DSC) study were measured in the range of 25–60 °C. The DSC plot can be founded in Fig. S9.† The free PNIPAAm single transition is observed at 28.07 °C, close to the reported values in the specialized literature.100 For the nanogel at pH = 5.0, two transitions are founded, the first at 28.19 °C, that can be attributed to the free NIPAAm, and the second is found at 45.16 °C and can be related to the CS–NIPAAm interaction. At pH = 7.4, a single transition at 49.76 °C was observed. For both examined pH values, the second cycle presents a similar behavior to the first one. The involved energy in the phase transition for pH = 7.4 is −1.20 W g−1 and for pH = 5.0 is −0.13 W g−1. Those results in addition to the observed increased temperature for the transition indicate that a stronger interaction between CS and NIPAAm at the higher pH value occurs.101
Furthermore, the scanning electron microscopy (SEM) technique was performed and the micrographs are presented in Fig. 3.
The micrographs reveal different morphologies for the different pH conditions. In the acidic situation, the nanogel presents rhombohedral structures with well-defined edges, its size varies between 3–5 μm. At neutral pH, laminar form is observed, the sheets are about 0.11 μm in thickness. While the size at pH = 7.4 correlates well with those reported in Table 3, those at pH = 5.0 are different. This could be explained because of morphology changes by the solvent evaporation. This reveals again, the nanogel pH dependence behavior. Finally, these results suggest that the synthesized nanogel can be used as a drug carrier with a pH dependence.
The encapsulation was performed in a HOAc/NaOAc buffer solution at pH = 5.0 and 20 °C, which are conditions where the hydrogel is swelled. The encapsulation efficiency (% EE) was defined as the ratio of the weight of [Cu(4,7-dimethyl-1,10-phenanthroline)(Indo)]NO3 contained into the hydrogel to the weight of [Cu(4,7-dimethyl-1,10-phenanthroline) (Indo)]NO3 added, loading capacity (% LC) was defined as the ratio of the weight of [Cu(4,7-dimethyl-1,10-phenanthroline)(Indo)]NO3 contained in the hydrogel to the weight of the hydrogel. Under the studied conditions, the hydrogel reached a % EE = 69.47 ± 1.41 and a % LC = 0.0496 ± 0.0010 (0.9899 ± 0.0201 mg of copper compound for each 20 mg of nanogel). The results are presented in Table S2.†
Fig. 4 shows the percentage of cumulative compound released as a function of time in both pH conditions. It can be observed that drug release depends on the pH value. In acidic conditions, the drug liberation is faster than for the pH = 7.4. In both cases, a Burst effect is observed in the first 6 hours. This effect is produced by the rapid release of the compound that interacts with the nanogel surface compared to that which is embodied in the polymeric membrane.
![]() | ||
| Fig. 4 Percentage of in vitro cumulative [Cu(4,7-dimethyl-1,10-phenanthroline)(Indo)]NO3 released as a time function. | ||
To understand the drug release mechanism, some kinetic drug release models were proved to establish the nanogel behavior. For both evaluated pH conditions, the Peppas–Sahlin model provided the best fit (Table 4 and Fig. S10†).103
| Parameter | pH = 5.0 | pH = 7.4 |
|---|---|---|
| k1 | 82.402 ± 0.942 | 40.907 ± 1.815 |
| k2 | −18.350 ± 0.382 | −6.397 ± 0.408 |
| m | 0.522 ± 0.010 | 0.665 ± 0.024 |
The magnitudes of k1 compared to k2 indicates that Fick diffusion controls the drug transport mechanism and the negative values of k2 suggest an almost inexistent contribution due to polymeric chains relaxation. Also, the m value shows a spherical material. The observations found here agree with the reports from other nanogels and soluble-water compounds.104 The increased release of this system in the acidic media allows us to propose it as a specific potential drug nanocarrier for the tumoral microenvironment.
Footnote |
| † Electronic supplementary information (ESI) available. See https://doi.org/10.1039/d2ra03346a |
| This journal is © The Royal Society of Chemistry 2022 |