Next Article in Journal
Analysis of Dynamic Characteristics of a 600 kW Storage Type Wind Turbine with Hybrid Hydraulic Transmission
Previous Article in Journal
Modeling of the Free Radical Copolymerization Kinetics of n-Butyl Acrylate, Methyl Methacrylate and 2-Ethylhexyl Acrylate Using PREDICI®
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Theoretical Study of the Adsorption Process of Antimalarial Drugs into Acrylamide-Base Hydrogel Model Using DFT Methods: The First Approach to the Rational Design of a Controlled Drug Delivery System

1
Grupo de Investigación en Ciencias Naturales y Exactas, Departamento de Ciencias Naturales y Exactas, Universidad de la Costa, Barranquilla 080002, Colombia
2
Grupo de Investigación en Química y Biología, Departamento de Química y Biología, Universidad del Norte, Barranquilla 081007, Colombia
3
Grupo de Química Computacional y Teórica (QCT-USFQ) and Instituto de Simulación Computacional (ISC-USFQ), Colegio Politécnico de Ciencias e Ingeniería, Diego de Robles y Vía Interoceánica, Universidad San Francisco de Quito, Quito 170901, Ecuador
4
Grupo de Investigación en Oxi/Hidrotratamiento Catalítico y Nuevos Materiales, Programa de Química-Ciencias Básicas, Universidad del Atlántico, Puerto Colombia 081001, Colombia
5
TecNM/Instituto Tecnológico de Aguascalientes-División de Estudios de Posgrado e Investigación, Ave. Adolfo López Mateos #1801Ote. Fracc. Bona Gens, Aguascalientes 20256, México
6
Grupo de Investigación en Compuestos Heterocíclicos, Programa de Química, Facultad de Ciencias Básicas, Universidad del Atlántico, Puerto Colombia 081001, Colombia
*
Author to whom correspondence should be addressed.
Processes 2019, 7(7), 396; https://doi.org/10.3390/pr7070396
Submission received: 11 May 2019 / Revised: 28 May 2019 / Accepted: 12 June 2019 / Published: 26 June 2019
(This article belongs to the Section Biological Processes and Systems)

Abstract

:
The interaction between three widely used antimalarial drugs chloroquine, primaquine and amodiaquine with acrylamide dimer and trimer as a hydrogel model, were studied by means of density functional theory calculation in both vacuum and water environments, using the functional wb97xd with 6-31++G(d,p) basis set and polarizable continuum model (C-PCM) of solvent. According to binding energy, around −3.15 to −11.91 kJ/mol, the interaction between antimalarial compounds and hydrogel model are exothermic in nature. The extent of interaction found is primaquine > amodiaquine > chloroquine. The natural bond orbital (NBO) calculation and application of second-order perturbation theory show strong charge transfer between the antimalarial and hydrogel model. In addition, the results suggest these interactions are polar in nature, where hydrogen bonds play a principal role in stabilization of the complex. Comparing with the gas-phase, the complexes in the water environment are also stable, with suitable values of Log P (Partition coefficient), and dipolar momentum. Consequently, these results encourage to test acrylamide hydrogels as antimalarial delivery systems.

1. Introduction

Malaria is a chronic infection caused by protozoans of the genus Plasmodium and it is considered as one of the deadliest diseases of the planet. Dramatically, 60% of the world’s population lives under risk of contracting this illness; in addition, according to the world Malaria report, nearly 800,000 people died in 2011 because of malaria, and most of them were children [1].
The Increasing antimalarial resistant plasmodium strains has gained the attention of the World Health Organization (WHO) and its effort for eradication or elimination of this illness [2]. This fact, and the toxicity degree of antimalarials, represent a challenge for the scientific community; consequently, overcoming these two facts is a WHO goal of the millennium; however, it is obvious this goal demands multidisciplinary fronts, mainly related to efficient and rational strategies for new delivery systems as well as the development of new antimalarial drugs [2,3].
Drug delivery systems that can control the release behaviors of multiple drugs have recently become an attractive route for enhanced therapeutic effects [4]. Hydrogels, lipids, and polymer-drug conjugate composites, have been studied as drug delivery systems on several illnesses [5,6].
Regarding malaria, several studies about antimalarial-drug delivery have been reported, however, the liposome was one of the most effective systems [7]. Combining primaquine (PQ), chloroquine (CQ) with liposomes and the polymeric matrix, have shown better behavior attaining a decreasing in drugs toxicity [7]. These results represent strong piece of evidence that these types of systems constitute a potential alternative to the conventional administration of drugs [7,8,9].
A hydrogel is a three-dimensional network polymer, with a size-porous structure depends on the degree of crosslinking [8]. The degree of crosslinking, in turn, determines the degree of swelling of the matrix. They are hydrophilic and biocompatible; thus, these properties allow its uses for biomedical applications mainly as drug-delivery carriers [8]. Nevertheless, the ability of a hydrogel to release a specific drug will depend almost exclusively on the compatibility between both, drug and hydrogel or polymeric matrix [7,8,9,10,11,12].
Compatibility between drugs and polymers concerns not only the miscibility but also the extent of interaction with no alteration in the chemical structure neither the polymer or the drug [13]. A good interaction (physisorption) between drug and delivery hydrogel/polymers is the first step in the design of drug delivery systems; however, no delivery carrier prepared from a particular polymer will serve as a universal carrier for all drugs. Hence, the assessment of the compatibility plays a pivotal role in successful design [13,14,15,16,17].
Several techniques are used to estimate the compatibility between the polymer and a specific drug. The interaction drug-polymer can be monitored by the construction of delivery profiles from the adsorbed drug into the polymer in order to obtain the kinetic parameters for the process and gain information about the interaction stability. Among the analytical techniques used for this aim, are differential scanning calorimetry (DSC), X-ray powder diffraction (XRD), nuclear magnetic resonance and Fourier transformed infrared spectroscopy (FTIR) [5,6]. However, these methods require a long analysis- time and cannot be used to predict compatibility.
Several kinds of molecular interactions could take place between drugs and polymer matrixes; however, the most important are dipole-dipole interactions and dispersion forces, where the first one is the strongest [18,19]. Generally, the presence of dipole-dipole forces into polymer-drugs systems guarantees the formation of a stable complex [19]. In addition, the dipole-dipole interaction between polymer and water makes the hydrogel swelling easier, then, the release of the drugs through the diffusion process takes place. In this sense, the study of intermolecular interactions between drugs and hydrogels allows adequate selection, but also suitable modification under a polymeric matrix to a specific drug delivery [20,21,22,23,24].
Energetic magnitudes of molecular interactions that come from drug-delivery systems could be estimated through thermodynamic parameters derived from Gibbs function, i.e., enthalpy (ΔH), entropy (ΔS), and free energy changes (ΔG) [19]. The associative interaction between drug-delivery systems could include hydrogen bonding, ionic aggregation, electrostatic interaction, and solvent complex formation. In this sense, the calculation of ΔH, ΔS, and ΔG, describe the entire interaction into the system [16,24,25,26].
Since the binding energy is of the order of thermal energy, bond formation (or interaction) is reversible; bonds are created and destroyed by the change in temperature or reactant concentration. Thus, thermodynamic function estimations not only represent the extent of interaction but also allows to estimate the dissociation-temperature of the complex as well [26,27,28,29].
Computational chemistry represents an excellent tool for estimating molecular interactions, as well as thermodynamic parameters and elucidating reaction mechanisms [30,31,32]. Nowadays, computational chemistry is the starting point in the design and rational synthesis of drug discovery, herbicides, and insecticides [33,34]. Therefore, using computational chemistry represents a useful tool to estimate the molecular interaction taking place in the drug delivery system, obtaining a predictive model for behavior of these systems. On the other hand, this will allow rational selection of the drug and the suitable system delivery. Moreover, computational simulation of drug delivery systems has emerged for use recently, mainly in controlled anticancer and antituberculosis drug delivery [32,35,36,37,38,39].
Despite the importance of eradicating or mitigating the malarial disease and the potential use of antimalarial delivery system drugs to reach this goal, there is no work reported where computational simulation is used with the purpose of gaining insight the molecular association in the antimalarial-delivery system. For this reason, the main objective of this work is to obtain the electronic and structural properties to determine if the non-covalent addition of drugs to a polymer administration system is favorable, in addition to determining the chemical reactivity using molecular descriptors.
This is the first theoretical study that evaluates the possibility of using polyacrylamide as a matrix for the antimalarial delivery system. Dimer (DAA) and trimer of acrylamide (TAA) were used as the models of polyacrylamide hydrogels; the molecular interactions with chloroquine (CQ), primaquine (PQ) and amodiaquine (AMQ) antimalarials were studied using Density Functional Theory (DFT) as the level of theory and the wb97xd/631++G(d,p) basis set. Electronic transferences energies and charge distribution were calculated by the means of the conceptual DFT jointly with Natural bond orbital, respectively [29,40].

2. Materials and Methods

Computational Methods

The minimum potential energy surface of three antimalarial drugs and their complexes with polyacrylamide models (DAA, TAA, DAA-CQ, DAA-PQ, DAA-AQ, TAA-CQ, TAA-PQ and TAA-AQ) were performed at Density Functional Theory (DFT) level, using functional exchange-correlation of wb97xd, taking into account both short-range and long-range interactions, combined with the basic set 6-31++G(d,p), from Gaussian 16 software [41]. This methodology has demonstrated to be accurate enough to evaluate the interaction between organic molecules [42,43,44].
The algorithm of Berny analytical gradient was used throughout as the optimization criterion. The requested convergence was density matrix 10−9 atomic units, maximum displacement threshold 0.0018 Å, and maximum force threshold 0.00045 Hartree Bohr-1. The nature of stationary points was recognized by calculating and diagonalizing the Hessian matrix (force constant matrix). Minimum energy geometry was characterized by means of normal-mode analysis, by having all positive frequencies.
Frequency calculations were used with the purpose to obtain some important parameters as zero-point vibrational energy (ZPVE), temperature corrections (E(T)) and absolute entropies (S(T)). When the calculation was carried out in the vacuum, the entire species i.e., reactant, complexes and products, were considered an ideal gas, with harmonic frequencies and inertia momentum estimated from the statistic mechanical standard method at 298.15 K and 1 atm [45].
Dimer and trimer of acrylamide (AA) were used as a model of polyacrylamide-based hydrogels. Some reports have proved the trimer of acrylamide (AA) represents a good model for simulation of polyacrylamide properties [30,46].
In order to achieve a better estimation of the binding energy [47,48], the basis-set superposition error (BSSE) correction was applied in the optimized structure of the various clusters by the standard counterpoise method in which the binding energy (BE) of a complex is defined as:
Δ E b = E C o m p l e x E M o d e l D A A o r T A A + E d r u g s + B S S E ,
where Ecomplex, EModel, and Edrugs are energies of DAA or TAA with drugs, energies of DAA or TAA and Free drugs molecules energies. BSSE is the basis set superposition error [49,50]. The negative value of ∆Eb indicates a minimization of energy; on the other hand, the interaction is exothermic [47].
Natural bond orbital (NBO) analysis [29] was used to understand the orbital interactions and changes in charge distribution as well as the bond order into the complex and free models and drugs. From the DFT point of view, the global interaction between drugs and polymers model is described by ΔN, which determines the number of electrons transferred from a molecule A to molecule B, and computed as:
Δ N = ( μ B μ A ) ( n B + n A ) ,
where µ is electronic potential, n is global hardness; A and B are acceptor and donor molecules, respectively. A negative value on ΔN indicates the electron flow will be spontaneous from A to B; the reverse direction will have positives values [40].
Solvation energy (ΔGs) and LogP (QSAR properties) were computed. ΔGs were calculated using the C-PCM method, with the wb97xd/6-31++G(d,p) theory level using Gaussian 16 [41]. The Log P was calculated using the Spartan software by selecting the Ghose–Crippen method. This method is independent of the wave energy function, that is, the calculation does not influence whether it obtained using a quantum mechanics, molecular mechanics, or a semi-empirical method. The Ghose–Crippen method was parameterized [51] taking in account 110 atoms/bonds, including the most common atoms such as C, H, O, N, S, and the halogens. Log P is a measure of solubility of the substance in water and into lipid phase (octanol). In this sense, it is necessary that the polymer chains interact with water, in order to swell and release the drugs through a diffusion process. Hence, the hydrophilicity and solvation energy are an important parameter in drug delivery system simulation.

3. Results

3.1. Geometrical Structures and Binding Energy

The study of polymer-antimalarial drugs was carried out using the wb97xd/6-31++G(d,p) level of theory. The minimum energy geometries of antimalarial drugs and complexes DAA-CQ, DAA-PQ, DAA-AQ, TAA-CQ, TAA-PQ and TAA-AQ were confirmed through frequency calculation.
The minimum energy structures are shown in Figure 1 and Figure 2. Application of second derives criterion using frequency calculation proved that all of them were minimum energy geometries.
Table 1 illustrates the geometric parameters for the structure studied in the vacuum. Nevertheless, when several possible interaction geometries of the model and drugs were studied, the structure of minimum energies was always the one in which the dipole-dipole interactions between the drug and polymer model takes place. Besides, Table 1 reveals that the interatomic bonds lengths are in the range (1.02 ≤ A ≤ 3.54) Å. The small interatomic length found for PQ and AQ suggest dipole-dipole interaction. Likewise, this result indicates a better interaction between hydrogel-PQ and AQ than hydrogel-CQ.
Table 2 shows the binding energy, corrected using BSEE described previously. It could be noted that the higher the acrylamide monomers number, the slower the binding energy. In general, the values of binding energy obtained vary from 62 to −55 kJ/mol and are in the non-covalent range of energy. These results and the equilibrium distances between drugs–polymer (2.00–4.3 Å) indicate these kinds of antimalarial drugs could physically adsorb into the hydrogel matrix; thus, they represent strong evidence to test the possibility of its use as a drug release system.
It could be noted that the binding energy follows the order: ΔEb (TAA-PQ) < ΔEb (TAA-AQ) < ΔEb (TAA-CQ) < 0, suggesting the polymeric matrix interacts better with primaquine and amodiaquine than chloroquine.
In order to understand the interactions between antimalarials and the model of hydrogels, the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbitals (LUMO) were studied. Figure 3 shows the frontier molecular orbital of antimalarial compounds and hydrogel model studies herein.
It could be noted for completely antimalarial drugs, i.e., chloroquine, primaquine and amodiaquine, HOMO orbitals are located into a specific molecular region. Presence of dipolar momentum in these molecules could be used in order to find the best polymeric matrix; consequently, permanent dipolar region or donor-acceptor region into molecules, could warrant the suitable interaction within the dipolar region of the hydrogel model. It can note that TAA has a similar characteristic of the HOMO-LUMO orbital, i.e., acceptor-donor regions, to the drugs frontiers orbitals. Therefore, there is a high probability of a better interaction between antimalarial and hydrogel as the monomer chain increases.
The dual DFT descriptor ΔN was calculated in order to understand the dipole interaction in nature. For this goal, polymeric models were considered as species B and antimalarial compounds as species A. Negatives values in all cases suggest the electrons flux goes from antimalarial to hydrogel model. Likewise, the hydrogel is acting as an acceptor species and drugs are the electronic donors (see Table 3). It can be seen that the electronic potential is negative, confirming the stability of all complexes between the hydrogel model and drugs. These results are strong evidence for the potential use of these complexes as an antimalarial drug delivery system.

3.2. Natural Bond Orbital (NBO) Analysis

The nature of the intermolecular bond for DAA-CQ, DAA-PQ, DAA-AQ, TAA-CQ, TAA-PQ and TAA-AQ complexes, were studied by means of NBO analysis. In this way, second-order perturbation theory was used in order to gain insights about the donor (i) acceptor (j) orbital binding. Associated stabilization energy was calculated using Equation (3), from Fock matrix analysis on NBO calculation [29,52,53,54].
E i j ( 2 ) = q i ( F i j ) 2 E j E i ,  
where qi is the donor electronic occupied orbital; Ej and Ei are orbital energies, and Fij is out-diagonal Fock matrix element. Table 3 shows the most important interactions between donor and acceptor elements in complexes DAA-CQ, DAA-PQ, DAA-AQ, TAA-CQ, TAA-PQ, and TAA-AQ. In all cases, the donor element is the lone pair from N, Cl, or O atoms, while the acceptor one is the antibonding, σ*, C-H, O-H, and N-H orbitals, confirming the molecular interactions are mainly dipolar in nature.
A close inspection of Table 4 reveals the most important interaction between PQ and AQ with TAA, occurs through a hydrogen bond with high stabilization energies in both PQ and AQ. Moreover, these results imply a greater extent for the interaction of the hydrogel model with primaquine and amodiaquine, but a little interaction with chloroquine. Likewise, these results are in agreement with the large bond energies obtained for TAA-PQ and TAA-AQ in comparison with the TAA-CQ complex.

3.3. Water Environment Behavior

Quantitative structure–activity relationship (QSAR) approaches, as progressive tools in modeling and prediction of many physiochemical properties, offers a fast measure of predictability in the absence of extensive experimental or computed data on compound properties. In this respect, the QSAR properties derived from Hyperchem software were calculated, with the purpose to estimate the performance of complexes in the water environment. Table 5 shows three fundamental properties that have a great influence on a drug-delivery system: hydrophilicity (Log P), solvation free energy (Gibbs free energy), and dipole momentum. These descriptors allow estimating the behavior of the drugs, once they are delivered into the receptor.
According to Table 5, all of the complexes have solvation energy <0, thus, all of them are stable in water. In addition, the values of Log P indicate all complexes are hydrophilic (Log P < 0) and polar in nature (µ > 0). These results are strong evidence about the possibility to test these models as a drug-delivery system; polarity and hydrophilicity are a warranty of the interaction of water and polymeric chain; in this sense, hydrogels would start to swell and then, release the antimalarial drugs.

4. Conclusions

Minimum energy structure of Chloroquine, (CQ), Primaquine (PQ) and Amodiaquine (AQ) and their Van der Waals complexes with dimer and trimer of acrylamide, were calculated using density functional theory (B3LYP/6-31++G(d,p) in both gas phase and water solution (C-PCM). The binding energy suggests all compounds could form a stable complex with acrylamide hydrogel in both vacuum and water. The second order perturbation theory suggests the complexes formed mainly through dipole-dipole interaction and hydrogen bonds play a pivotal role in the stability of these complexes. Additionally, QSAR properties as Log P, solvation free energy, and dipolar momentum, µ, show the complexes could be soluble, stable and could interact with water molecules, allowing the swelling of the hydrogel and release of the antimalarial drugs. The results show herein suggest amodiaquine (AQ) and primaquine (PQ) are suitable antimalarial drugs for delivery from an acrylamide-based hydrogel.

Author Contributions

Calculation and analysis, E.C., E.P., conceptualization and funding acquisition, J.T.; Investigation, methodology, validation, formal analysis, J.R.M., N.R., E.M., writing—original draft preparation, writing—review and editing, E.M; Just Software; A.D.M.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Organisation Mondiale de la Santé. World Malaria Report 2012 WHO Global Malaria Programme; World Health Organization: Geneva, Swizerland, 2012; ISBN 978-92-4-156453-3. [Google Scholar]
  2. Al Qaraghuli, M.M.; Obeid, M.A.; Aldulaimi, O.; Ferro, V.A. Control of malaria by bio-therapeutics and drug delivery systems. J. Med. Microbiol. Diagn. 2017, 6, 260. [Google Scholar] [CrossRef]
  3. Liu, J.; Xiao, Y.; Allen, C. Polymer–drug compatibility: A guide to the development of delivery systems for the anticancer agent, ellipticine. J. Pharm. Sci. 2004, 93, 132–143. [Google Scholar] [CrossRef] [PubMed]
  4. Tang, H.; Zhao, W.; Yu, J.; Li, Y.; Zhao, C. Recent Development of pH-Responsive Polymers for Cancer Nanomedicine. Molecules 2019, 24, 4. [Google Scholar] [CrossRef] [PubMed]
  5. Yang, L.; Sun, H.; Liu, Y.; Hou, W.; Yang, Y.; Cai, R.; Cui, C.; Zhang, P.; Pan, X.; Li, X.; et al. Self-Assembled Aptamer-Grafted Hyperbranched Polymer Nanocarrier for Targeted and Photoresponsive Drug Delivery. Angew. Chem. Int. Ed. 2018, 57, 17048–17052. [Google Scholar] [CrossRef] [PubMed]
  6. Externally Triggered Heat and Drug Release from Magnetically Controlled Nanocarriers|ACS Applied Polymer Materials. Available online: https://pubs.acs.org/doi/pdf/10.1021/acsapm.8b00100 (accessed on 27 May 2019).
  7. Murambiwa, P.; Masola, B.; Govender, T.; Mukaratirwa, S.; Musabayane, C.T. Anti-malarial drug formulations and novel delivery systems: A review. Acta Trop. 2011, 118, 71–79. [Google Scholar] [CrossRef] [PubMed]
  8. Ganji, F.; Vasheghani-Farahani, S.; Vasheghani-Farahani, E. Theoretical description of hydrogel swelling: A review. Iran. Polym. J. 2010, 19, 375–398. [Google Scholar]
  9. Huang, Y.; Jin, X.; Liu, H.; Hu, Y. A molecular thermodynamic model for the swelling of thermo-sensitive hydrogels. Fluid Phase Equilibria 2008, 263, 96–101. [Google Scholar] [CrossRef]
  10. Cai, S.; Suo, Z. Mechanics and chemical thermodynamics of phase transition in temperature-sensitive hydrogels. J. Mech. Phys. Solids 2011, 59, 2259–2278. [Google Scholar] [CrossRef]
  11. Du, X.; Zhou, J.; Shi, J.; Xu, B. Supramolecular Hydrogelators and Hydrogels: From Soft Matter to Molecular Biomaterials. Chem. Rev. 2015, 115, 13165–13307. [Google Scholar] [CrossRef]
  12. Huynh, L.; Neale, C.; Pomès, R.; Allen, C. Computational approaches to the rational design of nanoemulsions, polymeric micelles, and dendrimers for drug delivery. Nanomed.: Nanotechnol. Biol. Med. 2012, 8, 20–36. [Google Scholar] [CrossRef]
  13. Shen, E.; Kipper, M.J.; Dziadul, B.; Lim, M.-K.; Narasimhan, B. Mechanistic relationships between polymer microstructure and drug release kinetics in bioerodible polyanhydrides. J. Control. Release 2002, 82, 115–125. [Google Scholar] [CrossRef]
  14. Park, J.; Ye, M.; Park, K. Biodegradable polymers for microencapsulation of drugs. Molecules 2005, 10, 146–161. [Google Scholar] [CrossRef] [PubMed]
  15. Weiss, R.G.; Terech, P. (Eds.) Molecular Gels: Materials with Self-Assembled Fibrillar Networks; Springer: Dordrecht, The Netherlands, 2006; ISBN 978-1-4020-3352-0. [Google Scholar]
  16. Zweep, N.; Hopkinson, A.; Meetsma, A.; Browne, W.R.; Feringa, B.L.; van Esch, J.H. balancing hydrogen bonding and van der waals interactions in cyclohexane-based bisamide and bisurea organogelators. Langmuir 2009, 25, 8802–8809. [Google Scholar] [CrossRef] [PubMed]
  17. Hoy, R.S.; Fredrickson, G.H. Thermoreversible associating polymer networks. I. Interplay of thermodynamics, chemical kinetics, and polymer physics. J. Chem. Phys. 2009, 131, 224902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Maniruzzaman, M.; Pang, J.; Morgan, D.J.; Douroumis, D. Molecular Modeling as a Predictive Tool for the Development of Solid Dispersions. Available online: https://pubs.acs.org/doi/abs/10.1021/mp500510m (accessed on 3 November 2018).
  19. Maniruzzaman, M.; Morgan, D.J.; Mendham, A.P.; Pang, J.; Snowden, M.J.; Douroumis, D. Drug–polymer intermolecular interactions in hot-melt extruded solid dispersions. Int. J. Pharm. 2013, 443, 199–208. [Google Scholar] [CrossRef] [PubMed]
  20. Schmaljohann, D. Thermo- and pH-responsive polymers in drug delivery. Adv. Drug Deliv. Rev. 2006, 58, 1655–1670. [Google Scholar] [CrossRef] [PubMed]
  21. Seiffert, S.; Sprakel, J. Physical chemistry of supramolecular polymer networks. Chem. Soc. Rev. 2012, 41, 909–930. [Google Scholar] [CrossRef] [PubMed]
  22. Tanaka, F. Theoretical Study of Molecular Association and Thermoreversible Gelation in Polymers. Polym. J. 2002, 34, 479–509. [Google Scholar] [CrossRef] [Green Version]
  23. Peppas, N.A.; Huang, Y.; Torres-Lugo, M.; Ward, J.H.; Zhang, J. Physicochemical Foundations and Structural Design of Hydrogels in Medicine and Biology. Annu. Rev. Biomed. Eng. 2000, 2, 9–29. [Google Scholar] [CrossRef] [Green Version]
  24. Chun, B.J.; Lu, J.; Weck, M.; Jang, S.S. Characterization of molecular association of poly(2-oxazoline)s-based micelles with various epoxides and diols via the Flory–Huggins theory: A molecular dynamics simulation approach. Phys. Chem. Chem. Phys. 2015, 17, 29161–29170. [Google Scholar] [CrossRef]
  25. Bahar, I.; Erbil, H.Y.; Baysal, B.M.; Erman, B. Determination of polymer-solvent interaction parameter from swelling of networks: The system poly(2-hydroxyethyl methacrylate)-diethylene glycol. Macromolecules 1987, 20, 1353–1356. [Google Scholar] [CrossRef]
  26. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 2001, 46, 3–26. [Google Scholar] [CrossRef]
  27. Marquez, E.; Domínguez, R.M.; Mora, J.R.; Córdova, T.; Chuchani, G. Experimental and theoretical studies of the homogeneous, unimolecular gas-phase elimination kinetics of trimethyl orthovalerate and trimethyl orthochloroacetate. J. Phys. Chem. A 2010, 114, 4203–4209. [Google Scholar] [CrossRef] [PubMed]
  28. Exner, K.S.; Over, H. Kinetics of electrocatalytic reactions from first-principles: A critical comparison with the ab initio thermodynamics approach. Acc. Chem. Res. 2017, 50, 1240–1247. [Google Scholar] [CrossRef] [PubMed]
  29. Reed, A.E.; Curtiss, L.A.; Weinhold, F. Intermolecular interactions from a natural bond orbital, donor-acceptor viewpoint. Chem. Rev. 1988, 88, 899–926. [Google Scholar] [CrossRef]
  30. Shirota, H.; Ushiyama, H. Hydrogen-bonding dynamics in aqueous solutions of amides and acids: Monomer, dimer, trimer, and polymer. J. Phys. Chem. B 2008, 112, 13542–13551. [Google Scholar] [CrossRef] [PubMed]
  31. Liu, H.; Chen, J.; Shen, Q.; Fu, W.; Wu, W. Molecular insights on the cyclic peptide nanotube-mediated transportation of antitumor drug 5-fluorouracil. Mol. Pharm. 2010, 7, 1985–1994. [Google Scholar] [CrossRef]
  32. Albertorio, F.; Hughes, M.E.; Golovchenko, J.A.; Branton, D. Base dependent DNA–carbon nanotube interactions: Activation enthalpies and assembly–disassembly control. Nanotechnology 2009, 20, 395101. [Google Scholar] [CrossRef]
  33. Sparks, T.C.; Lorsbach, B.A. Agrochemical Discovery—Building the Next Generation of Insect Control Agents. In ACS Symposium Series; Gross, A.D., Ozoe, Y., Coats, J.R., Eds.; American Chemical Society: Washington, DC, USA, 2017; Volume 1264, pp. 1–17. ISBN 978-0-8412-3257-0. [Google Scholar]
  34. Baldi, A. Computational approaches for drug design and discovery: An overview. Syst. Rev. Pharm. 2010, 1, 99. [Google Scholar] [CrossRef]
  35. Gallo, M.; Favila, A.; Glossman-Mitnik, D. DFT studies of functionalized carbon nanotubes and fullerenes as nanovectors for drug delivery of antitubercular compounds. Chem. Phys. Lett. 2007, 447, 105–109. [Google Scholar] [CrossRef]
  36. Karataş, D.; Tekin, A.; Bahadori, F.; Çelik, M.S. Interaction of curcumin in a drug delivery system including a composite with poly(lactic-co-glycolic acid) and montmorillonite: A density functional theory and molecular dynamics study. J. Mater. Chem. B 2017, 5, 8070–8082. [Google Scholar] [CrossRef]
  37. Kaur, J.; Singla, P.; Goel, N. Adsorption of oxazole and isoxazole on BNNT surface: A DFT study. Appl. Surf. Sci. 2015, 328, 632–640. [Google Scholar] [CrossRef]
  38. Liu, Z.; Fan, A.C.; Rakhra, K.; Sherlock, S.; Goodwin, A.; Chen, X.; Yang, Q.; Felsher, D.W.; Dai, H. Supramolecular stacking of doxorubicin on carbon nanotubes for in vivo cancer therapy. Angew. Chem. Int. Ed. Engl. 2009, 48, 7668–7672. [Google Scholar] [CrossRef] [PubMed]
  39. Qin, W.; Li, X.; Bian, W.-W.; Fan, X.-J.; Qi, J.-Y. Density functional theory calculations and molecular dynamics simulations of the adsorption of biomolecules on graphene surfaces. Biomaterials 2010, 31, 1007–1016. [Google Scholar] [CrossRef] [PubMed]
  40. Geerlings, P.; De Proft, F.; Langenaeker, W. Conceptual density functional theory. Chem. Rev. 2003, 103, 1793–1874. [Google Scholar] [CrossRef] [PubMed]
  41. 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. Gaussian16 Revision B.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  42. Sun, H.; Kabb, C.P.; Dai, Y.; Hill, M.R.; Ghiviriga, I.; Bapat, A.P.; Sumerlin, B.S. Macromolecular metamorphosis via stimulus-induced transformations of polymer architecture. Nat. Chem. 2017, 9, 817–823. [Google Scholar] [CrossRef] [PubMed]
  43. Souza, B.S.; Mora, J.R.; Wanderlind, E.H.; Clementin, R.M.; Gesser, J.C.; Fiedler, H.D.; Nome, F.; Menger, F.M. Transforming a Stable Amide into a Highly Reactive One: Capturing the Essence of Enzymatic Catalysis. Angew. Chem. Int. Ed. 2017, 56, 5345–5348. [Google Scholar] [CrossRef] [PubMed]
  44. Mora, J.R.; Cervantes, C.; Marquez, E. New Insight into the Chloroacetanilide Herbicide Degradation Mechanism through a Nucleophilic Attack of Hydrogen Sulfide. Int. J. Mol. Sci. 2018, 19, 2864. [Google Scholar] [CrossRef] [PubMed]
  45. McQuarrie, D.A. Statistical Mechanics; University Science Books: Sausalito, CA, USA, 2000; ISBN 978-1-891389-15-3. [Google Scholar]
  46. Tan, T.T.M.; Rode, B.M. Molecular modelling of polymers, 3. Prediction of glass transition temperatures of poly(acrylic acid), poly(methacrylic acid) and polyacrylamide derivatives. Macromol. Theory Simul. 1996, 5, 467–475. [Google Scholar] [CrossRef]
  47. Scaranto, J.; Mallia, G.; Harrison, N.M. An efficient method for computing the binding energy of an adsorbed molecule within a periodic approach. The application to vinyl fluoride at rutile TiO2(110) surface. Comput. Mater. Sci. 2011, 50, 2080–2086. [Google Scholar] [CrossRef]
  48. Chunsrivirot, S.; Trout, B.L. Free Energy of Binding of a Small Molecule to an Amorphous Polymer in a Solvent. Langmuir 2011, 27, 6910–6919. [Google Scholar] [CrossRef] [PubMed]
  49. Gutowski, M.; Van Lenthe, J.H.; Verbeek, J.; Van Duijneveldt, F.B.; Chałasinski, G. The basis set superposition error in correlated electronic structure calculations. Chem. Phys. Lett. 1986, 124, 370–375. [Google Scholar] [CrossRef]
  50. Kruse, H.; Grimme, S. A geometrical correction for the inter- and intra-molecular basis set superposition error in Hartree-Fock and density functional theory calculations for large systems. J. Chem. Phys. 2012, 136, 154101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Ghose, A.K.; Crippen, G.M. Atomic Physicochemical Parameters for Three-Dimensional Structure-Directed Quantitative Structure-Activity Relationships I. Partition Coefficients as a Measure of Hydrophobicity. J. Comput. Chem. 1986, 7, 565–577. [Google Scholar] [CrossRef] [Green Version]
  52. Zheng, Y.-Z.; Zhou, Y.; Liang, Q.; Chen, D.-F.; Guo, R.; Lai, R.-C. Hydrogen-bonding Interactions between Apigenin and Ethanol/Water: A Theoretical Study. Sci. Rep. 2016, 6, 34647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Magdaline, J.D.; Chithambarathanu, T. Natural bond orbital analysis and vibrational spectroscopic studies of 2-furoic acid using density functional theory. Appl. Phys. 2012, 50, 7. [Google Scholar]
  54. Niklasson, A.M.N.; Challacombe, M. Density Matrix Perturbation Theory. Phys. Rev. Lett. 2004, 92. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Geometrical structure for the minimum found at the wb97xd/6-31++G(d,p) level of theory for antimalarial compounds as well as the acrylamide dimer and trimer used as model of the polymer.
Figure 1. Geometrical structure for the minimum found at the wb97xd/6-31++G(d,p) level of theory for antimalarial compounds as well as the acrylamide dimer and trimer used as model of the polymer.
Processes 07 00396 g001
Figure 2. Minimum energy geometry for complexes antimalarial-hydrogel model at the wb97xd/6-31++G(d,p) theory level. Dashed line shows the short-range interactions.
Figure 2. Minimum energy geometry for complexes antimalarial-hydrogel model at the wb97xd/6-31++G(d,p) theory level. Dashed line shows the short-range interactions.
Processes 07 00396 g002
Figure 3. Frontier orbital for the all molecule studies herein. HOMO: High-occupied molecular orbital; LUMO: low unoccupied molecular orbital.
Figure 3. Frontier orbital for the all molecule studies herein. HOMO: High-occupied molecular orbital; LUMO: low unoccupied molecular orbital.
Processes 07 00396 g003
Table 1. Geometry parameters for complexes DAA-CQ, DAA-PQ and DAA-AQ at wb97xd/6-31++G(d,p).
Table 1. Geometry parameters for complexes DAA-CQ, DAA-PQ and DAA-AQ at wb97xd/6-31++G(d,p).
ComplexBond Length Å
DAA-CQ53O-20H53O-29H53O-14H53O-2C49N-14H49N-50H
2.4503.1813.3641.2303.5421.89
DAA-PQ5O-59H5O-31H5O=4C58O-59H19N-20H55N-20H
2.8772.8532.2330.9771.0252.104
DAA-AQ5O-50H5O-48H49O-50H49O-2H1N-2H1N-50H
2.7382.9960.9902.2921.0142.679
TAA-CQ76O-13H76O-75C17N-78H77N-78H77N-13H77N-75C
2.3392.2332.0041.0273.1491.356
TAA-PQ65O-37H65O-9H650-64C36O-62H36O-37H36O-6C
2.7382.7521.2422.6820.9891.354
TAA-AQ75O-29H75O-27H28O-29H28O-77H76N-29H76N-77H
2.6923.1770.9942.2972.6871.021
Table 2. Geometry energy values from drugs (EF), hydrogel model (EH), complex (EC) and binding energy (ΔE) corrected with basis-set superposition error (BSSE) at the wb97xd/6-31++G(d,p) level of theory.
Table 2. Geometry energy values from drugs (EF), hydrogel model (EH), complex (EC) and binding energy (ΔE) corrected with basis-set superposition error (BSSE) at the wb97xd/6-31++G(d,p) level of theory.
Complex10−5 EF (kcal/mol)10−5 EH (kJ/mol)10−5 EC (kJ/mol)ΔEb (kJ/mol)
DAA-CQ−5.21−3.11−11.43−03.15
DAA-PQ−4.92−3.11−8.030−06.51
DAA-AQ−9.25−3.11−1.230−09.78
TAA-CQ−8.31−4.66−12.98−05.96
TAA-PQ−4.92−4.66−22.78−13.20
TAA-AQ−9.25−4.66−13.91−11.51
Table 3. Molecular descriptors from frontier orbitals, for the complex drugs-hydrogel model at wb97xd/6-31++G(d,p).
Table 3. Molecular descriptors from frontier orbitals, for the complex drugs-hydrogel model at wb97xd/6-31++G(d,p).
DescriptorsDAA-CQTAA-CQDAA-PQTAA-PQDAA-AQTAA-AQ
HOMO (kJ/mol)−125.330−135.991−120.711−116.332−124.690−129.202
LUMO (kJ/mol)−40.420−36.712−29.243−26.882−33.200−35.202
-(HOMO-LUMO) (kJ/mol)84.90199.28191.47289.45391.49194.002
Global Hardness(n) (kJ/mol)42.45149.64545.73144.72145.75047.001
Electronic potential (µ)−82.881−86.353−74.971−71.612−78.941−82.201
ΔN (Hydrogel-Drugs) (kJ/mol)−0.0403−0.036−0.111−0.099−0.040−0.032
Table 4. The second order perturbation theory for the drugs-Acrylamide trimer by using Fock matrix from Natural Bond Orbital (NBO) calculation at the wb97xd /6-31G++(d,p) theory level.
Table 4. The second order perturbation theory for the drugs-Acrylamide trimer by using Fock matrix from Natural Bond Orbital (NBO) calculation at the wb97xd /6-31G++(d,p) theory level.
ComplexDonor (i)Acceptor (j)E2ij (kJ/mol)
TAA-CQLP(1) 76NBD* 77N-H7814.39
LP(1) ClBD* 11C-H134.140
LP(2) ClBD* 11C-H138.870
TAA-PQLP(1) 65OBD* 36O-H3729.12
LP(2) 65OBD* 36O-H3780.50
LP(2) 36OBD* 61C-H622.170
TAA-AQLP(1) 65OBD* 36O-H3729.46
LP(2) 65OBD* 36O-H3797.74
LP(2) 36OBD* 61C-H622.180
Table 5. Quantitative structure–activity relationship (QSAR) properties of complex TAA-antimalarial, in water environment.
Table 5. Quantitative structure–activity relationship (QSAR) properties of complex TAA-antimalarial, in water environment.
DescriptorTAA-AQTAA-CQTAA-PQ
LogP−7.550−10.00−13.48
Gibbs free energy in water (kJ/mol)−146.54−135.75−125.23
Dipole momentum, µ, (Debye)6.94006.51907.4930
Stabilization energy (EH2O-Egas) kJ/mol−24.96−31.32−22.27

Share and Cite

MDPI and ACS Style

Cortes, E.; Márquez, E.; Mora, J.R.; Puello, E.; Rangel, N.; De Moya, A.; Trilleras, J. Theoretical Study of the Adsorption Process of Antimalarial Drugs into Acrylamide-Base Hydrogel Model Using DFT Methods: The First Approach to the Rational Design of a Controlled Drug Delivery System. Processes 2019, 7, 396. https://doi.org/10.3390/pr7070396

AMA Style

Cortes E, Márquez E, Mora JR, Puello E, Rangel N, De Moya A, Trilleras J. Theoretical Study of the Adsorption Process of Antimalarial Drugs into Acrylamide-Base Hydrogel Model Using DFT Methods: The First Approach to the Rational Design of a Controlled Drug Delivery System. Processes. 2019; 7(7):396. https://doi.org/10.3390/pr7070396

Chicago/Turabian Style

Cortes, Eliceo, Edgar Márquez, José R. Mora, Esneyder Puello, Norma Rangel, Aldemar De Moya, and Jorge Trilleras. 2019. "Theoretical Study of the Adsorption Process of Antimalarial Drugs into Acrylamide-Base Hydrogel Model Using DFT Methods: The First Approach to the Rational Design of a Controlled Drug Delivery System" Processes 7, no. 7: 396. https://doi.org/10.3390/pr7070396

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