Open access peer-reviewed chapter

Aggregation of Cyclodextrins: Fundamental Issues and Applications

Written By

Tânia F.G.G. Cova, Sandra M.A. Cruz, Artur J.M. Valente, Paulo E. Abreu, Jorge M.C. Marques and Alberto A.C.C. Pais

Submitted: 24 May 2017 Reviewed: 05 January 2018 Published: 18 April 2018

DOI: 10.5772/intechopen.73532

From the Edited Volume

Cyclodextrin - A Versatile Ingredient

Edited by Poonam Arora and Neelima Dhingra

Chapter metrics overview

1,375 Chapter Downloads

View Full Metrics

Abstract

The aggregation of cyclodextrins (CD) in aqueous solution is an old, yet still vastly unexplored topic that has been studied at least since the 1980s. At that time, few authors took into consideration the possibility of formation of aggregates for the interpretation of thermodynamic and thermophysical properties of CDs in aqueous solution. The aggregates appear at quite low CD concentrations and seem to encompass only a small number of CD molecules. They also occur in water in the presence of hydrophobic or amphiphilic moieties, including surfactants, assuming a preassembled state with the hydrophobic chains threading through one or two CDs. After a long period in which it has been neglected, CD aggregation is now a hot topic and one far from gathering consensus. In this chapter, a timely and critical review on the phenomenon of CD aggregation and the respective supramolecular properties, including some computational rationales, will be presented. A comprehensive summary of CD aggregates studied to date, indicating the formation conditions, characterization techniques, and applications, is also provided.

Keywords

  • cyclodextrins
  • self-aggregation
  • molecular dynamics simulations

1. Introduction

The aggregation of amphiphilic molecules in solution, including cyclodextrins (CD), can be considered as a microphase separation between polar and nonpolar phases. This involves both the hydrophobic and hydrophilic moieties of such molecules and induces the formation and growth of the so-called self-assembled structures. This is a spontaneous process in which the system components form ordered aggregates, typically involving an enthalpy gain in solvation, due to hydrogen bond formation, and possibly a gain in the entropy of bulk water (hydrophobic effect) [1, 2, 3]. The process is promoted by balanced attractive and repulsive interactions between molecules, or specific moieties, occurring from a less organized state (e.g., a solution or a disordered aggregate) to a final ordered state (e.g., a crystal). These soft interactions are generally weak, comparable to thermal energies, and noncovalent, with typical strengths varying from less than 5 kJ mol−1 for van der Waals forces to ca. 120 kJ mol−1 for hydrogen bonds [4]. Other relevant factors, such as the reversibility and flexibility, the interaction with the respective environment, and the mobility of the system components, determine the success of the molecular self-assembly process.

Despite a substantial amount of data is available on the thermodynamic aspects of assembly processes, the information on the CD aggregation behavior and detailed mechanisms of formation and dissociation is scarce. In particular, little is known about the cooperative nature of individual contributions of components for the CD aggregation/assembly process. This is clearly a topic requiring fundamental research, with direct implications for improving the supramolecular assembly design and fostering new formulation opportunities in this type of systems.

Table 1 summarizes some relevant physical-chemical parameters of the three most common natural cyclodextrins (α-, β-, and γ-CDs). CDs are able to solubilize, protect, or transport a large variety of compounds, e.g., inorganic nonassociated salts [5, 6], surfactants [7, 8, 9], and nonpolar drugs, by forming inclusion and noninclusion complexes. These CD-guest interactions are facilitated by the unique structure of CD, which combines a hydrophilic outer surface with an internal lipophilic cavity. The extensive development of CD-based formulations has resulted in a strongly increasing amount of studies, in which the exceptional features of CDs have been described (see e.g., [10, 11, 12, 13, 14, 15, 16]). In the last decades, CDs have also been used for functionalization of organic [17] and inorganic [18, 19] materials and for the synthesis of nanosized and microsized aggregates [20, 21, 22]. This process is, in general, driven by hydrophobic interactions and is affected by a number of factors such as the concentration of reactants, temperature, nature of solvent, and the addition of neutral or ionic co-solutes [23, 24, 25].

Propertyα-CDβ-CDγ-CD
N° glucose unites678
MW (g mol−1)97211351297
Solubility in water (g L−1)14518.5232
Outer diameter cavity, wide end (Å)13.715.316.9
Inner diameter cavity, wide end (Å)5.77.89.5
Volume cavity (Å3)174262427
No. water molecules* inside cavity5.88.714.2

Table 1.

Some physical properties of α-, β- and γ-cyclodextrins [26, 33].

Assuming a volume of a water molecule equal to 30 Å3.


One of the most intriguing properties of CDs is the respective solubility. It should be noted that the solubility in water does not follow a common trend, with β-CD being much less soluble than the other two [26]. Such behavior has attracted researchers for several years, and different explanations have been proposed, based on the rigidity of the CD ring and intramolecular H-bonding or CD self-aggregation [27, 28]. The formation of aggregates can be enhanced upon formation of inclusion complexes, but in general, negligible amounts of aggregates (of ca. 200–300 nm in size) are formed in pure CD solutions. The extent of aggregation may also increase with increasing CD concentration [29].

A wide range of experimental data generated on CD aggregation have been reviewed (see e.g., Refs. [2, 3, 30]). However, the mechanisms of CD aggregation and CD complex aggregation have not yet been fully understood.

CD aggregation was originally recognized based on random observations and deviations from theoretical expectations. Several analytical techniques are now being optimized to specifically detect and characterize CD aggregates. In CD solutions, dynamic light scattering (DLS) [31, 32] and transmission electron microscopy (TEM) [32] have provided reliable evidence of the aggregation phenomenon. The latter has been recognized by several research groups that identified aggregates of different shapes, including spherical and elongated particles, welded fibers, and rods [1, 2, 19, 29].

A significant amount of microscopic data [1] and information from dynamic light scattering (DLS), nuclear magnetic resonance spectroscopy (NMR), and dialysis membrane permeation methods have been reported [1, 3]. However, the full characterization of the CD aggregation behavior by these techniques has been hampered by the poor stability of the aggregates, which makes them sensitive to disruptive side effects [2, 32].

Computational methods, including molecular dynamics (MD) simulations, have been increasingly recognized as a valuable tool for the interpretation of experimental data [34, 35, 36, 37, 38, 39]. The simulation of host-guest systems involving CD molecules and aggregates is already offering ways to explore a set of favorable events at different scales. The aggregation behavior and the interaction patterns of CDs, modified with both hydrophobic and hydrophilic groups, have been investigated by atomistic simulations in nonpolar solvents and in water [34, 35]. Although most studies involving amphiphilic CDs have been conducted in aqueous solutions, the use of nonpolar solvents is also important in preparative and characterization procedures, as suggested in Ref. [34]. In the latter, dynamic light scattering observations demonstrated that CDs can form well-defined aggregates in nonpolar solvents (e.g., dichloromethane), displaying a hydrodynamic radius of ca. 80 nm and low polydispersity [34]. Electrostatic, dipolar, and dispersion forces have been recognized as playing a definite role in the formation and stabilization of these aggregates. The formation of even larger structures (such as micelles, vesicles, membranes, and nanoparticles) based on these smaller aggregates and their potential uses in pharmaceutical formulations have also been suggested.

Advertisement

2. General aspects

2.1. Controversial experimental evidence

The aggregation of single CDs and CD-guest complexes, in water, has been described in a number of recent studies [2, 3, 13, 31, 37, 40, 41]. Despite the efforts for understanding the factors that govern inclusion/binding with guest molecules, the precise manner in which CD molecules aggregate and the cooperative effects underlying this phenomenon are much less studied and still far from consensual.

The classical assumption states that inclusion complexes between CDs and guest molecules are always formed in “ideal” solutions, with individual complexes, independent of each other. However, the treatment of the interaction between system components is oversimplified, since CDs can form both inclusion and noninclusion complexes and water-soluble aggregates. In some cases, the aggregation results in the opalescence of aqueous CD solutions. However, the reduced diameter of the aggregates is smaller than the wavelength of visible light, and more often, the formation of clear solutions is promoted.

The first reference to the occurrence of self-aggregation dates back to 1983, when Koichiro and co-workers [42], based on viscosity and activity coefficient data of aqueous solutions of α- and γ-cyclodextrins, suggested the occurrence of dimers or larger aggregates. These authors have also pointed out that such aggregates are acting as “structure making” [43] by tightening water-water hydrogen bonding. Ten years later, Häusler and Müller-Goymann [44] have observed that at concentrations above 50% (w/w), hydroxypropyl (HP)-β-CD self-aggregates, leading to an increase in the solution viscosity. They also found that the addition of chaotropic solutes (e.g., urea and NaCl) tends to decrease the viscosity of solutions as a consequence of CD disaggregation. A similar observation was found for γ-CD (10% m/v) solutions [45], at physiological pH. In both works, the self-aggregation of CD monomers is justified by intermolecular H-bonding interactions promoted by the hydrophilic rims [44, 45].

Coleman et al. [46] extended their studies to β-CD, the less soluble of all native CDs, and found that the addition of water structure-breaking solutes or an increase of the solution pH (at values higher than 12.5) in order to ionize the ─OH groups leads to an increase in solubility. This evidence led them to hypothesize that the solubility of CDs is related with interactions between CD aggregates and the water through the formation of two chains of hydrated β-CD molecules forming a rod-like aggregate. These authors also argue that the structure of water has an important role on the aggregation/disaggregation of CDs.

The quantification of size and mass percentage of aggregates was only made possible, in a systematic manner, during the last decade. It has been concluded, by using photon correlation spectroscopy (PCS), that in a 12 mM CD aqueous solution, large polydisperse aggregates of 200–300 nm in size were formed [29]; however, the respective mass percentage is quite small when compared to that of free CD. For example, the mass contribution of α-CD aggregates is ca. 0.8% (0.096 mM) assuming coils or 0.001% (0.12 μM) considering spheres [29]. For other CDs, the mass contribution of aggregates is also residual, i.e., 0.0011% (0.13 μM) for β-CD [47] and 0.02% (0.154 mM) for γ-CD solutions [45]. He et al. [48] found a bimodal distribution in dynamic light scattering (DLS) data of α-, β-, and γ-CD aqueous solutions, with mean hydrodynamic radii of less than 1 nm and higher than 60 nm for the fast and slow components (attributed to monomer and aggregated CDs, respectively). This has been revisited by Puskás et al. [49] by using the same technique. However, contrarily to previous work, only γ-CD aggregates have been confirmed. In other studies [50, 51], the formation of CD aggregates of globular shape, at [β-CD] = 3 mM, with a minimum hydration radius of ca. 90 nm, was reported, based on data from different techniques, including DLS, cryo-TEM, and electron spin resonance (ESR) probe spectroscopy. However, at higher CD concentrations, these particles coexist with other structures as large as a few micrometers. These aggregates tend to increase with CD concentration, suggesting cooperative aggregate-aggregate interactions. In addition, Rao and Geckeler [52] have concluded that the formation of β-CD supramolecules can also occur in aqueous solutions at room temperature. By increasing the stirring time and the concentration from 4 to 10 mM, it was possible to follow the formation of cage-like structures (with an average size of 7.5 nm), after 5 h stirring, evolving to channeled structures (with 260 nm length) after 72 h stirring.

The thermodynamic parameters of β-CD self-assembly have been studied, assuming that CDs behave as colloids [36]. For that, the scattering intensity of a set of β-CD solutions of different concentrations has been measured; by plotting the scattering intensity as a function of β-CD concentration, two different regimes were observed below and above the critical aggregation concentration (cac = 1.6 mM): at c < cac, and only monomers (the occurrence of small aggregates cannot be also neglected) exist in solution. However, at c > cac, the scattering intensity increases linearly with β-CD concentration, indicating the presence of two different set of aggregates of 60.0 nm and 120 nm in size. Using the pseudo-phase model, the free energy of aggregation was calculated to be −15.95 kJ mol−1 at 298.15 K. From the dependence of the cac value on the temperature, the variation of the aggregation enthalpy and, consequently, the aggregation entropy (at 298.15 K) were estimated to be −26.48 kJ mol−1 and −35.32 J K−1 mol−1, respectively. These values for thermodynamic parameters suggest that (a) aggregation is enthalpy driven, and (b) aggregates are formed by noncovalent interactions [53]. Following the same approach, the cac values of α-, β-, and γ-CDs were also measured by using a permeability technique, based on dialysis membranes with molecular weight cut off higher than 2 kDa [3, 54]. Inspecting the flux of CDs as a function of concentration, two different regimes have been found which were attributed to the formation of aggregates. The calculated values for the cac of α-, β-, and γ-CDs are 12.2, 6.1, and 7.2 mM, respectively.

Recently, a thermodynamic study [55] using isodesmic and K2-K self-assembly models was performed on CD derivatives (HP-β-CD, HP-γ-CD, and sulfobutylether(SBE)-β-CD). The isodesmic model assumes that the Gibbs energy and equilibrium constant (K) are equal in each addition of a monomer to an aggregate (for details see Ref. [55]). The K2-K model is a modified version of the former, in which the K value of the first step of the self-assembly is different from those of the remaining steps. The respective cac values and aggregate sizes were also determined, using DLS and 1H NMR and TEM, respectively. These CDs displayed similar cac values of ca. 2% (m/v). Three different groups of particle sizes were also identified based on correlation functions: (i) a group with ca. 1 nm corresponds to a single CD molecule, (ii) a group with size values ranging from ca. 30 nm to 60 nm for HP-γ-CD and from ca. 10–70 nm for SBE-β-CD and from ca. 100 nm to 200 nm for HP-β-CD, and (iii) a small group with larger aggregates attaining 1 μm, for β-CD derivatives, and from ca. 140 nm to 1 μm for HP-γ-CD. It was shown that aggregation results from some cooperative contributions, with the first step of the aggregation being less favorable than the subsequent ones. A thermodynamic analysis indicated that the aggregation process was spontaneous, exothermic, and associated to an entropy loss. The calculated standard free energies range from −7.1 kJ mol−1 for SBE-β-CD to −10.6 for HP-γ-CD, and the enthalpy values were −20.6, −27.5, and −46.3 kJ mol−1 for HP-β-CD, SBE-β-CD, and HP-γ-CD, respectively [55].

The dynamics of CDs in aqueous solution has been fully assessed using different NMR techniques by Valente et al. [56, 57]. The analysis of 1H NMR self-diffusion of deuterium solutions of CDs shows that, for all three natural CDs, diffusion coefficients depend linearly on the CD volume fraction, suggesting that these molecules are behaving as nonaggregate hard spheres. A similar conclusion was reached from the analysis of the dependence of mutual diffusion coefficients on CDs concentration [58, 59, 60].

A more sensitive parameter, related to the volume of the diffusing particle, is the transverse magnetization relaxation time T2 or the spin–spin relaxation rate R2 (=1/T2). In a perfectly homogeneous magnetic field, the R2 relaxation rate can be measured directly from the free induction decay in the time domain or the full width at half height of the resonance in the frequency domain. The dependence of R2 as a function of CD volume showed that no aggregation occurs. However, the presence of more transient aggregates cannot be excluded for cases in which the lifetime of the aggregate is short compared to the respective tumbling time. The presence of very large aggregates, not visible in the NMR spectra on account of their slow rotational tumbling, cannot also be ruled out. A different but complementary aspect is related to the mechanism of interaction. For further insight, the aggregation of α-, β-, and γ-CDs in aqueous solutions was addressed by focusing on the CD-CD interactions using deuterium relaxation rates (R1) for deuterium-labeled CDs. In this particular case, the dependence of T1 (=1/R1) on the CD concentration, for all CDs, was explained by the equilibrium between monomeric and dimeric CDs and, again, no evidence in favor of large aggregates of CDs involving a nonnegligible fraction was found [57].

The formation of aggregates in aqueous solutions containing CDs can be promoted by the presence of guest compounds, which upon inclusion can also contribute to understand and predict the CD aggregation behavior. The structure and nature of the guest molecule can thus affect the CD aggregation process. CD-guest complexes are, most often, simply formed by one guest molecule and one CD molecule. However, ternary complexes are also frequently described [61], where water-soluble polymers [62, 63], metal ions, or organic salts [64] are used to potentiate some CD effect.

The coexistence of inclusion and noninclusion complexes in aqueous solutions containing CDs has been documented and associated with the formation of aggregates based on these complexes. The first evidence for aggregation with complexes involving CDs and lipophilic guests was reported by Mele and co-workers [65] in 1998. Later studies [32], with contradictory or ambiguous results, fostered further investigation on this phenomenon.

The formation of CD-guest-based aggregates in the nanosize range has been confirmed by DLS and TEM analyses and associated to a negative deviation from linearity. This type of aggregation occurred for a CD concentration of ca. 10% (wt/vol) [32]. The self-association of CD complexes can explain the observed decrease in the activity coefficient with increasing CD concentration and the dependence of the complex stoichiometry on the method applied.

Several model compounds [32] have been recently used to investigate the effect of the physicochemical properties of the guest molecules on the CD aggregation behavior. For instance, the impact of a set of esters of para-hydroxybenzoic acid, differing in the side chain length, on HP-β-CD aggregation was recently evaluated resorting to permeation experiments, DLS and MS. The number and size of CD aggregates (<200 nm) increased in the presence of longer guests. However, no clear relation was found between the extent of aggregate formation and the CD concentration.

The premicellar association of inclusion complexes of cationic surfactants and β-CD followed by micellar association of the inclusion complexes has also been suggested based on NMR studies. In addition, micelle-like assemblies with diameters exceeding 200 nm have been observed in aqueous solutions containing trans-β-carotene and β-CD and γ-CD [2].

The structure and size of these CD aggregates are clearly affected by water molecules and hydration shells [2]. Although the ability of CDs to self-assemble to form aggregates is well documented, it has also been shown that the aggregates are very unstable. Attempts to stabilize nanosize self-assembled CD aggregates of the native CDs, and their hydrophilic and monomeric derivatives have not been successful so far.

2.2. Common arrangements in CD aggregates

There are two typical crystal structures for native CDs: cages and channels (see Figure 1) [22, 66]. The cage arrangement occurs when CDs are grouped crosswise, displaying a herringbone pattern (Figure 1a), or are aligned in adjacent layers leading to a brick-like pattern (Figure 1b). In both cases, the formation of inclusion complexes is prevented, as the CD cavities are blocked on both portals by neighboring CDs.

Figure 1.

Schematic representation of packing structures of (a) cage-type; (b) layer-type CD; and (c) head-to-tail channel-type CD crystals. Adapted from Ref. [66].

The channel-type assembly is observed if the CDs are stacked in columns so that cavities are aligned to produce channels (Figure 1c). While the channel arrangement can be converted into a cage structure by water sorption/desorption cycles, with an intermediated amorphous state, the latter usually results from rapid recrystallization of CDs. After reaching sorption equilibrium, CD molecules undergo a slow rearrangement to the cage structure with defined water content [66]. Similar patterns are generally found in the crystal structures of CD inclusion complexes [66]. When the entire guest molecule is small enough to be included inside a single CD cavity, cage-type structures are formed. On the other hand, a channel-type structure is observed in the presence of long-chain molecules as guests (e.g., polymers). Rusa and co-workers [2, 67] have reported the encapsulation of poly(ε-caprolactone) and poly(l-lactic acid) into α- and γ-CDs. The inclusion of a polymeric structure inside the CD cavity induces the formation of channel-like structures, being those with α-CD more stable due to the increase of hydrophobic CD-polymer interactions. By using an appropriate experimental procedure, the authors were able to produce a solid-state channel packing of CDs containing only water molecules inside the cavities [68]. The same group has also observed the inclusion of poly(vinyl alcohol) (PVA) into CD by taking advantage of the freezing–thawing process for PVA gelation [69]. In this process, the gelation of PVA-containing composites occurs, taking into account two different types of cross-linking: (i) the hydrogel-bonding naturally observed during the freezing-thawing process and (ii) the CD-CD aggregation resulting from channel-type arrangements [69].

2.3. Applications of CD aggregates

As already mentioned in previous sections, a variety of CD-based aggregates can be formed under different conditions (e.g., concentration, solvent medium, and temperature). These include native and modified CDs, inclusion complexes and the respective aggregates and also rotaxanes and polyrotaxanes, nanotubes, and other high-order structures, such as nanospheres and network aggregates [70]. The potential uses of these self-assembled nanomaterials have been explored for advanced applications, ranging from drug solubilization and drug delivery [71], selective binding [72], and controlled adsorption [72]. In pharmaceutical and biomedical fields, it is expected that such applications may include (i) the nanoencapsulation of drugs in the hydrophobic interchain volumes and nanocavities of modified CDs, which can be used as drug carriers or pharmaceutical excipients, (ii) anticancer phototherapy, (iii) gene delivery, and (iv) protection of unstable active components through the formation of inclusion complexes [70]. Several interesting examples of these potential applications have been focused on amphiphilic CDs, which allow to easily modulate both hydrophobic/hydrophilic and self-assembly properties, by grafting different substituents on the portals of native CDs [70, 73]. For instance, supramolecular assemblies based on CD/porphyrin nanoassemblies have been studied in vitro [74] as potential nanotherapeutics in A375 human melanoma cells. Other micellar structures and spherical vesicles based on CD-perylene conjugates have been designed to be included in fluorescence sensory and photoresponsive materials, photoinduced electron transfer systems, and organic electronic devices [75]. The self-aggregation of amphiphilic CDs has also been explored for drug delivery applications, as they are able to capture selectively drug molecules, displaying enhanced solubilization capacity [71]. The affinity of amphiphilic CDs for incorporation in model and biological membranes has also been investigated and explored for preparation of functionalized lipid membranes and improved biomimetic systems [73, 76]. A broader range of potential applications of CD aggregates are compiled in recent publications such as Refs. [70, 73] and references therein.

Advertisement

3. Computational observations

Irrespective of the disparate observations, the self-aggregation capability must be affected by the type of CD present in the aqueous solution. While the aggregates of α-CD and γ-CD can be completely removed by standard filtering procedures, the formation of β-CD aggregates (at least as dimers or possibly larger aggregates) persists in solution, displaying fast aggregation kinetics. This suggests that the hydrophilic CD portals play a definite role in the aggregation process [29]. The disruption of hydrogen bond networks by ionization (or functionalization) increases solubility and may suppress aggregation.

The presence of aggregates in solutions containing structure-breaking solutes, in which the solubility of β-CD is enhanced, has provided new insights into this unusual behavior. The low solubility of β-CD (see Table 1) has been explained by the presence of aggregates and the respective unfavorable interaction with the hydrogen bond network of bulk water [46]. Note that aspects pertaining to the binding energy in the solid state cannot be disregarded. The relevance of understanding the mechanistic details of the CD aggregation phenomena encompasses either controlling/preventing the formation of aggregates (that preclude the development of specific formulations and the product development) or designing novel formulation strategies.

Computer simulations have been used to rationalize the experimental findings concerning recognition [77], inclusion [36, 77, 78, 79], and aggregation [34, 35, 37, 39, 80]. The cooperative binding of at least two CD monomers to a guest molecule has been considered the driving force responsible for self-assembly processes in the construction of CD-based nanoarchitectures [78]. For aggregation processes without using guests, the assembly is usually driven by the hydrophilic portals (in native CDs) and by interactions between substituent chains of CD derivatives with the neighboring cavities of other CDs [78].

The orientational patterns of inter-glucopyranose hydrogen bonds at the secondary portal of β-CD and the respective effect on the CD structure and dimer binding/stability in polar and nonpolar solvents have been explored by van der Spoel et al. [77] in the presence of various guest molecules. It was demonstrated, based on MD simulations and free energy calculations, that polar solvents with stronger hydrogen bond accepting abilities can easily disrupt intermolecular hydrogen bonds, resulting in less stable dimers. Also, the guest models included in the channel-type cavity increase the binding affinity between CD monomers, particularly in polar solvents [77]. Using a similar computational approach, the authors have explored the effect of three different dimerization modes of β-CD molecules and the presence of isoflavone drug analogues in the construction of CD-based nanostructured materials. It was demonstrated that the cooperative binding of CD cavities to guest molecules favors the dimerization process and, consequently, the overall stability and assembly of the CD nanostructures. It was also proved that the desolvation of CD dimers and entropy changes upon complexation cooperatively contributes to the binding process [78].

Another study [37] focused on the spontaneous adsorption of native CDs and the respective aggregates and the related dependence on temperature. It was found that the adsorption of both individual CDs and small CD aggregates (ca. 20 molecules) to the solution/air interface is negligible. The solute-solute interactions were significantly larger for β-CD than for α-CD at 298 K, and the dependence of these interactions on temperature was more relevant for the smaller CD, which displayed a more favorable aggregation at 283 K than at 298 K. The dynamic exchange of hydrogen bonds between the CD hydroxyl groups and the neighboring water molecules indicated a much larger occupancy for individual intramolecular H-bonds in β-CD.

In what follows, the CD-CD interactions [57], for deuterium labeled CDs, in aqueous solutions are further explored by atomistic simulations. Two types of systems are defined, one in which the β-CD is free in water and three others in which two β-CDs are present and may form dimers. In what concerns the latter, these include initial arrangements with proximity of one primary portal and one secondary portal (PS), two primary portals (PP), and two secondary portals (SS), as shown in Figure 2.

Figure 2.

Schematic illustration of the pairwise initial arrangements of β-CDs, with facing primary and secondary portals (PS), two primary portals (PP), and two secondary portals (SS), with the colors used to distinguish β-CD molecules.

The molecular dynamics simulations were performed with Gromacs (version 4.6.5), using the all-atom amber99sb [81] force field and the TIP3P water model. The initial coordinates of the β-CD were extracted from the RCSB protein data bank (PDB code: 1DMB), and partial charges were generated using the R.E.D.D. Server [82].

In each system, the molecules were accommodated in a cubic box (7.5 nm edge-length) containing approximately 13,000 explicit TIP3P water molecules. To obtain a starting configuration, each system was firstly subjected to an energy minimization step. All the calculations were carried out in NPT ensemble with periodic boundary conditions at a constant temperature of 300 K and a pressure coupling of 1.0 bar, respectively, to V-rescale and Berendsen external baths. A standard time step of 2 fs was used for both equilibration and production runs. A cut-off of 0.9 nm was used for calculating the Lennard-Jones interactions. Electrostatic interactions were evaluated using the particle mesh Ewald method [83]. Constraints were applied for bond lengths with the LINCS algorithm [84].

Equilibrium properties, structure, and dynamics of β-CD systems were calculated for the simulation runs of 50 ns after the systems were equilibrated for 2 ns. Geometric clustering was performed to identify dominant CD-CD structures, sampled during the MD simulations. The algorithm for cluster analysis is based on the hierarchical (top-down) approach [85] and allows evaluating the conformational prevalence of each CD-CD structure, by determining dominant clusters based on the root mean square deviation of the atom positions between all pairs of structures. For each CD-CD arrangement, the number of neighboring structures is calculated for RMSD values of 0.35 nm. Figure 3 presents the behavior of CD-CD structures in different simulation runs, each corresponding to a different initial arrangement. For each CD backbone, the center of mass (COM) of the oxygen atoms at the secondary portal (S) was defined as the reference point for evaluating aggregated and nonaggregated structures and the possible rotation or tilt of the CD molecules.

Figure 3.

Distribution of distances between the centers of mass of β-CD molecules, defined by the oxygen atoms of the secondary portals. Right panels illustrate the final conformations for the imposed PS, PP, and SS initial arrangements of β-CD molecules, in aqueous solution, sampled during the MD simulations at 300 K and identified by geometric cluster analysis. The color codes for CD molecules are as in Figure 1, while the initial arrangements are represented in black, green, and red, for PS, PP, and SS, respectively.

For PS and PP as initial arrangements, a significant evolution is observed in the relative positioning of the two molecules. The PS and PP initial arrangements display an almost complete rotation or a tilt of one molecule with respect to the other, leading to most favorable SS and PS arrangements, respectively. The “intermediate” PS arrangement increases the CD-CD interactions through partial inclusion of some P groups in the hydrophobic cavity of the other CD molecule. The initial SS arrangement prevails over the course of the simulation with a typical COM distance of 0.46 nm. In addition to intramolecular hydrogen bonds, the two CD molecules can form additional intermolecular hydrogen bonds, optimizing the CD-CD interaction. PS is an intermediate arrangement between the most (SS) and the least stable (PP) arrangements. The PP arrangement of the CD pair involves weaker interactions between P groups of the two molecules, producing a low prevalence, relatively open aggregate (COM distance of 0.8 nm), suggesting a relatively poor clustering of this dimeric aggregate.

The rotational autocorrelation functions (ACF) corresponding to the motion of each CD, free or mostly in dimer arrangement, were also inspected. Two alternatives were tested for the definition of the molecule fixed rotating vector defined by the C2-D (see Figure 4) bond (vector) and resulting from the use of three atoms (triplet), those of the C2-D bond and the adjacent carbon C3, and defined as the cross product of vectors C3-C2 and C2-D.

Figure 4.

Cyclodextrin structures (α-CD, n = 6; β-CD, n = 7; and γ-CD, n = 8).

The ACF curves are similar for these two cases and are represented in Figure 5. The curves were fitted from 0 up to 500.0 ps to a one-parameter exponential and reflect the slower decay for the dimer situation, τ0 = 1250 ps, much larger than for the free CD, τ0 = 448 ps. These values are of order of magnitude of those experimentally obtained.

Figure 5.

Average autocorrelation function of C2-D bond for CD molecules in the monomeric and dimeric states, in water at 300 K (black and orange curves, respectively).

The organization of CD molecules in aggregates when in the presence of guest-entities clearly deserves further attention. As an example of the behavior found for more complex systems, simulations were performed to study the inclusion complex of β-CD and poly(vinyl alcohol) (PVA) molecules in water. The importance of this polymer is related to the ability to form hydrogels exhibiting a high degree of swelling in water that has demonstrated a great potential to act as a matrix for many applications, including drug delivery [86], wound dressing [87], and sensors [88]. More recently, it has been found that such broad applications of PVA result from its ability to behave as an amphiphilic polymer [89]. This latter feature is relevant for studying the ability of CDs for forming host-guest or aggregate complexes. In this example, three molecular dynamics simulations were performed in systems containing two β-CD (denoted as A and B) molecules with 1 PVA, 2 PVA (A and B), and 10 PVA (A to J) oligomers, respectively. The main results are illustrated in Figure 6.

Figure 6.

Summary of the MD simulations between β-CD and PVA molecules. Left panels present the measured distances between the center of mass of the groups formed by (i) each β-CD in the dimer (DmAB), (ii) the secondary-secondary portals (SS), the primary-primary portals (DmAB/PP), the secondary-primary portals (PS), DmAB-PVA, DmAB-PVB, CDA-PVA, CDA-PVB, CDB-PVA, CDB-PVB, DmAB-PVA aggregate (Ag(4PVA)), CDA-Ag(4PVA), CDB-Ag(4PVA), and Ag(4PVA)—6 separated PVA molecules (Sg(6PVA)). Right panels illustrate the 3 dominant conformations (accounting for more than 50%, panel a, and 90%, panels b and c, of the occurrences) sampled during the equilibrated parts of the production runs, at 300 K, in aqueous solution. The force field F-85 [90] was used for the CDs. The electrostatic charges of PVA were calculated using the R.E.D. server and the remaining terms generated by ACPYPE [91].

CDs are able to form aggregates at early stages of the simulations, and PVA seems to promote the formation of CD dimers (DmAB). Indeed, PVA contains both hydrophilic and hydrophobic groups that may interact either with the outside part of the CDs or form inclusion complexes; the latter is shown by the snapshot in Figure 6(b), while the former appears for simulations of 2 CDs with 1 PVA (Figure 6(a)) and 2 CDs with 10 PVAs (Figure 6(c)). It is also noted that the prevalence of the type of configurations in Figure 6(b) and Figure 6(c) across the simulation is very high (above 95%), which is a strong indication that such complexes are stable; for 2 CDs with 1 PVA, even though the prevalence of the typical configuration represented in Figure 6(a) is above 50%, it is much less than the other two cases. Nonetheless, it is apparent from the distance versus time plot in Figure 6(b) that the PVA is able to exit the CD pocket in some instances and, then, reform the inclusion complex. In turn, it is particularly interesting to notice from the simulation of 2 CDs with 10 PVAs that the concomitant aggregation of the CDs with various molecules of PVA appears to be relatively stable complex and likely to delay or even prevent the formation of inclusion complexes (not observed during the course of the simulation).

Advertisement

4. Concluding remarks

CD aggregation is still a fertile ground for research, both from the experimental and computational points of view. It was seen that conflicting evidences are provided experimentally, calling for a comprehensive explanation of the involved phenomena, and that detailed, interaction-based, mechanistic rationales from computation are still much needed. It is expected that, in the next years, this subject will benefit from substantial efforts on both grounds.

Advertisement

Acknowledgments

The Coimbra Chemistry Centre is supported by the Fundação para a Ciência e a Tecnologia (FCT), Portuguese Agency for Scientific Research, through the Projects PEst-OE/QUI/UI0313/2014 and POCI-01-0145-FEDER-007630. Sandra M.A. Cruz thanks the postdoc grant from Coimbra Chemistry Centre (Contract DPDRH-56-15-401). Tânia F.G.G. Cova also acknowledge the PhD research Grant SFRH/BD/95459/2013 assigned by FCT.

References

  1. 1. Kurkov SV, Loftsson T. Cyclodextrins. International Journal of Pharmaceutics. 2013;453:167-180
  2. 2. Ryzhakov A, Do Thi T, Stappaerts J, Bertoletti L, Kimpe K, Sá Couto AR, Saokham P, Van den Mooter G, Augustijns P, Somsen GW, Kurkov S, Inghelbrecht S, Arien A, Jimidar MI, Schrijnemakers K, Loftsson T. Self-assembly of cyclodextrins and their complexes in aqueous solutions. Journal of Pharmaceutical Sciences. 2016;105:2556-2569
  3. 3. Saokham P, Sá Couto A, Ryzhakov A, Loftsson T. The self-assemble of natural cyclodextrins in aqueous solutions: Application of miniature permeation studies for critical aggregation concentration (cac) determinations. International Journal of Pharmaceutics. 2016;505:187-193
  4. 4. Whitesides GM, Boncheva M. Beyond molecules: Self-assembly of mesoscopic and macroscopic components. Proceedings of the National Academy of Sciences. 2002;99:4769-4774
  5. 5. Assaf KI, Ural MS, Pan F, Georgiev T, Simova S, Rissanen K, Gabel D, Nau WM. Water structure recovery in chaotropic anion recognition: High-affinity binding of dodecaborate clusters to γ-cyclodextrin. Angewandte Chemie International Edition. 2015;54:6852-6856
  6. 6. Terekhova I, Chibunova E, Kumeev R, Kruchinin S, Fedotova M, Kozbiał M, Wszelaka-Rylik M, Gierycz P. Specific and nonspecific effects of biologically active inorganic salts on inclusion complex formation of cyclodextrins with aromatic carboxylic acids. Chemical Engineering Science. 2015;122:97-103
  7. 7. Valente AJM, Söderman O. The formation of host–guest complexes between surfactants and cyclodextrins. Advances in Colloid and Interface Science. 2014;205:156-176
  8. 8. Martín VI, Ostos FJ, Angulo M, Márquez AM, López-Cornejo P, López-López M, Carmona AT, Moyá ML. Host-guest interactions between cyclodextrins and surfactants with functional groups at the end of the hydrophobic tail. Journal of Colloid and Interface Science. 2017;491:336-348
  9. 9. Shang C, Chen Y, Zhang C, Deng Z, Wu Z. Destruction of chitosan capsules based on host-guest interaction and controllable release of encapsulated dyes. Journal of Applied Polymer Science. 2017;134:45229
  10. 10. Venuti V, Rossi B, Mele A, Melone L, Punta C, Majolino D, Masciovecchio C, Caldera F, Trotta F. Tuning structural parameters for the optimization of drug delivery performance of cyclodextrin-based nanosponges. Expert Opinion on Drug Delivery. 2017;14:331-340
  11. 11. Liao R, Lv P, Wang Q, Zheng J, Feng B, Yang B. Cyclodextrin-based biological stimuli-responsive carriers for smart and precision medicine. Biomaterials Science. 2017;5:1736-1745
  12. 12. Abdolmaleki A, Ghasemi F, Ghasemi JB. Computer-aided drug design to explore cyclodextrin therapeutics and biomedical applications. Chemical Biology & Drug Design. 2017;89:257-268
  13. 13. Alsbaiee A, Smith BJ, Xiao L, Ling Y, Helbling DE, Dichtel WR. Rapid removal of organic micropollutants from water by a porous β-cyclodextrin polymer. Nature. 2016;529:190-194
  14. 14. Peng L, Liu S, Feng A, Yuan J. Polymeric nanocarriers based on cyclodextrins for drug delivery: Host–guest interaction as stimuli responsive linker. Molecular Pharmaceutics. 2017;14:2475-2486
  15. 15. Schmidt BVKJ, Barner-Kowollik C. Dynamic macromolecular material design—The versatility of cyclodextrin-based host–guest chemistry. Angewandte Chemie International Edition. 2017;56:8350-8369
  16. 16. Harada A, Kobayashi R, Takashima Y, Hashidzume A, Yamaguchi H. Macroscopic self-assembly through molecular recognition. Nature Chemistry. 2011;3:34-37
  17. 17. Medronho B, Andrade R, Vivod V, Ostlund A, Miguel MG, Lindman B, Voncina B, Valente AJM. Cyclodextrin-grafted cellulose: Physico-chemical characterization. Carbohydrate Polymers. 2013;93:324-330
  18. 18. Krings JA, Vonhoren B, Tegeder P, Siozios V, Peterlechner M, Ravoo BJ. Light-responsive aggregation of [small beta]-cyclodextrin covered silica nanoparticles. Journal of Materials Chemistry A. 2014;2:9587-9593
  19. 19. Polarz S, Smarsly B, Bronstein L, Antonietti M. From cyclodextrin assemblies to porous materials by silica Templating. Angewandte Chemie International Edition. 2001;40:4417-4421
  20. 20. Loftsson T. Self-assembled cyclodextrin nanoparticles and drug delivery. Journal of Inclusion Phenomena and Macrocyclic Chemistry. 2014;80:1-7
  21. 21. Messner M, Kurkov SV, Jansook P, Loftsson T. Self-assembled cyclodextrin aggregates and nanoparticles. International Journal of Pharmaceutics. 2010;387:199-208
  22. 22. He Y, Fu P, Shen X, Gao H. Cyclodextrin-based aggregates and characterization by microscopy. Micron. 2008;39:495-516
  23. 23. Lo Nostro P, Lopes JR, Ninham BW, Baglioni P. Effect of cations and anions on the formation of polypseudorotaxanes. The Journal of Physical Chemistry B. 2002;106:2166-2174
  24. 24. Becheri A, Lo Nostro P, Ninham BW, Baglioni P. The curious world of polypseudorotaxanes: Cyclodextrins as probes of water structure. The Journal of Physical Chemistry B. 2003;107:3979-3987
  25. 25. De Sousa FB, Lima AC, Denadai AML, Anconi CPA, De Almeida WB, Novato WTG, Dos Santos HF, Drum CL, Langer R, Sinisterra RD. Superstructure based on [small beta]-CD self-assembly induced by a small guest molecule. Physical Chemistry Chemical Physics. 2012;14:1934-1944
  26. 26. Sabadini E, Cosgrove T, Egídio FdC. Solubility of cyclomaltooligosaccharides (cyclodextrins) in H2O and D2O: A comparative study. Carbohydrate Research. 2006;341:270-274
  27. 27. Naidoo KJ, Chen JY-J, Jansson JLM, Widmalm G, Maliniak A. Molecular properties related to the anomalous solubility of β-cyclodextrin. The Journal of Physical Chemistry B. 2004;108:4236-4238
  28. 28. Cai W, Sun T, Shao X, Chipot C. Can the anomalous aqueous solubility of [small beta]-cyclodextrin be explained by its hydration free energy alone? Physical Chemistry Chemical Physics. 2008;10:3236-3243
  29. 29. González-Gaitano G, Rodríguez P, Isasi JR, Fuentes M, Tardajos G, Sánchez M. The aggregation of cyclodextrins as studied by photon correlation spectroscopy. Journal of Inclusion Phenomena and Macrocyclic Chemistry. 2002;44:101-105
  30. 30. Chen G, Jiang M. Cyclodextrin-based inclusion complexation bridging supramolecular chemistry and macromolecular self-assembly. Chemical Society Reviews. 2011;40:2254-2266
  31. 31. Zaman H, Bright AG, Adams K, Goodall DM, Forbes RT. Characterisation of aggregates of cyclodextrin-drug complexes using Taylor dispersion analysis. International Journal of Pharmaceutics. 2017;522:98-109
  32. 32. Stappaerts J, Do Thi T, Dominguez-Vega E, Somsen GW, Van den Mooter G, Augustijns P. The impact of guest compounds on cyclodextrin aggregation behavior: A series of structurally related parabens. International Journal of Pharmaceutics. 2017;529:442-450
  33. 33. Szejtli J. Introduction and general overview of cyclodextrin chemistry. Chemical Reviews. 1998;98:1743-1754
  34. 34. Raffaini G, Mazzaglia A, Ganazzoli F. Aggregation behaviour of amphiphilic cyclodextrins: The nucleation stage by atomistic molecular dynamics simulations. Beilstein Journal of Organic Chemistry. 2015;11:2459-2473
  35. 35. Raffaini G, Ganazzoli F, Mazzaglia A. Aggregation behavior of amphiphilic cyclodextrins in a nonpolar solvent: Evidence of large-scale structures by atomistic molecular dynamics simulations and solution studies. Beilstein Journal of Organic Chemistry. 2016;12:73-80
  36. 36. Cova TFGG, Nunes SCC, Pais AACC. Free-energy patterns in inclusion complexes: The relevance of non-included moieties in the stability constants. Physical Chemistry Chemical Physics. 2017;19:5209-5221
  37. 37. Mixcoha E, Campos-Terán J, Piñeiro Á. Surface adsorption and bulk aggregation of cyclodextrins by computational molecular dynamics simulations as a function of temperature: α-CD vs β-CD. The Journal of Physical Chemistry B. 2014;118:6999-7011
  38. 38. López CA, de Vries AH, Marrink SJ. Molecular mechanism of cyclodextrin mediated cholesterol extraction. PLoS Computational Biology. 2011;7:e1002020
  39. 39. Zhu X, Wu G, Chen D. Molecular dynamics simulation of cyclodextrin aggregation and extraction of anthracene from non-aqueous liquid phase. Journal of Hazardous Materials. 2016;320:169-175
  40. 40. Lucio D, Irache JM, Font M, Martínez-Ohárriz MC. Nanoaggregation of inclusion complexes of glibenclamide with cyclodextrins. International Journal of Pharmaceutics. 2017;519:263-271
  41. 41. Ghosh P, Maity A, Das T, Dash J, Purkayastha P. Modulation of small molecule induced architecture of cyclodextrin aggregation by guest structure and host size. The Journal of Physical Chemistry C. 2011;115:20970-20977
  42. 42. Koichiro M, Masahiro S, Masayuki N. Viscosity B-coefficients, apparent molar volumes, and activity coefficients for α- and γ-cyclodextrins in aqueous solutions. Bulletin of the Chemical Society of Japan. 1983;56:3556-3560
  43. 43. Marcus Y. Effect of ions on the structure of water: Structure making and breaking. Chemical Reviews. 2009;109:1346-1370
  44. 44. Häusler O, Müller-Goymann CC. Properties and structure of aqueous solutions of hydroxypropyl-beta-cyclodextrin. Starch—Stärke. 1993;45:183-187
  45. 45. Szente L, Szejtli J, Kis GL. Spontaneous opalescence of aqueous γ-cyclodextrin solutions: Complex formation or self-aggregation? Journal of Pharmaceutical Sciences. 1998;87:778-781
  46. 46. Coleman AW, Nicolis I, Keller N, Dalbiez JP. Aggregation of cyclodextrins: An explanation of the abnormal solubility ofβ-cyclodextrin. Journal of Inclusion Phenomena and Molecular Recognition in Chemistry. 1992;13:139-143
  47. 47. Wu A, Shen X, He Y. Micrometer-sized rodlike structure formed by the secondary assembly of cyclodextrin nanotube. Journal of Colloid and Interface Science. 2006;302:87-94
  48. 48. Wu A, Shen X, He Y. Investigation on γ-cyclodextrin nanotube induced by N,N′-diphenylbenzidine molecule. Journal of Colloid and Interface Science. 2006;297:525-533
  49. 49. Puskás I, Schrott M, Malanga M, Szente L. Characterization and control of the aggregation behavior of cyclodextrins. Journal of Inclusion Phenomena and Macrocyclic Chemistry. 2013;75:269-276
  50. 50. Bonini M, Rossi S, Karlsson G, Almgren M, Lo Nostro P, Baglioni P. Self-assembly of β-cyclodextrin in water. Part 1: Cryo-TEM and dynamic and static light scattering. Langmuir. 2006;22:1478-1484
  51. 51. Rossi S, Bonini M, Lo Nostro P, Baglioni P. Self-assembly of β-cyclodextrin in water. 2. Electron spin resonance. Langmuir. 2007;23:10959-10967
  52. 52. Rao JP, Geckeler KE. Cyclodextrin supramacromolecules: Unexpected formation in aqueous phase under ambient conditions. Macromolecular Rapid Communications. 2011;32:426-430
  53. 53. De M, You C-C, Srivastava S, Rotello VM. Biomimetic interactions of proteins with functionalized nanoparticles: A thermodynamic study. Journal of the American Chemical Society. 2007;129:10747-10753
  54. 54. Saokham P, Loftsson T. A new approach for quantitative determination of γ-cyclodextrin in aqueous solutions: Application in aggregate determinations and solubility in hydrocortisone/γ-cyclodextrin inclusion complex. Journal of Pharmaceutical Sciences. 2015;104:3925-3933
  55. 55. Do TT, Van Hooghten R, Van den Mooter G. A study of the aggregation of cyclodextrins: Determination of the critical aggregation concentration, size of aggregates and thermodynamics using isodesmic and K2–K models. International Journal of Pharmaceutics. 2017;521:318-326
  56. 56. Valente AJM, Carvalho RA, Söderman O. Do cyclodextrins aggregate in water? Insights from NMR experiments. Langmuir. 2015;31:6314-6320
  57. 57. Valente AJM, Carvalho RA, Murtinho D, Söderman O. Molecular dynamics of cyclodextrins in water solutions from NMR deuterium relaxation: Implications for cyclodextrin aggregation. Langmuir. 2017;33:8233-8238
  58. 58. Ribeiro ACF, Leaist DG, Esteso MA, Lobo VMM, Valente AJM, Santos CIAV, Cabral AMTDPV, Veiga FJB. Binary mutual diffusion coefficients of aqueous solutions of β-cyclodextrin at temperatures from 298.15 to 312.15 K. Journal of Chemical & Engineering Data. 2006;51:1368-1371
  59. 59. Ribeiro ACF, Santos CIAV, Valente AJM, Ascenso OS, Lobo VMM, Burrows HD, Cabral AMTDPV, Veiga FJB, Teijeiro C, Esteso MA. Some transport properties of γ-cyclodextrin aqueous solutions at (298.15 and 310.15) K. Journal of Chemical & Engineering Data. 2008;53:755-759
  60. 60. Ribeiro ACF, Valente AJM, Santos CIAV, Prazeres PMRA, Lobo VMM, Burrows HD, Esteso MA, Cabral AMTDPV, Veiga FJB. Binary mutual diffusion coefficients of aqueous solutions of α-cyclodextrin, 2-hydroxypropyl-α-cyclodextrin, and 2-hydroxypropyl-β-cyclodextrin at temperatures from (298.15 to 312.15) K. Journal of Chemical & Engineering Data. 2007;52:586-590
  61. 61. He J, Chipot C, Shao X, Cai W. Cooperative recruitment of amphotericin B mediated by a cyclodextrin dimer. The Journal of Physical Chemistry C. 2014;118:24173-24180
  62. 62. Figueiras A, Sarraguça JMG, Carvalho RA, Pais AACC, Veiga FJB. Interaction of omeprazole with a methylated derivative of β-cyclodextrin: Phase solubility, NMR spectroscopy and molecular simulation. Pharmaceutical Research. 2007;24:377-389
  63. 63. Figueiras A, Nunes SCC, Simões S, Pais AACC, Veiga F. Molecular interaction governing solubility and release profiles in supramolecular systems containing fenbufen, pluronics and cyclodextrins. Journal of Inclusion Phenomena and Macrocyclic Chemistry. 2015;81:395-407
  64. 64. Li K, Wu X, Gu Q, Zhao X, Yuan M, Ma W, Ni W, Hou Z. Inclusion complexes of organic salts with [small beta]-cyclodextrin as organocatalysts for CO2 cycloaddition with epoxides. RSC Advances. 2017;7:14721-14732
  65. 65. Mele A, Mendichi R, Selva A. Non-covalent associations of cyclomaltooligosaccharides (cyclodextrins) with trans-β-carotene in water: Evidence for the formation of large aggregates by light scattering and NMR spectroscopy. Carbohydrate Research. 1998;310:261-267
  66. 66. Hernández R, Rusa M, Rusa CC, López D, Mijangos C, Tonelli AE. Controlling PVA hydrogels with γ-cyclodextrin. Macromolecules. 2004;37:9620-9625
  67. 67. Rusa CC, Fox J, Tonelli AE. Competitive formation of polymer−cyclodextrin inclusion compounds. Macromolecules. 2003;36:2742-2747
  68. 68. Rusa CC, Bullions TA, Fox J, Porbeni FE, Wang X, Tonelli AE. Inclusion compound formation with a new columnar cyclodextrin host. Langmuir. 2002;18:10016-10023
  69. 69. Papancea A, Valente AJM, Patachia S, Miguel MG, Lindman B. PVA−DNA cryogel membranes: Characterization, swelling, and transport studies. Langmuir. 2008;24:273-279
  70. 70. Zerkoune L, Angelova A, Lesieur S. Nano-assemblies of modified cyclodextrins and their complexes with guest molecules: Incorporation in nanostructured membranes and amphiphile nanoarchitectonics design. Nanomaterials. 2014;4:741
  71. 71. Auzély-Velty R, Djedaïni-Pilard F, Désert S, Perly B, Zemb T. Micellization of hydrophobically modified cyclodextrins. 1. Micellar structure. Langmuir. 2000;16:3727-3734
  72. 72. Wang T, Chipot C, Shao X, Cai W. Structural characterization of micelles formed of cholesteryl-functionalized cyclodextrins. Langmuir. 2011;27:91-97
  73. 73. Sliwa W, Girek T. Cyclodextrins: Properties and Applications. Weinheim, Germany: John Wiley & Sons; 2016
  74. 74. Mazzaglia A, Bondì ML, Scala A, Zito F, Barbieri G, Crea F, Vianelli G, Mineo P, Fiore T, Pellerito C, Pellerito L, Costa MA. Supramolecular assemblies based on complexes of nonionic amphiphilic cyclodextrins and a meso-tetra(4-sulfonatophenyl)porphine tributyltin(IV) derivative: Potential nanotherapeutics against melanoma. Biomacromolecules. 2013;14:3820-3829
  75. 75. Jiang B-P, Guo D-S, Liu Y. Self-assembly of amphiphilic perylene−cyclodextrin conjugate and vapor sensing for organic amines. The Journal of Organic Chemistry. 2010;75:7258-7264
  76. 76. Bauer M, Charitat T, Fajolles C, Fragneto G, Daillant J. Insertion properties of cholesteryl cyclodextrins in phospholipid membranes: A molecular study. Soft Matter. 2012;8:942-953
  77. 77. Zhang H, Tan T, Feng W, van der Spoel D. Molecular recognition in different environments: β-cyclodextrin dimer formation in organic solvents. The Journal of Physical Chemistry B. 2012;116:12684-12693
  78. 78. Zhang H, Tan T, Hetényi C, Lv Y, van der Spoel D. Cooperative binding of cyclodextrin dimers to isoflavone analogues elucidated by free energy calculations. The Journal of Physical Chemistry C. 2014;118:7163-7173
  79. 79. Zhang H, Tan T, Hetényi C, van der Spoel D. Quantification of solvent contribution to the stability of noncovalent complexes. Journal of Chemical Theory and Computation. 2013;9:4542-4551
  80. 80. Zhang H, Tan T, van der Spoel D. Generalized born and explicit solvent models for free energy calculations in organic solvents: Cyclodextrin dimerization. Journal of Chemical Theory and Computation. 2015;11:5103-5113
  81. 81. van der Spoel D, van Maaren PJ, Caleman C. GROMACS molecule & liquid database. Bioinformatics. 2012;28:752-753
  82. 82. Vanquelef E, Simon S, Marquant G, Garcia E, Klimerak G, Delepine JC, Cieplak P, Dupradeau F-Y. R.E.D. Server: A web service for deriving RESP and ESP charges and building force field libraries for new molecules and molecular fragments. Nucleic Acids Research. 2011;39:W511-W517
  83. 83. Wong-ekkabut J, Karttunen M. The good, the bad and the user in soft matter simulations. Biochimica et Biophysica Acta (BBA)—Biomembranes. 2016;1858:2529-2538
  84. 84. Hess B, Bekker H, Berendsen HJC, Fraaije JGEM. LINCS: A linear constraint solver for molecular simulations. Journal of Computational Chemistry. 1997;18:1463-1472
  85. 85. Daura X, Gademann K, Jaun B, Seebach D, van Gunsteren WF, Mark AE. Peptide folding: When simulation meets experiment. Angewandte Chemie International Edition. 1999;38:236-240
  86. 86. Das S, Subuddhi U. Exploring poly(vinyl alcohol) hydrogels containing drug-cyclodextrin complexes as controlled drug delivery systems. Journal of Applied Polymer Science. 2014;131:40318(1-13)
  87. 87. Zhou J, Wang Q, Lu H, Zhang Q, Lv P, Wei Q. Preparation and characterization of electrospun polyvinyl alcoholstyrylpyridinium/β-cyclodextrin composite nanofibers: Release behavior and potential use for wound dressing. Fibers and Polymers. 2016;17:1835-1841
  88. 88. Ouyang L, Zhu L, Ruan Y, Tang H. Preparation of a native [small beta]-cyclodextrin modified plasmonic hydrogel substrate and its use as a surface-enhanced Raman scattering scaffold for antibiotics identification. Journal of Materials Chemistry C. 2015;3:7575-7582
  89. 89. Knaapila M, Stewart B, Costa T, Rogers SE, Pragana J, Fonseca SM, Valente AJM, Ramos ML, Murtinho D, Pereira JC, Mallavia R, Burrows HD. Incorporation of a cationic conjugated polyelectrolyte CPE within an aqueous poly(vinyl alcohol) sol. Macromolecules. 2016;49:9119-9131
  90. 90. Cezard C, Trivelli X, Aubry F, Djedaini-Pilard F, Dupradeau F-Y. Molecular dynamics studies of native and substituted cyclodextrins in different media: 1. Charge derivation and force field performances. Physical Chemistry Chemical Physics. 2011;13:15103-15121
  91. 91. Sousa da Silva AW, Vranken WF. ACPYPE—AnteChamber PYthon parser interfacE. BMC Research Notes. 2012;5:367

Written By

Tânia F.G.G. Cova, Sandra M.A. Cruz, Artur J.M. Valente, Paulo E. Abreu, Jorge M.C. Marques and Alberto A.C.C. Pais

Submitted: 24 May 2017 Reviewed: 05 January 2018 Published: 18 April 2018