Open access peer-reviewed chapter

Interactions between Bio-Based Compounds and Cyclodextrins

Written By

Bruno Filipe Figueiras Medronho, Sandra Gonçalves, Raquel Rodríguez-Solana, Artur J.M. Valente and Anabela Romano

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

DOI: 10.5772/intechopen.73531

From the Edited Volume

Cyclodextrin - A Versatile Ingredient

Edited by Poonam Arora and Neelima Dhingra

Chapter metrics overview

1,318 Chapter Downloads

View Full Metrics

Abstract

Bio-based compounds, such as “green” surfactants and phytochemicals, are regarded as future sustainable resources for a vast range of applications in a modern society increasingly demanding economical, social, and environmental awareness. Natural compounds from plants (phytochemicals) are very sought by the pharmaceutical, cosmetic, and food industries. On the other hand, the growing interest in “green” surfactants (e.g., carbohydrate-based) is due to, inter alia, their preparation from renewable raw materials, ready biodegradability, and biocompatibility, among other reasons of fundamental, practical, economical, and environmental orders. Despite the wide range of potential applications of these bio-based compounds, their practical use is still limited due to many reasons such as poor aqueous solubility, volatility, reactivity, etc. Generally, when complexed with cyclodextrins, these bio-based compounds enhance considerably their performance and potential applications. Thus, this chapter aims at recalling some general fundamental aspects of phytochemicals and “green” surfactants, such as structure, function, and applications. In addition, their interactions with cyclodextrins are discussed from a physicochemical point of view with special focus on the techniques, mathematic modeling, and thermodynamic parameters (e.g., interactions, stoichiometries, association constants, etc.).

Keywords

  • sugar-based surfactants
  • phytochemicals
  • essential oils
  • polyphenols
  • cyclodextrins
  • host-guest complex

1. Cyclodextrins: general considerations

Due to its structure, cyclodextrins (CDs) readily form inclusion complexes through noncovalent interactions with molecular guests. The lipophilic cavity of CDs provides a microenvironment into which appropriately sized nonpolar moieties can enter. The hydrophobicity of the cavity enables the accommodation of a broad range of hydrophobic guests such as the alkyl chains of surfactants or different phytochemicals [1]. The hydrophilic exterior usually imparts CDs and their complexes, considerable solubility in water. The charge and polarity of the guest molecule play also an important role in the CD-substrate host-guest interaction [2]. However, this aspect is obviously less important than the geometric fitting. In the case of the charge, the complexation of neutral molecules is easier than the ionized counterpart. In general, molecules can be encapsulated by CDs when they are less hydrophilic or less polar than the solvent and when the formed complex is stable.

The main driving force for the formation of the complex is the release of enthalpy-rich water molecules from the cavity; water molecules are displaced by more hydrophobic guest molecules present in the solution to achieve the apolar-apolar interactions and decrease of CD ring strain resulting in a favorable lower energy state. The beneficial modification of guest molecular properties after the formation of the inclusion complex leads to a large number of applications in areas as diverse as encapsulation of active substances (i.e., flavoring agents, metallic cations, fragrances, and pesticides), enzymatic synthesis, catalysis, and energy transfer studies [3, 4]. Additionally, CDs also find important uses in cosmetics, environment protection, bioconversion, packing, textiles, and food domain [1, 5].

Less than 10% of all produced CDs and CD derivatives are used by the pharmaceutical industry. The largest CD users are the food and the cosmetic industry. CDs have a high level of biocompatibility, are absorbed in the gastrointestinal tract, and are completely metabolized by the colon microflora [6]. Some of them are approved by the Food and Drug Administration or have been accredited as being “generally recognized as safe” (GRAS) [7]. In the cosmetic area, CD performance stands out in the following: solubilize and stabilize specific sensitive components, stabilize emulsions, improve the absorption of active components onto the skin, reduce or eliminate bad aromas from certain components, and reduce the loss of the active components through volatilization, rapid oxidation, destruction by light, etc. [8].

Advertisement

2. Plant phytochemicals: brief introduction to polyphenols and essential oils

Plants are a remarkable source of biologically active compounds with potential applications in cosmetic, pharmaceutical, and food industries. Among the bioactive compounds synthesized as secondary metabolites by plants, phenolic compounds are probably the most relevant ones. Physiologically, they play a vital role in plant protection and contribute to plant odors, plant pigmentation, and/or their flavors. Structurally, phenolic compounds have, at least, one aromatic ring, with one or more hydroxyl groups attached [9]. The great diversity of phenolic compounds present in nature (i.e., more than 8000 different structures have been identified up to now) results from variations in the basic chemical skeleton (e.g., degree of oxidation, hydroxylation, methylation, glycosylation, and conjugation with further molecules, particularly lipids, proteins, other phenolics, and biomolecular metabolites) [9]. Phenolic compounds are grouped by the number of phenol rings they contain and the structural elements that bind these rings to another; flavonoids, phenolic acids, tannins, stilbenes, and lignans are examples of important representatives of these groups [10].

On the other hand, plant essential oils (EOs) are mixtures of numerous highly complex volatile compounds (hydrogenated and oxygenated monoterpenes, sesquiterpenes, phenols, simple alcohols, ketones, coumarins, etc.) present in variable concentrations whose aroma depends on the individual constituents present [11]. Due to their natural properties, EOs have been used as therapeutic remedies and flavoring agents since ancient times. In the last decades, many investigations showed that EOs have a wide range of valuable biological activities, such as antimicrobial, herbicidal, insecticidal, antioxidant, etc.

Although many investigations demonstrated the broad range of biological activities of many phytochemicals, they still have restricted applicability as pharmaceuticals or in food products due to their poor water insolubility and bioavailability, high volatility, rapid oxidation, or degradation when exposed to environmental factors. New approaches have been developed to overcome these drawbacks, and among them, CDs have been suggested as excellent vehicles for the protection of phytochemicals for food and drug delivery proposes [12, 13, 14].

2.1. Interaction between phytochemicals and cyclodextrins

According to a recent review by Suvarna et al., there are many phytochemicals whose solubility, bioavailability, or therapeutic activity is significantly improved by complexation with CDs (e.g., quercetin, curcumin, artemisinin, resveratrol, naringenin, etc.) [12]. The methods used for the formation of inclusion complexes between CDs and bioactive compounds are essentially neutralization, slurry, solution, coprecipitation, kneading, and grinding [15].

The encapsulation of phytochemicals with CDs usually involves the formation of 1:1 inclusion complexes with the most versatile CD, the β-CDs, and its derivatives. These derivatives can be classified according to their interaction with the water molecules in hydrophilic, hydrophobic, and ionizable derivatives [13]. Examples of used hydrophilic β-CDs are the methylated β-CDs—2,6-dimethyl-β-CD (DM-β-CD) and 2,3,6-trimethyl-β-CD (TM-β-CD)—the hydroxyalkylated β-CDs such as 2-hydroxypropyl-β-CD (HP-β-CD), and the branched β-CDs, glycosyl-β-CD (G-β-CD). These molecules are suitable for the formation of host-guest inclusion complexes with poor water-soluble compounds. On the other hand, the hydrophobic derivatives, such as the alkylated β-CD 2,6-diethyl-β-CD (DE-β-CD), are used to decrease and modulate the released rate of water-soluble molecules. Finally, the ionizable β-CD can enhance the dissolution rate and the inclusion capacity and even decrease the side effects of some molecules [16, 17]. Among the ionizable CDs, O-carboxymethyl-β-CD (CM-β-CD), O-carboxymethyl-O-ethyl-β-CD (CME-β-CD), and sulfate and sulfobutylether-β-CD (SBE-β-CD) should be highlighted.

Owing to their potential health promotion effects particularly the antioxidant, anti-inflammatory, and antimicrobial properties, one of the actual promising applications of phenolic compounds is their use in the food industry as additives, e.g., in the development of functional foods. Nevertheless, the efficacy of these natural compounds is dependent on the preservation or improvement of their stability, bioactivity, and bioavailability [18]. Inclusion complexation with CDs improves water solubility of phenolics and enhances their shelf life and biological activity [15]. Additionally, it has been shown that the inclusion of phenolic compounds (e.g., hydroxycinnamic and chlorogenic acids) with CD (β-CD) strongly limited their interactions with proteins, which is important regarding the use of phenolics as food additives [19, 20]. Note that the interactions of these compounds with proteins, frequently added to functional foods to improve nutritional value and proper texture characteristics, often decrease the bioavailability of both proteins and phenolics [21].

There are many examples showing that the complexation of β-CD, or some of its derivatives, increases the biological activity of phenolics. For instance, Shao et al. observed that the complexation of chlorogenic acid (CGA) with CDs (β-CD and HP-β-CD) improved its antioxidant activity [22]. Moreover, the addition of CGA-CD complexes to grape juice reduced the degradation of anthocyanins due to copigmentation effect with the CGA/HP-β-CD complex showing the superior activity and copigmentation effects. Gabaldon et al. also used HP-β-CD to increase the aqueous solubility of kaempferol, quercetin, and myricetin and to improve their antioxidant activity due to the protection toward free radical attack [23]. The complexation of curcumin with an ionizable β-CD (SBE-β-CD) enhanced its water solubility and, thus, improved the in vitro cytotoxic (on HepG-2 cells) and antioxidant activity of these compounds [24]. This β-CD derivative and the HP-β-CD are the most used derivatives on the pharmaceutical industry due to their low toxicity and high solubility [16, 25, 26].

EOs can be regarded as mixtures of phytochemicals, and there are several studies reporting the complexation of EOs or their components with CDs mainly to overcome problems related with EO water insolubility, high volatility, rapid oxidation, heat damage, and degradation on exposure to air [14]. Although many studies focus the complexation of EO components with β-CDs or its derivatives, e.g., eugenol/HP-β-CD [27] and linalool/HP-β-CD [28], it has been observed that sometimes γ-CD is a better complexing agent. Ciobanu et al. showed that menthol, menthone, and pulegone are capable to form stable 1:1 inclusion complexes with β-CD, but eucalyptol forms a more stable inclusion complex with γ-CD due to the size of its cavity [29]. Polymeric CDs, which can be synthesized using cross-linking agents such as epichlorohydrin, also revealed promising results in some specific cases but are not matter of discussion in this chapter [29].

The inclusion complexes of EOs (or their components) with CDs have been mainly tested for food and pharmaceutical applications, but they could be an efficient tool to improve the use of EOs in aromatherapy, cosmetic, and household cleaning products. An interesting application of EOs is related to their incorporation in food packaging systems or edible films due to their antimicrobial, antioxidant, and insect repellent capacity. However, this is often limited due to flavoring and organoleptic considerations. CD inclusion complexes could overcome these limitations allowing EOs to reach effective concentrations in the food matrices without exceeding organoleptically acceptable levels and even providing controlled release-rate kinetics. Recent studies encourage the use of CD-EO (e.g., β-CD/Satureja montana EO; γ-CD inclusion complex encapsulated electrospun zein nanofibrous webs/thymol) complexes as part of active packaging systems [30, 31] as well as promising candidates to be used as safe and effective antimicrobial agents (β-CD/eugenol) to control postharvest diseases in fruits [27].

There are scientific evidences that inclusion complexation with CDs improves the pharmacological effects of EOs or their components [28, 32, 33]. For instance, the complexation of Hyptis pectinata L. EO with β-CD improved its analgesic effect in a mice model [34]. Also Bomfim et al. observed that β-CD complexation increased in vivo tumor growth inhibition capacity of Annona vepretorum EO [32]. Recently, Lima et al. reviewed the preclinical and clinical studies published on complexes between CDs and terpenes [33]. These are the major components of EOs that exhibit a wide range of biological activities on the human body. Their survey shows that there is robust experimental evidence that CDs improve the oral absorption and pharmacological properties of terpenes. Nevertheless, more pharmacokinetic and clinical studies are required before they can be effectively used in clinical targets.

Advertisement

3. Natural surfactants: brief introduction to sugar-based amphiphiles

Despite the production of surfactants based on fats, oils, and carbohydrates, being a known area for several decades, on an industrial scale, this is a relatively new issue [35]. These amphiphilic molecules that have one of the main building blocks from a natural source are often called “natural surfactants” [36, 37]. For example, alkyl glycosides which are synthesized from a “natural” sugar unit and a “nonnatural” fatty alcohol are often regarded as natural surfactants. Considering their amphiphilic nature, it has been always a challenge to attach a carbohydrate molecule, such as the hydrophilic group (due to the numerous hydroxyl groups) to a fat and oil derivative, such as a fatty acid or a fatty alcohol. However, nowadays, several successful synthesis routes are well established, and numerous types of natural surfactants are known and available, even on a commercial scale [38]. Nowadays, carbohydrate-based surfactants (CBS) are among the most important classes of amphiphilic compounds [39, 40, 41]. Their structure results from the combination of sugar and lipids, naturally biosynthesized within living cells or, alternatively, synthetically prepared by sequential reactions using carbohydrate and fatty materials. The growing interest in such compounds is due to, inter alia, their preparation from renewable raw materials, biodegradability, mildness to the skin, and biocompatibility, among other reasons [42, 43]. In particular, CBS can be relatively easily prepared from the most abundant renewable vegetable raw materials (e.g., cellulose, pectin, hemicellulose, starch, etc.) in a wide range of structures and geometries by modular synthesis thanks to the presence of numerous reactive hydroxyl groups. Such structural diversity makes CBS excellent models to get insight on the surfactant mechanisms in modifying interfacial properties. This knowledge is crucial for the control of the formation and stability of diverse colloidal systems such as micelles, vesicles, foams, and emulsions [44]. An important structural feature of these surfactants is the typical sugar headgroup, a voluminous and relatively rigid moiety that can be functionalized by a myriad of reagents and synthetic schemes. Numerous properties and functionalities can be expected from such almost unlimited number of different compounds that can find specific applications in different industrial areas [45, 46]. CBS also present great advantages on the environmental side; their higher biodegradability and lower toxicity profile are important reasons to consider CBS as valid alternatives to the more common petrochemical-based surfactants.

3.1. Case study: alkyl polyglycosides

In recent years there has been a growing focus on three classes of surfactants with sugar or a polyol derivative as polar headgroup: alkyl polyglycosides (APGs), alkyl glucamides, and sugar esters. In this chapter, we will focus on APGs, which are regarded as “perfect amphiphilic structures” with excellent surface activity as well as solubility. APGs have been synthesized, for the first time, more than 100 years ago [47]. APGs are completely based on renewable resources and combine very good performance, multifunctionality, and competitive price with mildness. This explains why APGs are the most successful sugar-based surfactants nowadays. It is important to mention that not pure alkyl monoglucosides but rather a complex mixture of alkyl mono-, di-, tri-, and oligoglycosides is produced in the industrial processes. Because of this, the industrial products are called alkyl polyglycosides. The surfactants are thus characterized by the length of the alkyl chain and the average number of glucose units linked to it, the degree of polymerization (DP). Alkyl glycosides are stable at high pH and sensitive to low pH where they hydrolyze to sugar and fatty alcohol. The sugar unit is more water-soluble and less soluble in hydrocarbons than the corresponding polyoxyethylene unit; hence, APGs and other polyol-based surfactants are more lipophobic than their polyoxyethylene-based surfactant counterparts [48]. This makes the physicochemical behavior of APG surfactants in oil/water systems distinct from that of conventional nonionic surfactants. Moreover, APGs do not show the pronounced inverse solubility vs. temperature relationship that normal nonionics do [49]. This makes an important difference in solution behavior between APGs and polyoxyethylene-based surfactants. The critical micelle concentration (cmc) values of the pure alkyl monoglycosides and the technical APG are comparable with those of typical nonionic surfactants and decrease distinctly with increasing alkyl chain length. The alkyl chain length has a far stronger influence on the cmc than the number of glycoside groups of the APG. The influence of the DP of APGs on their phase behavior has been described by Fukuda et al. [50]. The region in which the liquid crystalline phases occur is only slightly dependent on the concentration with a greater expansion in the case of APGs with a higher DP.

Regarding their applications, APGs are mainly used in personal care products such as cosmetics, manual dishwashing, and detergents [51]. APGs have also been used in more advanced applications such as the extraction and purification of membrane proteins, which plays a major role in the determination of protein structures and functions [52]. This is because APGs have reduced protein-denaturing properties in comparison to conventional surfactants. Generally, the environmental fate of surfactants is inextricably linked with their biodegradation behavior. Thus, fast and complete biodegradability is the most important requirement for an environmentally compatible surfactant. The general environmental impact of chemicals lies mainly in their ecotoxicity, which is relatively high in the case of surfactants because of their surface activity and the resulting effects on biological membranes [47]. In the case of CBS (APGs in particular), they present a quite favorable environmental profile: the rate of biodegradation is usually high, and the aquatic toxicity is low. In addition, APGs exhibit favorable dermatological properties, being very mild when exposed to the skin and eye [53]. The interested reader can find excellent overviews on APGs elsewhere [45, 46, 47, 51, 54].

3.2. Interaction between alkyl polyglycosides and cyclodextrins

Saenger and Mullerfahrnow [55] and Casu et al. [56] were pioneers in the study of the interactions between APGs and CDs. By surface tension measurements, it was shown that the addition of CDs leads to an increase of surfactant critical micelle concentration. Furthermore, the interaction is most pronounced when the CD cavity and hydrophobic part of the surfactant exhibit the tightest fit. Such a view was also supported by 1H and 13C NMR spectroscopy. By the analysis of NMR chemical shifts of octyl- (C8G1) and dodecyl (C12G1) α- and β-d-glucopyranoside and octadecyl β-d-glucopyranoside (C18G1), in the absence and presence of α-CD, two main conclusions were taken: the presence of α-CD only affects the chemical shifts of the surfactant alkyl chain, and the chemical shift of the protons H-3 and H-5, located inside the CD cavity, increases by increasing the length of the alkyl chain [57]. Furthermore, it was observed that no chemical shift is observed for γ-CD/C12G1 mixed systems. These conclusions corroborate previous and subsequent studies, showing that the CD cavity is protruded just by the surfactant tail and the magnitude of the interaction is dependent on the relationship between the volumes of the CD cavity and the surfactant hydrocarbon chain [56, 58]. A better characterization of these complexes was accomplished by the quantification of the binding process. The effect of the length of the alkyl chain and surfactant head group on the association constant of APGs with different CDs was studied by 1H NMR spectroscopy, NMR self-diffusion, and surface tension (Table 1) [58, 59, 60]. Although the comparison between association constants computed on the basis of different physical parameters is a difficult task [5], the analysis of the data allowed concluding that by increasing the alkyl chain length and decreasing the CD cavity volume (from β- to α-CD), the association constant increases. Furthermore, some authors stated that no interactions were observed when γ-CD was used [59, 58]. It is also worth noticing that for all mentioned systems, by subtracting the critical aggregation concentration (cac) of the surfactants from the CD concentration, the critical micelle concentration value is obtained. This clearly suggests a 1:1 APG/CD complexation [4], and actually it finds support by the Job’s plot [61] reported by Bernat et al. [60]. Another important issue is to understand the reliability of the binding constants reported in Table 1. Rymdén et al. found that an increase of a methylene group for a series of alcohols decreases the standard free energy of the alcohol: β-CD binding for ca. 3.0 kJ/mol [62]. This variation is similar to that observed for K1,1 values obtained by self-diffusion coefficients (Table 1). On the other hand, comparing K1,1 values for C8G1 with those from other monoalkyl surfactants, one can conclude that the APG/CD complex is more stable [63, 64]. This has been justified by the occurrence of hydrogen bonds between the sugar structure and the hydroxyl groups located at the rim of the CD. However, studying the effect of the number of sugar moieties in the surfactant head on the free energy of binding, an algebraic increase in the Gibbs free energy is observed. Indeed, comparing the binding constants for the interactions between β-C12G1 and β-C12G2 with β-CD, it is possible to conclude that the addition of an extra sugar moiety in the surfactant head decreases the K values for the supramolecular association. Thus, it can be hypothesized that no significant sugar-sugar interactions are involved in the interaction with CD, as it was previously discussed. Another hypothesis arises from the effect of carbohydrates on the water structure, for example, Ribeiro et al. have found that the presence of carbohydrates leads to an increase of the entropy in water [65], also called a structure “breaking effect” [66]. Consequently, an increase of the concentration of the sugar molecules in solution may contributes for a decrease in the binding entropy change and, consequently, to an increase in the binding Gibbs free energy.

α-CDβ-CDObs.
β-C8G13.68 (±1.6) × 1030.99 (±0.17) × 103Self-diffusion NMR [58]
1.85 (±0.35) × 103Surface tension [60]
β-C9G176 (±750) × 103275 (±5300) × 103Self-diffusion NMR [58]
β-C10G1340 ± 30[CD] = 1 mM, 1H NMR [59]
β-C12G1440 ± 40[CD] = 2 mM, 1H NMR [59]
410 ± 40[CD] = 1 mM, 1H NMR [59]
β-C12G2125 ± 10[CD] = 1 mM, 1H NMR [59]

Table 1.

Binding constants, K1,1 in dm3 mol−1, for the inclusion complexes CD/APG.

Up to now, we have been discussing the binding process assuming a 1:1 APG/CD (α- and β-) binding stoichiometry. However, it should be stressed that from the study of the interactions between C12G1, and C18G1, and CDs, there are strong evidences for the occurrence of other species consistent with 1:2 complexes [56]. More recently, Haller and Kaatze, studying the interaction between C8G1 and α-CD by ultrasonic attenuation spectroscopy, concluded that besides 1:1 (APG/CD) complexes, the formation of 1:2 and 2:1 complexes (although in very low concentration) should not be ruled out [67].

Advertisement

4. On the methods to follow the interactions between cyclodextrins and phytochemicals and sugar-based surfactants

As it must have already been understood from the previous subsections, an accurate choice of the technique to follow the host-guest association is a key issue for a reliable thermodynamic and kinetic characterization of the association process. In general, the experimental techniques can be subdivided into two different categories, labeled as I and II [68]. Methods from group I (e.g., surface tension) are measuring changes in physical properties that are proportional, in some ways, to the extent of binding, while those from group II (e.g., 1H NMR spectroscopy) rely on direct measurements of the free and bound ligand in a solution containing a known amount of the CD and guest molecule. Comments on such a division can be found in a couple of reviews (see, e.g., [69]), and it is outside of the scope of this chapter. The same is valid for computational techniques as relevant tools to infer on the structure of the supramolecular compounds [70, 71, 72].

In this section, the most relevant techniques used to study the interactions between CDs and sugar-based surfactants (e.g., APGs) and phytochemicals (e.g., polyphenols, EOs, and their components) will be highlighted (Table 2).

Experimental methodsSystemObs.
NMRNerolidol + β-CD[74]
UV-visNerolidol + CDs1[74]
Phase solubility studiesCabreuva essential oil + HP-β-CD[74]
Phase solubility studiesβ-caryophyllene + HP-β-CD[76]
UV-visBlack pepper essential oil + HP-β-CD[76]
Phase solubility studiesCDs1 + PPs2[77]
Phase solubility studies, NMR, TGA, DSCβ-CD + estragole[78]
NMRβ-CD + eugenol[27]
NMRβ-CD + rosmarinic acid[75]
NMRCyclohexylacetic acid + β-CD[79]
NMRCholic acid + β-CD[79]
UV-vis, NMRThymol and carvacrol + CDs1[80]
NMRn-Octyl-β-d-glucoside and n-nonyl-β-d-glucoside + α-CD and β-CD[58]
UAS3Octyl-β-d-glucopyranoside + α-CD[67]
Surface tension, NMRAPGs4 + β-CD[59]
Surface tensionOctyl-β-d-glucopyranoside + α-CD[60]
Phase solubility studies, ITC, NMRNootkatone + β-CD and HP-β-CD[71]
TGACinnamon essential oil + β-CD[90]
NMR, FTIR, release kineticsMonochlorotriazinyl β-CD + EOs5[91]
XRD, NMR, TGAThymol + γ-CD[30]
DSC, TGA, FTIR, XRD, GC/MS, NMRIsopulegol + α-CD and β-CD[72]
Phase solubility studies, NMR, HPLCPolymethoxyflavones + HP-β-CD[93]
GC, total organic carbon, phase-solubility studiesSBE-β-CD, SBE-γ-CD and HP-β-CD + EOS6[94, 95]

Table 2.

Compilation of the most relevant techniques used to study the interactions between CDs and bio-based compounds.

CDs: alpha-cyclodextrin (α-CD), beta-cyclodextrin (β-CD), gamma-cyclodextrin (γ-CD), 2-hydroxypropyl-β-cyclodextrin (HP-β-CD), randomly methylated-beta-cyclodextrins (RAMEB), low methylated beta-cyclodextrin (CRYSMEB), and sulfobutylether β-cyclodextrin (SBE-β-CD)


PPs: trans-anethole, estragole, eugenol, isoeugenol (phenylpropenes), caffeic acid, p-coumaric acid, and ferulic acid (hydroxycinnamic acids)


UAS: ultrasonic attenuation spectroscopy


APGs: glucopyranosides (octyl G8, decyl G10, dodecyl G12, tetradecyl G14) and two maltosides (decyl M10, dodecyl M12)


EOs: essential oils of cedarwood, clove, eucalyptus, and peppermint


EOS: essential oils of Artemisia dracunculus, Citrus reticulata Blanco, Citrus aurantifolia, Melaleuca alternifolia, Melaleuca quinquenervia, and Rosmarinus officinalis cineoliferum


NMR spectrometry falls in the group II techniques, and it is used to determine association constants through the chemical shift changes noticed either by the guest or by the CD [73, 74, 75]. Focusing on EOs, and as previously discussed, they may present undesired features, such as volatility, poor aqueous solubility, and stability. Therefore, host-guest supramolecular complexes are often obtained by using solid-state-based methods [14] such as freeze-drying [76, 77], coprecipitation [78], and the saturated approach [27], improving the solubility of the EO and thus allowing the use of NMR techniques for the quantitative and qualitative assessment of the complexation process [74]. For example, the complexation of eugenol with β-CD was obtained by using the saturation method, and the obtained complex, in the solid state, was characterized by either 1H, 13C, or 2D NMR techniques, confirming the thread of CD’s cavity by the aromatic ring of the eugenol [27]. Other techniques will be mentioned later since they fall on the so-called group II.

DOSY 1H NMR has been used to study inclusion complexes between CD and different sugar-based substrates [79]. Kfoury et al. report a comprehensive study on the complexation between two phenol isomers (thymol and carvacrol) and CDs by using different NMR techniques, including DOSY [80]. The data allowed concluding that those isomers have a binding constant of 1344 M−1 and 1336 M−1, respectively.

The self-diffusion measurements are, in principle, applicable to any systems as long as the free and complexed guests are soluble to an extent that allows for a good signal-to-noise ratio. It is important to note that on account of the rapid exchange on the NMR time scale, average diffusion coefficients for both the guest and for the CD are typically obtained. This method, as well as that involving chemical shift changes analysis, is also limited to systems where no overlapping of well-defined resonances is observed. The method relies on the fact that the self-diffusion coefficients of the uncomplexed guest are higher than the self-diffusion of the host-guest complex, as defined by the Stokes-Einstein equation. The change in the self-diffusion coefficient of the CD upon complexation is often small since the complex is often of the same size as the CD molecule, and thus the information from the CD self-diffusion is rather limited [58].

Ultrasonic relaxation technique falls into group II techniques and is based on the application of ultrasound to a given solution, with a frequency ranging from 20 kHz to several GHz, and subsequently measuring the molecular structural relaxation. The relaxation is sensitive to molecular volume changes [81], and thus, it may convey information on the stability constants of the host-guest complexes [82]. Furthermore, the use of a large frequency range allows to follow processes with relaxation times in the range from 20 ps to 20 μs [83], and thus the kinetics of the CD-surfactant association can be investigated. Haller and Kaatze, by using ultrasonic attenuation spectroscopy, were able to quantify the dynamics of unimer-micelle exchange of a sugar-based surfactant (i.e., octyl-β-d-glucopyranoside (C8G1)) in the presence of α-CD [67].

Also from group II, surface tension has also been used to follow the effect of CDs on the aggregation and interfacial properties of surfactants in CD-surfactant-containing solutions [59, 60]. Surface tension is a measure of cohesive forces between liquid molecules present at the surface, and it represents the quantification of force per unit length of free energy per unit area [84]. In general, the presence of CDs will increase the surface tension of an APG solution. Knowing that natural CDs are not surface-active, they cannot replace APG at the air interface [85]. Therefore, CD molecules contribute for the depletion of APG unimers from the interface due to the great interaction between these unimers and CDs. There are several examples where surface tension measurements have been used to assess the stoichiometry and stability constants of host-guest complexes [60, 86].

Isothermal titration calorimetry (ITC) is a sensitive and powerful technique to study host-guest interactions by measuring the enthalpy and the free energy of binding [86, 87]. There are also some cases where the kinetic constants of the binding process can be obtained by ITC (see, e.g., [88]). For example, the heat produced by a stepwise addition of HP-β- and β-CD solution to a nootkatone allowed to characterize the complexation process with a binding enthalpy and binding constant of −6.99 kJmol−1 and 4838 M−1, and −14.38 kJmol−1, and 5801 M−1, respectively [71]. Unfortunately, the strict conditions required by this technique do not allow its routine implementation on a large scale [89].

As has been pointed out before, some phytochemicals (EOs, in particular) are, in general, poorly soluble in aqueous solutions; therefore, the formation of complexes with CD in solid state is a strategy for further applications. Consequently, there are several available techniques used to evaluate the complexation. Thermal techniques, such as thermal degradation analysis and differential scanning calorimetry, are classical examples of methods used to assess complexation. Moreover, thermal degradation also allows evaluating the thermal stability of the EO upon complexation [90, 78].

Other spectroscopic techniques, such as FTIR and XRD, which can be included in group II, have also been used, but the information is, in our opinion, rather qualitative [30, 72, 78, 91]. Another interesting approach to learn about the formation of host-guest complexes, in solid state, is to study the release kinetics of the EO. These studies, although do not allow to quantify the total amount of EO incorporated into the CDs, are of utmost importance to evaluate the presence of the EO in the complex as well as to provide hints on the release mechanism; the latter is quite relevant for EOs used as fragrances [91, 92].

For such poor soluble compounds, the complexation can also be evaluated by carrying out phase-solubility studies. These can be performed by using complexes in solid state or by checking the ability of increasing concentrations of CD to solubilize saturated solutions of EO. This allows assessing how much the solubility of the EO is improved upon complexation as well as the corresponding complex binding constant. The details on the quantitative determination of those parameters will be given in the next section. Different techniques can be applied to obtain the phase-solubility profiles. For instance, the solubility of black pepper EO in the presence of hydroxypropyl-beta-CD (HP-β-CD) was evaluated by UV-visible spectroscopy [76]. On the other hand, phase-solubility profiles for the encapsulation of polymethoxyflavones, obtained from mandarin EO, into HP-β-CD were obtained by using HPLC [93]. Kfoury et al. have used gas chromatography to study the ability of sulfobutylether-β- and sulfobutylether-γ-CD to encapsulate EOs components, such as limonene, estragole, and α- and β-pinene; their solubility in water was improved more than one order of magnitude [94]. It is worth noticing that, recently, a technique based on the total organic carbon determination has been reported and validated to follow the solubility improvement of EOs when increasing the concentration of CD [95].

Advertisement

5. Methods for computation of binding constants: the case of EO-CD association

In this section a rather simple and straight overview of the most used model equations to compute binding constants for host-guest association is provided. To do so, EO will be used as the guest compound.

As discussed before, the phase-solubility plots are a widely applied method to get knowhow on the improvement of the EO solubility driven by complexation and also for computation of the EO-CD association constant. Assuming that the most common type of EO/CD complex has a 1:1 stoichiometry, the corresponding reaction and equilibrium (binding) constant, K1, can be written as

CD+GK1CDGE1
K1=CDG/CDfGfE2

where G represents the guest (here the EO) and [CD]f and [G]f are the concentrations of uncomplexed (free) species in the system. Assuming that the change in the aqueous solubility of the EO (ΔS) is only due to the formation of the complex, we can write

ΔS=STS0=CDGE3

where ST is the measurable total solubility and S0 is the solubility of the EO in water in the absence of CD (i.e., the intrinsic solubility). Thus, it follows that

Gf=S0E4
CDf=CDTSTS0E5

where [CD]T is the total concentration of CD in the solution.

Substituting Eqs. (3)(5) in Eq. (2) and after algebraic manipulation, we obtain

ST=S0+K1S01+K1S0CDTE6

Fitting Eq. (6) to experimental data of ST = f([CD]T) allows the calculation of the intercept (S0) and the association constant, K1. As discussed by Loftsson et al., the determination of K is highly dependent on the intercept accuracy [96]. In order to overcome this drawback, the authors have established the concept of “complexation efficiency” (CE) which can be obtained independently of S0, according to the following relation:

CE=CDGCDf=STS0/CDT1STS0/CDTE7

where the term ((STS0)/[CD]T) represents the slope of the phase-solubility profile.

The application of Eq. (7) is useful but limited by the dependence on S0 and to 1:1 complexes.

Let us now assume the following reaction:

nCD+GKnCDnGE8

where n is a stoichiometry coefficient and Kn is the corresponding binding constant. Thus, from the conservation of mass equations:

Gf=GTCDnGE9

and

CDf=CDTnCDGE10

the binding constant equation can be written as

Ka=CDnGCDTnCDGnSTCDGE11

By measuring a physical parameter, known as ΔA, directly related with the formation of the complex (CDnG), and performing the experiment in such a way that [CD]T> > [CDG], Eq. (11) takes the form of the so-called “Benesi-Hildebrand” equation [97]:

Ka=ΔACDTnSTΔAE12

or its linear form

1ΔA=1ST+1KnSTCDTnE13

However, it should be pointed out that nowadays, with the available software, such as Origin® and MatLab®, there is no need to linearize Eq. (12) once such methodology brings some restrictions to the computation of K and n.

It should be stressed that a key point in all these procedures is the accurate previous knowledge of the stoichiometry of complexation, but this is not always a simple task. The most common method for the determination of stoichiometry is the method of continuous variation or the Job plot; the virtues and limitations of this method were recently reviewed [98], and thus it is not our intention to further discuss it in the present chapter.

Going back to the determination of the binding constants, the most accurate way to compute K is by using the first principles. Here, for the sake of simplicity, the 1:1 and 1:2 (G/CD) stoichiometric ratios will be focused. Additionally, these examples also correspond to the large majority of complexes formed between CDs and EOs. For more complex stoichiometries, the computational treatment of the resulting equations (not shown) is not straightforward as a consequence of multicollinearity [99]. Multicollinearity causes larger standard errors in the quantities calculated and lowers statistical significance of the results. In limiting cases, several local minima may be obtained by iteration; these correspond to noticeably different combinations of the quantities calculated and may be the reason why different K values are reported for the same host-guest systems.

Assuming that a 1:1 complex (CD-S) is formed, the binding constant (Eq. (2)) can be rewritten as

K1=f1fCDTfGTE14

where f is defined as [CD − G]/[G]T.

Despite the binding process being followed by ΔA (e.g., for 1H NMR, ΔA will be equal to the chemical shift of a given 1H resonance), the observed ΔA for a host molecule is expressed as

Aobs=1fACD,f+fACDGE15

where ACD,f and ACDG represent the measurable physical parameter related to CD in free and complexed states, respectively.

The variation of the physical parameter in the presence and absence of a guest molecule, ΔAobs = ΔAobs − ΔACD, can be expressed as

ΔAobs=ΔACDGCDTCDGE16

which, after some algebraic manipulation and simplification, results in [100, 101]:

ΔAobs=ΔACDG2CDTGT+CDT+1K1GT+CDT+1K124GTCDT1/2E17

It should be stressed that the application of Eq. (17) shows some drawbacks when the total concentrations of CD and guest are low and/or the binding constant is very weak, i.e., for the simplest 1:1 case, when y is sufficiently small, xx2yy/2x, and, consequently, Eq. (17) reduces to

ΔAobs=ΔACDGT+1K1GTE18

where T=CDT+ST. If T is kept constant in the experiments, as is common practice when Job plots are used to obtain stoichiometries, the observed displacement varies linearly with [S]T or [CD]T, but the fitting parameters are present in the form of a ratio that generates an infinite number of acceptable solutions. Consequently, it is suggested that T should be chosen in such a way that its value should be of the same order of magnitude than K1−1 [102].

Another approach lays on the assumption of a 2:1 (CD/G) complexation, in a two-step mechanism. In these circumstances, the complexation process is defined by two binding constants K1 and K2, and the corresponding mass balances are defined as

Gf=GTCDGCD2GE19

and

CDf=CDTCDG2CD2GE20

From the equilibrium constants and Eqs. (19) and (20), we can write

Aobs=CDfACD+CDGACDG+2CD2GACD2GCDf+CDG+2CD2GE21

where ACD, ACD−G, and ACD2−G are the contributions of the CD and 1:1 and 2:1 CD/G complexes, with concentrations [CD], [CDG], and [CD2G], respectively, for the observed (experimental) physical parameter A. Using a similar procedure to that used for a 1:1 complexation, it is possible to write Eq. (21) as a function of [CD], that is

Aobs=ACD+CDK1ACDG+K1K2CDACD2G1+K1CDf+K1K2CD2E22

On the other hand, the free CD concentration is given by

CD3+1K2,1CDT+2STCD2+1K1,1K2,1CDTK2,1+STK2,1CDCDTK1,1K2,1=0E23

One method for estimation of the free CD concentration is through an analytical solution of the real solution of a third-degree equation [103]:

fx=x3+ax2+bx+cE24

using a Cardin-Tartaglia formulae

x=r13aqrE25

where

q=13b19a2E26

and

r=16ab12c127a3+127b316abc+14c2+127a3c1108a2b23E27
Advertisement

6. Conclusions

Despite the huge potential of many bio-based compounds in several diverse areas, their use is still limited due to different reasons such as poor aqueous solubility, volatility, reactivity, etc. Therefore, advanced strategies have to be developed in order to minimize some of these weaknesses to make bio-based molecules usable on a larger scale. It became patent in this chapter that CD interaction with bio-based compounds, such as different phytochemicals or sugar-based surfactants, has generally a remarkable positive impact on their performance, improving their aqueous solubility and availability and decreasing their degradation rate, etc. Moreover, it is clear that this is not only interesting and beneficial from an application point of view but also very stimulating from a fundamental perspective where thermodynamics, modeling, and different experimental methodologies get together for a deep and challenging characterization of the systems. Eventually, such knowledge will be crucial for the future development of improved formulations and make use at full extent of the exciting properties of bio-based compounds.

Advertisement

Acknowledgments

The Portuguese Foundation for Science and Technology (FCT) is acknowledged through the projects PTDC/AGR-TEC/4814/2014, UID/QUI/UI0313/2013, and POCI-01-0145-FEDER-07630. FCT is also acknowledged through the researcher grants IF/01005/2014, SFRH/BPD/103086/2014, and SFRH/BPD/84112/2012 financed by POPH-QREN and subsidized by the European Science Foundation.

References

  1. 1. Del Valle EMM. Cyclodextrins and their uses: A review. Process Biochemistry. 2004;39(9):1033-1046. DOI: 10.1016/S0032-9592(03)00258-9
  2. 2. Jiang LX, Yan Y, Huang JB. Versatility of cyclodextrins in self-assembly systems of amphiphiles. Advances in Colloid and Interface Science. 2011;169(1):13-25. DOI: 10.1016/j.cis.2011.07.002
  3. 3. 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(1):324-330. DOI: 10.1016/j.carbpol.2012.08.109
  4. 4. Nilsson M, Cabaleiro-Lago C, Valente AJM, Soderman O. Interactions between gemini surfactants, 12-s-12, and beta-cyclodextrin as investigated by NMR diffusometry and electric conductometry. Langmuir. 2006;22(21):8663-8669. DOI: 10.1021/la061220e
  5. 5. Valente AJM, Soderman O. The formation of host-guest complexes between surfactants and cyclodextrins. Advances in Colloid and Interface Science. 2014;205:156-176. DOI: 10.1016/j.cis.2013.08.001
  6. 6. Dias HMAM, Berbicz F, Pedrochi F, Baesso ML, Matioli G. Butter cholesterol removal using different complexation methods with beta-cyclodextrin, and the contribution of photoacoustic spectroscopy to the evaluation of the complex. Food Research International. 2010;43(4):1104-1110. DOI: 10.1016/j.foodres.2010.02.002
  7. 7. Lakkakula JR, Krause RWM. A vision for cyclodextrin nanoparticles in drug delivery systems and pharmaceutical applications. Nanomedicine-UK. 2014;9(6):877-894. DOI: 10.2217/Nnm.14.41
  8. 8. Szejtli J. Past, present, and future of cyclodextrin research. Pure and Applied Chemistry. 2004;76(10):1825-1845. DOI: 10.1351/pac200476101825
  9. 9. Crozier A, Jaganath IB, Clifford MN. Dietary phenolics: Chemistry, bioavailability and effects on health. Natural Product Reports. 2009;26(8):1001-1043. DOI: 10.1039/b802662a
  10. 10. Manach C, Scalbert A, Morand C, Remesy C, Jimenez L. Polyphenols: Food sources and bioavailability. The American Journal of Clinical Nutrition. 2004;79(5):727-747
  11. 11. Vishwakarma GS, Gautam N, Babu JN, Mittal S, Jaitak V. Polymeric encapsulates of essential oils and their constituents: A review of preparation techniques, characterization, and sustainable release mechanisms. Polymer Reviews. 2016;56(4):668-701. DOI: 10.1080/15583724.2015.1123725
  12. 12. Suvarna V, Gujar P, Murahari M. Complexation of phytochemicals with cyclodextrin derivatives—An insight. Biomedicine & Pharmacotherapy. 2017;88:1122-1144. DOI: 10.1016/j.biopha.2017.01.157
  13. 13. Pinho E, Grootveld M, Soares G, Henriques M. Cyclodextrins as encapsulation agents for plant bioactive compounds. Carbohydrate Polymers. 2014;101:121-135. DOI: 10.1016/j.carbpol.2013.08.078
  14. 14. Marques HMC. A review on cyclodextrin encapsulation of essential oils and volatiles. Flavour and Fragrance Journal. 2010;25(5):313-326. DOI: 10.1002/ffj.2019
  15. 15. Esfanjani AF, Jafari SM. Biopolymer nano-particles and natural nano-carriers for nano-encapsulation of phenolic compounds. Colloids and Surfaces B: Biointerfaces. 2016;146:532-543. DOI: 10.1016/j.colsurfb.2016.06.053
  16. 16. Loftsson T, Duchene D. Cyclodextrins and their pharmaceutical applications. International Journal of Pharmaceutics. 2007;329(1–2):1-11. DOI: 10.1016/j.ijpharm.2006.10.044
  17. 17. Matsuda H, Arima H. Cyclodextrins in transdermal and rectal delivery. Advanced Drug Delivery Reviews. 1999;36(1):81-99. DOI: 10.1016/S0169-409x(98)00056-8
  18. 18. Fang ZX, Bhandari B. Encapsulation of polyphenols—A review. Trends in Food Science & Technology. 2010;21(10):510-523. DOI: 10.1016/j.tifs.2010.08.003
  19. 19. Budryn G, Palecz B, Rachwal-Rosiak D, Oracz J, Zaczynska D, Belica S, Navarro-Gonzalez I, Meseguer JMV, Perez-Sanchez H. Effect of inclusion of hydroxycinnamic and chlorogenic acids from green coffee bean in beta-cyclodextrin on their interactions with whey, egg white and soy protein isolates. Food Chemistry. 2015;168:276-287. DOI: 10.1016/j.foodchem.2014.07.056
  20. 20. Budryn G, Zaczynska D, Rachwal-Rosiak D, Oracz J. Changes in properties of food proteins after interaction with free and beta-cyclodextrin encapsulated hydroxycinnamic acids. European Food Research and Technology. 2015;240(6):1157-1166. DOI: 10.1007/s00217-015-2419-9
  21. 21. Rawel HM, Rohn S. Nature of hydroxycinnamate-protein interactions. Phytochemistry Reviews. 2010;9(1):93-109. DOI: 10.1007/s11101-009-9154-4
  22. 22. Shao P, Zhang JF, Fang ZX, Sun PL. Complexing of chlorogenic acid with beta-cyclodextrins: Inclusion effects, antioxidative properties and potential application in grape juice. Food Hydrocolloids. 2014;41:132-139. DOI: 10.1016/j.foodhyd.2014.04.003
  23. 23. Mercader-Ros MT, Lucas-Abellan C, Fortea MI, Gabaldon JA, Nunez-Delicado E. Effect of HP-beta-cyclodextrins complexation on the antioxidant activity of flavonols. Food Chemistry. 2010;118(3):769-773. DOI: 10.1016/j.foodchem.2009.05.061
  24. 24. Cutrignelli A, Lopedota A, Denora N, Iacobazzi RM, Fanizza E, Laquintana V, Perrone M, Maggi V, Franco M. A new complex of curcumin with sulfobutylether-beta-cyclodextrin: Characterization studies and in vitro evaluation of cytotoxic and antioxidant activity on HepG-2 cells. Journal of Pharmaceutical Sciences. 2014;103(12):3932-3940. DOI: 10.1002/jps.24200
  25. 25. Davis ME, Brewster ME. Cyclodextrin-based pharmaceutics: Past, present and future. Nature Reviews. Drug Discovery. 2004;3(12):1023-1035. DOI: 10.1038/nrd1576
  26. 26. Stella VJ, Rajewski RA. Cyclodextrins: Their future in drug formulation and delivery. Pharmaceutical Research. 1997;14(5):556-567. DOI: 10.1023/A:1012136608249
  27. 27. Gong L, Li TT, Chen F, Duan XW, Yuan YF, Zhang DD, Jiang YM. An inclusion complex of eugenol into beta-cyclodextrin: Preparation, and physicochemical and antifungal characterization. Food Chemistry. 2016;196:324-330. DOI: 10.1016/j.foodchem.2015.09.052
  28. 28. Aytac Z, Yildiz ZI, Kayaci-Senirmak F, Tekinay T, Uyar T. Electrospinning of cyclodextrin/linalool-inclusion complex nanofibers: Fast-dissolving nanofibrous web with prolonged release and antibacterial activity. Food Chemistry. 2017;231:192-201. DOI: 10.1016/j.foodchem.2017.03.113
  29. 29. Ciobanu A, Mallard I, Landy D, Brabie G, Nistor D, Fourmentin S. Retention of aroma compounds from Mentha piperita essential oil by cyclodextrins and crosslinked cyclodextrin polymers. Food Chemistry. 2013;138(1):291-297. DOI: 10.1016/j.foodchem.2012.10.106
  30. 30. Aytac Z, Ipek S, Durgun E, Tekinay T, Uyar T. Antibacterial electrospun zein nanofibrous web encapsulating thymol/cyclodextrin-inclusion complex for food packaging. Food Chemistry. 2017;233:117-124. DOI: 10.1016/j.foodchem.2017.04.095
  31. 31. Kfoury M, Auezova L, Greige-Gerges H, Fourmentin S. Promising applications of cyclodextrins in food: Improvement of essential oils retention, controlled release and antiradical activity. Carbohydrate Polymers. 2015;131:264-272. DOI: 10.1016/j.carbpol.2015.06.014
  32. 32. Bomfim LM, Menezes LRA, Rodrigues ACBC, Dias RB, Rocha CAG, Soares MBP, Neto AFS, Nascimento MP, Campos AF, Silva LCRCE, Costa EV, Bezerra DP. Antitumour activity of the microencapsulation of Annona vepretorum essential oil. Basic and Clinical Pharmacology. 2016;118(3):208-213. DOI: 10.1111/bcpt.12488
  33. 33. Lima PSS, Lucchese AM, Araujo HG, Menezes PP, Araujo AAS, Quintans LJ, Quintans JSS. Inclusion of terpenes in cyclodextrins: Preparation, characterization and pharmacological approaches. Carbohydrate Polymers. 2016;151:965-987. DOI: 10.1016/j.carbpol.2016.06.040
  34. 34. Menezes PD, Araujo AAD, Doria GAA, Quintans LJ, de Oliveira MGB, dos Santos MRV, de Oliveira JF, Matos JD, Carvalho FMD, Alves PB, de Matos IL, dos Santos DA, Marreto RN, da Silva GF, Serafini MR. Physicochemical characterization and analgesic effect of inclusion complexes of essential oil from Hyptis pectinata L. Poit leaves with beta-cyclodextrin. Current Pharmaceutical Biotechnology. 2015;16(5):440-450
  35. 35. Falbe J. Surfactants in Consumer Products—Theory, Technology and Application. Berlin: Springer; 1986
  36. 36. Holmberg K, Jonsson B, Kronberg B, Lindman B. Surfactants and Polymers in Aqueous Solution. Hoboken: Wiley; 2002
  37. 37. von Rybinski W. Natural surfactants. Current Opinion in Colloid & Interface Science. 2001;6(2):146-147. DOI: 10.1016/S1359-0294(01)00081-4
  38. 38. Hill K, Rhode O. Sugar-based surfactants for consumer products and technical applications. Fett/Lipid. 1999;101:25-33
  39. 39. Dembitsky VM. Astonishing diversity of natural surfactants: 1. Glycosides of fatty acids and alcohols. Lipids. 2004;39(10):933-953. DOI: 10.1007/s11745-004-1316-1
  40. 40. Queneau Y, Chambert S, Besset C, Cheaib R. Recent progress in the synthesis of carbohydrate-based amphiphilic materials: The examples of sucrose and isomaltulose. Carbohydrate Research. 2008;343(12):1999-2009. DOI: 10.1016/j.carres.2008.02.008
  41. 41. Ruiz C. Sugar-based Surfactants Fundamentals and Applications. CRC Press; 2009
  42. 42. Hill K, LeHen-Ferrenbach C. Sugar-based surfactants for consumer products and technical applications. In: Sugar-Based Surfactants Fundamentals and Applications. CRC Press; 2009. pp. 1-20
  43. 43. Kitamoto D, Morita T, Fukuoka T, Konishi M, Imura T. Self-assembling properties of glycolipid biosurfactants and their potential applications. Current Opinion in Colloid & Interface Science. 2009;14(5):315-328. DOI: 10.1016/j.cocis.2009.05.009
  44. 44. Razafindralambo H, Blecker C, Paquot M. Screening of basic properties of amphiphilic molecular structures for colloidal system formation and stability: The case of carbohydrate-based surfactants. In: Nagarajan R, editor. Amphiphiles: Molecular Assembly and Applications. Vol. 1070. Washignton: ACS; 2011
  45. 45. Soderman O, Johansson I. Polyhydroxyl-based surfactants and their physico-chemical properties and applications. Current Opinion in Colloid & Interface Science. 1999;4(6):391-401. DOI: 10.1016/S1359-0294(00)00019-4
  46. 46. vonRybinski W. Alkyl glycosides and polyglycosides. Current Opinion in Colloid & Interface Science. 1996;1(5):587-597
  47. 47. von Rybinski W, Hill K. Alkyl polyglycosides—Properties and applications of a new class of surfactants. Angewandte Chemie International Edition. 1998;37(10):1328-1345. DOI: 10.1002/(Sici)1521-3773(19980605)37:10<1328::Aid-Anie1328>3.0.Co;2-9
  48. 48. Shinoda K, Carlsson A, Lindman B. On the importance of hydroxyl groups in the polar head-group of nonionic surfactants and membrane lipids. Advances in Colloid and Interface Science. 1996;64:253-271
  49. 49. Lindman B, Medronho B, Karlström G. Clouding of nonionic surfactants. Current Opinion in Colloid & Interface Science. 2016;22:23-29. DOI: 10.1016/j.cocis.2016.01.005
  50. 50. Fukuda K, Soderman O, Lindman B, Shinoda K. Microemulsions formed by alkyl polyglucosides and an alkyl glycerol ether. Langmuir. 1993;9(11):2921-2925. DOI: 10.1021/La00035a032
  51. 51. Hill K, Von Rybinski W, Stoll G. Alkyl Polyglucosides: Technology, Properties and Applications. New York: VCH; 1997
  52. 52. Costes F, Elghoul M, Bon M, Ricolattes I, Lattes A. Synthesis and structural-analysis of long-chain N-acetyl-N-alkyllactosylamines, a new series of surfactants derived from unprotected lactose. Langmuir. 1995;11(10):3644-3647. DOI: 10.1021/La00010a010
  53. 53. Holmberg K. Natural surfactants. Current Opinion in Colloid & Interface Science. 2001;6(2):148-159. DOI: 10.1016/S1359-0294(01)00074-7
  54. 54. von Rybinski W, Hill K. Alkyl polyglycosides. In: Holmberg K, editor. Novel Surfactants. New York: Marcel Dekker; 1998
  55. 55. Saenger W, Mullerfahrnow A. Cyclodextrins increase surface-tension and critical micelle concentrations of detergent solutions. Angewandte Chemie-International Edition in English. 1988;27(3):393-394. DOI: 10.1002/anie.198803931
  56. 56. Casu B, Grenni A, Naggi A, Torri G, Virtuani M, Focher B. Interaction of cyclodextrins (cyclomalto-oligosaccharides) with glycolipids-NMR-studies of aqueous systems of cyclomaltohexaose and alkyl glycosides. Carbohydrate Research. 1990;200:101-109. DOI: 10.1016/0008-6215(90)84185-W
  57. 57. Carvalho RA, Correia HA, Valente AJM, Soderman O, Nilsson M. The effect of the head-group spacer length of 12-s-12 gemini surfactants in the host-guest association with beta-cyclodextrin. Journal of Colloid and Interface Science. 2011;354(2):725-732. DOI: 10.1016/j.jcis.2010.11.024
  58. 58. AJM V, Nilsson M, Soderman O. Interactions between n-octyl and n-nonyl beta-d-glucosides and alpha- and beta-cyclodextrins as seen by self-diffusion NMR. Journal of Colloid and Interface Science. 2005;281(1):218-224. DOI: 10.1016/j.jcis.2004.08.018
  59. 59. Reinsborough VC, Stephenson VC. Inclusion complexation involving sugar-containing species: Beta-cyclodextrin and sugar surfactants. Canadian Journal of Chemistry. 2004;82(1):45-49. DOI: 10.1139/V03-180
  60. 60. Bernat V, Ringard-Lefebvre C, Le Bas G, Perly B, Djedaini-Pilard F, Lesieur S. Inclusion complex of n-octyl beta-d-glucopyranoside and alpha-cyclodextrin in aqueous solutions: Thermodynamic and structural characterization. Langmuir. 2008;24(7):3140-3149. DOI: 10.1021/la7034906
  61. 61. Plot J. Formation and stability of inorganic complexes in solution. Annali di chimica. 1928;9:113-203
  62. 62. Rymdén R, Carlfors J, Stilbs P. Substrate binding to cyclodextrins in aqueous solution: A multicomponent self-diffusion study. Journal of Inclusion Phenomena. 1983;1:159-167
  63. 63. Castronuovo G, Elia V, Niccoli M, Velleca F, Viscardi G. Role of the functional group in the formation of the complexes between alpha-cyclodextrin and alkanols or monocarboxylic acids in aqueous solutions. A calorimetric study at 25 degrees C. Carbohydrate Research. 1998;306(1–2):147-155. DOI: 10.1016/S0008-6215(97)10046-5
  64. 64. Castronuovo G, Elia V, Iannone A, Niccoli M, Velleca F. Factors determining the formation of complexes between alpha-cyclodextrin and alkylated substances in aqueous solutions: A calorimetric study at 25 degrees C. Carbohydrate Research. 2000;325(4):278-286. DOI: 10.1016/S0008-6215(99)00328-6
  65. 65. Ribeiro ACF, Esteso MA, Lobo VMM, Valente AJM, Simoes SMN, Sobral AJFN, Burrows HD. Interactions of copper (II) chloride with sucrose, glucose, and fructose in aqueous solutions. Journal of Molecular Structure. 2007;826(2–3):113-119. DOI: 10.1016/j.molstruc.2006.04.035
  66. 66. Marcus Y. Effect of ions on the structure of water: Structure making and breaking. Chemical Reviews. 2009;109(3):1346-1370. DOI: 10.1021/cr8003828
  67. 67. Haller J, Kaatze U. Octyl glucopyranoside and cyclodextrin in water. Self-aggregation and complex formation. The Journal of Physical Chemistry. B. 2009;113(7):1940-1947. DOI: 10.1021/jp808733p
  68. 68. Mwakibete H, Cristantino R, Bloor DM, Wynjones E, Holzwarth JF. Reliability of the experimental methods to determine equilibrium-constants for surfactant cyclodextrin inclusion complexes. Langmuir. 1995;11(1):57-60. DOI: 10.1021/La00001a013
  69. 69. Brocos P, Banquy X, Diaz-Vergara N, Perez-Casas S, Pineiro A, Costas M. A critical approach to the thermodynamic characterization of inclusion complexes: Multiple-temperature isothermal titration calorimetric studies of native cyclodextrins with sodium dodecyl sulfate. The Journal of Physical Chemistry. B. 2011;115(49):14381-14396. DOI: 10.1021/jp208740b
  70. 70. Lawtrakul L, Inthajak K, Toochinda P. Molecular calculations on beta-cyclodextrin inclusion complexes with five essential oil compounds from Ocimum basilicum (sweet basil). ScienceAsia. 2014;40(2):145-151. DOI: 10.2306/scienceasia1513-1874.2014.40.145
  71. 71. Kfoury M, Landy D, Ruellan S, Auezova L, Greige-Gerges H, Fourmentin S. Nootkatone encapsulation by cyclodextrins: Effect on water solubility and photostability. Food Chemistry. 2017;236:41-48. DOI: 10.1016/j.foodchem.2016.12.086
  72. 72. Menezes PD, Doria GAA, Araujo AAD, Sousa BMH, Quintans LJ, Lima RN, Alves PB, Carvalho FMS, Bezerra DP, Mendonca FJB, Scotti L, Scotti MT, da Silva GF, de Aquino TM, Sabino AR, do Egito EST, Serafini MR. Docking and physico-chemical properties of alpha- and beta-cyclodextrin complex containing isopulegol: A comparative study. Journal of Inclusion Phenomena and Macrocyclic Chemistry. 2016;85(3–4):341-354. DOI: 10.1007/s10847-016-0633-0
  73. 73. Schneider HJ, Hacket F, Rudiger V, Ikeda H. NMR studies of cyclodextrins and cyclodextrin complexes. Chemical Reviews. 1998;98(5):1755-1785. DOI: 10.1021/Cr970019t
  74. 74. Azzi J, Danjou PE, Landy D, Ruellan S, Auezova L, Greige-Gerges H, Fourmentin S. The effect of cyclodextrin complexation on the solubility and photostability of nerolidol as pure compound and as main constituent of cabreuva essential oil. Beilstein Journal of Organic Chemistry. 2017;13:835-844. DOI: 10.3762/bjoc.13.84
  75. 75. Medronho B, Valente AJM, Costa P, Romano A. Inclusion complexes of rosmarinic acid and cyclodextrins: Stoichiometry, association constants, and antioxidant potential. Colloid & Polymer Science. 2014;292(4):885-894. DOI: 10.1007/s00396-013-3124-5
  76. 76. Rakmai J, Cheirsilp B, Mejuto JC, Torrado-Agrasar A, Simal-Gandara J. Physico-chemical characterization and evaluation of bio-efficacies of black pepper essential oil encapsulated in hydroxypropyl-beta-cyclodextrin. Food Hydrocolloids. 2017;65:157-164. DOI: 10.1016/j.foodhyd.2016.11.014
  77. 77. Kfoury M, Sahraoui ALH, Bourdon N, Laruelle F, Fontaine J, Auezova L, Greige-Gerges H, Fourmentin S. Solubility, photostability and antifungal activity of phenylpropanoids encapsulated in cyclodextrins. Food Chemistry. 2016;196:518-525. DOI: 10.1016/j.foodchem.2015.09.078
  78. 78. Yang ZJ, Huang LY, Yao XD, Ji HB. Host-guest complexes of estragole with beta-cyclodextrin: An experimental and theoretical investigation. Flavour and Fragrance Journal. 2017;32(2):102-111. DOI: 10.1002/ffj.3358
  79. 79. Cameron KS, Fielding L. NMR diffusion spectroscopy as a measure of host-guest complex association constants and as a probe of complex size. The Journal of Organic Chemistry. 2001;66(21):6891-6895. DOI: 10.1021/Jo010081x
  80. 80. Kfoury M, Landy D, Ruellan S, Auezova L, Greige-Gerges H, Fourmentin S. Determination of formation constants and structural characterization of cyclodextrin inclusion complexes with two phenolic isomers: Carvacrol and thymol. Beilstein Journal of Organic Chemistry. 2016;12:29-42. DOI: 10.3762/bjoc.12.5
  81. 81. Kato S, Nomura H, Miyahara Y. Ultrasonic relaxation study of aqueous-solutions of cyclodextrins. Journal of Physical Chemistry. 1985;89(25):5417-5421. DOI: 10.1021/J100271a021
  82. 82. Hall D, Bloor D, Tawarah K, Wynjones E. Kinetic and equilibrium studies associated with the formation of inclusion-compounds involving normal-butanol and normal-pentanol in aqueous cyclodextrin solutions. Journal of the Chemical Society, Faraday Transactions. 1986;1(82):2111-2121. DOI: 10.1039/F19868202111
  83. 83. Kaatze U. Acoustical spectroscopy of carbohydrate aqueous solutions: Saccharides; alkyl glycosides; Cyclodextrins. Part I. Conformer variations. Archives of Acoustics. 2010;35(4):715-738. DOI: 10.2478/v10168-010-0054-9
  84. 84. Tariq M, Freire MG, Saramago B, Coutinho JAP, Lopes JNC, Rebelo LPN. Surface tension of ionic liquids and ionic liquid solutions. Chemical Society Reviews. 2012;41(2):829-868. DOI: 10.1039/c1cs15146k
  85. 85. Szejtli J. Utilization of cyclodextrins in industrial products and processes. Journal of Materials Chemistry. 1997;7(4):575-587. DOI: 10.1039/A605235e
  86. 86. Pineiro A, Banquy X, Perez-Casas S, Tovar E, Garcia A, Villa A, Amigo A, Mark AE, Costas M. On the characterization of host-guest complexes: Surface tension, calorimetry, and molecular dynamics of cyclodextrins with a non-ionic surfactant. The Journal of Physical Chemistry. B. 2007;111(17):4383-4392. DOI: 10.1021/jp0688815
  87. 87. Wadso L, Li YJ, Li X. Isothermal titration calorimetry in the student laboratory. Journal of Chemical Education. 2011;88(1):101-105. DOI: 10.1021/ed100649e
  88. 88. Nilsson M, Valente AJM, Olofsson G, Soderman O, Bonini M. Thermodynamic and kinetic characterization of host-guest association between bolaform surfactants and alpha- and beta-cyclodextrins. The Journal of Physical Chemistry. B. 2008;112(36):11310-11316. DOI: 10.1021/jp802963x
  89. 89. Russell DJ, Hansen LD. Calorimeters for biotechnology. Thermochimica Acta. 2006;445(2):151-159. DOI: 10.1016/j.tca.2005.08.023
  90. 90. Wen P, Zhu DH, Feng K, Liu FJ, Lou WY, Li N, Zong MH, Wu H. Fabrication of electrospun polylactic acid nanofilm incorporating cinnamon essential oil/beta-cyclodextrin inclusion complex for antimicrobial packaging. Food Chemistry. 2016;196:996-1004. DOI: 10.1016/j.foodchem.2015.10.043
  91. 91. Khanna S, Chakraborty JN. Optimization of monochlorotriazine beta-cyclodextrin grafting on cotton and assessment of release behavior of essential oils from functionalized fabric. Fash Text. 2017;4:1-18. DOI: 10.1186/S40691-017-0089-X
  92. 92. Yang ZJ, Yao XD, Xiao ZB, Chen HY, Ji HB. Preparation and release behaviour of the inclusion complexes of phenylethanol with beta-cyclodextrin. Flavour and Fragrance Journal. 2016;31(3):206-216. DOI: 10.1002/ffj.3302
  93. 93. Russo M, Rigano F, Arigo A, Sciarrone D, Calabro ML, Farnetti S, Dugo P, Mondello L. Rapid isolation, reliable characterization, and water solubility improvement of polymethoxyflavones from cold-pressed mandarin essential oil. Journal of Separation Science. 2016;39(11):2018-2027. DOI: 10.1002/jssc.201501366
  94. 94. Kfoury M, Pipkin JD, Antle V, Fourmentin S. Captisol (R): An efficient carrier and solubilizing agent for essential oils and their components. Flavour and Fragrance Journal. 2017;32(5):340-346. DOI: 10.1002/ffj.3395
  95. 95. Kfoury M, Auezova L, Greige-Gerges H, Fourmentin S. Development of a total organic carbon method for the quantitative determination of solubility enhancement by cyclodextrins: Application to essential oils. Analytica Chimica Acta. 2016;918:21-25. DOI: 10.1016/j.aca.2016.03.013
  96. 96. Loftsson T, Hreinsdottir D, Masson M. Evaluation of cyclodextrin solubilization of drugs. International Journal of Pharmaceutics. 2005;302(1–2):18-28. DOI: 10.1016/j.ijpharm.2005.05.042
  97. 97. Benesi HA, Hildebrand JH. A spectrophotometric investigation of the interaction of iodine with aromatic hydrocarbons. Journal of the American Chemical Society. 1949;71(8):2703-2707. DOI: 10.1021/Ja01176a030
  98. 98. Ulatowski F, Dabrowa K, Balakier T, Jurczak J. Recognizing the limited applicability of job plots in studying host-guest interactions in supramolecular chemistry. The Journal of Organic Chemistry. 2016;81(5):1746-1756. DOI: 10.1021/acs.joc.5b02909
  99. 99. Draper NR, Smith H. Applied Regression Analysis. New York: Wiley-Interscience; 1998
  100. 100. Cabrer PR, Azvarez-Parrilla E, Meijide F, Seijas JA, Nunez ER, Tato JV. Complexation of sodium cholate and sodium deoxycholate by beta-cyclodextrin and derivatives. Langmuir. 1999;15(17):5489-5495
  101. 101. Calderon V, Schwarz G, Garcia F, Tapia MJ, Valente AJM, Burrows HD, Garcia JM. Synthesis and characterization of new aromatic polyamides bearing crown ethers and acyclic ethylene oxide units in the pendant structure. III. Benzo-18-crown-6 systems and their open-chain counterparts. Journal of Polymer Science, Part A: Polymer Chemistry. 2006;44(21):6252-6269. DOI: 10.1002/pola.21710
  102. 102. Tapia MJ, Burrows HD, Garcia JM, Garcia F, Pais AACC. Lanthanide ion interaction with a crown ether methacrylic polymer, poly(1,4,7,10-tetraoxacyclododecan-2-ylmethyl methacrylate), as seen by spectroscopic, calorimetric, and theoretical studies. Macromolecules. 2004;37(3):856-862. DOI: 10.1021/ma0353888
  103. 103. Valente AJM, Dinis CJS, Pereira RFP, Ribeiro ACF, Lobo VMM. Interactions between β-cyclodextrin and some sodium alkyl sulfates and sulfonates as seen by electrical conductivity measurements. Portugaliae Electrochimica Acta. 2006;24:129-136

Written By

Bruno Filipe Figueiras Medronho, Sandra Gonçalves, Raquel Rodríguez-Solana, Artur J.M. Valente and Anabela Romano

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