Abstract
Inclusion complexation of the sunscreen ingredient avobenzone (AVB) with β-cyclodextrin (β-CD) was investigated to improve its aqueous solubility and photostability; another ultraviolet (UV) filter, oxybenzone (OXB), and the phytochemical antioxidant curcumin (CUR) served as a comparison. In this study, the 1-octanol/water partition coefficients, acid dissociation constants, phase-solubility diagrams with β-CD, and ultraviolet–visible (UV–vis) spectral changes induced by UVA1 (365 nm) irradiation were evaluated. β-CD at concentrations 50–100 times that of AVB most effectively protected the photostability of AVB. Additionally, an UVA1-insensitive species with a diketo tautomer, which has an UVC-absorbing band and the potential to cause photodegradation, was stored in the inclusion complex. Acetonitrile–water mixtures at various volume ratios were screened to mimic the internal cavity of β-CD for the AVB tautomeric species using nuclear magnetic resonance (NMR) spectral integrals for the components. The results indicated that β-CD provides a hydrophobic environment similar to that of a 40–50% acetonitrile aqueous solution and enhances the photostability of AVB. However, excess β-CD induced a hyperchromic effect on the diketo tautomer. Aggregation of the AVB/β-CD inclusion complexes at β-CD concentrations of ≥2 mM enhances UVC band absorption. To avoid excess β-CD, a molar ratio of 50–100 of β-CD to AVB is recommended as the optimal composition. This study newly exhibited that the cavity of β-CD mitigates the reactivity of UVA1 toward AVB by inducing the diketo tautomer form of AVB within the cavity.
1. Introduction
The United States Food and Drug Administration (U.S. FDA) began to use the term “generally recognized as safe and effective” (GRASE) for over-the-counter (OTC) drugs and ingredients in therapeutic products.1−3 The GRASE ingredients numbered in the hundreds and included many familiar products, such as sunscreen, pain relievers, and medicated lotions. On the basis of scientific evidence and clinical studies, the U.S. FDA reverted to the 2019 proposal that an OTC drug product is safe and effective for its intended use.4
The U.S. FDA proposed its latest regulatory update for the products to oversee sunscreen safety in 2019.4 On the basis of the available information, the agency reviewed 16 ingredients and reported that only two physical/inorganic/mineral products, ZnO and TiO2, were GRASE. In contrast, the following chemical/organic/synthetic ultraviolet (UV) filters are non-GRASE ingredients because of insufficient data (see Chart 1): the benzophenone derivatives, e.g., avobenzone [AVB, p-tert-butyl-p-methoxydibenzoylmethane (BMDM)], oxybenzone [OXB, benzophenone-3 (BP-3), 2-hydroxy-4-methoxyphenylbenzophenone (HMBP)], dioxybenzone (benzophenone-8), and sulisobenzone (benzophenone-4), the salicylate esters, e.g., homosalate and octisalate, the cinnamoyl esters, e.g., octocrylene (OCR), octinoxate (OCX), and cinoxate, and others, e.g., padimate O, meradimate, and ensulizole. para-Aminobenzoic acid and trolamine salicylate are no longer used in sunscreen marketed in the U.S.5
Chart 1. Chemical Structures and Properties of the Dominant Chemical UV Filter Ingredients5.
AVB, OXB, and homosalate are considered endocrine-disruptors.6 The U.S. FDA has reported that they and other ingredients, such as octisalate, OCX, and OCR, are systemically absorbed into the body and can be detected in the skin, blood plasma, breast milk, and urine samples weeks after use.7 The European Commission has ruled on the safety of homosalate and OCR, and, have proposed to limit allowable usage concentrations.8,9 This ruling does not imply that the other aforementioned UV filters are excluded.
These rules may not be well received by those with darker colored skin. The demand for organic UV filters is a testament to public concerns and preferences. Customers avoid sunscreen products containing inorganic UV-scattering agents, claiming that whiteness affects their makeup and appearance. The demand for organic UV filters is a significant factor in the sunscreen market. The maximum allowed concentration of OCX is 7.5% in the U.S. and Canada, whereas OCX is regulated at 10% in the European Union (EU), People’s Republic of China (PRC), and Australia.10−13 In Japan, an OCX concentration as high as 20% is accepted. In contrast, OCX-based products are banned in Hawaii because of their toxicity to marine ecosystems.14
We aimed to explore a strategy that enhances the aqueous solubility of hydrophobic ingredients or topical antidrugs to avoid transdermal adsorption into the blood.15−25 In addition, we aimed to clarify the molecular mechanism by which UVA (320–400 nm, frequently pouring regardless of climate) absorption and photodegradation of the benzophenone-type UV filter AVB are protected and sustained by the aqueous solubilizing reagent, cyclodextrins (CDs), which encapsulate ligands into their hydrophobic internal cavity and restrict tautomerization and degradation.26−29
AVB and OXB are favorable examples of UV filters because their spectroscopic traceability allows for physicochemical experiments.29 AVB, the most prominent UVA filter, is known for its ability to absorb a wide range of UVA rays, particularly in the UVA1 band (340–400 nm). Sunscreen formulations containing AVB provide broad-spectrum protection from UV exposure and are key ingredients in popular products. AVB is approved for use with maximum concentrations ranging from 3% in the U.S. and Canada to 5% in the EU and Australia.10−14
Given Lipinski’s rule of five,30 AVB with a partition coefficient of log KOW = 4.51 and molecular weight of 310.4 g/mol is expected, with a high skin permeability from 1.8 to 4.3 ng/mL.31 The keto-enol form with intramolecular hydrogen bonding (the chelated keto-enol form) is coplanar, photostable, and absorbs in the UVA1 band. In contrast, the diketo tautomer absorbs light in the UVC range (200–280 nm) and is prone to degradation.32 The efficacy of the keto-enol form as a barrier for UVA exposure and protection against skin cancer is reduced because tautomerism between the keto-enol and diketo forms leads to instability and harm.33,34 Notably, the photodegradation products of avobenzone, which are highly reactive radical species, potentially induce inflammation in skin tissue and have harmful effects on human health.35
Modifications to the scaffold are required to overcome these drawbacks of AVB; the modifications include swapping AVB aromatic groups, prohibiting aggregation and avoiding diketo formation with alternatives.36 Other methods for retaining the AVB structure include quenching the triplet excited state using chemical additives,37,38 scavenging radicals using antioxidants,39 and encapsulation in micelles/cyclodextrin/metal complexes.40 Electron density at the diketo group can be an efficient target to regulate the photodegradation process of AVB owing to α-cleavage of the diketo form with a triplet excited state via a Norrish type I reaction.34,35,37
2. Materials and Methods
2.1. Materials
AVB (CAS Registry Number 70356-09-1) was supplied by Fujifilm Wako Pure Chemical Industries (Osaka, Japan). OXB (131-57-7), curcumin (CUR, 458-37-7), β-CD (7585-39-9), hydroxypropyl-β-cyclodextrin (HP-β-CD, 128446-35-5), HPLC-grade 1-octanol, and deuterated solvents (D2O, methanol-d4, acetonitrile-d3, and other solvents) were obtained from Tokyo Chemical Industry (Tokyo, Japan). All of the other materials and solvents were of analytical grade.
The aqueous-phase solvents were prepared by mixing 25 mM KH2PO4 and 25 mM Na2HPO4 solutions in 25 mM Pi buffer. The pH was adjusted before fixing the prescribed concentrations of the acid components of the buffer. The JP1 and JP2 buffer solutions were 10 mM sodium citrate/HCl buffer (pH 1.2) and Pi buffer (pH 6.8), which were prepared in compliance with the description for solution media for the dissolution test in Japanese Pharmacopeia (JP)-implemented international harmonization. The modified Britton–Robinson universal pH buffer comprised 28.6 mM KH2PO4, 28.6 mM HBO2, and 28.6 mM NaCl, titrated with 1 M NaOH (here, modified means elimination of barbital to cancel its interaction with organic solutes). For the partition equilibrium examination, the aqueous saturated 1-octanol phase comprised 1-octanol, which was flooded with a modified Britton–Robinson buffer adjusted to the appropriate pH in advance.
2.2. Reversed-Phase High-Performance Liquid Chromatography (HPLC) Measurements
The sample solution was filtered with a membrane filter (Minisart RC 4, 0.22 μm pore size, Sartorius, Göttingen, Germany). Fractionation of the sample in the filtrate was performed by HPLC (SPD-20A, Shimadzu Co., Kyoto, Japan) with a mobile phase of 25 mM citric acid buffer (pH 3.0)/methanol (3:7) or 25 mM Pi buffer (pH 2.5)/acetonitrile-d3 (1:1) at a flow rate of 1 mL/min, using a reversed-phase column (Capcell Pak C18, Shiseido, 5 μm, 250 mm × ⌀ 4.6 mm) at a temperature of 313 K. The concentrations of AVB, OXB, and CUR were determined by monitoring with a photodiode array (PDA) detector at wavelengths between 200 and 600 nm. A single-beam detector was used at wavelengths of 360 and 270 nm for AVB and 430 and 230 nm for CUR.
2.3. Flask-Shaking Method for the 1-Octanol/Water Partition Experiment
The hydrophobic drug dissolved in the aqueous saturated 1-octanol phase was mixed with an isochoric solution of the modified Britton–Robinson buffer (pH 5–11.5), saturated with 1-octanol. Thereafter, the samples in the screw-capped vials were machine-shaken at 298 K in a thermostatic chamber for 60 min and subsequently kept stationary for 60 min; the procedure corresponds to the conventional “flask-shaking” method.41−43 After settling, the pH in the aqueous phase was confirmed using pH indicator strips (MColoHast pH 0–6.0, MQuant pH 7.5–14, and MQuant pH 0–14, Merck, Germany) to be that of the adjusted pH value. Aliquots (10 μL) from the 1-octanol and the aqueous layers were injected into the HPLC instrument. The 1-octanol/water partition coefficients, log P (equivalent to log KOW), and acid dissociation constants, pKa, were optimized using curve fitting, approximated by eq 1
![]() |
1 |
which is Avdeef’s diagram44 derived from the Henderson–Hasselbalch expression for the equilibrium between the protonated and deprotonated species. The curve-fitting procedures used the solver module of Microsoft Excel 2016 with the implemented GRG nonlinear option. Approximate curves corresponding to these optimized parameters were drawn, and the pKa and log P values for AVB, OXB, and CUR were obtained.
2.4. Phase-Solubility Isotherm Diagram of Drug to CDs
An excess amount of drug powder, coordinated by groping trials, was added to screw-capped vials containing the Pi buffer (5 mL, pH 6.8) in the absence or presence of β-CD or HP-β-CD. The solutions were mechanically shaken at 298 K in a thermostatic chamber, and the supernatant was sampled and filtered after passage for appropriate periods. The concentration of the drug in the supernatant was determined using HPLC.
The phase-solubility diagram consists of the equilibrium concentration of the guest drug as the ordinate and the total concentration of the host CD as the abscissa.45 If a straight line and parabolic curve can be approximated on the phase-solubility diagram, then, the curves are classified as AL- and AP-types in the Higuchi and Connors catalogs, respectively.45,46 For the linear correlation corresponding to the AL-type, regression analysis of the experimental data set using eq 2 provides the stability constant K1:1 for equimolar drug/CD complexes
![]() |
2 |
where D0 is the solubility of the drug in the absence of CD, which is comparable to the ordinate intercept, and the slope is the gradient of the ordinate, referred to as the abscissa.
The parabolic curve corresponding to the AP-type indicated that the complexes were associated with a stoichiometry of equimolar guest/host and that of single guest/double hosts (or further convoluted proportions).45,46 Multiple regression analysis of the measured data set to the total concentration of the host ([CD]) and its squares ([CD]2) using eq 3 can be used to obtain the stability constants for a combination of the equimolar complex K1:1 and double-host associating complex K1:2.45,46
![]() |
3 |
This is the sum of the dissociated (free) and associated (complexed) drug concentrations. Although the Benesi–Hildebrand equation with the square of the host concentration has been used for parabolic correlations, its validity remains unclear.
Occasionally, a phase-solubility diagram exhibits a hyperbolic curve, referred to as a saturation curve. This curve is classified as the AN-type on the Higuchi and Connors catalog.45−48 The curve is described by eq 4′ according to the Langmuir adsorption isotherm model.47−51 To process with the linear regression analysis of ordinate ([CD]/γ) referred to abscissa ([CD]), the Hanes–Woolf expression in eq 4′ may be used
![]() |
4 |
![]() |
4′ |
where n indicates the asymptotic concentration of the drug associated with CD (saturation, maximal γ) and K is equivalent to the drug/CD complex stability constant. If the concentration of CD is comparable to reciprocal K, then the obtained γ is accorded to the half amount of n. Notably, the saturated curve illustrates that the solubility of the guest–host complex is less potent and that the association is more complicated.46 The stoichiometry of a complex of drugs with CD is difficult to determine; the prevailing opinion is that more than one guest molecule incorporates a host molecule and that hydration or ionization of the guest regulates the solubility of the complex.46
Szejtili46 stated that, in rare cases, the ascending region at lower host concentrations is followed by a plateau region, similar to that of the AN-type saturation curve. Simultaneously, it terminates at any host concentration threshold. It is named the BS-type in the Higuchi and Connors catalog,45−48 and for concentrations higher than the threshold, the guest concentration decreases along a hyperbola, probably according to the solubility product KSP.49 To analyze this type of pattern, a hypothesis that is yet to be discussed is required.52 For the concentration of the drug in the plateau region, a soluble complex of the drug and CD is enlarged by adhesion of the excess CD, similar to a snowball. At the threshold concentration of CD, the scale of the aggregated dispersoids transcends an upper limit, and the aggregation induces subsequent sedimentation.52 Thus, the solubility product may be considered a property of poorly soluble aggregated dispersoids containing a tolerable amount of the drug and an excess amount of CD.
2.5. UV–Vis Spectroscopy for CD Inclusion and UVA1 Photodegradation
UVA1 (365 nm) was dissolved in the 25 mM Pi buffer (pH 6.8) in the absence or presence of β-CD and was uniformly irradiated onto samples using an UV lamp (GL15, Toshiba Co., Tokyo, Japan, 15 W) at an optical path length of 15 cm. The samples were obtained after passage for appropriate periods and measured using an UV–vis spectrophotometer in the maximum range between 200 and 700 nm (arranged on demand) from long wavelengths toward short wavelengths to avoid unnecessary irradiation of the UV region. To obtain HPLC chromatograms, aliquots of the UV-irradiated solution were obtained by pipetting at 0, 1/5, 1/2, 1, 2, 3, 4, 24, 48, 72, 96, 120, 144, and 168 h for AVB and at 0, 1, 2, 3, and 18 h for CUR. To obtain those of OXB in the absence or presence of β-CD, the invariable samples were certified at 0 and 48 h. For the samples in acetonitrile aqueous solution or methanol aqueous solution at a ratio of 7:3, the time evolution of the UVA1 irradiated solution was examined under similar conditions.
The photostability of the drug under solar light was evaluated using UV–vis spectral measurements. Sample solutions were exposed to solar light outdoors at a sunny location on a latitude of 35.92° N and a longitude of 139.91° E for 4 h, from 12:00 to 16:00, at an average temperature of 284.55 K.
2.6. Singular Value Decomposition (SVD) Computation for Spectrometric Data
The ith observed spectrum {Φ⃗i|1 ≤ i ≤ n} of the sample represents an m-dimensional vertical vector measured at a specific wavelength. The wavelength range spans 230–700 nm with an interval of 1 nm, resulting in m = 471. Matrix M, defined in eq 5, consists of a horizontal sequence of the obtained spectral vectors (with dimensions m × n = 471 × 64).
![]() |
5 |
M and Mt represent the real and transposed matrixes, respectively. The products MtM and MMt form orthogonal matrixes. The matrixes describing M can be transformed into eq 6.
![]() |
6 |
The matrix ∑ comprises the diagonal elements {σi|1 ≤ i ≤ r} (positive absolute values ordered in descending order representing singular values denoting dispersion). The ith column of the orthogonal matrix Λ is the coefficient vector corresponding to the singular value (σi) and vector (λ⃗i) is a specific singular vector. The rows of matrix Ψ are considered basis function vectors; the principal component vector (ω⃗i) results from the product of λ⃗i and the corresponding σi.
![]() |
7 |
Matrix Ψ comprises rows that are basis function vectors. We applied SVD to a 471 × 64 spectral data matrix. The dimensionality was determined based on the logarithm of the singular value against the index, establishing the minimum dimensionality required to replicate the vector space of the document spectrum. The value can be almost negligible if the singular value is much smaller than a certain threshold (e.g., several hundredths of the highest singular value). The chosen dimensionality r enables the principal components to approximately reproduce the vector space, including the document spectrum as the j–h vector (x⃗j) composed of the ith elements (xi,j), as described in eq 8.53−60
![]() |
8 |
2.7. Nuclear Magnetic Resonance (NMR) Spectroscopy
1H NMR measurements were conducted using a 400 MHz NMR spectrometer (JNM-ECZ 400 S, Japan Electronics Co., Ltd., Tokyo, Japan). Sample solutions were prepared using D2O and methanol-d4 as the protic solvent, acetonitrile-d3 as the aprotic solvent, and their mixtures, with sample concentrations exceeding 1.5% (w/v). The chemical shifts of the samples were calibrated using the internal tetramethylsilane (TMS) signal as the zero point, whereas the solvent signals in the literature served as reference points. When sodium salts were required, the neutral AVB species in D2O were dissolved in equimolar amounts of aqueous NaOH and methanol-d4, followed by drying under reduced pressure in a rotary evaporator and heating to dryness below Tm of the sodium salts. The formation of sodium salts was confirmed by measuring Tm using a differential scanning calorimeter (DSC8230, Rigaku Co., Ltd., Tokyo, Japan).
1H–1H homonuclear correlation spectroscopy (COSY) involves scanning electromagnetic radiation pulses through hydrogen nuclei, eliciting responses from resonant hydrogen atoms with geminal, vicinal, or long-range coupling. The diagonal signal corresponds to the hydrogen response to the scanned radio waves at a specific frequency, whereas cross peaks that do not align with the diagonal reveal adjacent hydrogens. 1H–1H homonuclear Overhauser effect spectroscopy (NOESY) identifies signals arising from hydrogen atoms in close spatial proximity, providing through-space correlations via spin–lattice relaxation. For optimal spectral assignment by NOESY, the mixing time should fall between half of T1 and T1, with increasing longitudinal relaxation time, enhancing NOESY sensitivity. This can be achieved by selecting a low-viscosity solvent (such as acetone-d6) and removing the dissolved oxygen from the sample.
1H–13C heteronuclear multiple quantum coherence (HMQC) measurements detect coupling cross peaks between carbon-13 nuclei adjacent to scanning protons. The experiments were conducted according to a previously described protocol.22 When the sample produced confusing signals in the HMQC experiments, 1H–13C heteronuclear single quantum coherence (HSQC) measurements were used to obtain a higher resolution. 1H–13C heteronuclear multiple bond coherence (HMBC) measurements identify coupling cross peaks between the carbon-13 nuclei and scanning protons through two or three bonds. This measurement followed the experimental protocol.
2.8. NMR Titration for the Tautomeric Species of AVB
In the one-dimensional (1D) 1H NMR spectra, the integral intensities correspond to the relative stoichiometry of protons for the species, except for the signals assigned to the stable isotopes of deuterated solvents. We could recognize the aromatic proton signals of AVB, but the corresponding signals sometimes overlapped between its keto-enol and diketo species, depending upon the solvent composition. The signals of the aromatic protons at the para positions to the keto/enol substituents were neglected because of the high probability of overlap, in which the chemical shift drifts, Δδ, related to the shielding effect are not noticeably accompanied by the electronic states in the keto/enol substituents.
We chose the signals of the aromatic protons at the ortho positions and the aliphatic protons at the ortho substituents (tert-butyl and methoxy groups) to keto/enol substituents and measured their integral intensities. Using the simultaneous equations built as linear combinations of integral values for the isolated and overlapping signals, we determined the molar ratio of the keto-enol and diketo species.
3. Results and Discussion
3.1. 1-Octanol/Water Partition Coefficients of AVB, OXB, and CUR
AVB absorbed a unique UVA1 band, whereas OBZ was comparatively verified as an UV filter absorbing UVB (280–320 nm) and UVA2 (320–340 nm) light with a benzophenone moiety (cf. Chart 1).36 AVB has a bibenzoylmethane scaffold, whereas CUR was used as an antioxidant reference for diketomethane and keto-enol tautomerization.61 We evaluated the hydrophobicity of these samples using the flask-shaking method to determine the partition coefficient between the 1-octanol and aqueous phases,41−44 in which the pH of the aqueous phase was regulated between 5.0 and 11.5 with a modified Britton–Robinson universal buffer. The limits of detection for the calibration curves of AVB, OXB, and CUR prepared using reversed-phase HPLC were 0.0055, 0.0323, and 0.0030 mM, respectively, while the limits of quantitation were 0.0142, 0.0881, and 0.0112 mM, respectively. Figure 1 shows the pH profiles of the apparent partition coefficients (distribution coefficients), and log D for AVB, OXB, and CUR. Curve fitting for Avdeef’s log D–pH diagram for acidic compounds was used to estimate the pKa and log P values for the neutral form.41 The reported pKa values of AVB and OXB are 9.7 and 7.6, and the log P values of AVB and OXB are 4.51 and 3.79.62 CUR has three pKa values of 7.8, 8.5, and 9.0, corresponding to the keto-enol, first phenolic OH, and second phenolic OH, respectively.63 The updated pKa values of CUR were 7.56, 8.72, and 10.17,64 and its log P was 3.2.65 Our obtained pKa and log P values were consistent with published values, but the order of log P values was AVB (4.27) > OXB (2.99) > CUR (2.65). Curcumin contains a keto-enol site and two phenolic hydroxyl groups. The acid dissociation constants of dissociable groups with the same structure change consecutively. In other words, the acid dissociation constant of one group changes depending upon whether or not the other group has dissociated. Cantor and Schimmel demonstrated that the acid dissociation constant of the terminal group in oligopeptides composed of alanine changes with peptide chain length.66 Therefore, it is challenging to distinguish the apparent pKa values derived from the two phenolic hydroxyl groups. In Figure 1c, only one pKa was calculated; however, the plot at pH 11.5 showed lower values compared to Avdeef’s log–pH plot. This suggests that the dissociation of the second phenolic hydroxyl group is occurring. This confirmed the presence of a few ionized forms of AVB, OXB, and CUR at pH 6.8.
Figure 1.
pH profiles of distribution coefficients, log D, for (a) AVB, (b) OXB, and (c) CUR. The 1-octanol/water (the modified Britton–Robinson buffers at pH 5–11.5) partition coefficients, log P, the acid dissociation constants, pKa, were optimized using curve fitting approximated to log D = (log P)/(1 + 10pH–pKa), which is Avdeef’s diagram41 derived from the Henderson–Hasselbalch expression for the equilibrium between the protonated and deprotonated species. Curve-fitting procedures employed the Solver module of Microsoft Excel 2016 with the implemented GRG nonlinear option. The approximated curves corresponding to these optimized parameters were drawn, and the pKa and log P values for AVB, OXB, and CUR were represented. The inset photograph shows the equilibrium CUR partitioned in 1-octanol/water phases at several pH values.
3.2. Phase-Solubility Diagrams of AVB, OXB, and CUR with CDs
The CDs are a macrocyclic hexamer (α-CD), heptamer (β-CD), and octamer (γ-CD), that consist of glucopyranosides linked by α1–4-glycoside bonds.46,67 Consisting of a cavity (with a diameter of 0.60–0.65 nm) appropriate to include various phenyl portions, β-CD has been widely used for the mechanochemical syntheses of pharmaceutics.67 Because its use is limited owing to its notable low aqueous solubility (16.3 mmol/L) and nephrotoxicity suspicion, especially in parenteral drug delivery,67 alternatively, the partially 2-hydroxypropylated derivative of β-CD (HP-β-CD) was developed, which is more highly soluble (more than 400 mmol/L) and can be chemically degraded and photodegraded.65 The derivative is considered to have higher solubility and greater nontoxicity at low to moderate oral and intravenous doses.67 Additionally, the derivative has generally been verified as safe and is administered parenterally to animals and humans.68
Figure 2 shows the phase-solubility diagrams of AVB, OXB, and CUR to β-CD or HP-β-CD. In panels a and b of Figure 2, the apparent solubility to β-CD of AVB and OXB increases linearly, but plateaus appear at concentrations greater than 5 or 6 mM of β-CD. We attempted to explain their profiles using Langmuir adsorption isotherms and Hanes–Woolf plots, which proved unsuitable. Vishwakarma et al.69 reported the Benesi–Hildebrand analyses of the phase-solubility diagram for ABZ and β-CD, and concluded that linear or parabolic relationships prevailed; that is, the stoichiometry of AVB and β-CD was 1:1 or 1:2, when the researchers examined less than 1 mM β-CD. Given that our results indicated a linear correlation, we confirmed the equimolar complex of AVB and β-CD in the 0–6 mM range. Precipitation was observed at the higher concentration of β-CD, which might have suspended the AVB/β-CD or OXB/β-CD inclusion complex. Sbora et al.70 determined the molar ratio of AVB with β-CD and HP-β-CD to be 1:2 in the solid lyophilized inclusion. The theoretical study of Vishwakarma et al.69 supported the probability of the 1:2 inclusion complex of AVB/β-CD.
Figure 2.
Phase-solubility diagrams of (a) AVB, (d) OXB, and (c) CUR with β-CD and (d) AVB, (e) OXB, and (f) CUR with HP-β-CD in 25 mM Pi/NaOH buffer at pH 6.8. The regression lines for panels a, b, and e provided the determination coefficients, 0.9223, 0.9994, and 0.9872. The parabolic curves for panels c, d, and f brought the determination coefficients, 0.9946, 0.9995, and 0.9901.
The pattern shown in Figure 2b is consistent with that reported by Sarveiya et al.71 Although the phase-solubility curves of AVB and OXB for β-CD appeared to correspond to the BS-type in the conventional catalog of Higuchi and Connors,45−48 they differed from the BS-pattern; the appearance of overhangs above the plateau region in the higher β-CD concentration, confirmed the results obtained by Simeoni et al.72 Assuming the suspended AVB/β-CD inclusion complex would transform from the equimolar form to the double-side-capped complex (with a stoichiometry of 1:2) at the higher β-CD concentration, as mentioned above, the gap between the highest solubility on the linear correlation (i.e., the formation of the equimolar complex) and plateau level (i.e., the double-side-capped complex or more coagulated particles) could be acceptable as an indication of the difference in the solubility (dispersion concentration) of the different forms. We observed that precipitation probably corresponded to this gap.
Figure 2e indicates that the apparent solubility of OXB increases linearly in the diagram for 0–100 mM HP-β-CD (r2 = 0.9872). Corresponding to the AL-type described by Higuchi and Conners,45 this phenomenon indicates the formation of the OXB/HP-β-CD equimolar complex. Despite resembling the results obtained for OXB/HP-β-CD (0–60 mM) by Sarveiya et al.,71 a notable difference is evident; the slope obtained by Sarveiya et al.71 overlapped the results for OXB/HP-β-CD approximately 4 times in Figure 2e. This difference was probably caused by differences in the experimental conditions (solute concentration, pH, or temperature). Simeoni et al.72 reported a lower OXB/HP-β-CD diagram stability constant. A generality of the AL-type correlations observed in the OXB inclusion complexes with other cyclodextrins: HP-α-CD, HP-γ-CD, and sulfobutyl ether (SBE)-β-CD was evident. Because the UV–vis spectrum of OXB remained almost unchanged in the absence and presence of 4 mM β-CD before UVA1 irradiation (Figure 3), the electronic structure of OXB was sustained in the solution or the equimolar inclusion complex.
Figure 3.
UV–vis spectra of 50 μM (a) AVB, (b) OXB, and (c) CUR irradiated with the UVA1 (365 nm) lamp in 25 mM Tris–HCl (pH 7.4). The spectra of panels d–f correspond to these drugs irradiated with the UVA1 lamp in 8 mM β-CD.
Figure 2d shows the parabolic curve obtained for AVB/HP-β-CD. The physicochemical properties, photodegradation, and chemical stability of many inclusion/encapsulated complexes of AVB with HP-β-CD have been studied.36−38 As Yang et al.73 and Yuan et al.74 published parabolic curves similar to those of the AVB/HP-β-CD phase-solubility diagrams obtained in this study, the AVB complex can be assumed to be double-side-capped with two HP-β-CDs at both benzoyl sides of AVB. Using traditional regression analysis with the 2-order equation, we obtained the stability constants for an equimolar and 1:2-ratio complex of K1:1 = 90.3 L/mol and K1:2 = 40.4 L2/mol2 (r2 = 0.9995). The stability constants indicate that the affinity of AVB for β-CD (K1:1 = 1023 L/mol, r2 = 0.9223) is approximately 10 times greater than that for HP-β-CD as the equimolar complex.
Panels c and f of Figure 2 illustrate the Higuchi and Conners AP-type curves45 obtained for the CUR/β-CD and CUR/HP-β-CD complexes. This suggests that the double-side-capped CUR complex comprises cyclodextrins with equivalent feruloyl (i.e., 4-hydroxy-3-methoxycinnamoyl) moieties.75,76 Arya and Raghav77 justified this 1:2 stoichiometric ratio for CUR/β-CD. Although Karpkird et al.78 appropriated an AL-type line (r2 = 0.9464), their plots appear to follow a downward convex curve. Jahed et al.79 proposed an AL-type relationship between CUR solubility and the β-CD concentrations. However, the diagram reported by Jahed et al.79 included a result for 20 mM on the abscissa, which was excess to the β-CD solubility. Reducing the plot to a maximum of 16.3 mM was expected to yield a gradual parabolic curve.
Ghanghoria et al.80 indicated that the phase-solubility diagram was of the BS-type. However, the researchers did not report the total amount of CUR in their screw-capped vials, and we speculated that CUR solubility was not greater than 0.016 mM. Mashaqbeh et al.81 also reported an AN-type diagram (saturated at most with 0.00012 mM CUR) that was examined under similar circumstances. Interpretations of the physicochemical and quantitative properties of dietary ingredients and CDs in literature require reasonable agreement. Although we considered the above reports did not essentially contradict our parabolic curves, we failed to rationalize that CUR/β-CD and CUR/HP-β-CD provide the AL-type phase-solubility diagrams reported by Jantarat et al.82 On the basis of our results, the diagrams for CUR/β-CD and CUR/HP-β-CD should be parabolic. The calculated K1:1 values for CUR with β-CD was 771.7 M–1. This K1:1 value is in the same order of magnitude as described by Mashaqbef et al.81 (487.3 M–1) and Chen et al.83 (198 M–1). In the same way, the K1:1 value of the CUR/HP-β-CD is 6619 M–1, it is in the same order of magnitude as described by Celebioglu and Uyar84 (3073 M–1) and Li et al.85 (2941 M–1). It was confirmed that the singly occupied complex of CUR with cyclodextrin is predominant, and the K1:1 value for CUR/HP-β-CD (r2 = 0.9901) is 8.6 times higher than that for CUR/β-CD (r2 = 0.9946).
In conclusion, the apparent solubility of AVB, OXB, and CUR increased in a linear or parabolic manner depending upon the cyclodextrin concentration up to 4 mM. Furthermore, the stability constants for the equimolar complexes of AVB, OXB, and CUR to β-CD were more significant than those to HP-β-CD. Although HP-β-CD was used in many studies to obtain the AVB or OXB inclusion complexes,70−74,82 we confirmed that the efficacy of forming the β-CD inclusion was higher than that of forming the HP-β-CD inclusion. The stability constants were independent of the hydrophobicity of AVB, OXB, and CUR. Further study focused on the effects of this CD because β-CD efficiently improved the AVB aqueous solubility and held no structural diversity.
3.3. UV–Vis Absorption Spectral Change of AVB with/without β-CD
Figure 3 shows the UVA1 photodegradation of 50 μM AVB, OXB, and CUR in the absence or presence of 4 mM β-CD. Absorption peaks of AVB in the aqueous phase were observed at 360 and 388 nm. According to the theoretical study by Sahoo et al.,86 the chelated (intramolecular cyclic hydrogen bonding) keto-enol forms of p-tert-butylbenzoyl-p-methoxyphenyl–enol (χ1, form A) and p-methoxylbenzoyl-p-tert-butylphenyl–enol (χ2, form B) correspond to the higher and lower wavelength peaks, respectively. In various solvent studies performed by Mturi and Martincigh,87 the maximum peaks (360 nm in the aqueous phase) were found at 365 nm in dimethyl sulfoxide (DMSO), 358 nm in methanol, 355 nm in ethyl acetate, and 350 nm in cyclohexane. As the chelated form is dominant in nonpolar solvents, the 360 and 388 nm peaks could be considered as corresponding to the chelated and nonchelated forms. In the presence of β-CD, the AVB spectrum lacks a shoulder at 388 nm. Assuming AVB intrudes into the hydrophobic internal cavity of β-CD, hydrogen bonding becomes dominant. This suggests the 388 nm peak corresponds to the nonchelated form. In a hydrophobic environment, the diketo form of AVB increases to enhance the 267 nm peak.
Figure 4a shows the titration results obtained for 50 μM AVB with 0–10 mM β-CD. The 360 and 388 nm peaks of AVB were observed in the absence of β-CD and gradually increased depending upon the β-CD concentration which ranged between 0 and 2 mM. However, when the β-CD concentration increased to more than 2 mM, the 388 nm peak disappeared, but the 267 and 360 nm peaks increased gradually. Vishwakarma et al.69 reported that the spectra of 10 μM AVB decreased for 0–1 mM β-CD and increased for 1–10 mM β-CD, indicating that the absorption change switched from decrease to increase at a threshold of 80–100 times the β-CD concentration to AVB. According to the researchers, the peak decrease at 360 and 388 nm for β-CD concentrations less than this threshold is induced owing to a polarity effect on the main absorption band of a π–π* nature in AVB. The switch to the increment of the 360 nm peak at a higher β-CD concentration than the threshold indicates the possibility of multiple host–guest complexation. This diversity is consistent with the possibility of the divalent binding of AVB to β-CDs at concentrations greater than 6 mM in the phase-solubility diagram (Figure 2a).
Figure 4.
(a) UV–vis spectral titration for 50 μM AVB with 0–10 mM β-CD. In 25 mM Pi buffer (pH 6.5). The spectra of AVB were represented with gradually diluted curves depending upon the concentration of β-CD. They peaked at 388 nm, and spectra gradually morphed and descended with 0, 1/40, 1/20, 1/10, 1, and 2 mM β-CD. Furthermore, twin peaks at 274 and 365 nm expanded with 2, 4, 6, 8, and 10 mM β-CD. The signal at 388 nm would be assigned to the keto-enol species in the aqueous phase, while the signals at 274 and 365 nm would be to the diketo and chelated keto-enol species in β-CD’s hydrophobic internal cavity, respectively. UVA1 irradiation insignificantly degraded AVB with (b) 2 mM and (c) 4 mM β-CD for 0–6 days.
Figure 5 shows the reversed-phase HPLC chromatogram of AVB that was obtained using an isochoric acetonitrile/water solvent as the mobile phase. A component with a 270 nm peak at a retention time of approximately 2 min and a 360 nm peak at 4 min is present. The mobile phase easily eluted the diketo form owing to its hydrogen-bonding acceptors and structural flexibility. In contrast, the stationary phase could retain the chelated keto-enol form because of the shielded acceptor and expanded planar structure.
Figure 5.
(a) Reversed-phase HPLC chromatogram of intact AVB. Stationary and mobile phases were the ODS column and the Pi buffer (pH 2.5) in H2O/acetonitrile = 1:1, respectively. Signals were observed with the photodiode array (PDA) at 200–600 nm wavelength. Panels b and c represented the cross sections of spectra at the retention times tR of 2 and 4 min, respectively. The keto-enol form is planar and absorbs in the UVA1 range (340–400 nm), while the diketo tautomer absorbs in the UVC range (200–280 nm)81.
The spectra of OXB and the OXB/β-CD shown in Figure 3, exhibit no noticeable difference in the peak heights at 243, 291, and 325 nm. The spectral peak at 432 nm of the CUR/β-CD complex is narrower than that of the neat CUR because CUR would intrude into the hydrophobic internal cavity of β-CD to locate in an apolar atmosphere and restrict its fluctuation. In contrast, the OXB spectra that exhibit no difference indicated that the complexation of OXB into β-CD afforded no changes. Al-Rawashdeh et al.88 confirmed little change in the NMR chemical shifts of OXB and its complexes. In their scheme, based on theoretical computations, the benzophenone moiety intrudes into the hydrophobic internal cavity of β-CD, and its keto-phenol, in which the hydrogen acceptor site interacts with one of the C6-OH groups in β-CD, is held at the interface between the internal cavity and external aqueous phase.
3.4. Photodegradation of AVB, OXB, and CUR in Aqueous Solution
As shown in Figure 3 and panels b and c of Figure 4, the UVA1 lamp irradiated at 365 nm. In the absence of β-CD, AVB gradually photodegraded during UVA1 irradiation for a week and exhibited a hypochromic effect on the absorbance at λ = 267, 360, and 388 nm; the byproducts increased as indicated by the peak at 250 nm. In the presence of β-CD, the 360 nm peak decreased slowly and partially, and the 267 nm peak decreased more slowly than the 360 nm and partially. The increase in the intensity of the peak or shoulder at 250 nm was insignificant, suggesting the protection from AVB photodegradation. Panels b and c of Figure 4 demonstrate UVA1 irradiation insignificantly degraded AVB upon adding 2 mM and 4 mM β-CD. β-CD with a molarity 80–100 times that of AVB induced the hypochromic effect on the absorbance of AVB and prevented its UVA1 photodegradation.
Figure S1 shows the RP-HPLC patterns measured at λ = 360 nm (panels a and c) and 270 nm (panels b and d). Without β-CD, the signals at the retention time tR of 2.2 min observed at λ = 360 and 270 nm decreased during 2-d irradiation. Thereafter, the signals at tR = 1 and 1.4 min observed at λ = 270 nm increased from 2 days of irradiation onward, suggesting AVB was photodegraded during this irradiation. With β-CD, the signal at tR = 2.2 min observed at λ = 360 nm decreased slightly during 7 days of irradiation. The signals at tR = 2.2 and 1.2 min observed at λ = 270 nm remained. The signal at 1 min increased marginally, indicating that AVB in its complex with β-CD was partially photodegraded. Although the signal heights could not be compared between the 360 and 270 nm signals, the 360 nm signal decreased gradually, but the 270 nm signal increased. The keto-enol species encapsulated in the β-CD complex transformed into the diketo species, certifying photostabilization under the 365 nm lamp. The keto-enol species encapsulated in the β-CD complex transformed into the diketo species as if AVB intended to escape photodegradation under UV irradiation. This metaphor reflects a crucial concept of this study.
Because OXB absorbs light in the UVB (280–320 nm) and UVA2 (320–340 nm) regions, the 365 nm lamp could not presumably affect the OXB and the OXB/β-CD complex. Panels b and e of Figure 3 show the maintenance of the OXB spectrum for 2 days. CUR exhibits an absorption spectrum with a peak at 432 nm, which increased by the formation of an inclusion complex with β-CD (panels c and f of Figure 3). When exposed to UV light for 4 h, the CUR solution containing β-CD displayed disruption of 31.90%, compared to that (40.95%) by the CUR solution without β-CD. These results indicate that the formation of the inclusion complex between CUR and β-CD helps suppress the photodegradation of CUR. Many scholars have complexed curcumin with CDs to protect the molecule from photodegradation.89 However, marked differences are evident in the photostabilization effects of various cyclodextrins.90 Complex formation has a destabilizing effect rather than a photostabilizing effect on the photodegradation process, and an increase has been observed in the degradation rate.89 In organic solvents, CUR decomposes upon exposure to light. Whether the hydrophobic internal cavity of β-CD renders CUR photostabilizing depends upon the situation.
Figure S2 shows the RP-HPLC chromatogram of CUR containing these three components. The mobile phase comprised 40% acetonitrile in phosphate buffer (pH 2.5). The spectra of the components exhibit single peaks at 430, 420, and 415 nm. In the aqueous phase, the spectra of CUR and the CUR/β-CD complex peaked at 430 nm and shouldered at 360 nm; the shoulder did not correspond to the components of CUR, thereby suggesting the diketo tautomer of CUR. The 360 nm peak of the chelated keto-enol tautomer of CUR could become narrower owing to the formation of the inclusion complex with β-CD.
Figure S3 shows the changes in UV–vis spectra of AVB, OXB, and CUR in the presence and absence of β-CD under solar light. The evaluation of drug stability under solar light was conducted by modifying a previously reported method.91 For AVB, no significant changes in stability were observed under solar light exposure, regardless of the presence or absence of β-CD, compared to UVA1 irradiation. In the case of OXB, the change in absorption peaks due to solar light exposure was more significant than that observed under UVA1 irradiation. Because OXB is reported to be unstable to UVB (280–320 nm),92 it is presumed that the slight UVB present in solar light accelerated the degradation of OXB. For CUR, significant changes in absorption peaks were observed under solar light in the presence of β-CD. These changes were not observed under UVA1 irradiation, suggesting that UVB and visible light present in solar light may affect the stability of CUR. Additionally, in AVB and CUR solutions containing β-CD, precipitation was observed under solar light exposure, resulting in an elevated baseline.
Conclusively, although β-CD encapsulation can occasionally enhance and, in contrast, suppress the stability of the guest,27,51 β-CD canceled the photodegradation of AVB.
3.5. Photodegradation of AVB, OXB, and CUR in 50% Acetonitrile Solution
In section 3.3, we discussed the effect of β-CD encapsulation on the spectrum of AVB, which indicated that the contribution of UVA1 irradiation on the AVB/β-CD complex cannot directly compare with that on neat AVB. Figure 5 indicates the multiple retention times at λ = 270 nm for the diketo forms corresponding to their rotamers and the single retention time at λ = 360 nm for the chelated keto-enol form. Therefore, the mobile phase (acetonitrile/H2O = 1:1) avoids the structural diversity of the keto-enol species in an aqueous solution. Hanson et al.33 reported that SDS micellar encapsulation transforms the AVB spectrum to a level similar to that of methanol. The researchers discovered that AVB photodegradation due to solar-mimicry light irradiation progressed in water or cyclohexane but was completely inhibited in dry methanol or 20 mM SDS aqueous solution (AVB to be quarantined from the water phase). Our preliminary experiments allowed the aqueous solvent containing isochoric acetonitrile to mimic the hydrophobic atmosphere in the internal cavity of β-CD.
In this study, we examined the effect of UVA1 irradiation on AVB, OXB, and CUR in the absence or presence of β-CD in an isochoric acetonitrile/water solvent, as shown in Figures 4, 5, and 6, respectively. In this solvent, the AVB photodegradation products were minimal. Adding equimolar or 4 mM β-CD did little to modify the observed 25 μM AVB, OXB, and CUR spectra of this solvent. UVA1 irradiation of AVB involved a 360 nm peak reduction and 270 nm peak increase. These spectral drifts were slightly attenuated in the presence of β-CD. We determined the AVB concentration based on the calibration line of the absorbance at λ = 360 nm of the AVB standard. Figure 6a shows the time course of the AVB concentration under UVA1 irradiation. Equimolar amounts and 4 mM β-CD reduced the photodegradation velocity of AVB. However, the obtained diagrams of the logarithm of the reactant concentration versus reaction time appeared to have downward convex curves instead of first-order linear functions. This implies that the photodegradation of AVB followed a higher-dimensional or composite reaction. Furthermore, as their intercepts decreased depending upon the β-CD concentration, the opposite equilibrium between the keto-enol (360 nm peak) and diketo (270 nm peak) species might compete with the photodegradation. In other words, the keto-enol species in this solvent were interconverted into diketo species, as if AVB intended to escape photodegradation under UVA1 irradiation.
Figure 6.
Time evolution of the degradation of (a) 25 μM AVB, (b) 50 μM OXB, and (c) 25 μM CUR in the absence (black circles) and presence (equimolar, Bordeaux triangles; 4 mM, indigo rhombuses) of β-CD in isochoric acetonitrile solvent under the UVA1 irradiation (365 nm). AVB, OXB, and CUR signals and their calibration lines were determined wavelengths at 360, 290, and 430 nm, respectively. The calibration lines provided r2 = 0.999 and 3σ = 0.022. The obtained diagrams of the logarithm of AVB concentration to the reaction time seemed to have downward convex curves rather than first-order linear functions. Figures S4–S6 represented the corresponding observed spectra.
UVA1 irradiation did not significantly influence the spectra of OXB in the isochoric acetonitrile/water solvent. Figure 6b indicates that the OXB concentration derived from the 290 nm absorbance was marginally reduced by UVA1 irradiation for up to 50 h, as expected. As shown in Figure S7, a longer duration caused peaks intensities in the CUR spectra to decrease. Adding β-CD delayed the CUR decrements under irradiation. Figure 6c shows the time course of the CUR concentration observed as the absorbance at λ = 430 nm and first-order reaction rate of the photodegradation of CUR, which decreased upon adding the equimolar and 4 mM β-CD. Mangolim et al.75 explained that the primary degradation product is the cyclization of curcumin due to the loss of two hydrogen atoms from the molecule, which is also derived from vanillin, vanillic acid, ferulic aldehyde, ferulic acid, and 4-vinyl guaiacol.63 These tendencies do not contradict our results for the first-order degradation of CUR.
We conclude that the photodegradation kinetics of AVB involve a composite reaction process, such as keto-enol and diketo tautomerization.
3.6. 1H NMR Titration of the Keto-Enol and Diketo Forms of AVB
UVA1 irradiation induces the photodegradation of AVB in an isochoric acetonitrile/water solvent. With the 360 nm peak height decreasing along a downward convex curve to the cumulative UV-absorbing energy, the 270 nm peak gradually increased. This indicates that the chelated keto-enol species absorbed UVA1 light and was subsequently photodegraded or partially converted to the diketo species. Figure 7 shows the absorbance of AVB at λ = 270 and 360 nm in a solvent with various proportions of acetonitrile/H2O and methanol/H2O. Although increasing the 360 nm absorbance induces a reduction in the 270 nm absorbance, the absorbances measured at different wavelengths cannot be quantitatively compared because of the different molar adsorption coefficients. The molar ratio between the coexisting keto-enol and diketo species was used to calibrate the absorbance to the molarity.
Figure 7.
Absorbance of AVB at 270 and 360 nm in a solvent with various proportions of (a) acetonitrile/H2O and (b) methanol/H2O. It represented intact ABZ with circles and the equimolar mixture of AVB and β-CD with triangles. Supposing the solvent molecules are homogeneous volume spheres, the geometrical kissing number in three dimensions is 12. It indicates a hexagonal close-packed lattice where a central sphere contacts 3–4 surrounding atoms. Hence, if the solvent component contains a higher volume ratio than 66–75%, the surroundings of a solute would be almost filled with these solvent molecules. As acetonitrile and methanol molecules are distorted, we considered that this threshold proportion would be attenuated. That could be why the bending points are at about 60% (v/v)24.
In the 1H NMR spectrum, the signal integrals of the assigned species provided a molar ratio. Figure S8 shows the 1H NMR spectrum of AVB dissolved in methanol-d4/D2O = 7:3 solvent. The signals at δ = 8.1205 ppm (J = 9.15 Hz) of the aromatic ortho-protons in the p-tert-butyl benzoyl moiety of the keto-enol species were used as an internal standard with an integral of 2. Their corresponding proton signals at the obvious 8.0782 ppm and hidden 8.0553 ppm (predicted by the J value) of the diketone species were partially imposed onto the signals b at δ = 8.0530 and 8.0313 ppm (central δ = 8.0422 ppm and J = 8.68 Hz) of the aromatic protons in the p-methoxyphenyl moiety with an integral intensity of 2.10. The meta-proton signals c at δ = 7.662 ppm (J = 8.24 Hz) of the keto-enol species had an integral intensity of 2.61, indicating contamination by the corresponding proton signals of the diketo species, where their doublet signals at the hidden 7.6640 ppm and obvious 7.6411 ppm partially overlapped onto the doublet signals at δ = 7.6620 and 7.6514 ppm for the keto-enol species. The methoxy-proton signals at δ = 3.9788 and 3.9700 ppm and the tert-butyl proton signals f at δ = 1.4252 and 1.4057 ppm were separated between species. Kumari et al.93 showed the singlet signal of the geminal protons in the dibenzoyl methane at δ = 4.7 ppm for AVB in DMSO-d6. However, the signal for AVB in the D2O mixed solvent appeared to be superimposed on the signal of water (and methanol–OH) in the spectra obtained in this study.93
By building simultaneous equations from these integral relationships, we resolved the molar ratio of the keto-enol and diketo species in 70% methanol to obtain 84.7 and 15.3%, respectively. Figure S9 shows molar ratios of 90.7 and 9.3% in the isochoric acetonitrile/water solvent. Figure S10 shows that the absence or presence of β-CD did not affect the chemical shift of AVB. Figure S11 summarizes the 1H NMR titrations for the keto-enol/diketo molar ratio of AVB in the acetonitrile-d3/D2O solvent with various proportions and methanol-d4/D2O = 7:3 solvent. The diketo population in 40% acetonitrile-d3 is 1:5, whereas that in 100% acetonitrile-d3 is 1:50. The diketo form was more favored in a protic solvent (i.e., D2O) than the keto-enol form, in which hydrogen bonds formed between the keto and enol moieties. The chelated keto-enol form (intermolecular cyclic hydrogen bonding) loses its pair of hydrogen bond donors and acceptors.
The keto-enol population decreased, the diketo population increased, and their absolute gradients became gentle, depending upon the proportion of acetonitrile in the solvent mixture. Assuming that the acetonitrile molecule is a sphere arranged in a hexagonal close-packed lattice, the critical probability of forming the internal linkage of the heterogeneous sphere was estimated at 1/3 in the three-dimensional percolation model.18,24 Hence, the protic solvent (i.e., water), which occupies 33% of the volume of the acetonitrile solvent, can build a hydrogen bond network. A solute dissolved in 66% or less aqueous acetonitrile can be linked through a hydrogen bond network in such a rough approximation. Therefore, more than 66% of the acetonitrile aqueous solvent has a similar nonpolarity, presumably because the molar ratios of the keto-enol and diketo species provided a saturation of more than 66% acetonitrile-d3.
3.7. Molar Absorbance Coefficient of 270 and 360 nm Peaks
Figure S12 shows that the spectral augmentation is proportional to the amount of AVB in the methanol/H2O = 7:3 solvent. From NMR titration, the diketo and enol forms were expected to be 15.3 and 84.7%, respectively. Therefore, the abscissae of the regression lines were condensed to the species contents, leading to the molar absorption coefficients (i.e., absorbance gradients to the molarity) for the diketo and keto-enol forms being approximately 8.1 × 107 and 4.3 × 107 L/mol. Figure S13 confirms that the molar absorption coefficients for the diketo and keto-enol forms in the isochoric acetonitrile/water solvent were approximately 13.8 × 107 and 3.6 × 107 L/mol, respectively. The obtained parameters allowed the absorbance to be converted into AVB molarity.
Figure 8 shows the time course of the keto-enol and diketo molarity increase/decrease in the absence or presence of β-CD. UVA1 (365 nm) irradiation gradually degraded part of the chelated keto-enol form; therefore, this form decreased. In contrast, the diketo form increased. Adding β-CD attenuated the molarity decrease of the chelated keto-enol species of AVB depending upon the amount of β-CD, whereas adding β-CD induced a marginal change in the molarity increase of the diketo species of AVB (Figure 8a). The diketo increase to the keto-enol decrease was 19.3, 20.1, and 23.9% for none, 25 μM β-CD, and 4 mM β-CD, respectively, suggesting that β-CD included the AVB into its internal cavity and converted the chelated keto-enol form to the diketo form. In this study, because the UVA1 lamp induced photodegradation, AVB in the diketo form was not photodegraded.
Figure 8.
Time evolution of the keto-enol and diketo molarity increase/decrease in the (a) aqueous buffer, (b) 40% acetonitrile, and (c) 70% MeOH upon adding none (circles), equimolar (25 μM, triangles), or excess (4 mM, rhombuses) β-CD. UVA1 (365 nm) irradiation gradually degraded a part of the keto-enol form so that the amount of the keto-enol form (closed signs) descended. Meanwhile, the diketo form (open signs) was elevated. The protecting effect of the equimolar or excess β-CD on the photodegradation seemed equivalent to 40% acetonitrile. The 70% methanol solution experiments were limited, but we obtained similar results within the examinable range.
The protecting effect of equimolar or excess β-CD on the AVB photodegradation appeared equivalent to that of 40% acetonitrile, and less than that of 70% MeOH. The specific permittivities of water, acetonitrile, methanol, and ethanol are 80.0, 37.5, 33.3, and 25.0, respectively.94 Predictions based on the mixed solvents owing to a simple linear combination afford specific permittivity values for 40% acetonitrile, 70% methanol, and 30% ethanol of 62.8, 47.3, and 63.5, respectively. The apparent permittivity corresponding to the internal cavity of β-CD would be comparable to these values. Figure S11 suggests that the molar ratio of keto-enol and the diketo species in the presence of β-CD corresponds to that between those in 40 and 50% acetonitrile and is less than that in 70% methanol.
According to Mturi and Martincigh87 and Henson et al.,33 UV irradiation does not affect AVB in dry methanol, which is a protic solvent. This suggests that excess methanol inhibited AVB photodegradation. The solvent permittivity did not precisely evaluate the hydrophobicity of the internal cavity of β-CD. Topological arguments for the percolation theory or kissing numbers suggest that intermolecular hydrogen bonding between AVB and solvents regulates the conformational or tautomeric rearrangements and the related photoreactivity of the keto-enol moiety. The photostability study conducted by Kumari et al.93 would be apt, as their use of 30% protic ethanol and the use of 40% aprotic acetonitrile in this study stabilized AVB to the same level as the internal cavity of the inclusion host.
Conclusively, the hydrophobicity in the internal cavity of β-CD is comparable to that in the 40–50% acetonitrile aqueous solvent.
3.8. Additional Discussion
The photostability of AVB under solar light irradiation showed no significant difference compared to UVA1 irradiation, whereas OXB and CUR became destabilized. Because most UVB in solar light is absorbed by the ozone layer, only a small amount reaches the Earth’s surface. OXB is sensitive to UVB, and this small fraction of UVB may cause its degradation. According to the data in this study, the degradation of CUR under UVA1 irradiation was milder than that observed under solar light irradiation. Yan et al.95 reported that CUR undergoes degradation 1.7 times faster under 425 nm light irradiation compared to 620 nm light irradiation. On the basis of these findings, it is possible that the solar light accelerated the degradation of CUR.
The solubilization of the CD-inclusion complexes involves equilibrated solubility and dissolution kinetics.47−49,60 In this study, we obtained irregular phase-solubility diagrams, contrary to the Higuchi–Connors catalog45,46 (see Figure 2). Previously, we discussed the relationship between the solid state of the ligand and the CD inclusion complex as observed by X-ray powder diffraction patterns, time courses of dissolution in aqueous solution, and phase-solubility diagrams.49 The programmed mixing ratio can accomplish the spring (i.e., fast dissolution) and parachute (i.e., stabilization of the supersaturated drug solution) effects.21,50,60
Although the inclusion complex of AVB and β-CD in the solid state was considered to form the guest and host ratio of 1:2,69,70 its phase-solubility diagram was similar to that of the BS-type, indicating a molar ratio of 1:1 in solution at less than 5 mM β-CD. In this case, the proportionally increasing phase in a low concentration of β-CD provided the equimolar complex; the plateau phase in ≥5 mM of β-CD induced the production of the 1:2 complex. Discrepancies may exist between the dissolved and solid states (aggregation). The solubility of the equimolar complex is higher than that of the 1:2 complex. This could be the reason the highest concentration in the proportionally increasing phase was higher than the plateau level of the apparent solubility of AVB.
Although the OXB/β-CD complex with a ratio of 1:2 has not been reported in the literature, to the best of our knowledge, the phase-solubility diagram of OXB to β-CD resulted in a similar curve. Plausibly, the plateau level would correspond to the formation of the 2:2 or more equimolar complex, of which the solubility could be lower than that of the equimolar OXB/β-CD complex generated in the proportionally increasing phase. Kumari et al.96 claimed a tandem supramolecular structure with equimolar complexes of guest OXB and the cyclic host molecule C-methylresorcin[4]arene. The researchers observed an unceasing drift in the NMR spectral chemical shift when the proportion of this host was 0–8 times that of the guest. The equimolar complex was not at the saturation point. Loftsson et al.97 stated that aqueous α-CD, β-CD, and γ-CD that contain poorly soluble drugs form microparticles with diameters ranging between 1 and 50 μm at relatively high CD concentrations. Aqueous parenteral solutions of HP-β-CD at a relatively high concentration sometimes aggregate as transient clusters (or transient particulate matter) with a diameter greater than 1000 μm.96
Adding β-CD beyond the proportional increasing phase for AVB and β-CD enables the AVB/β-CD complex to form microparticle aggregations. If the β-CD concentration is more than 40–100 times that of the AVB concentration, the greater the increase in the 267 and 360 nm peaks, as shown in Figure 4a. Excess β-CD induces the corresponding diketo and chelated keto-enol species98 included in β-CD lamination. The chelated keto-enol species is the most stable isomer, which absorbs the UVA1 band owing to π–π* transitions.87 In contrast, the diketo species is a metastable isomer in which intramolecular hydrogen bonding is cleaved and the conjugation of the keto-enol bridge is lost to eliminate the protective effect of UVA1 irradiation.99,100 Instead, this species absorbs the UVC band due to the n−π* transitions to generate the triplet state, which induces an electronic transition to a singlet state; the cleavage of the α-carbon produces two radical species owing to a Norrish type I reaction.101 Therefore, a valid way to improve the photostability and efficacy of AVB would be to transfer the excitation energy of the diketo species to any other quencher.102 OCX and OCR, which absorb UVB and are poorly biodegradable, have achieved excellent results as candidate quenchers.102
In this study, the hyperchromism of the 267 nm peak was more dominant than that of the 360 nm peak. As the diketone species is relatively hydrophilic, the environment in the internal cavity of β-CD favors including the diketo species instead of the chelated keto-enol species with hydrophobicity. To identify the solvent with hydrophobicity, polarity, or permittivity equivalent to those of the internal cavity of β-CD, 70% methanol (protic) and various proportions of acetonitrile (aprotic) aqueous solutions were used. Because the UV–vis spectral pattern, peak wavelengths, and molar extinction coefficient of AVB vary depending upon the solvent components,33,87 the molar populations of the chelated keto-enol and diketo species were evaluated using the signal integral intensity in the NMR spectra.
In dry acetonitrile, the molar proportion of the relatively hydrophilic diketo species was approximately 1:50, and that of the hydrophobic chelated keto-enol species was the most significant. Increasing the water content led to an increase in the proportion of the diketo species, and the obvious UV–vis spectral pattern in the 40% acetonitrile aqueous solution became consistent with that of the AVB/β-CD complex in an aqueous solution. In the 40% acetonitrile aqueous solution, the effect of UVA1 irradiation on AVB was suppressed in the absence and presence of β-CD, as shown in Figure 8b. The 70% methanol solution experiments were limited; however, similar results were obtained within the examined range. Kumari et al.93 reported a fundamental investigation using a similar approach, which resulted in a 30–40% ethanol solution providing an equivalent atmosphere to that of the internal cavity of β-CD.
The internal cavity of β-CD retains the favorable hydrophobicity to induce hypochromicity for AVB, but excess β-CD induces hyperchromicity because of aggregation to form microparticles. When β-CD was used to improve the solubility of AVB, β-CD with an appropriate ratio of 40–100 times to AVB favors inducing photostability in AVB. An improvement in AVB solubility enables a decrease in the transdermal transfer of this hydrophobic ingredient to the blood. A local anesthetic, dibucaine, protects the analgesic ketoprofen from photodegradation,26 but β-CD accelerates its photodegradation.27 Although CD inclusion complexes generally improve aqueous solubility, chemical and metabolic stabilization, unpleasant smell, and taste, cases in which CDs enhance decomposition, defeasance, and disablement do occur. From β-CD, AVB suppresses UVA1 absorption and transforms into a diketo species that is not degradable by UVA1 irradiation. In this study, we did not verify the general disadvantage of the low photostability of AVB;33 therefore, these observations cannot contribute to commercially available sunscreen products. This study has illustrated the storage of diketo species in the internal cavity of β-CD.
4. Conclusion
According to Avdeef’s diagram, AVB, OXB, and CUR have log P values of 4.27, 2.99, and 2.65, respectively, and pKa values of 8.59, 7.97, and 8.81, respectively. The phase-solubility diagrams for AVB/β-CD and OXB/β-CD showed a proportional increase at less than 5–6 mM β-CD and plateaued at higher concentrations. Adding β-CD resulted in a hypochromic effect at less than 2 mM β-CD and a hyperchromic effect at ≥2 mM β-CD in the UV–vis spectra of AVB. However, this did not affect the OXB spectrum. At low β-CD concentrations, the hydrophobic internal cavity of β-CD reduced AVB absorbance. At ≥2 mM β-CD concentration, the AVB/β-CD complex aggregated as soluble particles, and the spectral pattern was magnified. The hypothermic effect at the 388 nm shoulder to 360 nm peak bordered on 2 mM β-CD to 50 μM AVB. This reflects the formation of intramolecular hydrogen bonds in the keto-enol moiety of the AVB tautomer in the hydrophobic cavity. These experimental observations suggest that the hydrophobic cavity of β-CD protects against AVB photodegradation by balancing keto-enol and diketo tautomerism. We artificially prepared an appropriate hydrophobic atmosphere for AVB and observed AVB photostabilization in a 40–50% acetonitrile aqueous solution. The molar ratio 1:50–100 of AVB and β-CD is optimum for photostabilization. The proposed condition to apply β-CD to the sunscreen AVB was expected to be efficient in solubilizing and avoiding harmful transdermal permeation into blood and organs.
Glossary
Abbreviations Used
- AVB
avobenzone
- BMDM
p-tert-butyl-p-methoxydibenzoylmethane
- BP
benzophenone
- CD
cyclodextrin
- CUR
curcumin
- DMSO
dimethyl sulfoxide
- U.S. FDA
United States Food and Drug Administration
- GRASE
generally recognized as safe and effective
- HMBP
2-hydroxy-4-methoxyphenylbenzophenone
- HP-β-CD
hydroxypropyl-β-CD
- JP
Japanese Pharmacopoeia
- OCR
octocrylene
- OCX
octinoxate
- OTC
over the counter
- OXB
oxybenzone
- PABA
para-aminobenzoic acid
- SBE-β-CD
sulfobutyl ether-β-CD
- UV
ultraviolet
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.langmuir.4c04108.
SVD analysis for the UV–vis spectra of AVB under UVA1 irradiation in the absence and presence of β-CD (Introduction), RP-HPLC chromatograms of AVB under UVA1 irradiation in the absence and presence of β-CD (Figure S1), RP-HPLC chromatogram of neat CUR (Figure S2), UV–vis spectra of drugs irradiated with solar light in the absence and presence β-CD (Figure S3), RP-HPLC chromatograms of CUR under UVA1 irradiation in the absence and presence of β-CD (Figure S4), UV–vis spectra of AVB under UVA1 irradiation in the absence or presence of β-CD (Figure S5), UV–vis spectra of OXB under UVA1 irradiation in the absence or presence of β-CD (Figure S6), UV–vis spectra of CUR under UVA1 irradiation in the absence or presence of β-CD (Figure S7), 1H NMR spectrum of AVB in the methanol-d4/D2O = 7:3 solvent (Figure S8), 1H NMR spectrum of AVB in the acetonitrile-d3/D2O = 1:1 solvent (Figure S9), 1H NMR spectrum of AVB/β-CD equimolar mixture in the acetonitrile-d3/D2O = 1:1 solvent (Figure S10), NMR titration for keto-enol and diketo molar ratio (Figure S11), results of UV–vis measurements of AVB in methanol/H2O = 7:3 solvent (Figure S12), UV–vis spectra of AVB in acetonitrile/H2O = 1:1 solvent (Figure S13), and three-dimensional plot of the trajectory analysis results for UV–vis spectra of AVB under UVA1 irradiation in the absence and presence of β-CD (Appendix) (PDF)
Three-dimensional plot of the trajectory analysis results for UV–vis spectra of AVB under UVA1 irradiation in the absence and presence of β-CD (Video) (MP4)
Author Contributions
† Chihiro Kuroda and Tomohiro Tsuchida contributed equally to this work as the co-first authors. Chihiro Kuroda investigated, visualized, and accomplished arrangements. Tomohiro Tsuchida reviewed and discussed the manuscript. Chihiro Tsunoda and Megumi Minamide preliminarily examined and discussed the manuscript. Ryosuke Hiroshige experimentally coordinated and instructed. Satoru Goto supervised the research and wrote the draft.
The authors declare no competing financial interest.
Supplementary Material
References
- United States Food and Drug Administration (U.S. FDA) . OTC Monographs@FDA; U.S. FDA: Silver Spring, MD, 2021; https://www.accessdata.fda.gov/drugsatfda_docs/omuf/OTCMonograph_M020-SunscreenDrugProductsforOTCHumanUse09242021.pdf (accessed Aug 9, 2024).
- United States Food and Drug Administration (U.S. FDA) . Questions and Answers: FDA Posts Deemed Final Order and Proposed Order for Over-the-Counter Sunscreen; U.S. FDA: Silver Spring, MD, 2022; https://www.fda.gov/drugs/understanding-over-counter-medicines/questions-and-answers-fda-posts-deemed-final-order-and-proposed-order-over-counter-sunscreen (accessed Aug 9, 2024).
- Dicken J. E.Over-the-counter drugs. Information of FDA’s regulation of most OTC drugs. Report to the Congressional Committees; U.S. Government Accountability Office (GAO): Washington, D.C., 2020; GAO-20-572, https://www.gao.gov/assets/gao-20-572.pdf (accessed Aug 9, 2024).
- Wang S. Q.; Lim H. W. Highlights and implications of the 2019 proposed rule on sunscreens by the US Food and Drug Administration. J. Am. Acad. Dermatol. 2019, 81, 650–651. 10.1016/j.jaad.2019.04.007. [DOI] [PubMed] [Google Scholar]
- Lim H. W.; Mohammad T. F.; Wang S. Q. Food and Drug Administration’s proposed sunscreen final administrative order: How does it affect sunscreens in the United States?. J. Am. Acad. Dermatol. 2022, 86, e83–e84. 10.1016/j.jaad.2021.09.052. [DOI] [PubMed] [Google Scholar]
- Lee S.; Ka Y.; Lee B.; Lee I.; Seo Y. E.; Shin H.; Kho Y.; Ji K. Single and mixture toxicity evaluation of avobenzone and homosalate to male zebrafish and H295R cells. Chemosphere 2023, 343, 140271 10.1016/j.chemosphere.2023.140271. [DOI] [PubMed] [Google Scholar]
- Matta M. K.; Florian J.; Zusterzeel R.; Pilli N. R.; Patel V.; Volpe D. A.; Yang Y.; Oh L.; Bashaw E.; Zineh I.; Sanabria C.; Kemp S.; Godfrey A.; Adah S.; Coelho S.; Wang J.; Furlong L.-A.; Ganley C.; Michele T.; Strauss D. G. Effect of sunscreen application on plasma concentration of sunscreen active ingredients a randomized clinical trial. JAMA 2020, 323, 256–267. 10.1001/jama.2019.20747. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ebert K. E.; Griem P.; Weiss T.; Brüning T.; Hayen H.; Koch H. M.; Bury D. Toxicokinetics of homosalate in humans after dermal application: Applicability of oral-route data for exposure assessment by human biomonitoring. Arch. Toxicol. 2024, 98, 1383–1398. 10.1007/s00204-024-03704-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nitulescu G.; Lupuliasa D.; Adam-Dima I.; Nitulescu G. M. Ultraviolet filters for cosmetic applications. Cosmetics 2023, 10, 101. 10.3390/cosmetics10040101. [DOI] [Google Scholar]
- Stiefel C.; Schwack W. Photoprotection in changing times—UV filter efficacy and safety, sensitization processes and regulatory aspects. Int. J. Cosmet. Sci. 2015, 37, 2–30. 10.1111/ics.12165. [DOI] [PubMed] [Google Scholar]
- Office of the Federal Register (OFR), National Archives and Records Administration (NARA) . Code of Federal Regulations (CFR), Title 21, Chapter I, Subchapter D, Part 352, Subpart B 352.10; OFR, NARA: Washington, D.C., 2024; https://www.ecfr.gov/current/title-21/chapter-I/subchapter-D/part-352/subpart-B/section-352.10 (accessed Aug 10, 2024).
- Health Products and Food Branch (HPFB), Health Canada . Primary Sunscreen Monograph; HPFB, Health Canada: Ottawa, Ontario, Canada, 2022; https://webprod.hc-sc.gc.ca/nhpid-bdipsn/atReq?atid=sunscreen-ecransolaire&lang=eng (accessed Aug 10, 2024).
- Therapeutic Goods Administration (TGA), Department of Health and Aged Care . Australian Regulatory Guidelines for Sunscreens (ARGS), Version 3; TGA, Department of Health and Aged Care: Woden, Australian Capital Territory, Australia, 2024; https://www.tga.gov.au/resources/resource/guidance/australian-regulatory-guidelines-sunscreens-args (accessed Aug 10, 2024).
- Suh S.; Pham C.; Smith J.; Mesinkovska N. A. The banned sunscreen ingredients and their impact on human health: A systematic review. Int. J. Dermatol. 2020, 59, 1033–1042. 10.1111/ijd.14824. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goto S.QSAR study for transdermal delivery of drugs and chemicals. Colloid and Interface Science in Pharmaceutical Research and Development; Elsevier, Inc.: Amsterdam, Netherlands, 2014; Chapter 6, pp 121–129, 10.1016/B978-0-444-62614-1.00006-5. [DOI] [Google Scholar]
- Tateuchi R.; Sagawa N.; Shimada Y.; Goto S. Enhancement of the 1-octanol/water partition coefficient of the anti-inflammatory indomethacin in the presence of lidocaine and other local anesthetics. J. Phys. Chem. B 2015, 119, 9868–9873. 10.1021/acs.jpcb.5b03984. [DOI] [PubMed] [Google Scholar]
- Shimada Y.; Tateuchi R.; Chatani H.; Goto S. Mechanisms underlying changes in indomethacin solubility with local anesthetics and related basic additives. J. Mol. Struct. 2018, 1155, 165–170. 10.1016/j.molstruc.2017.10.101. [DOI] [Google Scholar]
- Kataoka H.; Sakaki Y.; Komatsu K.; Shimada Y.; Goto S. Melting process of the peritectic mixture of lidocaine and ibuprofen interpreted by site percolation theory model. J. Pharm. Sci. 2017, 106, 3016–3021. 10.1016/j.xphs.2017.04.010. [DOI] [PubMed] [Google Scholar]
- Kasai T.; Shiono K.; Otsuka Y.; Shimada Y.; Terada H.; Komatsu K.; Goto S. Molecular recognizable ion-paired complex formation between diclofenac/indomethacin and famotidine/cimetidine regulates their aqueous solubility. Int. J. Pharm. 2020, 590, 119841 10.1016/j.ijpharm.2020.119841. [DOI] [PubMed] [Google Scholar]
- Chatani H.; Goto S.; Kataoka H.; Fujita M.; Otsuka Y.; Shimada Y.; Terada H. Effects of phosphate on drug solubility behavior of mixture ibuprofen and lidocaine. Chem. Phys. 2019, 525, 110415 10.1016/j.chemphys.2019.110415. [DOI] [Google Scholar]
- Fujita M.; Goto S.; Chatani H.; Otsuka Y.; Shimada Y.; Terada H.; Inoo K. The function of oxybuprocaine: A parachute effect that sustains the supersaturated state of anhydrous piroxicam crystals. RSC Adv. 2020, 10, 1572–1579. 10.1039/C9RA09952B. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kinoshita T.; Tsunoda C.; Goto S.; Hasegawa K.; Chatani H.; Fujita M.; Kataoka H.; Katahara Y.; Shimada Y.; Otsuka Y.; Komatsu K.; Terada H. Enthalpy–entropy compensation in the structure-dependent effect of nonsteroidal anti-inflammatory drugs on the aqueous solubility of diltiazem. Chem. Pharm. Bull. 2022, 70, 120–129. 10.1248/cpb.c21-00834. [DOI] [PubMed] [Google Scholar]
- Hasegawa K.; Goto S.; Kataoka H.; Chatani H.; Kinoshita T.; Yokoyama H.; Tsuchida T. Quantification of crystallinity during indomethacin crystalline transformation from α- to γ-polymorphic forms and of the thermodynamic contribution to dissolution in aqueous buffer and solutions of solubilizer. RSC Adv. 2024, 14, 4129. 10.1039/D3RA08481G. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hasegawa K.; Ogawa S.; Chatani H.; Kataoka H.; Tsuchida T.; Goto S. Thermodynamic and kinetic analysis of the melting process of S-ketoprofen and lidocaine mixtures. RSC Pharmaceut. 2024, 1, 536–547. 10.1039/D4PM00039K. [DOI] [Google Scholar]
- Suenaga S.; Kataoka H.; Hasegawa K.; Koga R.; Tsunoda C.; Kuwashima W.; Tsuchida T.; Goto S. How does the powder mixture of ibuprofen and caffeine attenuate the solubility of ibuprofen? Comparative study for the xanthine derivatives to recognize their intermolecular interactions using Fourier-transform infrared (FTIR) spectra, differential scanning calorimetry (DSC), and X-ray powder diffractometry (XRPD). Mol. Pharmaceutics 2024, 21 (9), 4524–4540. 10.1021/acs.molpharmaceut.4c00429. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takara A.; Kobayashi K.; Watanabe S.; Okuyama K.; Shimada Y.; Goto S. Dibucaine inhibits ketoprofen photodegradation via a mechanism different from that of antioxidants. J. Photochem. Photobiol., A 2017, 333, 208–212. 10.1016/j.jphotochem.2016.10.026. [DOI] [Google Scholar]
- Hiroshige R.; Goto S.; Tsunoda C.; Ichii R.; Shimizu S.; Otsuka Y.; Makino K.; Takahashi H.; Yokoyama H. Trajectory of the spectral/structural rearrangements for photo-oxidative reaction of neat ketoprofen and its cyclodextrin complex. J. Inclusion Phenom. Macrocyclic Chem. 2022, 102, 791–800. 10.1007/s10847-022-01160-3. [DOI] [Google Scholar]
- Tsurushima M.; Kurosawa Y.; Goto S. Propionic acid groups and multiple aromatic rings induce binding of ketoprofen and naproxen to the hydrophobic core of bovine serum albumin. Mol. Pharmaceutics 2023, 20, 3549–3558. 10.1021/acs.molpharmaceut.3c00169. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Minamide M.; Tsurushima M.; Koga R.; Hasegawa K.; Kurosawa Y.; Tsuchida T.; Goto S. Distinguishing the transitions of fluorescence spectra of tryptophan-134 and 213 in BSA induced by bindings of UV filters, oxybenzone-3, and avobenzone. Bull. Chem. Soc. Jpn. 2024, 97, uoae058 10.1093/bulcsj/uoae058. [DOI] [Google Scholar]
- Lipinski C. A. Lead- and drug-like compounds: the rule-of-five revolution. Drug Discovery Today: Technologies. 2004, 1, 337–341. 10.1016/j.ddtec.2004.11.007. [DOI] [PubMed] [Google Scholar]
- Matta M. K.; Zusterzeel R.; Pilli N. R.; Patel V.; Volpe D. A.; Florian J.; Oh L.; Bashaw E.; Zineh I.; Sanabria C.; Kemp S.; Godfrey A.; Adah S.; Coelho S.; Wang J.; Furlong L.-A.; Ganley C.; Michele T.; Strauss D. G. Effect of sunscreen application under maximal use conditions on plasma concentration of sunscreen active ingredients: A randomized clinical trial. JAMA 2019, 321, 2082–2091. 10.1001/jama.2019.5586. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bonda C. A.; Lott D.. Sunscreen Photostability. In Principles and Practice of Photoprotection; Wang S. Q., Lim W., Eds.; Springer International Publishing: Cham, Switzerland, 2016; pp 247–273, 10.1007/978-3-319-29382-0_14. [DOI] [Google Scholar]
- Hanson K. M.; Cutuli M.; Rivas T.; Antuna M.; Saoub J.; Tierce N. T.; Bardeen C. J. Effects of solvent and micellar encapsulation on the photostability of avobenzone. Photochem. Photobiol. Sci. 2020, 19, 390–398. 10.1039/c9pp00483a. [DOI] [PubMed] [Google Scholar]
- Berenbeim J. A.; Wong N. G. K.; Cockett M. C. R.; Berden G.; Oomens J.; Rijs A. M.; Dessent C. E. H. Unravelling the keto–enol tautomer dependent photochemistry and degradation pathways of the protonated UVA filter avobenzone. J. Phys. Chem. A 2020, 124, 2919–2930. 10.1021/acs.jpca.0c01295. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moi S.; Hosamani B.; Kumar K.; Gunaga S.; Raghothama S.; Gowd K. H. Photochemical studies of new synthetic derivatives of avobenzone under sunlight using UV-spectroscopy. J. Photochem. Photobiol., A 2021, 420, 113488 10.1016/j.jphotochem.2021.113488. [DOI] [Google Scholar]
- Gholap A. D.; Sayyad S. F.; Hatvate N. T.; Dhumal V. V.; Pardeshi S. R.; Chavda V. P.; Vora L. K. Drug delivery strategies for avobenzone: A case study of photostabilization. Pharmaceutics 2023, 15, 1008. 10.3390/pharmaceutics15031008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paris C.; Lhiaubet-Vallet V.; Jiménez O.; Trullas C.; Miranda M. A. A blocked diketo form of avobenzone: photostability, photosensitizing properties and triplet quenching by a triazine-derived UVB-filter. Photochem. Photobiol. 2009, 85, 178–84. 10.1111/j.1751-1097.2008.00414.x. [DOI] [PubMed] [Google Scholar]
- Afonso S.; Horita K.; Sousa e Silva J. P.; Almeida I. F.; Amaral M. H.; Lobão P. A.; Costa P. C.; Miranda M. S.; Esteves da Silva J. C. G.; Sousa Lobo J. M. Photodegradation of avobenzone: Stabilization effect of antioxidants. J. Photochem. Photobiol., B 2014, 140, 36–40. 10.1016/j.jphotobiol.2014.07.004. [DOI] [PubMed] [Google Scholar]
- Pasuch Gluzezak A. J.; Dos Santos J. L.; Maria-Engler S. S.; Gaspar L. R. Evaluation of the photoprotective and antioxidant potential of an avobenzone derivative. Front. Physiol. 2024, 15, 1347414. 10.3389/fphys.2024.1347414. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aree T.; Jongrungruangchok S. Structure-antioxidant activity relationship of β-cyclodextrin inclusion complexes with olive tyrosol, hydroxytyrosol, and oleuropein: Deep insights from X-ray analysis, DFT calculation and DPPH assay. Carbohydr. Polym. 2018, 199, 661–669. 10.1016/j.carbpol.2018.07.019. [DOI] [PubMed] [Google Scholar]
- Avdeef A.Absorption and Drug Development, Solubility, Permeability, and Charge State; John Wiley & Sons, Inc.: Hoboken, NJ, 2003. [Google Scholar]
- Hansch C.; Leo A.. Substituent Constants for Correlation Analysis in Chemistry and Biology; John Wiley & Sons, Inc.: Hoboken, NJ, 1979. [Google Scholar]
- Hansch C.; Leo A.; Hoekman D.. Exploring QSAR—Hydrophobic, Electronic, and Steric Constants; American Chemical Society (ACS), Washington, D.C., 1995. [Google Scholar]
- Sangster J.Octanol–Water Partition Coefficients: Fundamentals and Physical Chemistry; John Wiley & Sons, Inc.: Hoboken, NJ, 1997. [Google Scholar]
- Higuchi T.; Connors K. A.. Phase Solubility Techniques. Advanced Analytical Chemistry of Instrumentation; Wiley-Interscience: New York, 1965; Vol. 4, pp 117–212. [Google Scholar]
- Szejtli J. Medicinal applications of cyclodextrins. Med. Res. Rev. 1994, 14, 353–386. 10.1002/med.2610140304. [DOI] [PubMed] [Google Scholar]
- Li Y.-P.; Goto S.; Shimda Y.; Komatsu K.; Yokoyama H.; Terada H.; Makino K. Study of intermolecular interaction of hydroperoxyl-β-cyclodextrin complex through phase diagrams of the fusion entropy: contrast between nifedipine and nicardipine hydrochloride. Phys. Chem. Biophys. 2015, 5 (5), 187. 10.4172/2161-0398.1000187. [DOI] [Google Scholar]
- Li Y.-P.; Goto S.; Shimada Y.; Makino K. Phase solution and solution recrystallization equilibrium constants of hydroxypropyl-β-cyclodextrin complexes with nifedipine and nicardipine hydrochloride. J. Pharm. Sci. Technol. Jpn. 2016, 76, 267–273. 10.14843/jpstj.76.267. [DOI] [Google Scholar]
- Shimizu S.; Wada-Hirai A.; Li Y.-P.; Shimada Y.; Otsuka Y.; Goto S. Relationship between phase solubility diagrams and crystalline structures during dissolution of cimetidine/cyclodextrin complex crystals. J. Pharm. Sci. 2020, 109, 2206–2212. 10.1016/j.xphs.2020.03.029. [DOI] [PubMed] [Google Scholar]
- Wada-Hirai A.; Shimizu S.; Ichii R.; Tsunoda C.; Hiroshige R.; Fujita M.; Li Y.-P.; Shimada Y.; Otsuka Y.; Goto S. Stabilization of the metastable α-form of indomethacin-induced by the addition of 2-hydroxypropyl-β-cyclodextrin, causing supersaturation (spring) and its sustaining deployment (parachute). J. Pharm. Sci. 2021, 110, 3623–3630. 10.1016/j.xphs.2021.07.002. [DOI] [PubMed] [Google Scholar]
- Hiroshige R.; Goto S.; Ichii R.; Shimizu S.; Wada-Hirai A.; Li Y.-P.; Shimada Y.; Otsuka Y.; Makino K.; Takahashi H. Protective effects of cyclodextrins on edaravone degradation induced by atmospheric oxygen or additive oxidant. J. Inclusion Phenom. Macrocyclic Chem. 2022, 102, 327–338. 10.1007/s10847-021-01122-1. [DOI] [Google Scholar]
- Schönbeck C.; Gaardahl K.; Houston B. Drug solubilization by mixtures of cyclodextrins: Additive and synergistic effects. Mol. Pharmaceut. 2019, 16, 648–654. 10.1021/acs.molpharmaceut.8b00953. [DOI] [PubMed] [Google Scholar]
- Goto S.; Komatsu K.; Terada H. Topology of the interconversion pathway networks of cycloheptane conformations and those of related n-membered rings. Bull. Chem. Soc. Jpn. 2013, 86, 230–242. 10.1246/bcsj.20110252. [DOI] [Google Scholar]
- Shiratori T.; Goto S.; Sakaguchi T.; Kasai T.; Otsuka Y.; Higashi K.; Makino K.; Takahashi H.; Komatsu K. Singular value decomposition analysis of the secondary structure features contributing to the circular dichroism spectra of model proteins. Biochem. Biophys. Rep. 2021, 28, 101153 10.1016/j.bbrep.2021.101153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takatsuka M.; Goto S.; Kobayashi K.; Otsuka Y.; Shimada Y. Evaluation of pure antioxidative capacity of antioxidants: ESR spectroscopy of stable radicals by DPPH and ABTS assays with singular value decomposition. Food Biosci. 2022, 48, 101714 10.1016/j.fbio.2022.101714. [DOI] [Google Scholar]
- Kurosawa Y.; Otsuka Y.; Goto S. Increased selectivity of sodium deoxycholate to around tryptophan213 in bovine serum albumin upon micellization as revealed by singular value decomposition for excitation-emission matrix. Colloids Surf., B 2022, 212, 112344 10.1016/j.colsurfb.2022.112344. [DOI] [PubMed] [Google Scholar]
- Tsunoda C.; Goto S.; Hiroshige R.; Kasai T.; Okumura T.; Yokoyama H. Optimization of the stability constants of the ternary system of diclofenac/famotidine/β-cyclodextrin by nonlinear least-squares method using theoretical equations. Int. J. Pharm. 2023, 638, 122913 10.1016/j.ijpharm.2023.122913. [DOI] [PubMed] [Google Scholar]
- Horizumi Y.; Goto S.; Takatsuka M.; Yokoyama H. Effects of local anesthetics on liposomal membranes determined by their inhibitory activity of lipid peroxidation. Mol. Pharmaceutics 2023, 20, 2911–2918. 10.1021/acs.molpharmaceut.2c01053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hasegawa K.; Goto S.; Tsunoda C.; Kuroda C.; Okumura Y.; Hiroshige R.; Wada-Hirai A.; Shimizu S.; Yokoyama H.; Tsuchida T. Using singular value decomposition to analyze drug/β-cyclodextrin mixtures: insights from X-ray powder diffraction patterns. Phys. Chem. Chem. Phys. 2023, 25, 29266–29282. 10.1039/D3CP02737F. [DOI] [PubMed] [Google Scholar]
- Oshite Y.; Wada-Hirai A.; Ichii R.; Kuroda C.; Hasegawa K.; Hiroshige R.; Yokoyama H.; Tsuchida T.; Goto S. Comparative study on the effects of the inclusion complexes of non-steroidal anti-inflammatory drugs with 2-hydroxypropyl-β-cyclodextrins on dissociation rates and supersaturation. RSC Pharmaceut. 2024, 1, 80–97. 10.1039/D3PM00039G. [DOI] [Google Scholar]
- Priyadarsini K. I. The chemistry of curcumin: From extraction to therapeutic agent. Molecules 2014, 19, 20091–20112. 10.3390/molecules191220091. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Santos C. I. R.Photodegradation of commonly used ultraviolet filters—Degradation kinetics in different matrices. M.S. Thesis, Faculdade de Engenharia, Universidade do Porto, Porto, Portugal, 2018; https://repositorio-aberto.up.pt/bitstream/10216/116485/2/296334.pdf (accessed Aug 13, 2024). [Google Scholar]
- Tønnesen H. H.; Karlsen J. Studies on curcumin and curcuminoids. VI. Kinetics of curcumin degradation in aqueous solution. Z. Lebensm. Unters. Forsch. 1985, 180, 402–404. 10.1007/BF01027775. [DOI] [PubMed] [Google Scholar]
- Zhao J.; Zhao Y.; Zheng W.; Lu Y.; Feng G.; Yu S. Neuroprotective effect of curcumin on transient focal cerebral ischemia in rats. Brain Res. 2008, 1229, 224–32. 10.1016/j.brainres.2008.06.117. [DOI] [PubMed] [Google Scholar]
- Kumari A.; Raina N.; Wahi A.; Goh K. W.; Sharma P.; Nagpal R.; Jain A.; Ming L. C.; Gupta M. Wound-healing effects of curcumin and its nanoformulations: A comprehensive review. Pharmaceuts 2022, 14, 2288. 10.3390/pharmaceutics14112288. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cantor C. R.; Schimmel P. R.. Biophysical Chemistry Part 1: The Conformation of Biolosical Macromolecules; W. H. Freeman and Company: New York, 1980; Chapter 2. [Google Scholar]
- Challa R.; Ahuja A.; Ali J.; Khar R. K. Cyclodextrins in drug delivery: an updated review. AAPS PharmSciTechnol. 2005, 6, E329–E357. 10.1208/pt060243. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Loftsson T.; Masson M.; Brewster M. E. Self-association of cyclodextrins and cyclodextrin complexes. J. Pharm. Sci. 2004, 93, 1091–1099. 10.1002/jps.20047. [DOI] [PubMed] [Google Scholar]
- Vishwakarma J.; Takkella D.; Sharma S.; Gavvala K. Modulation of excimer formation and photostability of avobenzone inside the nanocavities of cyclodextrins. J. Photochem. Photobiol., A 2024, 449, 115411 10.1016/j.jphotochem.2023.115411. [DOI] [Google Scholar]
- Sbora R.; Budura E. Q.; Nitulescu G. M.; Balaci T.; Lupuleasa D. Preparation and characterization of inclusion complexes formed by avobenzone with β-cyclodextrin, hydroxypropyl-β-cyclodextrin and hydroxypropyl-α-cyclodextrin. Farmacia 2015, 63, 548–555. [Google Scholar]
- Sarveiya V.; Templeton J. F.; Benson H. A. E. Inclusion complexation of the sunscreen 2-hydroxy-4-methoxy benzophenone (oxybenzone) with hydropropyl-β-cyclodextrin: Effect on membrane diffusion. J. Inclusion Phenom. Macrocyclic Chem. 2004, 49, 275–281. 10.1007/s10847-004-6098-6. [DOI] [Google Scholar]
- Simeoni S.; Scalia S.; Benson H. A. E. Influence of cyclodextrins on in vitro human skin absorption of the sunscreen, butyl-methoxydibenzoylmethane. Int. J. Pharm. 2004, 280, 163–171. 10.1016/j.ijpharm.2004.05.021. [DOI] [PubMed] [Google Scholar]
- Yang J.; Wiley C. J.; Godwin D. A.; Felton L. A. Influence of hydroxypropyl-β-cyclodextrin on transdermal penetration and photostability of avobenzone. Eur. J. Pharmaceut. Biopharmaceut. 2008, 69, 605–612. 10.1016/j.ejpb.2007.12.015. [DOI] [PubMed] [Google Scholar]
- Yuan L.; Li S.; Huo D.; Zhou W.; Wang X.; Bai D.; Hu J. Studies on the preparation and photostability of avobenzone and (2-hydroxy)propyl-β-cyclodextrin inclusion complex. J. Photochem. Photobiol., A 2019, 369, 174–180. 10.1016/j.jphotochem.2018.09.036. [DOI] [Google Scholar]
- Mangolim C. S.; Moriwaki C.; Nogueira A. C.; Sato F.; Baesso M. L.; Neto A. M.; Matioli G. Curcumin-β-cyclodextrin inclusion complex: stability, solubility, characterization by FT-IR, FT-Raman, X-ray diffraction and photoacoustic spectroscopy, and food application. Food Chem. 2014, 153 (15), 361–370. 10.1016/j.foodchem.2013.12.067. [DOI] [PubMed] [Google Scholar]
- Ferreira L.; Mascarenhas-Melo F.; Rabaça S.; Mathur A.; Sharma A.; Giram P. S.; Pawar K. D.; Rahdar A.; Raza F.; Veiga F.; Mazzola P. G.; Paiva-Santos A. C. Cyclodextrin-based dermatological formulations: Dermopharmaceutical and cosmetic applications. Colloids Surf., B 2023, 221, 113012 10.1016/j.colsurfb.2022.113012. [DOI] [PubMed] [Google Scholar]
- Arya P.; Raghav N. In-vivo studies of curcumin-β-cyclodextrin inclusion complex as sustained release system. J. Mol. Struct. 2021, 1228, 129774 10.1016/j.molstruc.2020.129774. [DOI] [Google Scholar]
- Karpkird T.; Khunsakorn R.; Noptheeranuphap C.; Midpanon S. Inclusion complexes and photostability of UV filters and curcumin with beta-cyclodextrin polymers: effect on cross-linkers. J. Inclusion Phenom. Macrocyclic Chem. 2018, 91, 37–45. 10.1007/s10847-018-0796-y. [DOI] [Google Scholar]
- Jahed V.; Zarrabi A.; Bordbar A.-K.; Hafezi M. S. NMR (1H, ROESY) spectroscopic and molecular modelling investigations of supramolecular complex of β-cyclodextrin and curcumin. Food Chem. 2014, 165, 241–246. 10.1016/j.foodchem.2014.05.094. [DOI] [PubMed] [Google Scholar]
- Ghanghoria R.; Kesharwani P.; Agashe H. B.; Jain N. K. Transdermal delivery of cyclodextrin-solubilized curcumin. Drug Delivery Transl. Res. 2013, 3, 272–285. 10.1007/s13346-012-0114-y. [DOI] [PubMed] [Google Scholar]
- Mashaqbeh H.; Obaidat R.; Al-Shar’I N. Evaluation and characterization of curcumin-β-cyclodextrin and cyclodextrin-based nanosponge inclusion complexation. Polymers 2021, 13, 4073. 10.3390/polym13234073. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jantarat C.; Sirathanarun P.; Ratanapongsai S.; Watcharakan P.; Sunyapong S.; Wadu A. Curcumin-hydroxypropyl-β-cyclodextrin inclusion complex preparation methods: Effect of common solvent evaporation, freeze drying, and pH shift on solubility and stability of curcumin. Trop. J. Pharm. Res. 2014, 13, 1215–1223. 10.4314/tjpr.v13i8.4. [DOI] [Google Scholar]
- Chen J.; Qin X.; Zhong S.; Chen S.; Su W.; Liu Y. Characterization of Curcumin/Cyclodextrin Polymer Inclusion Complex and Investigation on Its Antioxidant and Antiproliferative Activities. Molecules 2018, 23, 1179. 10.3390/molecules23051179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Celebioglu A.; Uyar T. Fast-dissolving antioxidant curcumin/cyclodextrin inclusion complex electrospun nanofibrous webs. Food Chem. 2020, 317, 126397 10.1016/j.foodchem.2020.126397. [DOI] [PubMed] [Google Scholar]
- Li N.; Wang N.; Wu T.; Qiu C.; Wang X.; Jiang S.; Zhang Z.; Liu T.; Wei C.; Wang T. Preparation of curcumin-hydroxypropyl-β-cyclodextrin inclusion complex by cosolvency-lyophilization procedure to enhance oral bioavailability of the drug. Drug Dev. Ind. Pharm. 2018, 44 (12), 1966–1974. 10.1080/03639045.2018.1505904. [DOI] [PubMed] [Google Scholar]
- Sahoo D. K.; Mohanty P.; Biswal H. S.; Gowd K. H. Conformers influence on UV-absorbance of avobenzone. J. Photochem. Photobiol., A 2024, 453, 115671 10.1016/j.jphotochem.2024.115671. [DOI] [Google Scholar]
- Mturi G. J.; Martincigh B. S. Photostability of the sunscreening agent 4-tert-butyl-4′-methoxydibenzoylmethane (avobenzone) in solvents of different polarity and proticity. J. Photochem. Photobiol., A 2008, 200 (2–3), 410–420. 10.1016/j.jphotochem.2008.09.007. [DOI] [Google Scholar]
- Al-Rawashdeh N. A. F.; Al-Sadeh K. S.; Al-Bitar M. B. Inclusion complexes of sunscreen agents with β-cyclodextrin: Spectroscopic and molecular modeling studies. J. Spectrosc. 2013, 2013, 841409 10.1155/2013/841409. [DOI] [Google Scholar]
- Tønnesen H. H. Solubility, chemical and photochemical stability of curcumin in surfactant solutions. Studies of curcumin and curcuminoids, XXVIII. Pharmazie 2002, 57, 820–824. [PubMed] [Google Scholar]
- Bortolus P.; Monti S.. Photochemistry in Cyclodextrin Cavities. In Advances in Photochemistry; Neckers D. C., Volman D. H., Von Bünau G., Eds.; Wiley: Hoboken, NJ, 1996, Vol 21, Chapter 1, pp 1–133, 10.1002/9780470133521.ch1. [DOI] [Google Scholar]
- Rajamohan R.; Muthuraja P.; Govindasamy C.; Allur Subramanian S.; Jae Kim S.; Murali Krishnan M.; Murugavel K.; Rok Lee Y. Enhanced photostability and biocompatibility of sunscreen formulation of 2-phenylbenzimidazole-5-sulfonic acid with methyl-beta-cyclodextrin. J. Mol. Liq. 2023, 390, 123013 10.1016/j.molliq.2023.123013. [DOI] [Google Scholar]
- Wong N. G. K.; Berenbeim J. A.; Hawkridge M.; Matthews E.; Dessent C. E. H. Mapping the intrinsic absorption properties and photodegradation pathways of the protonated and deprotonated forms of the sunscreen oxybenzone. Phys. Chem. Chem. Phys. 2019, 21, 14311–14321. 10.1039/C8CP06794E. [DOI] [PubMed] [Google Scholar]
- Pajoubpong J.; Mayhan C. M.; Dar A. A.; Greenwood A. I.; Klebba K. C.; Cremer M. L.; Kumari H. Dynamic macromolecular material design: The versatility of cucurbituril over cyclodextrin in host-guest chemistry. Nanoscale Adv. 2024, 6, 4376–4384. 10.1039/D4NA00324A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mohsen-Nia M.; Amiri H.; Jazi B. Dielectric Constants of Water, Methanol, Ethanol, Butanol and Acetone: Measurement and Computational Study. J. Solution Chem. 2010, 39, 701–708. 10.1007/s10953-010-9538-5. [DOI] [Google Scholar]
- Yan S.; Liao X.; Xiao Q.; Huang Q.; Huang X. Photostabilities and anti-tumor effects of curcumin and curcumin-loaded polydopamine nanoparticles. RSC Adv. 2024, 14, 13694–13702. 10.1039/D4RA01246A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kang X.; Eisenhart A.; Dar A. A.; Mittapalli R. R.; Greenwood A.; Alsheddi L.; Beck T. L.; Li S. K.; Kumari H. A novel inclusion complex of oxybenzone with C-methylresocin[4]arene deters skin permeation. RSC Adv. 2023, 13, 25846–25852. 10.1039/D3RA01890C. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Loftsson T.; Brewster M. E. Pharmaceutical applications of cyclodextrins. 1. Drug solubilization and stabilization. J. Pharm. Sci. 1996, 85, 1017–1025. 10.1021/js950534b. [DOI] [PubMed] [Google Scholar]
- Kojić M.; Petković M.; Etinski M. A new insight into the photochemistry of avobenzone in the gas phase and acetonitrile from ab initio calculations. Phys. Chem. Chem. Phys. 2016, 18, 22168–22178. 10.1039/C6CP03533G. [DOI] [PubMed] [Google Scholar]
- Huong S. P.; Andrieu V.; Reynier J.-P.; Rocher E.; Fourneron J.-D. The photoisomerization of the sunscreen ethylhexyl p-methoxy cinnamate and its influence on the sun protection factor. J. Photochem. Photobiol., A 2007, 186, 65–70. 10.1016/j.jphotochem.2006.07.012. [DOI] [Google Scholar]
- Dubois M.; Gilard P.; Tiercet P.; Deflandre A.; Lefebvre M. A. Photoisomerization of the sunscreen filter PARSOL 1789. J. Chim. Phys. 1998, 95, 388–394. 10.1051/jcp:1998149. [DOI] [Google Scholar]
- Gonzenbach H.; Hill T. J.; Truscott T. G. The triplet energy levels of UVA and UVB sunscreens. J. Photochem. Photobiol., B 1992, 16, 377–379. 10.1016/1011-1344(92)80025-Q. [DOI] [PubMed] [Google Scholar]
- d’Agostino S.; Azzali A.; Casali L.; Taddei P.; Grepioni F. Environmentally friendly sunscreens: Mechanochemical synthesis and characterization of β-CD inclusion complexes of avobenzone and octinoxate with improved photostability. ACS Sustainable Chem. Eng. 2020, 8, 13215–13225. 10.1021/acssuschemeng.0c02735. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.