Next Article in Journal
Adapting Reverse Mentoring Strategy to SMEs: A New Pilot Model Implemented in Brazil
Next Article in Special Issue
Biogeographic Variability in Kernel Oil and Press Cake Content of Beauty Leaf Tree (Calophyllum inophyllum L.), as Determined by Chemical and Near-Infrared Spectroscopy Analysis
Previous Article in Journal
Structural Evolution and Sustainability of Agricultural Trade between China and Countries along the “Belt and Road”
Previous Article in Special Issue
Evaluation of Hesperidin on Sausages Stuffed in a New Modified Casing during Long-Term Storage—A Preliminary Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation, Optimization, and Characterization of Inclusion Complexes of Cinnamomum longepaniculatum Essential Oil in β-Cyclodextrin

1
Faculty of Agriculture, Forestry and Food Engineering, Yibin University, Yibin 644001, China
2
Sichuan Province Engineering Technology Research Center of Oil Cinnamon, Yibin 610039, China
*
Authors to whom correspondence should be addressed.
Sustainability 2022, 14(15), 9513; https://doi.org/10.3390/su14159513
Submission received: 13 June 2022 / Revised: 28 July 2022 / Accepted: 29 July 2022 / Published: 3 August 2022
(This article belongs to the Special Issue Innovative Detection Technology to Meet Food and Fuel Needs)

Abstract

:
Cinnamomum longepaniculatum essential oil (CLEO) possesses antibacterial, anti-inflammatory, and antioxidant activities. However, CLEO shows volatilization and poor solubility, which limits its application field. In this research, inclusion complexes of β-cyclodextrin (β-CD) with CLEO were produced, and its physicochemical properties were characterized. Response surface methodology was used to obtain optimum preparation conditions. A statistical model was generated to define the interactions among the selected variables. Results show that the optimal conditions were an H2O/β-CD ratio of 9.6:1 and a β-CD/CLEO ratio of 8:1, with the stirring temperature of 20 °C for the maximal encapsulation efficiency values. The physicochemical properties of CLEO/β-CD inclusion complexes (CLEO/β-CD-IC) were investigated. Fourier transform infrared spectroscopy showed that correlative characteristic bands of CLEO disappeared in the inclusion complex. X-ray diffraction presented different sharp peaks at the diffraction angle of CLEO/β-CD-IC. The thermogravimetric analysis demonstrated the thermal stability of CLEO was enhanced after encapsulation. Tiny aggregates with a smaller size of CLEO/β-CD-IC particles were observed by scanning electron microscopy. The comparison of β-CD, CLEO, and physical mixtures with CLEO/β-CD-IC confirmed the formation of inclusion complexes.

1. Introduction

The Cinnamomum genus tree Cinnamomum longepaniculatum (Gamble) N. Chao ex H. W. Li (C. longepaniculatum) is an endemic species in China [1]. Many parts of C. longepaniculatum contain volatile essential oil, which provides raw materials for the essence, food, medicine, cosmetics, and other chemical industries. The economic value of C. longepaniculatum is abundant essential oil content in the leaves and twigs, approximately 3.8 to 4.5%. The main ingredient of C. longepaniculatum essential oil (CLEO) includes 1,8-cineole, α-terpineol, terpinen-4-ol, α-pinene, α-phellandrene, γ-terpinene, and β-pinene (Figure 1) [2,3].
Various chemical constituents of C. longepaniculatum essential oil were inferred and verified its multiple biological activities by previous studies. It has an excellent antibacterial effect and anti-inflammatory actions by in vitro and in vivo experiments. Moreover, anticancer, analgesic, anthelmintic, and antioxidant effects were also confirmed [4,5]. Terpenes and sterols of CLEO as traditional pharmaceuticals or food additives were discovered in many medicinal and aromatic plants.
CLEO, as an essential oil, is sensitive and readily decomposed by heating and exposure to light and oxygen. Degradation of essential oil accompanies chemical group modifications and can cause color changes, unpleasant odor, and decreased activity [6]. Furthermore, the physical properties of essential oil, such as low water solubility and volatilization, limit its application potential in more fields [7]. In order to protect essential oil and improve physicochemical properties, it can be encapsulated into host molecules cyclodextrins (CDs).
CDs are cyclic oligosaccharides consisting of D-glucopyranose units and connected by α-1,4 glucosidic bond, issued from starch enzymatic hydrolysis. CDs have truncated cone shapes and hydrophobic cavities. Hydroxyl groups are faced outwards, while hydrogen atoms and glycosidic oxygen bridges rotate inwards to form lipophilicity [8]. Geometric hydrophobic compounds can be partly or fully enveloped in cavity interiors [9]. The amount of glucose units decides the cavity size and the ability to accommodate hydrophobic guest molecules [10]. CDs are non-poisonous agents and have been utilized as wall materials to control the core materials’ release and prolong their shelf life. Frequently-used inclusion material was β-CD, with advantages such as suitable cavity size, ease of production, low oral poisonousness, and easy-to-form inclusion complexes with organic solvents [11]. β-CD can be used to encapsulate the essential oil due to the 0.6–0.8 nm cavity size suitable for essential oil compounds (relative molecular mass from 80 to 250) [12]. When β-CD and guest compounds form inclusion complexes, physicochemical properties of the complex can be changed compared with guest compounds, including improved solubility, stability, and bioavailability; meanwhile, reduced adverse reactions. Therefore, in pharmaceutical, agriculture, food, and cosmetic industries, β-CD is utilized to manufacture inclusion complexes [13,14,15,16,17]. Several studies have demonstrated the process and performance of inclusion complexes based on β-CD and various essential oil [18,19,20].
Response surface methodology (RSM), as a multivariate statistical technique, is applied for optimizing complex processes [21,22,23,24]. Although previous studies have studied the preparation of CLEO inclusion complexes, the application of RSM to the encapsulation and physicochemical characterization of inclusion complexes has not been exploited [25]. The main purpose of this research is to produce C. longepaniculatum essential oil/β-cyclodextrin inclusion complexes (CLEO/β-CD-IC), establish an optimum process parameter condition, and estimate the physicochemical characterization of products.

2. Materials and Methods

2.1. Materials

C. longepaniculatum volatile essential oil (CLEO), extracted from C. longepaniculatum leaf by hydrodistillation method, was provided by Shiping Spice Co., Ltd. (Yibin, China), β-CD was purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Ethanol was analytical grade and obtained from KeLong Chemicals Co., Ltd. (Chengdu, China). Other reagents were analytical grade.

2.2. Preparation of Inclusion Complexes

CLEO/β-CD-IC was produced by the saturated aqueous solution method by the literature with slight modifications [26]. Appropriate quality of β-CD was deliquesced in preheated deionized water to form the β-CD saturated solution. A specified volume of CLEO ethanolic solution (1:10, v/v) was tardily introduced into the β-CD solution. The mixed solution was stirred on a thermostatic agitator at a particular temperature. After stirring for a preset time, the solution was cooled down and stored at 4 °C for 24 h. The sediments were vacuum percolated, and then powders were collected and put into the hot air circulating oven at 60 °C until dry at a constant weight.
To prepare physical mixtures, β-CD and CLEO were homogenized in the mortar and pestle at room temperature [11]. The mixtures were placed in a hermetically sealed glass for further use.

2.3. Encapsulation Efficiency (EE) and Loading Efficiency (LE)

The total essential oil in CLEO/β-CD-IC was distilled through the Clevenger-type hydrodistillation method. Obtained inclusion complexes were added into deionized water and distilled for 2 h. The volatile oil layer divided and located in the arm of the Clevenger’s units. Subsequently, the weight of the volatile oil was determined.
The EE and LE were calculated by the following equation, respectively.
EE   ( % ) = M E M I   ×   100 %
LE   ( % ) = M E M T   ×   100 %
where ME represented the mass of CLEO encapsulated, MI was the mass of CLEO added initially, and MT referred to the total mass of CLEO/β-CD-IC.

2.4. The Response Surface Methodology

2.4.1. Single-Factor Experiments

Before designing the response surface parameter, single-factor experiments were carried out to decide the appropriate range of various factors [27]. The ranges of independent variables were as follows: ratios of H2O to β-CD (5:1, 10:1, 15:1, 20:1, 25:1 mL/g), stirring time (0.5, 1, 1.5, 2, 2.5 h), ratios of β-CD to CLEO (4:1, 6:1, 8:1, 10:1, 12:1 g/mL) and stirring temperature (20, 30, 40, 50, 60 °C) [28]. The effect of each factor was determined by changing it in defined ranges while all other factors were constant, where the H2O/β-CD ratio was 5:1, stirring time was 1 h, β-CD/CLEO ratio was 10:1, and stirring temperature was 20 °C; each experiment was repeated three times.

2.4.2. Optimization of Experimental Design

RSM combined with the Box–Behnken design was carried out for optimization experiments on the encapsulation efficiency (Y) as the response variable [29]. According to data of the single-factor tests, three factors, including CLEO/β-CD ratio (A), stirring temperature (B), and stirring time (C), were selected as variables to conduct runs of RSM. Three levels of variates were confirmed on the basis of preliminary tests and coded as −1, 0, and 1 [30]. Factors codes and levels are shown in Table 1.
The experimental results were fitted to a second-order polynomial formula to survey the effects of autonomous factors and the correlation with the response variable.
Y = α0 + α1A + α2B + α3C + α11A2 + α22B2 + α33C2 + α12AB + α13AC + α23BC
where Y represented the response variable; α0 was the constant; α1–α3 were linear coefficients; α11, α22, and α33 were quadratic coefficients; and α12, α13, and α23 were interactive coefficients.

2.5. Fourier Transform Infrared (FTIR) Spectroscopy

FTIR spectra of β-CD, CLEO, physical mixtures, and CLEO/β-CD-IC were obtained on an FTIR spectrometer (Nicolet iS 5, Thermo Scientific, Waltham, MA, USA) to detect the variation in typic peaks of β-CD after the encapsulation. Spectra were performed at the region of 4000 to 400 cm−1 with 32 scans at an optical resolution of 4 cm−1.

2.6. X-ray Diffractogram (XRD) Analysis

The XRD analysis was performed by an X-ray diffractometer (Ultima IV, Rigaku, Tokyo, Japan). The operation was processed with Cu Kα radiation at 40 KV voltage and of 30 mA current. Each scan was performed in a 2θ angle range from 5° to 60° with a step size of 0.02° at the scanning speed of 5°/min.

2.7. Thermogravimetry/Derivative Thermogravimetry (TG/DTG) Analysis

The thermal stability of β-CD, CLEO, physical mixtures, and CLEO/β-CD-IC were measured on a thermogravimetric analyzer (TQ5000IR, TA Instruments, New Castle, DE, USA). Each sample was heated from 30 °C up to 300 °C, with a heating rate of 10 °C/min under a 50 mL/min flow nitrogen atmosphere. DTG results of respective targets were calculated by the first derivative of weight (%) versus time (min).

2.8. Differential Scanning Calorimetry (DSC)

A differential scanning calorimeter (DSC 214, NETZSCH, Selb, Germany) was used to conduct DSC thermograms of the samples. In this process, the temperature range was performed between 30 °C and 500 °C, and the heating rate was increased by 10 °C/min with a continuous nitrogen flow.

2.9. Morphological Examination

The morphological structure of β-CD, physical mixtures, and CLEO/β-CD-IC were evaluated using a scanning electron microscope (SEM, JSM-7200F, JEOL, Tokyo, Japan) at a voltage of 5 kV and different levels of magnification from 1000× to 5000×. Each sample was covered with thin golden layer before scanning to ensure complete conductivity.

2.10. Statistical Analysis

Results were presented as the average and standard deviation values. Statistical analysis was executed by one-way ANOVA using IBM SPSS Statistics 22 (IBM Co., New York, NY, USA), followed by the Bonferroni post hoc test for specific comparisons. p values < 0.05 were deemed to be significant difference. Design Expert 12 (Stat-Ease Inc., Minneapolis, MN, USA) was used in the response surface tests design, mathematical modeling, and optimization of the data.

3. Results and Discussion

3.1. Effects of Single Factors

Single factor results certified that the H2O/β-CD ratio, β-CD/CLEO ratio, and stirring temperature significantly affected the EE and LE. Figure 2a shows that different H2O/β-CD ratios had an effect on the EE and LE. The EE improved with the rising concentration of the β-CD solution from H2O/β-CD ratio 25:1 to 10:1 and was highest at 10:1 ratio. Then, the EE decreased at 5:1. Meanwhile, the LE expressed a similar trend and was highest at a 10:1 ratio. It demonstrated that water-occupied β-CD non-polar groups were easily replaced by CLEO molecules and caused a high embedding rate at a certain concentration of host molecules solution.
Figure 2b displays that the impact of different times on the EE and LE was not significant. After a definite stirring time, the encapsulation and disassembly rate achieved balance. With the extension of time, volatilization and degradation of essential oil may be more serious, making the EE and LE steady or even decline.
Figure 2c shows β-CD/CLEO ratio had a dramatic influence on the EE and LE. As the β-CD/CLEO ratio increased, the EE continuously improved until the 10:1 ratio. However, LE was significantly reduced when β-CD/CLEO ratio was more than 8:1. This is determined by the molecular structure of β-CD, where seven glucose residues determine the size of its hydrophobic inner cavity diameter, which can only hold a limited number of guest molecules. Thus, the EE can increase based on an adequate amount of wall materials. Nevertheless, when most cyclodextrin molecules were filled with essential oil, with the increase in cyclodextrin content, the EE could not be improved and accompanied LE visibly decreased. On the other hand, excessive CLEO could lead to an inadequate envelope, with CLEO adhering to the external β-CD and causing a decrease in the EE and LE. In other inclusion complexes studies, the concentrated orange oils/β-CD ratio of 12:88 showed a greater encapsulation capacity [31]. Retention of cinnamon essential oil attained the peak at the essential oil to the β-CD ratio of 10:90 [32].
Figure 2d reveals the EE was not continuously added with increasing stirring temperature. The higher EE did not occur at a high temperature exceeding 40 °C. The stirring temperature influenced the stability of CLEO/β-CD-IC. While the temperature increases to an extremely high degree, the disorder of molecules movement may occur, resulting in a partially disintegrate of the inclusion complexes. In addition, more volatilization of CLEO occurs at high temperatures, which reduces the EE [29].

3.2. Optimization of CLEO/β-CD-IC

Single-factor tests indicated that H2O/β-CD ratios, β-CD/CLEO ratio, and stirring temperature affected the EE. In order to optimize the three variables for preparing CLEO/β-CD-IC, RSM with a BBD design was applied based on three factors and three levels on the basis of single-factor trial results. A total of 17 measurements were performed. The EE ranged from 34.56% to 67.98%, as shown in Table 2.
A quadratic model was elected for regression analysis, and the resulting model fitted to a second-order polynomial regression formula was represented following:
Y = 53.55 − 0.63A + 14.00B − 1.70C − 2.71A2 − 1.71B2 − 1.02C2 − 0.45AB − 0.65AC − 1.75BC
As shown in Table 3, the ANOVA presents that the model was significant (p < 0.05), and simultaneously a lack of fit was non-significant (p > 0.05). This suggested that the variables in the model can explain the experimental variation of the EE. In this model, the R2 was 0.9856, and the adjusted R2 was 0.967, manifesting relevance existed between the observed and predicted values.
The most significant effect was the linear term of β-CD/CLEO ratio (B) (p < 0.01), and the linear term of stirring temperature (C) was also significant (p < 0.05). Quadratic terms of A presented a significant influence (p < 0.05); however, quadratic terms of B and C were not significant.
3D response surface plots were applied to visualize the interaction between two factors and the maximum EE on each variable. Figure 3 displays two variables kept constant and another variable altered in a specified scope. In Figure 3a, when the temperature was fastened at the center point (30 °C), the EE was increased by improving the β-CD/CLEO ratio from 4:1 to 8:1 and reached the maximum value at 8:1. Figure 3b indicates that keeping β-CD/CLEO ratio at 6:1, the maximum with the highest desirability lies towards the intermediate values of H2O/β-CD ratio. When at a higher temperature, volatilization of CLEO will lead to a decrease in EE [33]. The contour plots are elliptical, suggesting that the two factors have some interaction.
Verification tests were applied to validate the recommendatory optimized criteria from the derivation of RSM. The optimum procedure for the EE was designed at an H2O/β-CD ratio of 9.614:1, a β-CD/CLEO ratio of 8:1, and a stirring temperature of 20 °C, which predicted an EE of 68.304%. The optimal conditions were modified according to the actual situation as follows: H2O/β-CD ratio of 9.6:1, β-CD/CLEO ratio of 8:1, and stirring temperature of 20 °C. The observed results were 68.31% ± 1.13%, comparable with the predictand. This indicated a superior correlation between predictand and practical data by the regression model. Consequently, the preparation conditions obtained by RSM were verified and can be used to predict the EE effectually.

3.3. FTIR Spectroscopy

FTIR technique allows the recognition of vibration motion and rotation motion of bonds in molecules. Therefore, FTIR is widely utilized to illustrate the guest molecules and cyclodextrins during the inclusion process through characteristic bands of guest and host molecules shifting, decreasing, enlarging, or disappearing [34]. The formation process of inclusion complexes between guest and host molecules relates to chemical interaction through the hydrophobic interactions and H-bonding forces in the internal cavity of the host molecules. For this reason, it can serve to demonstrate the complex formation between CLEO and β-CD. In Figure 4, the representative bands of the CLEO, β-CD, corresponding physical mixtures, and CLEO/β-CD-IC are displayed.
The FTIR spectrum of CLEO illustrates representative bands following: stretching vibration of the methyl (CH3) groups at 2964 cm−1 and 2878 cm−1, stretching vibration of methylene (CH2) groups at 2923 cm−1 and 2853 cm−1, CH3 groups bending vibration at 1373 cm−1 [35].
The infrared spectrum of β-CD depicts the broad bands of the stretching vibration of the O-H group at 3380 cm−1. While other bands can also be related: 2925 cm−1 for CH2 groups stretching vibration, 1640 cm−1 for H-O-H bending vibration, 1156 cm−1 for C-O-C stretching vibration, 1079cm−1 for C-C stretching vibrations, and 1028 cm−1 for C-O stretching vibration. Similar results can be found reported previously [9,36].
The physical mixture spectrum presents the characteristic bands of β-CD: 3384 cm−1 (OH), 2925 cm−1 (CH2), and 1028 cm−1 (C-O). Moreover, the representative bands of CLEO, 2853 cm−1 (CH2), are also observed, suggesting CLEO was not covered up and did not interact with β-CD by being mixed physically.
The spectrum of the CLEO/β-CD-IC is practically completely derived from wide and intense bands of β-CD [37]. The representative wavenumbers of O-H stretching vibration from β-CD decreased, while the wavenumber values of C-O stretching vibrations were slightly increased after encapsulation. This can be interpreted as hydrophobic interaction between essential oil components in CLEO has been partially displaced by interactions between essential oil molecules and β-CD cavity, and hydrogen bonds were also established between hydroxyl groups of essential oil with cyclodextrins, which altered the characteristic bands after complexation. The characteristic absorption bands of CLEO at 2964 cm−1, 2878 cm−1, 2923 cm−1, and 2853 cm−1 vanished from the spectrum of the CLEO/β-CD-IC since the active CLEO was incorporated into the β-CD cavity, forming a strong physical cross-link, limiting the vibrations of aforementioned groups. Similar results for encapsulation can be found in the literature [38].

3.4. XRD Analysis

The crystalline state of the samples can be reflected by X-ray diffraction analysis. Figure 5 represents the XRD patterns of β-CD, physical mixtures, and CLEO/β-CD-IC. The β-CD crystalline profile exhibits intense and narrow peaks. The main sharp diffraction peaks were between 9° and 27°at 2θ values, and less intense peaks were located in the range of 32° to 44° [39]. No visible difference was perceived in the diffractograms of the physical mixtures compared with β-CD, demonstrating a novel phase was not established, which proved the inclusion complexes can not be acquired by a primitively physical blend of the raw materials. Conversely, in comparison to β-CD, the diffraction maximum of CLEO/β-CD-IC was obviously weakened or disappeared at 9.1°, 10.7°, 12.6°, 22.8°, and 27.1°, while peaks of 7.2°, 14.5°, 15.2°, 17.5° emerged. The diffractogram demonstrated a new crystalline structure, which may be issued from CLEO entrapped into the cavity interior of β-CD. Qualitatively, the partial diffraction peaks approximate to β-CD revealed CLEO/β-CD-IC have a higher proportion of β-CD participating in the complexation and demined the interaction. The above results supported that CLEO and β-CD formed inclusion complexes in solid-phase with unique characteristics.

3.5. TG/DTG and DSC

Thermogravimetric analysis has been employed as a proven technique for evaluating the guest molecule interacting with host materials in a kinetic control regime. The TG/DTG profiles of β-CD, CLEO, physical mixtures, and CLEO/β-CD-IC were exhibited in Figure 6a,b. The thermal decomposition of β-CD exited two degradation stages [40]. The first stage at the range of 50 to 105 °C was the evaporation of the water-binding β-CD, while the second stage was above 310 °C due to molecular main chain decomposition. The TG curves of CLEO showed an obvious mass loss from beginning to 130 °C, correlated with the volatilization of CLEO. For the physical mixtures thermograms, two weight loss steps were located at 30 to 110 °C and 320 to 360 °C. The former presented evaporation of water in β-CD and high volatilization of essential oil, while the latter presented degradation of β-CD. The TG curves of CLEO/β-CD-IC were inconsistent with those of β-CD or physical mixtures. The thermograms revealed an attenuated weight loss under 310 °C. This may be interpreted as the reduction of the proportion of water-binding β-CD, which was partially replaced by CLEO-based appropriate hydrophobic ligands. A protracted weight loss of CLEO/β-CD-IC was also related to the evaporation of endogenic CLEO, which continuously heated to 310 °C. Since the decomposition temperature of CLEO/β-CD-IC exceeds those of β-CD and physical mixtures, it demonstrated forming inclusion complexes increased the thermal stability of CLEO.
DSC was also utilized to investigate the thermal characteristics of each sample, and the results are exhibited in Figure 6c. The DSC curve of CLEO indicated a wide endothermic peak from 140.1 °C to 188.4 °C, corresponding to its decomposition process. For the thermogram of β-CD, the endothermic peak of 73.3–127.4 °C is concerned with water loss in the cavity interior of β-CD. The semblable DSC curve of physical mixtures manifested that disengaged β-CD subsist in the physical mixtures. After encapsulation, changes in the DSC curve were discovered in the thermograms of CLEO/β-CD-IC. The shift of the endothermic peak and decline of the area in the inclusion complexes represented the disorder of the water inside the β-CD cavity [41]. Strikingly, no typical endothermic peak of CLEO appeared in the CLEO/β-CD-IC curve, confirming that CLEO had been defended in the hydrophobic cavity of β-CD. Therefore, combined with the findings of FTIR and XRD, it was demonstrated the inclusion complexes formed successfully [42].

3.6. Morphological Detection

The topography of β-CD, physical mixtures, and CLEO/β-CD-IC were compared through scanning by SEM to acquire microscopic pictures (Figure 7). β-CD particles displayed amorphous, irregular crystals with different sizes and comparative dense structures [43]. The morphology of physical mixtures did not show remarkable variation compared to β-CD on shapes and sizes, certifying no interaction with each other. In contrast, the SEM images of CLEO/β-CD-IC were unlike those of β-CD and physical mixtures. CLEO/β-CD-IC particles appeared in the form of tiny aggregates with smaller sizes and altered shapes. In the inclusion complexes, typical structures of cyclodextrins disappeared.
In other works of literature, Croton rhamnifolioides essential oil [44], Lippia pedunculosa essential oil [45], and Exocarpium Citri Grandis essential oil [6] complexed with cyclodextrins, respectively, the morphology and particle sizes of inclusion complexes were transformed compared to pure β-CD. Multiple morphologies of various inclusion complexes can be observed, which may be based on the admitting ability of the β-CD cavity on various guest molecules to encapsulation. This behavior revealed the interaction between essential oil and β-CD and the accomplishment of encapsulation.

4. Conclusions

The inclusion complexes of CLEO with β-CD were produced using the saturated solution method. The quadratic model was established for the EE, and the following optimal procedure parameters were acquired by response surface methodology: 9.6:1 of H2O/β-CD ratio, 8:1 of β-CD/CLEO ratio, 20 °C of stirring temperature. The corresponding EE was 68.31%, comparable with the predictive value of 68.30%. Physicochemical properties suggested the successful formation of inclusion complexes by CLEO displacing the internal water molecules. FTIR revealed changes in wavenumbers and peak intensity of inclusion complexes compared with CLEO, supporting CLEO entered in the cavity of β-CD. XRD spectroscopy reflected the transformation of crystal characteristics of inclusion complexes after complexation. The morphological characterizations observed by SEM exhibited that CLEO/β-CD-IC particles were smaller and formless agglomerates compared with physical mixtures. For TG analysis, disparate mass loss curves proved that the inclusion complexes had better thermal stability than that of free CLEO. CLEO/β-CD-IC supplied thermostability to CLEO compounds, and it was beneficial to expand the application of CLEO in the food, pharmaceutical, and daily chemical industries. CLEO/β-CD-IC can be a prospective novel material as a food preservative and pharmaceutical form based on the biological activities of CLEO. Consequently, the antimicrobial properties of CLEO/β-CD-IC should be investigated in further research.

Author Contributions

Conceptualization, Q.W.; methodology, Y.Y., C.W. and L.Z.; investigation, Y.Y. and C.W.; resources, X.Z.; writing—original draft preparation, Y.Y.; writing—review and editing, Q.W. and X.Z.; visualization, Y.Y.; supervision, Q.F.; project administration, Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Doctor Launch Project of Yibin University (2020QH05) and Yibin High-end Innovation and Entrepreneurship Project (2021YGC03).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhao, X.; Yan, Y.; Zhou, W.H.; Feng, R.Z.; Shuai, Y.K.; Yang, L.; Liu, M.J.; He, X.Y.; Wei, Q. Transcriptome and metabolome reveal the accumulation of secondary metabolites in different varieties of Cinnamomum longepaniculatum. BMC Plant Biol. 2022, 22, 243. [Google Scholar] [CrossRef] [PubMed]
  2. Huang, J.; Yang, L.; Zou, Y.; Luo, S.; Wang, X.; Liang, Y.; Du, Y.; Feng, R.; Wei, Q. Antibacterial activity and mechanism of three isomeric terpineols of Cinnamomum longepaniculatum leaf oil. Folia Microbiol. 2021, 66, 59–67. [Google Scholar] [CrossRef]
  3. Yin, H.; Wang, C.; Yue, J.; Jiao, S.S.; Wang, X.H.; Jiang, L.L. Extraction of Cinnamomum longipaniculatum Essential Oil and Its Chemical Composition, Antioxidation and Antibacterial Properties. Stored Proced. 2020, 20, 183–191. [Google Scholar]
  4. Li, L.; Li, Z.W.; Yin, Z.Q.; Wei, Q.; Jia, R.Y.; Zhou, L.J.; Xu, J.; Song, X.; Zhou, Y.; Du, Y.H.; et al. Antibacterial activity of leaf essential oil and its constituents from Cinnamomum longepaniculatum. Int. J. Clin. Exp. Med. 2014, 7, 1721–1727. [Google Scholar]
  5. Du, Y.H.; Feng, R.Z.; Li, Q.; Wei, Q.; Yin, Z.Q.; Zhou, L.J.; Tao, C.; Jia, R.Y. Anti-inflammatory activity of leaf essential oil from Cinnamomum longepaniculatum (Gamble) N. Chao. Int. J. Clin. Exp. Med. 2014, 7, 5612–5620. [Google Scholar]
  6. Su, Z.; Qin, Y.; Zhang, K.; Bi, Y.; Kong, F. Inclusion Complex of Exocarpium Citri Grandis Essential Oil with β-Cyclodextrin: Characterization, Stability, and Antioxidant Activity. J. Food Sci. 2019, 84, 1592–1599. [Google Scholar] [CrossRef]
  7. Cao, C.; Wei, D.; Xu, L.; Hu, J.; Qi, J.; Zhou, Y. Characterization of tea tree essential oil and large-ring cyclodextrins (CD(9)-CD(22)) inclusion complex and evaluation of its thermal stability and volatility. J. Sci. Food Agric. 2021, 101, 2877–2883. [Google Scholar] [CrossRef]
  8. Mura, P. Advantages of the combined use of cyclodextrins and nanocarriers in drug delivery: A review. Int. J. Pharm. 2020, 579, 119181. [Google Scholar] [CrossRef] [PubMed]
  9. Hădărugă, N.G.; Szakal, R.N.; Chirilă, C.A.; Lukinich-Gruia, A.T.; Păunescu, V.; Muntean, C.; Rusu, G.; Bujancă, G.; Hădărugă, D.I. Complexation of Danube common nase (Chondrostoma nasus L.) oil by β-cyclodextrin and 2-hydroxypropyl-β-cyclodextrin. Food Chem. 2020, 303, 125419. [Google Scholar] [CrossRef]
  10. Marques, C.S.F.; Barreto, N.S.; Oliveira, S.S.C.; Santos, A.L.S.; Branquinha, M.H.; Sousa, D.P.; Castro, M.; Andrade, L.N.; Pereira, M.M.; Silva, C.F.D.; et al. β-Cyclodextrin/Isopentyl Caffeate Inclusion Complex: Synthesis, Characterization and Antileishmanial Activity. Molecules 2020, 25, 4181. [Google Scholar] [CrossRef]
  11. Kringel, D.H.; Antunes, M.D.; Klein, B.; Crizel, R.L.; Wagner, R.; de Oliveira, R.P.; Dias, A.R.G.; Zavareze, E.D.R. Production, Characterization, and Stability of Orange or Eucalyptus Essential Oil/β-Cyclodextrin Inclusion Complex. J. Food Sci. 2017, 82, 2598–2605. [Google Scholar] [CrossRef]
  12. Shrestha, M.; Ho, T.M.; Bhandari, B.R. Encapsulation of tea tree oil by amorphous beta-cyclodextrin powder. Food Chem. 2017, 221, 1474–1483. [Google Scholar] [CrossRef] [PubMed]
  13. Cunha, F.V.M.; do Nascimento Caldas Trindade, G.; da Silva Azevedo, P.S.; Coêlho, A.G.; Braz, E.M.; Pereira de Sousa Neto, B.; de Rezende, D.C.; de Sousa, D.P.; de Assis Oliveira, F.; Nunes, L.C.C. Ethyl ferulate/β-cyclodextrin inclusion complex inhibits edema formation. Mater. Sci. Eng. C Mater. Biol. Appl. 2020, 115, 111057. [Google Scholar] [CrossRef] [PubMed]
  14. Mu, K.; Jiang, K.; Wang, Y.; Zhao, Z.; Cang, S.; Bi, K.; Li, Q.; Liu, R. The Biological Fate of Pharmaceutical Excipient β-Cyclodextrin: Pharmacokinetics, Tissue Distribution, Excretion, and Metabolism of β-Cyclodextrin in Rats. Molecules 2022, 27, 1138. [Google Scholar] [CrossRef]
  15. Lin, L.; Liao, X.; Li, C.; Abdel-Samie, M.A.; Siva, S.; Cui, H. Cold nitrogen plasma modified cuminaldehyde/β-cyclodextrin inclusion complex and its application in vegetable juices preservation. Food Res. Int. 2021, 141, 110132. [Google Scholar] [CrossRef]
  16. Fuentes Campo, A.; Sancho, M.I.; Melo, G.; Dávila, Y.A.; Gasull, E. In vitro and in vivo inhibition of Hass avocado polyphenol oxidase enzymatic browning by paeonol, β-cyclodextrin, and paeonol:β-cyclodextrin inclusion complex. J. Biosci. Bioeng. 2019, 127, 703–709. [Google Scholar] [CrossRef]
  17. Kringel, D.H.; Lang, G.H.; Dias, Á.R.G.; Gandra, E.A.; Valente Gandra, T.K.; da Rosa Zavareze, E. Impact of encapsulated orange essential oil with β-cyclodextrin on technological, digestibility, sensory properties of wheat cakes as well as Aspergillus flavus spoilage. J. Sci. Food Agric. 2021, 101, 5599–5607. [Google Scholar] [CrossRef]
  18. Hogenbom, J.; Istanbouli, M.; Faraone, N. Novel β-Cyclodextrin and Catnip Essential Oil Inclusion Complex and Its Tick Repellent Properties. Molecules 2021, 26, 7391. [Google Scholar] [CrossRef]
  19. Hogenbom, J.; Jones, A.; Wang, H.V.; Pickett, L.J.; Faraone, N. Synthesis and Characterization of β-Cyclodextrin-Essential Oil Inclusion Complexes for Tick Repellent Development. Polymers 2021, 13, 1892. [Google Scholar] [CrossRef] [PubMed]
  20. Marques, C.S.; Carvalho, S.G.; Bertoli, L.D.; Villanova, J.C.O.; Pinheiro, P.F.; Dos Santos, D.C.M.; Yoshida, M.I.; de Freitas, J.C.C.; Cipriano, D.F.; Bernardes, P.C. β-Cyclodextrin inclusion complexes with essential oils: Obtention, characterization, antimicrobial activity and potential application for food preservative sachets. Food Res. Int. 2019, 119, 499–509. [Google Scholar] [CrossRef] [PubMed]
  21. Zhu, Y.; Yu, J.; Jiao, C.; Tong, J.; Zhang, L.; Chang, Y.; Sun, W.; Jin, Q.; Cai, Y. Optimization of quercetin extraction method in Dendrobium officinale by response surface methodology. Heliyon 2019, 5, e02374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Feng, C.H.; Drummond, L.; Sun, D.W.; Zhang, Z.-H. Evaluation of natural hog casings modified by surfactant solutions combined with lactic acid by response surface methodology. LWT-Food Sci. Technol. 2014, 58, 427–438. [Google Scholar] [CrossRef]
  23. Feng, C.H. Optimizing Procedures of Ultrasound-Assisted Extraction of Waste Orange Peels by Response Surface Methodology. Molecules 2022, 27, 2268. [Google Scholar] [CrossRef] [PubMed]
  24. Feng, C.H.; Sun, D.W. Optimisation of immersion vacuum cooling operation and quality of Irish cooked sausages by using response surface methodology. Int. J. Food Sci. Technol. 2014, 49, 1850–1858. [Google Scholar] [CrossRef]
  25. Liu, Y.; Yin, Z.Q.; Wei, Q.; Jia, R.Y.; Fan, J.; Zhou, L.J.; Du, Y.H. Study on preparation process and stability of beta-cyclodextrin inclusion compound in volatile oil of Cinnamomum longepaniculatum leaves. Zhongguo Zhong Yao Za Zhi 2013, 38, 2105–2108. [Google Scholar]
  26. Yang, Y.H.; Li, X.Z.; Zhang, S. Preparation methods and release kinetics of Litsea cubeba essential oil microcapsules. RSC Adv. 2018, 8, 29980–29987. [Google Scholar] [CrossRef] [Green Version]
  27. Wu, J.; Zhang, J.; Yu, X.; Shu, Y.; Zhang, S.; Zhang, Y. Extraction optimization by using response surface methodology and purification of yellow pigment from Gardenia jasminoides var. radicans Makikno. Food Sci. Nutr. 2021, 9, 822–832. [Google Scholar] [CrossRef]
  28. Wang, Y.; Yin, C.; Cheng, X.; Li, G.; Zhu, X. β-Cyclodextrin Inclusion Complex Containing Litsea cubeba Essential Oil: Preparation, Optimization, Physicochemical, and Antifungal Characterization. Coatings 2020, 10, 850. [Google Scholar] [CrossRef]
  29. Bai, C.; Xu, P.; Qi, W.; Kong, Q.; Xiong, G. Preparation and characterization of a sustained-release bio-preservative based on β-cyclodextrin encapsulated eugenol. Food Biosci. 2021, 42, 101192. [Google Scholar] [CrossRef]
  30. He, D.; Yan, L.; Hu, Y.; Wu, Q.; Wu, M.; Choi, J.I.; Tong, H. Optimization of Porphyran Extraction from Pyropia yezoensis by Response Surface Methodology and Its Lipid-Lowering Effects. Mar. Drugs 2021, 19, 53. [Google Scholar] [CrossRef] [PubMed]
  31. Torres-Alvarez, C.; Castillo, S.; Sánchez-García, E.; Aguilera González, C.; Galindo-Rodríguez, S.A.; Gabaldón-Hernández, J.A.; Báez-González, J.G. Inclusion Complexes of Concentrated Orange Oils and β-Cyclodextrin: Physicochemical and Biological Characterizations. Molecules 2020, 25, 5109. [Google Scholar] [CrossRef] [PubMed]
  32. Petrović, G.M.; Stojanović, G.S.; Radulović, N.S. Encapsulation of cinnamon oil in β-cyclodextrin. J. Med. Plants Res. 2010, 4, 1382–1390. [Google Scholar]
  33. Ma, S.; Zhao, Z.; Liu, P. Optimization of preparation process of β-cyclodextrin inclusion compound of clove essential oil and evaluation of heat stability and antioxidant activities in vitro. J. Food Meas. Charact. 2018, 12, 2057–2067. [Google Scholar] [CrossRef]
  34. Li, H.; Chang, S.L.; Chang, T.R.; You, Y.; Zhao, B. Inclusion complexes of cannabidiol with β-cyclodextrin and its derivative: Physicochemical properties, water solubility, and antioxidant activity. J. Mol. Liq. 2021, 334, 116070. [Google Scholar] [CrossRef]
  35. Sampath, S.; Subramani, S.; Janardhanam, S.; Subramani, P.; Yuvaraj, A.; Chellan, R. Bioactive compound 1,8-Cineole selectively induces G2/M arrest in A431 cells through the upregulation of the p53 signaling pathway and molecular docking studies. Phytomedicine 2018, 46, 57–68. [Google Scholar] [CrossRef]
  36. Biernacka, M.; Ilyich, T.; Zavodnik, I.; Pałecz, B.; Stepniak, A. Studies of the Formation and Stability of Ezetimibe-Cyclodextrin Inclusion Complexes. Int. J. Mol. Sci. 2021, 23, 455. [Google Scholar] [CrossRef]
  37. 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. [Google Scholar] [CrossRef]
  38. Franco, P.; De Marco, I. Formation of Rutin-β-Cyclodextrin Inclusion Complexes by Supercritical Antisolvent Precipitation. Polymers 2021, 13, 246. [Google Scholar] [CrossRef]
  39. Gao, S.; Liu, Y.; Jiang, J.; Ji, Q.; Ye, F. Physicochemical properties and fungicidal activity of inclusion complexes of fungicide chlorothalonil with β-cyclodextrin and hydroxypropyl-β-cyclodextrin. J. Mol. Liq. 2019, 293, 111513. [Google Scholar] [CrossRef]
  40. Sbârcea, L.; Udrescu, L.; Ledeţi, I.; Szabadai, Z.; Fuliaş, A.; Sbârcea, C. β-Cyclodextrin inclusion complexes of lisinopril and zofenopril. J. Therm. Anal. Calorim. 2016, 123, 2377–2390. [Google Scholar] [CrossRef]
  41. Zhang, W.; Li, X.; Yu, T.; Yuan, L.; Rao, G.; Li, D.; Mu, C. Preparation, physicochemical characterization and release behavior of the inclusion complex of trans-anethole and β-cyclodextrin. Food Res. Int. 2015, 74, 55–62. [Google Scholar] [CrossRef]
  42. Tao, F.; Hill, L.E.; Peng, Y.; Gomes, C.L. Synthesis and characterization of β-cyclodextrin inclusion complexes of thymol and thyme oil for antimicrobial delivery applications. LWT-Food Sci. Technol. 2014, 59, 247–255. [Google Scholar] [CrossRef]
  43. Fauziah, C.I.; Zaibunnisa, A.H.; Osman, H.; Aida, W. Thermal Analysis and Surface Morphology Study of Cholesterol: Β-Cyclodextrin Inclusion Complex. Adv. Mater. Res. 2013, 812, 221–225. [Google Scholar] [CrossRef]
  44. Oliveira Brito Pereira Bezerra Martins, A.; Wanderley, A.G.; Alcântara, I.S.; Rodrigues, L.B.; Cesário, F.; Correia de Oliveira, M.R.; Castro, F.F.E.; Albuquerque, T.R.; da Silva, M.S.A.; Ribeiro-Filho, J.; et al. Anti-Inflammatory and Physicochemical Characterization of the Croton Rhamnifolioides Essential Oil Inclusion Complex in β-Cyclodextrin. Biology 2020, 9, 114. [Google Scholar] [CrossRef] [PubMed]
  45. Dos Passos Menezes, P.; de Oliveira Araujo, F.; Andrade, T.A.; Trindade, I.A.S.; de Araujo-Filho, H.G.; de Souza Siqueira Quintans, J.; Quintans-Junior, L.J.; Menezes, L.R.A.; de Almeida, R.N.; Braga, R.M.; et al. Physicochemical Characterization and Antinociceptive Effect of β-cyclodextrin/Lippia pedunculosa Essential Oil in Mice. Curr. Top. Med. Chem. 2018, 18, 797–807. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of representative major compounds for C. longepaniculatum essential oil.
Figure 1. Chemical structures of representative major compounds for C. longepaniculatum essential oil.
Sustainability 14 09513 g001
Figure 2. Effect of independent variable on the EE and LE of CLEO/β-CD-IC. H2O/β-CD ratios (a); stirring time (b); β-CD/CLEO ratios (c); stirring temperature (d). Note: Values labelled by diverse letters (a–c) means significant difference (p < 0.05); ns indicates not significant.
Figure 2. Effect of independent variable on the EE and LE of CLEO/β-CD-IC. H2O/β-CD ratios (a); stirring time (b); β-CD/CLEO ratios (c); stirring temperature (d). Note: Values labelled by diverse letters (a–c) means significant difference (p < 0.05); ns indicates not significant.
Sustainability 14 09513 g002
Figure 3. Response surface plots revealing effects of H2O/β-CD ratio (a), β-CD/CLEO ratio (b), and stirring temperature (c) on the EE.
Figure 3. Response surface plots revealing effects of H2O/β-CD ratio (a), β-CD/CLEO ratio (b), and stirring temperature (c) on the EE.
Sustainability 14 09513 g003
Figure 4. FTIR spectra of CLEO, β-CD, physical mixtures and CLEO/β-CD-IC.
Figure 4. FTIR spectra of CLEO, β-CD, physical mixtures and CLEO/β-CD-IC.
Sustainability 14 09513 g004
Figure 5. XRD patterns of β-CD, physical mixtures, and CLEO/β-CD-IC.
Figure 5. XRD patterns of β-CD, physical mixtures, and CLEO/β-CD-IC.
Sustainability 14 09513 g005
Figure 6. Thermogravimetry curves (a), DTG curves (b) and DSC thermogram (c) of CLEO, β-CD, physical mixtures and CLEO/β-CD-IC.
Figure 6. Thermogravimetry curves (a), DTG curves (b) and DSC thermogram (c) of CLEO, β-CD, physical mixtures and CLEO/β-CD-IC.
Sustainability 14 09513 g006
Figure 7. SEM micrographs of β-CD × 1.0 k (a), physical mixtures × 1.0 k (b), CLEO/β-CD-IC × 2.0 k (c).
Figure 7. SEM micrographs of β-CD × 1.0 k (a), physical mixtures × 1.0 k (b), CLEO/β-CD-IC × 2.0 k (c).
Sustainability 14 09513 g007
Table 1. Range of independent variables in RSM.
Table 1. Range of independent variables in RSM.
VariableSymbolLevel of Variation
−101
H2O/β-CD ratio (g/mL)A5:110:115:1
β-CD/CLEO ratio (g/mL)B6:18:110:1
Stirring temperature (°C)C203040
Table 2. Box–Behnken design scheme and the response.
Table 2. Box–Behnken design scheme and the response.
RunsABCY
H2O/β-CD Ratio (mL/g)β-CD/CLEO Ratio (g/mL)Stirring Temperature (°C)Encapsulation Efficiency (%)
1543036.10
21562052.65
31583061.27
41082067.98
51084063.59
61063054.02
71063052.84
8583063.09
91543036.07
101063054.58
111063054.17
12564048.30
131042034.56
141044037.19
15562052.92
161564045.42
171063052.16
Table 3. ANOVA for the EE response of preparation of CLEO/β-CD-IC.
Table 3. ANOVA for the EE response of preparation of CLEO/β-CD-IC.
Variance OriginSum of SquaresdfMean SquareF-Valuep-ValueSignificance
Model1661.619184.6253.15<0.0001significant
A3.1313.130.89960.3745-
B1568.2811568.28451.46<0.0001significant
C23.15123.156.670.0364significant
A231.03131.038.930.0203significant
B212.27112.273.530.1023-
C24.3514.351.250.2998-
AB0.801010.80100.23060.6457-
AC1.7011.700.49020.5064-
BC12.32112.323.550.1017-
Residual24.3273.47---
Lack of Fit20.2136.746.570.0503not significant
Pure Error4.1041.03---
Cor Total1685.9316----
R20.9856-----
Adjusted R20.9670----significant
Model1661.619184.6253.15<0.0001significant
Note: A: H2O/β-CD ratio, B: β-CD/CLEO ratio, C: stirring temperature. A2, B2, and C2 mean square of A, B, and C.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yan, Y.; Zhao, X.; Wang, C.; Fang, Q.; Zhong, L.; Wei, Q. Preparation, Optimization, and Characterization of Inclusion Complexes of Cinnamomum longepaniculatum Essential Oil in β-Cyclodextrin. Sustainability 2022, 14, 9513. https://doi.org/10.3390/su14159513

AMA Style

Yan Y, Zhao X, Wang C, Fang Q, Zhong L, Wei Q. Preparation, Optimization, and Characterization of Inclusion Complexes of Cinnamomum longepaniculatum Essential Oil in β-Cyclodextrin. Sustainability. 2022; 14(15):9513. https://doi.org/10.3390/su14159513

Chicago/Turabian Style

Yan, Yue, Xin Zhao, Chao Wang, Qiong Fang, Lu Zhong, and Qin Wei. 2022. "Preparation, Optimization, and Characterization of Inclusion Complexes of Cinnamomum longepaniculatum Essential Oil in β-Cyclodextrin" Sustainability 14, no. 15: 9513. https://doi.org/10.3390/su14159513

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop