GC-MS analysis of volatiles in cinnamon essential oil extracted by different methods

T. Yu, H. Yao, S. Qi and J. Wang*

South China University of Technology, School of Food Science and Engineering, Wushan Road 381, Guangzhou, 510641, China

*Corresponding author: wangjuan@scut.edu.cn

 

SUMMARY

Cinnamon essential oil (CEO) was extracted by three different methods: steam distillation (SD), ultrasound-assisted steam distillation (UASD) and microwave-assisted steam distillation (MASD). The volatiles in CEO were separated and identified by gas chromatography–mass spectrometry (GC-MS), and the differences in volatiles among the three different methods were further analyzed through principal component analysis. The results showed that 36 individual volatile components were present in the CEO from the three different methods. In general, the numbers of aldehydes, esters, alcohols, terpenes, aromatics and ketones were 6, 3, 7, 17, 2, and 1, respectively. The most abundant volatile component was determined to be cinnamic aldehyde. The content of total cinnamic aldehydes, which determines the price of CEO, was the highest among the three methods in the UASD sample (85.633%). Moreover, the highest yield (8.33‰) of essential oil was extracted by the UASD method. Therefore, UASD was the best way for CEO extraction in this research and was recommended for future industrial applications.

 

RESUMEN

Análisis de volátiles mediante GC-MS de aceites esenciales de canela extraídos por diferentes métodos. El aceite esencial de canela (AEC) se extrajo mediante tres métodos diferentes: destilación al vapor (DV), destilación al vapor asistida por ultrasonido (DVAU) y destilación al vapor asistida por microondas (DVAM). Los volátiles del AEC se separaron e identificaron mediante cromatografía de gases-espectrometría de masas (GC-MS), las diferencias de los volátiles entre los tres métodos se analizaron adicionalmente a través del análisis de componentes principales. Los resultados mostraron la presencia de 36 componentes volátiles en el AEC mediante los tres métodos diferentes. En general, el número de aldehídos, ésteres, alcoholes, terpenos, aromáticos y cetonas presentes fue de 6, 3, 7, 17, 2 y 1, respectivamente. Se determinó que el componente volátil más abundante era el aldehído cinámico. El contenido de aldehído cinámico total, el cual decide el precio del AEC, en la muestra de DVAU (85,633%), fue el más alto entre tres métodos. Además, el mayor rendimiento (8,33‰) de aceite esencial se encontró mediante el método DVAU. Por lo tanto, DVAU fue la mejor forma de extracción de AEC en esta investigación y se recomienda en futuras aplicaciones industriales.

 

Submitted: 30 April 2019; Accepted: 19 September 2019; Published online: 26 August 2020

ORCID ID: Yu T https://orcid.org/0000-0003-3700-8199, Yao H https://orcid.org/0000-0002-5459-2289, Qi S https://orcid.org/0000-0002-6578-1464, Wang J https://orcid.org/0000-0002-5015-0176

KEYWORDS: Cinnamic aldehyde; Cinnamon essential oil; Extraction method; GC-MS; Volatiles

PALABRAS CLAVE: Aceite esencial de canela; Aldehído cinámico; GC-MS; Método de extracción; Volátiles.

Citation/Cómo citar este artículo: Yu T, Yao H, Qi S, Wang J. 2020. GC-MS analysis of volatiles in cinnamon essential oil extracted by different methods. Grasas Aceites 71 (3), e372. https://doi.org/10.3989/gya.0462191

Copyright: ©2020 CSIC. This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International (CC BY 4.0) License.


 

CONTENT

1. INTRODUCTIONTOP

Cinnamon (Cinnamomum cassia Presl), a tropical evergreen tropical tree from the Lauraceae family, is widely distributed in Southeast Asia (Chang, Chen, Chang, 2010). As the biggest producer of cinnamon in the world, China has produced more than 80% of the cinnamon in the world, especially in the Guangdong and Guangxi provinces, which accounted for 95% of the total production in China (Li et al., 2013). Cinnamon is often used as a traditional medicine in India and China due to its unique medicinal and aromatic values, and it is mainly used for the treatment of anorexia, heart disease, intestinal disease and helminthic infections. Cinnamon bark and cinnamon essential oil (CEO) are included in pharmacopoeias in many countries (India, Britain, China, Australia, Belgium, France, Germany, Hungary, Japan, Portugal and Switzerland), and they have also been used as food additives, condiments and flavoring agents due to their carminative, antioxidant and preservative properties (Nabavi et al., 2015). Rui et al., (2009) found that the number of volatiles identified in Cinnamomum cassia, Cinnamomum zeylanicum, Cinnamomum tamala, Cinnamomum burmannii, Cinnamomum pauciflorum were 22, 22, 13, 6 and 21, respectively. Cinnamaldehyde, with good inhibitory effects on many food spoilage microorganisms, is the characteristic volatile component of CEO, which exists in all kinds of cinnamon species (Matan et al., 2006). In addition, in international trade, the higher the content of total cinnamic aldehydes in CEO, the higher price of CEO is. Therefore, as the main volatile, the cinnamic aldehyde level should be obtained to the greatest extent during the extracting procedure of CEO.

Until now, the hydrodistillation, steam distillation (SD), steam and water distillation, maceration, hollow distillation, and expression methods have been widely used to obtain essential oils or extracts from plant materials (Jeyaratnam et al., 2016). SD is a promising extraction method as its outstanding advantages such as solvent-free, easy to operate, and safe. In addition, SD can prevent volatile oils from decomposition because steam is able to reduce the boiling point of the oils (Wong et al., 2014).

Furthermore, the ultrasound-assisted and microwave-assisted extraction techniques are also recognized as efficient methods with a short extraction time, increased yield and good quality (Cravotto et al., 2008). Azlina et al., (2013) extracted gaharu essential oil via ultrasonic assisted steam distillation (UASD), which increased extraction efficiency and reduced production costs. Microwave-assisted steam distillation (MASD) combines the advantages of both conventional and modern technologies, and is probably the leading technology in the essential oil production industry. Golmakani and Rezaei (2010) increased extraction efficiency by 4 times using MASD compared to the traditional SD method during the extraction of Zataria multiflora essential oil.

At present, the SD method is used to produce CEO in factories. Since its disadvantages are low yield of essential oil, low total cinnamic aldehyde content in the products and high energy consumption, it is necessary to find a new extraction method with higher yield and cinnamic aldehyde content. In the present research, The UASD and MASD methods were used for comparison with the traditional SD method in the yield of essential oil and cinnamic aldehyde content as determined by Gas chromatography– mass spectrometry (GC–MS).

2. MATERIALS AND METHODSTOP

2.1. Materials TOP

Cinnamon (Cinnamomum cassia Presl) was collected from Yunfu City, (Guangdong Province, China) and naturally air-dried, then ground and screened into powder using a shaker with a 40-mesh sieve, and placed in a desiccator prior to use. Three samples were prepared for each treatment, for a total of 9 CEO samples.

2.2. Preparation of cinnamon essential oil by SDTOP

Each cinnamon sample (150 g) was placed in a glass distillation flask (3.6 L). According to our former optimized process, where the power of a steam generator (self-made) was set at 600 W, and the vapor generated in the steam generator passed through the material for 2 hours, the CEO was separated from the mixture of water-oil by static stratification.

2.3. Preparation of cinnamon essential oil by UASDTOP

The ultrasonic pre-treatment process was carried out in an ultrasonic cell crusher (JY92-IIN model, SCIENTZ, China). Each cinnamon sample (150 g) was put in a stoppered flask, and subjected to ultrasonic processing under settled conditions. Based on our former optimized process, ultrasonic power of 250 W and distillation temperature of 40 °C with a water to raw material ratio (w/w) of (16:1) for 25 min were applied. Then the samples were placed in glass distillation flasks (3.6 L) and the vapor generated in the steam generator passed through the material for 2 h, with the same conditions as SD.

2.4. Preparation of cinnamon essential oil by MASDTOP

The microwave pre-treatment process was carried out in a microwave oven (P70D20N1P-G5 (W0) model, SCIENTZ, China). The essential oil was obtained by MASD extraction according to our former optimized processing method. Briefly, each cinnamon sample (150 g) was put in a plugged glass flask, and subjected to microwave under settled conditions: a power of 400 W with a water to raw material ratio (w/w) (16:1) for 5 min. Then, the samples were put in the glass distillation flask (3.6 L) and the steam generated in the steam generator passed through the material for 2 h with the same parameters as SD.

2.5. The yield calculationTOP

The yield of CEO was calculated by eq. 1.

2.6. GC–MS analysisTOP

The volatile components were analyzed by GC-MS (7890A-5975C, Agilent, USA) using a HP-5 capillary column (The column was 30 meters in length with an inner diameter of 0.25 mm and 0.25 μm thickness) with 95% methyl and 5% dimethyl poly siloxane as the stationary phase. Helium was used as the carrier gas at a flow rate of 1.0 mL/min. The split ratio was 30:1 and the injected quantity was 0.2 µL. The program temperature conditions were: the oven temperature was maintained at 40 °C for 5 min, increased to 260 °C at a rate of 3 °C/min and held for 10 min, and then increased to 280 °C at 10 °C/min and maintained for 2 min. The mass spectrometer was operated in the electron ionization mode at 70 eV and electron multiplier voltage was adopted at 1823.5 V. The ion source was established at a temperature of 230 °C, the maximum temperature was set at 250 °C, and the quadrupole rod temperature was employed at 150 °C, with the maximum temperature of 200 °C. The mass range for this scanning was 50.0-550.0 amu. All volatile components were identified by matching the recorded mass spectra with the standard mass spectra provided by NIST11.L database.

2.7. Principal component analysis (PCA)TOP

PCA was performed by SPSS 21.0 (SPSS Inc., Chicago, IL, USA) to identify the most important volatile components in the samples extracted by the three different methods explained above. p < 0.05 was considered statistically significant.

3. RESULTS AND DISCUSSIONTOP

3.1. The yield of essential oil TOP

According to Table 1, the UASD method showed the highest yield of CEO, followed by MASD. And the lowest yield was found for the SD method. Significant differences in yield were seen among the three methods. The cell wall may have been damaged by the ultrasonic treatment so that the CEO was able to escape more easily; microwave treatment was beneficial for CEO extraction, but the effect was not as good as the ultrasonic treatment.

Table 1. The yield of Cinnamon essential oil (CEO)*.
Extraction Method Yield of CEO (‰)
Steam distillation (SD) 3.91±0.09a
Ultrasound-assisted steam distillation (UASD) 8.33±0.02c
Microwave -assisted steam distillation (MASD) 5.53±0.03b
*Different superscript letters mean significant difference (p < 0.05), n=3.

3.2. Volatiles separated and identified by GC–MSTOP

The volatile components collected from the different methods were separated and identified by GC-MS. Total ionisation chromatograms (TIC) of the volatile constituents of CEO extracted via SD, UASD and MASD are shown in Figure 1(a–c) and the identification results of volatile components are listed in Table 2. Six kinds of compounds were identified, including aldehydes, esters, alcohols, terpenes, aromatics and ketones.

Table 2. Volatile components of cinnamon essential oil extracted by SD, UASD and MASD*
Code Components Molecular formula Relative content (%)a
SD UASD MASD
1 α-Pinene C10H16 0.092±0.006 0.042±0.004 0.054±0.003
2 Camphene C10H16 0.078±0.007 0.037±0.003 0.048±0.002
3 Benzaldehyde C7H6O 0.513±0.009 0.455±0.041 0.536±0.007
4 β-Pinene C10H16 0.107±0.005 0.044±0.004 0.090±0.001
5 6-Methyl-5-hepten-2-one C8H14O 0.016±0.002 0.008±0.001 0.012±0.002
6 p-Cymene C10H14 0.017±0.003 0.013±0.002 0.029±0.002
7 D-limonene C10H16 0.053±0.006 0.034±0.006 0.047±0.002
8 Salicylaldehyde C7H6O2 0.122±0.002 0.131±0.001 0.165±0.001
9 Phenethyl alcohol C8H10O 0.202±0.005 0.168±0.003 0.151±0.004
10 Phenylpropyl aldehyde C9H10O 0.205±0.005 0.180±0.001 0.217±0.005
11 Borneol C10H18O 0.268±0.004 0.242±0.002 0.255±0.002
12 Cinnamic alcohol C9H10O 0.319±0.002 0.311±0.002 0.435±0.002
13 Alpha terpineol C10H18O 0.031±0.005 0.038±0.003 0.020±0.004
14 Cinnamic aldehyde C9H8O 73.345±0.005 72.371±0.004 67.211±0.010
15 o-Anisaldehyde C8H8O2 0.376±0.005 0.391±0.006 0.531±0.010
16 Phenyl ethyl acetate C10H12O2 0.478±0.003 0.413±0.006 0.460±0.005
17 Trans -2-decenal C10H18O 0.137±0.004 0.103±0.004 0.104±0.003
18 Cyclosativene C15H24 0.025±0.002 0.018±0.002 0.028±0.003
19 α-Ylangene C15H24 0.044±0.001 0.043±0.003 0.045±0.005
20 α-Copaene C15H24 0.341±0.001 0.209±0.002 0.266±0.002
21 Eugenol C10H12O2 0.202±0.003 0.234±0.006 0.298±0.007
22 β-Elemene C15H24 0.027±0.001 0.022±0.003 0.025±0.002
23 Caryophyllene C15H24 0.227±0.007 0.170±0.001 0.250±0.007
24 α-Himachalene C15H24 0.109±0.001 0.087±0.002 0.091±0.007
25 Cinnamyl acetate C11H12O2 8.583±0.353 8.210±0.009 9.553±0.006
26 2-Methoxycinnamaldehyde C10H10O2 10.61±0.271 13.262±0.006 15.900±0.353
27 γ- Muurolene C15H24 0.451±0.002 0.365±0.002 0.407±0.001
28 α-Muurolene C15H24 0.232±0.003 0.186±0.004 0.244±0.005
29 α-Curcumene C15H24 0.194±0.002 0.160±0.002 0.205±0.007
30 β-Bisabolene C15H24 0.136±0.002 0.084±0.002 0.127±0.003
31 δ-Cadinene C15H24 0.324±0.001 0.258±0.003 0.283±0.007
32 α-Longipinene C15H24 0.190±0.001 0.146±0.004 0.198±0.004
33 Nerolidol C15H26O 0.537±0.007 0.496±0.006 0.591±0.007
34 Spathulenol C15H24O 0.738±0.035 0.590±0.030 0.656±0.036
35 Globulol C15H26O 0.576±0.007 0.355±0.006 0.316±0.007
36 Benzyl benzoate C14H12O2 0.094±0.006 0.124±0.003 0.152±0.004
*SD (steam distillation); UASD (ultrasound-assisted steam distillation); MASD (microwave-assisted steam distillation). aEach value was expressed as mean ± standard deviation (n=3)

Figure 1. Total ion chromatograms (TIC) of cinnamon essential oil extracted by steam distillation (a), ultrasound-assisted steam distillation (b) and microwave -assisted steam distillation(c).

 

The TIC results (Figure 1) showed that the retention time of the volatile components ranged between 10 and 56 min, most of which were concentrated between 22 mins to 40 mins. Referring to the NIST11.L map library, 36 identical volatile components were confirmed in all the tested samples extracted by SD, UASD and MASD, indicating that extraction method had a minor influence on the varieties of volatiles in the CEO. Moreover, there were three major components in the tested samples, and the remaining components were minor.

Table 2 shows that there were six aldehydes in the volatile components. The relative contents of total aldedydes were 85.104%, 86.713% and 84.447% for the samples extracted by SD, UASD and MASD, respectively, indicating that aldehydes were the major components in the essential oil. Among the aldehydes, cinnamic aldehyde accounted for the majority of the contents (73.345%, 72.371% and 67.211% for SD, UASD and MASD, respectively). Cinnamic aldehyde plays a crucial role in imparting the characteristic flavor to cinnamon (Jayatilaka et al., 1995), and it is a potent natural food preservative due to its antibacterial properties against five common foodborne pathogenic bacteria (Bacillus cereus, Listeria monocytogenes, Staphylococcus aureus, Escherichia coli, and Salmonella anatum) (Rui et al., 2009; Ooi et al., 2006). Apart from cinnamic aldehyde, 2-methoxycinnamaldehyde (10.611%, 13.262% and 15.900%, respectively), benzaldehyde (0.513%, 0.455% and 0.536% for SD, UASD and MASD, respectively), and o-anisaldehyde (0.376%, 0.391% and 0.531% for SD, UASD and MASD, respectively) were the other aldehydes were present in high relative contents. Among the above aldehyde components, 2-methoxycinnamaldehyde is considered a crucial component of the unique antibacterial property in CEO (Chang et al., 2001) and a potential agent for anticancer therapy (Wong et al., 2016). The contents of total cinnamic aldehydes (including cinnamic aldehyde and 2-methoxycinnamaldehyde) were 83.956% (SD), 85.633% (UASD) and 83.110% (MASD). Therefore, the highest total content of cinnamic aldehydes was obtained from UASD.

The analytical results were basically consistent with previous works, although there were differences in the number of volatile aldehydes between the present study and earlier studies (Rui et al., 2009; Li et al., 2010), which might be due to different cultivars. Furthermore, it was shown that the MASD extraction method produced the lowest cinnamic aldehyde, but yielded the highest 2-methoxycinnamaldehyde compared with the SD and UASD extraction methods. It might be due to the heating effect of the microwave, which led to the conversion of cinnamic aldehyde in the essential oil to 2-methoxycinnamaldehyde.

Yu et al., found that many esters in cinnamon were valuable packaging materials (Yu et al., 2007). Table 2 shows that three kinds of esters were identified in the CEO extracted by the three methods, i.e., phenylethyl acetate (0.478%, 0.413%, and 0.460% for SD, UASD and MASD, respectively), cinnamyl acetate (8.583%, 8.210%, and 9.553% for SD, UASD and MASD, respectively) and benzyl benzoate (0.094%, 0.124% and 0.152% for SD, UASD and MASD, respectively). Some researchers found that cinnamyl acetate and phenylethyl acetate were the main volatile substances that impart sweetness, fruitiness, floral aroma, and honey aroma to cinnamon essence ointments (Liet al., 2013). Moreover, the sample with the highest content of total esters (10.013%) and cinnamyl acetate (9.553%) was the CEO extracted by MASD. According to the report described by Jeyaratnam et al., (2016), it might be attributed to the molecular polarization of the microwave, which leads to the extraction of oxygenates from CEO.

The results showed that there were 7 alcohols and 17 terpenes in the CEO, which is basically consistent with the work by Singh et al., (2007). In summary, 7 alcohols identified from the CEO extracted by SD, UASD and MASD accounted for 2.300%, 2.025% and 2.325% of the total peak area, respectively; while 17 terpenes accounted for 3.206%, 2.260% and 2.724%, respectively. It was noteworthy that the content of each substance was less than 1%.

Other volatiles in the CEO were aromatics and ketones. The total amount of these compounds was less than 0.3% in the CEO, but they play vital roles in the aroma, antibacterial and antioxidant properties of CEO. For example, 6-methyl-5-hepten-2-one can endow CEO with the spicy flavor (Luna et al., 2006), which is significant for the aroma of the commercial essence (Alonso et al., 2009). As aromatics have phenolic groups, p-cymene and eugenol could be provided with definite antioxidant activities. In addition, researchers have also discovered that p-cymene is effective in suppressing conidial germination (Hong et al., 2015) and the p-cymene-treated fruit can reduce levels of several anthocyanins without any phytotoxic effect (Kordali et al., 2008), which might contribute to its overall antioxidant capacity. Moreover, eugenol is the main substance in C. zeylanicum leaf oil, C. pauciflorum leaf oil and C. burmannii leaf oil, which were found to exhibit strong antibacterial effects (Ali et al.,2005) and nematicidal activity (Park et al., 2007).

3.3. Results of principal component analysis TOP

The principal component analysis (PCA) has been widely applied in data mining to investigate the underlying structure and to extract the maximum information from large data matrices, so as to preserve as much complete data as possible (Lopez et al., 2007). In this work, PCA was carried out to interpret the differences in the volatile components of CEO extracted by SD, UASD and MASD. Table 3 lists 22 major components of the tested samples. Table 4 shows that the cumulative contribution percent of variance of the first two principal components (Eigen values >1) was 97.744%, where the first principal component accounted for 55.666% and the second principal component accounted for 42.078%. The first principal component was found to be significantly relevant to phenyl ethyl acetate (0.997), borneol (0.973) and γ-muurolene (0.966), while the second principal component was strongly characterized by salicylaldehyde (0.996), cinnamic aldehyde (0.995), o-anisaldehyde (0.993) and eugenol (0.991).

Table 3. Loadings of volatiles in the first two principal components
Compounds Component matrix Component Rotated component Matrixb
1 2 1 2
Benzaldehyde 0.916 0.065 0.760 0.516
Salicylaldehyde 0.615 -0.784 -0.002 0.996
Phenethyl alcohol -0.130 0.983 0.506 -0.853
Phenylpropyl aldehyde 0.993 0.025 0.796 0.595
Borneol 0.647 0.751 0.973 -0.190
Cinnamic alcohol 0.789 -0.604 0.247 0.963
Cinnamic aldehyde -0.651 0.754 -0.045 -0.995
o-Anisaldehyde 0.706 -0.707 0.117 0.993
Phenyl ethyl acetate 0.818 0.573 0.997 0.056
α-Copaene 0.574 0.808 0.951 -0.280
Eugenol 0.520 -0.852 -0.118 0.991
Caryophyllene 0.994 0.062 0.819 0.566
Cinnamyl acetate 0.894 -0.380 0.467 0.851
2-Methoxy cinnamaldehyde 0.358 -0.918 -0.287 0.942
γ- Muurolene 0.625 0.768 0.966 -0.216
α-Muurolene 0.987 0.147 0.866 0.496
α-Curcumene 0.998 0.108 0.843 0.527
δ-Cadinene 0.539 0.839 0.942 -0.326
α-Longipinene 0.973 0.197 0.886 0.447
Nerolidol 0.968 -0.251 0.605 0.796
Spathulenol 0.603 0.743 0.933 -0.211
Globulol 0.044 0.999 0.653 -0.757
b Rotation converged in triplicate.
Table 4. Percentage of variance and cumulative variance explained by the principal components
Component Initial eigen values Extraction sums squared loadings Rotation sums of squared loadings
Total Percentage of variance Cumulative percentage Total Percentage of variance Cumulative percentage Total Percentageof variance Cumulative percentage
1 12.247 55.666 55.666 12.247 55.666 55.666 11.102 50.464 50.454
2 9.257 42.078 97.744 9.257 42.078 97.744 10.402 47.281 97.744

Figure 2 (loading plot) and Figure 3 (scores plot) show the score distribution of the first two principal components. Both figures show clear separations among the CEO extracted by SD, UASD and MASD. The PCA figure of SD was positioned at the lower right quadrant of the biplot, which was characterized by globulol, phenethylalcohol and δ-cadinene, compared to UASD and MASD. The MASD was positioned at the upper left area of the score plot and close to the positive half of second principal component, dominated by the presence of cinnamyl alcohol and o-anisaldehyde. The components obtained by UASD appeared on the score map in the lower left area with cinnamic aldehyde.

Figure 2. Loading plots after principal components analysis of the variables in the plane defined by the two first principal components.

 

Figure 3. Scatter plot of scores via principal component analysis on the individuals in the plane defined by the two first principal components*
*SD (steam distillation); UASD (ultrasound-assisted steam distillation); MASD (microwave-assisted steam distillation)

 

4. CONCLUSIONSTOP

36 kinds of volatile components were identified in CEO. UASD contained the highest content of total cinnamic aldehydes and the highest yield compared to the SD and MASD methods. Therefore, the UASD method is recommended for future industrial application.

ACKNOWLEDGEMENTSTOP

This project was financially supported by the Project of Science and Technology of the Guangdong Province of China (Grant No. 2016A020210010, 2016B090923016); Natural Science Foundation of Guangdong Province, China (2018A030313026); the Fundamental Research Funds for the Central Universities (2019MS098). The authors gratefully acknowledge the support of Jinxuan Cao, Huihua Huang and Jin Liao during the development of this scientific work.

 

REFERENCESTOP


Ali SM, Khan AA, Ahmed I, Musaddiq M, Ahmed KS, Polasa, H. 2005. Antimicrobial activities of Eugenol and Cinnamaldehyde against the human gastric pathogen Helicobacter pylori. Ann. Clin. Microbiol. Antimicrob. 4, 20. https://doi.org/10.1186/1476-0711-4-20
Alonso A, Vázquez-Araújo L, García-Martínez S, Ruiz JJ, Carbonell-Barrachina ÁA. 2009. Volatile compounds of traditional and virus-resistant breeding lines of Muchamiel tomatoes. Eur. Food Res. Technol. 230, 315–323. https://doi.org/10.1007/s00217-009-1173-2
Azlina, MF, Hasfalina CM, Zurina ZA, Hishamuddin, J. 2013. Optimization and kinetic study of gaharu oil extraction. Int. Sch. Sci. Res. Innov. 7, 454–457.
Chang KS, Tak JH, Kim SI, Lee WJ, Ahn YJ. 2010. Repellency of Cinnamomum cassia bark compounds and cream containing cassia oil to Aedesaegypti (diptera: culicidae) under laboratory and indoor conditions. Pest. Manag. Sci. 62, 1032–1038. https://doi.org/10.1002/ps.1268
Chang ST, Chen PF, Chang SC. 2001. Antibacterial ­activity of leaf essential oils and their constituents from Cinnamomum osmophloeum. J. Ethnopharmacol. 77, 123–127. https://doi.org/10.1016/S0378-8741(01)00273-2
Cravotto G, Boffa L, Mantegna S, Perego P, Avogadro M, Cintas P. 2008. Improved extraction of vegetable oils under high-intensity ultrasound and/or microwaves. Ultrason. Sonochem. 15, 898–902. https://doi.org/10.1016/j.ultsonch.2007.10.009
Golmakani MT, Rezaei K. 2010. Microwave-assisted hydrodistillation of essential oil from zataria multiflora boiss. Eur. J. Lipid Sci. Technol. 110, 448–454. https://doi.org/10.1002/ejlt.200700239
Hong JK, Yang HJ, Jung H, Yoon DJ, Sang MK, Jeun YC. 2015. Application of volatile antifungal plant essential oils for controlling pepper fruit anthracnose by Colletotrichum gloeosporioides. Plan. Pathol. J. 31, 269–277. https://doi.org/10.5423/PPJ.OA.03.2015.0027
Jayatilaka A, Poole SK, Poole CF, Chichila TM. 1995. Simultaneous micro steam distillation/solvent extraction for the isolation of semivolatile flavor compounds from cinnamon and their separation by series coupled-column gas chromatography. Anal.Chim. Acta 302, 147–162. https://doi.org/10.1016/0003-2670(94)00445-R
Jeyaratnam N, Nour AH, Kanthasamy R, Nour AH, Yuvaraj AR, Akindoyo JO. 2016. Essential oil from Cinnamomum cassia bark through hydrodistillation and advanced microwave assisted hydrodistillation. Ind. Crops Prod. 92, 57–66. https://doi.org/10.1016/j.indcrop.2016.07.049
Kordali S, Cakir A, Ozer H, Cakmakci R, Kesdek M, Mete E. 2008. Antifungal, phytotoxic and insecticidal properties of essential oil isolated from Turkish Origanum acutidens and its three components, carvacrol, thymol and p-cymene. Bioresour. Technol. 99, 8788-8795. https://doi.org/10.1016/j.biortech.2008.04.048
Li R, Wang Y, Jiang ZT, Jiang S. 2010. Chemical composition of the essential oils of Cinnamomum loureirii Nees. From China obtained by hydrodistillation and microwave-assisted hydrodistillation. J. Essent. Oil Res. 22, 129–131. https://doi.org/10.1080/10412905.2010.9700281
Li YQ, Kong DX, Wu H. 2013. Analysis and evaluation of essential oil components of cinnamon barks using GC–MS and FTIR spectroscopy. Ind. Crops Prod. 41, 269–278. https://doi.org/10.1016/j.indcrop.2012.04.056
Lopez P, Sanchez C, Batlle R, Nerín C. 2007. Vapor-phase activities of cinnamon, thyme, and oregano essential oils and key constituents against foodborne microorganisms. J. Agric. Food Chem. 55, 4348–4356. https://doi.org/10.1021/jf063295u
Luna G, Morales MT, Aparicio R. 2006. Characterisation of 39 varietal virgin olive oils by their volatile compositions. Food Chem. 98, 243–252. https://doi.org/10.1016/j.foodchem.2005.05.069
Matan N, Rimkeeree H, Mawson AJ, Chompreeda P, Haruthaithanasan V, Parker M. 2006. Antimicrobial activity of cinnamon and clove oils under modified atmosphere conditions. Int. J. Food Microbiol. 107, 180–185. https://doi.org/10.1016/j.ijfoodmicro.2005.07.007
Nabavi S, Di LA, Izadi M, Sobarzo-Sánchez E, Daglia, M. 2015. Antibacterial effects of cinnamon: From farm to food, cosmetic and pharmaceutical industries. Nutrients 7, 7729–7748. https://doi.org/10.3390/nu7095359
Ooi LS, Li Y, Kam SL, Wang H, Wong EY, Ooi VE. 2006. Antimicrobial activities of cinnamon oil and cinnamaldehyde from the Chinese medicinal herb Cinnamomum cassia Blume. Am. J. Chin. Med. 34, 511–522. https://doi.org/10.1142/S0192415X06004041
Park IK, Kim J, Lee SG, Shin SC. 2007. Nematicidal activity of plant essential oils and components from ajowan (Trachyspermum ammi), allspice (Pimentadioica) and litsea (Litseacubeba) essential oils against pine wood nematode (Bursaphelenchus xylophilus). J. Nematol. 39, 275–279.
Rui W, Wang R, Bao Y. 2009. Extraction of essential oils from five cinnamon leaves and identification of their volatile compound compositions. Innovative Food Sci. Emerg. Technol. 2, 289–292. https://doi.org/10.1016/j.ifset.2008.12.002
Singh G, Maurya S, Delampasona MP, Catalan CAN. 2007. A comparison of chemical, antioxidant and antimicrobial studies of cinnamon leaf and bark volatile oils, oleoresins and their constituents. Food Chem. Toxicol. 45, 0–1661. https://doi.org/10.1016/j.fct.2007.02.031
Wong HY, Tsai KD, Liu, YH, Yang SM, Chen TW, Cherng J, Cherng JM. 2016. Cinnamomum verum component 2-methoxycinnamaldehyde: A novel anticancer agent with both anti-topoisomerase I and II activities in human lung adenocarcinoma a549 cells in vitro and in vivo. Phytother. Res. 30, 331–340. https://doi.org/10.1002/ptr.5536
Wong YC, Ahmad-Mudzaqqir MY, Wan-Nurdiyana WA. 2014. Extraction of essential oil from cinnamon (Cinnamomum zeylanicum). Orient J. Chem. 30, 37–47.
Yu LI, Huang GC. 2007. Microwave assisted synthesis and application of cinnamic esters on packaging materials. Packaging Eng. 28, 34–36.