NC State
BioResources
Gu, H., Liu, Y., Yan, J., Yu, X., Li, Y., Li, C., and Peng, W. (2022). "Chemical component characterization and potential medicinal utilization of extracts and pyrolyzates from Jasminum nudiflorum Lindl. wood," BioResources 17(2), 2525-2546.

Abstract

Utilization of lignocellulosic biomass is receiving increasing attention lately. In this study, Jasminum nudiflorum Lindl. (JNL) wood were extracted using methanol, ethanol, and benzene/ethanol (2:1, v:v) separately. Fourier transform-infrared spectroscopy (FTIR) and gas chromatography-mass spectroscopy (GC-MS) were used to study the chemical components of extracts. A thermogravimetric (TG) analyzer and pyrolysis (Py)-GC-MS investigated the characteristics of thermal loss law and pyrolyzates of JNL wood, respectively. The FTIR results showed that many functional groups were detected from the extracts of JNL wood, which were consistent with the chemical structures in the components detected by GC-MS. There were two obvious stages of thermal loss for removing moisture and decomposition of the organic constituents. The components of the extracts and pyrolyzates were esters, acids, aldehydes, alcohols, inositol, furfural, alkanes, phenols, ketones, antibiotics, saccharides, and glycosides. Among them, some components, such as ethyl iso-allocholate, scopoletin, isosorbide dinitrate, and idebenone, have high medicinal value. This study revealed the chemical component characterization and potential medicinal utilization of JNL wood. It provides the scientific basis for enhancing the utilization value of JNL wood.


Download PDF

Full Article

Chemical Component Characterization and Potential Medicinal Utilization of Extracts and Pyrolyzates from Jasminum nudiflorum Lindl. Wood

Haiping Gu,a Yuhao Liu,b Jie Yan,a Xuewei Yu,a Yiyang Li,a Cheng Li,a,* and Wanxi Peng a,*

Utilization of lignocellulosic biomass is receiving increasing attention lately. In this study, Jasminum nudiflorum Lindl. (JNL) wood were extracted using methanol, ethanol, and benzene/ethanol (2:1, v:v) separately. Fourier transform-infrared spectroscopy (FTIR) and gas chromatography-mass spectroscopy (GC-MS) were used to study the chemical components of extracts. A thermogravimetric (TG) analyzer and pyrolysis (Py)-GC-MS investigated the characteristics of thermal loss law and pyrolyzates of JNL wood, respectively. The FTIR results showed that many functional groups were detected from the extracts of JNL wood, which were consistent with the chemical structures in the components detected by GC-MS. There were two obvious stages of thermal loss for removing moisture and decomposition of the organic constituents. The components of the extracts and pyrolyzates were esters, acids, aldehydes, alcohols, inositol, furfural, alkanes, phenols, ketones, antibiotics, saccharides, and glycosides. Among them, some components, such as ethyl iso-allocholate, scopoletin, isosorbide dinitrate, and idebenone, have high medicinal value. This study revealed the chemical component characterization and potential medicinal utilization of JNL wood. It provides the scientific basis for enhancing the utilization value of JNL wood.

DOI: 10.15376/biores.17.2.2525-2546

Keywords: Jasminum nudiflorum Lindl.; Wood; Extract; Chemical component; Pyrolysis; High-value application

Contact information: a: School of Forestry, Henan Agricultural University, Zhengzhou 450002, China; b: School of Environmental and Municipal Engineering, North China University of Water Resources and Electric Power, Zhengzhou 450045, China;

* Corresponding authors: licheng@henau.edu.cn; pengwanxi@163.com

INTRODUCTION

Biorefining is an environmentally friendly method for developing chemicals by achieving the highest possible yield from limited and precious natural resources (Rosdiana et al. 2017). The concept of biorefining has attracted widespread attention. Many researchers have explored plants and their corresponding by-products as sources of value-added components to improve their utilization efficiency and promote sustainable development of ecology (Fernando et al. 2006; Dessbesell et al. 2017; Dugmore et al. 2017; Xie et al. 2017). Wood is an abundant renewable resource mainly composed of lignin, cellulose, and hemicellulose. Some applications of wood focus on energy and fuel production (Fitzpatrick et al. 2010), while others focus on wood extracts as a resource for value-added components, such as chemicals, biomedicines, and biologically active components (Rosdiana et al. 2017).

Jasminum nudiflorum Lindl. (JNL) is a type of deciduous shrub. It was originally planted in the middle and Northern provinces of China and has now been planted throughout the country (Takenaka et al. 2002). The flowers and leaves of JNL have been used as crude drugs in Chinese folk medicine (Tanahashi et al. 1999). Li et al. (2010) reported that a leaf extract of JNL is non-poisonous and is rich in flavonoids, secoiridoids, and fatty acids to treat inflammatory swelling, purulent eruptions, bruises, and traumatic bleeding. However, little detailed information is available about the chemical composition and corresponding utilization value of JNL wood. Therefore, the lack of a systematic and in-depth analysis on the chemical composition of JNL wood has slowed the development of high-value-added products with suitable processing efficiency.

Extraction is usually used to investigate the composition of biomass. For example, a total of 54 active ingredients were detected in Pterocarpus santalinus wood extracts, which were extracted by ethanol, ethanol/benzene (1:1), and methanol/ethanol (1:1) solvents (Jiang et al. 2020). The compounds from the leaves of Acacia nilotica extracted successively with 70% of acetone, methanol, ethanol, and chloroform were mainly tannins, saponins, glycosides, and flavonoids by gas chromatography-mass spectrometry (GC-MS) analysis (Revathi et al. 2017).

Pyrolysis (Py) has gained more attention because it has proved to be a much cheaper, more effective and efficient thermochemical pathway than other methods such as gasification and fermentation (Xing et al. 2016). Moreover, Py provides lots of information about the pyrolysis properties and thermal decomposition process of biomass through thermogravimetric (TG) analysis and biomass compositions, including the organic matter and gaseous products by Py-GC-MS analysis (Wen et al. 2019). According to Huang et al. (2020), the main pyrolytic by-products were phenols (19.2%), and furans (12.4%) for water hyacinth roots and nitrides (11.9%), and phenols (10%) for water hyacinth stems and leaves. The Py-GC-MS results in the research of Calixto et al. (2021) indicated that phenols, furans, and C1-C4 oxygenated compounds are the main pyrolysis products of corn stover, bean pods, sugarcane bagasse, and pineapple crown leaves.

Fig. 1. Experimental flow chart

In this study, the authors used methanol, ethanol, and benzene/ethanol (2:1, v:v) to extract the chemical components of JNL wood (without bark) separately. In addition, the components of JNL wood and the characteristics of its extracts were analyzed by Fourier transform-infrared spectroscopy (FTIR), GC-MS, Py-GC-MS, and thermogravimetric techniques (Fig. 1). This study provides useful information on the chemical composition, pyrolysis characteristics, and potential medicinal utilization of JNL wood. Moreover, it is of great value to provide the basis for sustainable development of high value-added applications of JNL wood.

EXPERIMENTAL

Materials and Reagents

The wood (without bark) of JNL was provided by the biotechnology laboratory of the College of Forestry, Henan Agricultural University (Zhengzhou, China). After drying at 40 °C, the wood was powdered by a plant disintegrator (FZ102, Tanjing Taisite Ins. Corp., Tanjing, China). Then, the powdered wood was sieved to pass through a 0.074-mm AS200 sieving instrument and stored in desiccators for further experiments of GC-MS, FTIR, Py-GC-MS, and TG analyses.

All reagents were purchased from Sigma Chemical Co. (St. Louis, MO, USA). All organic solvents used were high performance liquid chromatography grade, and inorganic chemicals were analytical grade or better.

Methods

JNL wood extraction

The JNL wood was extracted by methanol, ethanol, and benzene/ethanol (2:1, v:v) separately based on the methods of Gu et al. (2021). Briefly, 10.0 g of the JNL wood samples were mixed separately with methanol, ethanol and benzene/ethanol (2:1, v:v) at a solid-liquid ratio of 1:20. After 12 h of immersion at 25 °C, the mixed solutions were fully extracted by Soxhlet extraction for 5 h at 70 °C. After being filtrated rapidly with filter paper that had previously been immersed into ethanol for 24 h, the extracts were evaporated at 45 °C under a vacuum of 0.01 MPa and concentrated to 20 mL. Then, the extracts were kept into sealed reagent bottles at 4 °C for the following determination. The extraction yields of methanol, ethanol, and benzene/ethanol (2:1, v:v) in this study were 3.41%, 2.27%, and 4.06%, respectively.

FTIR analysis of the JNL wood extracts

According to the research of Wang et al. (2020), the three extract samples were dripped onto the KBr tablet, and the FTIR spectra of the tablets were recorded with an IR AFFINITY-1 spectrophotometer (Shimadzu, Kyoto, Japan) over a wavenumber range of 400 to 4000 cm−1 and scanned 32 times.

GC-MS analysis of the JNL wood extracts

The chemical components present in the extract samples were analyzed by a GC-MS instrument (Agilent 7890B-5977A; Agilent Technologies Inc., Santa Clara, CA, USA) equipped with a column of HP-5MS (30 m × 250 μm × 0.25 μm). The carrier gas used was high purity helium (99.99%) at a constant flow rate of 1 mL min-1. The split ratio was 2:1. The temperature program of the GC started at 50 °C, rose to 250 °C at a rate of 8 °C min-1, and then rose to 300 °C at a rate of 5 °C min-1.

The program of MS was scanned over the range 30 to 600 AMU (m/z), the ionization voltage was 70 eV, and the ionization current was 150 μA for electron ionization (EI). The ion source and the quadrupole temperature were set at 230 and 150 °C, respectively (Liu et al. 2017).

TG analysis of JNL wood

The samples of JNL wood were pyrolyzed using a TG analyzer (TGA Q50 V20.8 Build 34; TA Instruments, New Castle, DE, USA.) Initial sample mass of 3.03 mg was placed in a ceramic TG crucible. The analysis temperature was performed at 30 ℃, and then increased to 300 °C (ensure more useful organic compositions in the JNL wood are obtained during pyrolysis process) at a linear heating rate of 5 °C min-1. High-purity nitrogen with a flow rate of 100 mL min-1 was used as the carrier gas. The TG experiments were performed in triplicate, and the average results were used for plotting weight-derivative weight (deriv. weight, % °C-1, weight loss of sample per percent against temperature). The empty crucible clamp was treated as a reference.

Py-GC-MS analysis of JNL wood

The samples of JNL wood were pyrolyzed in a pyrolysis apparatus (PY-2020iS; Frontier Co., Ltd., Japan). The carrier gas was high purity helium, and the pyrolysis temperature was from room temperature to 300 °C at a heating rate of 20 ℃ ms-1. Pyrolysis products were analyzed by GC-MS (CDS5000-Agilent 7890B-5977A ISQ; Agilent Technologies Inc., Santa Clara, CA, USA), which was carried out with a HP-5MS column coated with a neutral phase (60 m × 250 μm i.d., 0.25 μm film thickness, Agilent Technologies Inc., Santa Clara, CA, USA). Helium at a constant flow rate (1.0 mL min-1) was used as carrier gas, and the injection valve temperature was set at 300 ℃. The temperature program of GC began at 40 °C and increased to 200 °C at the rate of 10 °C min-1, followed by a split injection at ratio of 60:1 (v:v). The program of MS was scanned over the 35 to 550 AMU (m/z), with an ionizing voltage of 70 eV and an ionization current of 150 µA of electron ionization (Gu et al. 2021).

RESULTS AND DISCUSSION

FTIR Analysis of JNL Wood Extracts

The technology of FTIR is rapid and sensitive to confirm the presence of chemical bonds and functional groups in samples (Suksuwan et al. 2015). FTIR is usually used as a qualitative technique. The specific absorbance bands in FTIR spectra provide direct molecular-level information allowing investigation of molecular structure and conformations (Movasaghi et al. 2008). Moreover, FTIR could be also used as a semi-quantitative indirect measure of the variation of functional groups through the analysis of second derivative curves (Koch et al. 1998). In this experiment, the FTIR spectra of methanol, ethanol, and benzene/ethanol extracts from JNL wood were recorded in the 400 to 4000 cm−1 region (Fig. 2).

Fig. 2. Infrared spectra of methanol, ethanol, and benzene/ethanol extracts from JNL wood

As shown in Fig. 2, the spectra of the three extracts were similar in appearance and various functional groups were observed. A broad band near 3379 cm-1 indicates the stretching vibrations of intermolecular OH (Lourençon et al. 2015). A group of absorbance peaks at 2835 to 2972 cm-1 was assigned mainly to the symmetric and asymmetric stretching vibrations of CH in alkanes. CH2 and CH3 bending vibrations are present at 1331 to 1454 cm-1 (Liu et al. 2013; Alekhina et al. 2015). The peak at 1655 cm-1 corresponded to C=C and C=O stretching, which originated from ester acids and aromatic components (Cruz et al. 2016). A small absorbance peak at 802 to 1257 cm-1 was related to the vibrations of C-C, C-O, and C-H in organic components (Gu et al. 2015). These results indicated that the similar chemical compounds present in the three extracts of JNL wood may contain phenols, polysaccharides, alcohols, acids, ketones, and ethers.

In contrast, some absorption peaks were not observed in the methanol extracts but existed in the ethanol and benzene/ethanol extracts (Fig. 2). For example, the obvious peak at 881 cm-1 and the small peak at 802 cm-1 (Fig. 2) were assigned to lone aryl C-H wag and two-adjacent aryl C-H wags of the aromatic components, respectively. A small absorption peak at 1257 cm-1 was related to the vibration of C-O-C and OH of polysaccharides. It indicated that some organic compounds can be extracted with ethanol and benzene/ethanol but not by methanol. This is probably because that some absorbance peaks may have not been detected because the chemical bonds were unstable or condensed during the extraction process with high temperature (Liu et al. 2017).

GC-MS Analysis of the JNL Wood Extracts

The methanol, ethanol, and benzene/ethanol extracts of JNL wood were analyzed by GC-MS to further characterize the chemical components of the extracts. The GC-MS chromatograms of the three kinds of extracts are shown in Figs. 3 through 5. The relative content of each product was calculated from the summed areas normalization of the peaks. According to Gu et al. (2021), the retention times and mass spectra of the extracts were compared with the GC-MS mass spectra from the National Institute Standard and Technology (NIST) library and published literatures to assign the components. The confidence was set at 80% in this study. The relative contents of the constituents higher than 2.50% were accepted as the main components (Table 1).

Fifty peaks were detected, and 47 compounds were identified in the methanol extract of JNL wood. The methanol extract of JNL wood was rich in secondary metabolites of inositol, which is important in plant biochemistry and physiology (Loewus and Murthy 2000; Stevenson et al. 2000). This result was different from the previous studies in which secoiridoids and saccharides were found to be the main components in the methanol extract of JNL leaf and bark, respectively (Takenaka et al. 2000; Gu et al. 2021).

In the ethanol extract of JNL wood, 30 peaks were detected and 22 compounds were identified and phenol was the main constituent in this extract (Table 1). This is notably different from the methanol extract of JNL bark, in which saccharide is the main component (Gu et al. 2021). Pourmorad et al. (2006) reported that the phenolic compounds are potent in radical scavenging for plants. It can be inferred that the ethanol extract of JNL wood has antioxidant capacity.

A total of 76 peaks were detected, and 76 compounds were identified in the benzene/ethanol extract of JNL wood. Alcohol, particularly 1-hexanol, 2-ethyl-, was the major component in the extract (Table 1). This result was consistent with the main constituent in the benzene/ethanol extract of JNL bark (Gu et al. 2021).

The spectral data show that the benzene/ethanol extract presented the most compounds, while the least compounds were found in the ethanol extract, which was consistent with the yields of the three extracts. This was because solvent choice is vital for the extracted compounds and the extraction efficiency Cowan (1999). In addition, the major components and the contents of each extract from JNL wood were completely different with those from JNL bark (Gu et al. 2021), which suggests different chemical structures of JNL wood and bark.

Based on the functional groups and peak areas, the compounds identified in the methanol, ethanol, and benzene/ethanol extracts of JNL wood can be classified as esters, acids, aldehydes, phenols, alcohols, ketones, inositol, furfural, alkanes, saccharides, and glycosides. Among the identified compounds, some of them are bioactive and used in medicines. For example, clindamycin, inositol, scopoletin, cinnamaldehyde, cyclotetrasiloxane, furanone, ester, steroids, and terpenoid substances that possess antimicrobial activity and are widely used to treat inflammatory diseases (Bharathy et al. 2012; Keskin et al. 2012; Sheela and Uthayakumari 2013; Anupama et al. 2014; Hameed et al. 2015; Angel et al. 2016; Narayanan et al. 2017). Specifically, 3,5-dimethoxy-4-hydroxycinnamaldehyde and ethyl iso-allocholate have been used as anticancer, anesthetic, antiulcer antiviral, and hypoglycemic drugs (Daffodil et al. 2012). Scopoletin inhibits thyroid function and prevents hyperglycemia without hepatotoxicity (Panda and Kar 2006).

Fig. 3. Total ion chromatogram of methanol extract from JNL wood

Fig. 4. Total ion chromatogram of ethanol extract from JNL wood

Fig. 5. Total ion chromatogram of benzene/ethanol extract from JNL wood

Table 1. Main Components of Methanol, Ethanol, and Benzene/Ethanol Extracts of JNL Wood by GC-MS (≥ 2.50%)

Isosorbide dinitrate is effective for slowing or preventing the progression of heart failure (Taylor et al. 2004). 4H-Pyran-4-one, 2,3-dihydro-3,5-dihydroxy-6-methyl- has an anti-cancer effect (Ban et al. 2007). DL-Arabinose has anti-tumor and antiviral activity (Hadi et al. 2016; Sosa et al. 2016). d-Mannose is effective for recurrent urinary tract infections (Kranjčec et al. 2014), and 7-methyl-Z-tetradecen-1-ol acetate can help with coughing (Lou et al. 2018). 5-hydroxymethylfurfural is reported to be a promising candidate for therapy of sickle cell disease (Rosatella et al. 2011). Acetamide N-methyl-N-[4-(3-hydroxypyrrolidinyl)-2-butynyl]- and 11,13-dihydroxy-tetradec-5-ynoic acid, methyl ester are new chemical compounds (Idan et al. 2015; Hussein et al. 2016,) and their effects need further research.

Thus, it appears that JNL wood is a potentially useful resource in the medical field. The highest total content of bioactive components was found in the methanol extract compared with the other two extracts. Therefore, the appropriate solvent can be chosen for extraction according to the need for different components. The authors’ results suggest that the methanol extract possessed the best prospects for development.

TG Analysis of JNL Wood

The pyrolysis processes of JNL wood from 30 to 300 °C at a linear heating rate of 5 °C min-1 were studied by TG analysis. The changes in wood weight and the derivative weight curves are shown in Fig. 6. The thermal degradation stages, weight loss, peak temperature and the corresponding residual weight of JNL wood in TG analysis are shown in Table 2. Based on the curve, there were two obvious stages of thermal loss in the wood (Fig. 6, Table 2). The first stage occurred up to 75 °C with a weight loss of 8.58%, and the maximum weight loss rate, which was shown in the first peak at 39.4 °C, occurred in this stage (Fig. 6, Table 2). According to some reports, this stage is responsible for the drying process, and the weight loss represents removal of moisture and light volatile matter (Liu et al. 2013; Mishra and Mohanty 2018). The decrease in weight (8.58%) indicated high content of moisture and small molecules in the wood. The curve between 75 and 200 °C was relatively flat, indicating possible occurrence of a small amount of polymer depolymerization and recombination in the sample (Gu et al. 2021).

The second stage of weight loss was 31.2% occurring in the temperature range of 200 to 300 °C (Fig. 6, Table 2). It was clear that the decomposition temperature for the wood biomass was about 200 °C, which was delayed compared with some reported plants with similar moisture content (Sait et al. 2012). There was a peak around 300 °C in the derivative weight curve (Fig. 6, Table 2), showing that the weight loss rate was fastest at this temperature. According to some reports, this stage is called the active pyrolytic zone, and weight loss is mainly due to decomposition of organic constituents (such as hemicellulose, cellulose, and lignin) into volatiles (Sait et al. 2012; Mishra and Mohanty 2018). In general, hemicellulose and cellulose (180 to 450 °C) are highly reactive and burn at a lower temperature than lignin, which has quite a broad burning range of 250 to 700 °C (Mishra and Mohanty 2018). The thermo-chemical decomposition route followed by a biomass is given as extract, cellulose, hemicelluloses, and lignin or char (Sait et al. 2012). At this stage, the weight has changed noticeably, which may be caused by remarkable chemical changes, such as macromolecules pyrolyzed into smaller more volatile molecules. In addition, the heating temperature in this study was set below 300 °C, which was far from carbonization temperature. Therefore, many volatile components can be obtained in the range 200 to 300 °C to make full use of JNL wood (Mishra and Mohanty 2018).

Fig. 6. Weight-derivative weight curves of JNL wood during pyrolysis at a heating rate of 5 °C min−1

Table 2. TG Analysis of JNL Wood during Pyrolysis

Py-GC-MS Analysis of JNL Wood

The pyrolysis temperature of the JNL wood sample was set at 300 °C, and the pyrolyzates were detected by GC-MS. The total ion chromatograms of pyrolyzates are shown in Fig. 7, and the results of pyrolyzates are listed in Table S1. A total of 203 peaks were detected, and 203 chemical compounds were identified. Among these chemical compounds, the main pyrolysis products from JNL wood were: acetaldehyde, hydroxy- (6.85%), acetic acid (6.63%), beta-D-glucopyranose, 1,6-anhydro- (3.41%), 4-(1-hydroxyallyl)-2-methoxyphenol (2.78%), methyl glyoxal (2.43%), acetic acid, methyl ester (2.23%), 2,4,7(1H, 3H, 8H)-pteridinetrione (2.22%), and ethylene oxide (2.12%).

According to functional groups, these compounds were classified as aldehydes, organic acids, phenols, esters, ketones, alkanes, and heterocyclic chemicals. Moreover, a large amount of valuable medicinal components were detected, such as n-hexadecanoic acid, which is an anti-inflammatory compound (Aparna et al. 2012). The cinnamaldehyde compound (3,5-dimethoxy-4-hydroxycinnamaldehyde) is widely used to synthesize various drugs, which has antimicrobial, anti-inflammatory, anticancer, anesthetic, antiulcer, antiviral, and hypoglycemic effects (Vadivel and Gopalakrishnan 2011). Benzene 1,2,3-trimethoxy-5-methyl- is an important chemical intermediate for intelligent stimulant medicine, such as idebenone (Kitajima el al. 1988). In addition, 6-azacytosine is a biologically active compound for antineurotic preparations of low toxicity and has immunotropic properties (Alekseeva et al. 1994). In conclusion, there were quite a few components in the pyrolysis products of JNL wood that are widely used in medicine. This suggests that JNL wood has potential for high-value applications in the field of pharmacy.

Fig. 7. Total ion chromatograms of pyrolyzates from JNL wood by Py-GC-MS

From the analysis results of GC-MS, TG, and Py-GC-MS, it is clearly shown that many components can be obtained from the extracts and pyrolyzates of JNL wood. Moreover, some components, such as 3,5-dimethoxy-4-hydroxycinnamaldehyde, ethyl iso-allocholate, scopoletin, isosorbide dinitrate, d-mannose, n-hexadecanoic acid, etc. can be used as medicines to combat various kinds of diseases. Thus, JNL wood is a potentially useful resource in the medical field. However, isolation and purification of the effective components to make full use of JNL wood are still a long ways away.

CONCLUSIONS

  1. A total of 47, 22, and 76 compounds were identified in the ethanol, methanol, and benzene/ethanol extracts of JNL wood, respectively. The main components were esters, alcohols, ethers, fatty acids, phenols, and hydrocarbons.
  2. A total of 203 peaks were detected and 203 chemical compounds were identified in the pyrolyzates at 300 °C of JNL wood. The main components were aldehydes, organic acids, phenols, esters, ketones, and alkanes.
  3. Many kinds of identified compounds, such as ethyl iso-allocholate, scopoletin, isosorbide dinitrate, and idebenone, are bioactive compounds with medicinal value.
  4. The TG analysis of the JNL wood resulted in two obvious stages of thermal loss for removal of moisture (up to 75 °C) and decomposition of the organic constituents (200 to 300 °C).
  5. This research learned the chemical composition, pyrolysis characteristics, and potential medicinal utilization of JNL wood, which provides the scientific basis for the JNL wood to become a resource with high application value.

ACKNOWLEDGMENTS

This work was supported by the National Natural Science Foundation of China (41701360) and Henan province science and technology research project (202102310608, 212102110109).

REFERENCES CITED

Alekhina, M., Ershova, O., Ebert, A., Heikkinen, S., and Sixta, H. (2015). “Softwood kraft lignin for value-added applications: Fractionation and structural characterization,” Industrial Crops and Products 66, 220-228. DOI: 10.1016/j.indcrop.2014.12.021

Alekseeva, I. V., Pal’chikovskaya, L. I., Shalamai, A. S., Ognyanik, S. S., Morgart, N. V., and Petrusha, N. A. (1994). “Synthesis of N 1-substituted 6-azacytosines and their biological activity,” Pharmaceutical Chemistry Journal 28(4), 227-230. DOI: 10.1007/BF02219791

Angel, L. P. L., Yusof, M. T., Ismail, I. S., Ping, B. T., Mohamed Azni, I. N. A., Kamarudin, N. H., and Sundram, S. (2016). “An in vitro, study of the antifungal activity of Trichoderma virens 7b and a profile of its non-polar antifungal components released against Ganoderma boninense,” The Journal of Microbiology 54(11), 732-744. DOI: 10.1007/s12275-016-6304-4

Anupama, N., Madhumitha, G., and Rajesh, K. S. (2014). “Role of dried fruits of Carissa carandas as anti-inflammatory agents and the analysis of phytochemical constituents by GC-MS,” Biomed Research International 2014, Article ID 512369. DOI: 10.1155/2014/512369

Aparna, V., Dileep, K. V., Mandal, P. K., Karthe, P., Sadasivan, C., and Haridas, M. (2012). “Anti‐inflammatory property of n‐hexadecanoic acid: Structural evidence and kinetic assessment,” Chemical Biology & Drug Design 80(3), 434-439. DOI: 10.1111/j.1747-0285.2012.01418.x

Ban, J. O., Hwang, I. G., Kim, T. M., Hwang, B. Y., Lee, U. S., Jeong, H. S., Yoon, Y. W., Kim, D. J., and Hong, J. T. (2007). “Anti-proliferate and pro-apoptotic effects of 2,3-dihydro-3,5-dihydroxy-6-methyl-4H-pyranone through inactivation of NF-κB in human colon cancer cells,” Archives of Pharmacal Research 30(11), article no. 1455. DOI: 10.1007/BF02977371

Bharathy, V., Sumathy, B. M., and Uthayakumari, F. (2012). Determination of phytocomponents by GC-MS in leaves of Jatropha gossypifolia L.,” Science Research Reporter 2(3), 286-290.

Calixto, G. Q., Melo, D., Melo, M., and Braga, R. M. (2021). “Analytical pyrolysis (Py–GC/MS) of corn stover, bean pod, sugarcane bagasse, and pineapple crown leaves for biorefining,” Brazilian Journal of Chemical Engineering 2021, 1-10. DOI: 10.1007/s43153-021-00099-1

Cowan, M. M. (1999). “Plant products as antimicrobial agents,” Clinical Microbiology Reviews 12(4), 564-582. DOI: 10.1128/CMR.12.4.564

Cruz, M. D., Van Schoors, L., Benzarti, K., and Colin, X. (2016). “Thermo‐oxidative degradation of additive free polyethylene. Part I. Analysis of chemical modifications at molecular and macromolecular scales,” Journal of Applied Polymer Science 133(18), article ID 43287. DOI: 10.1002/app.43287

Daffodil, E. D., Uthayakumari, F. K., and Mohan, V. R. (2012). “GC-MS determination of bioactive compounds of Curculigo orchioides gaertn,” Science Research Reporter 2(3), 198-201.

Dessbesell, L., Xu, C., Pulkki, R., Leitch, M., and Mahmood, N. (2017). “Forest biomass supply chain optimization for a biorefinery aiming to produce high-value bio-based materials and chemicals from lignin and forestry residues: A review of literature,” Canadian Journal of Forest Research 47(3), 277-288. DOI: 10.1139/cjfr-2016-0336

Dugmore, T. I. J., Clark, J. H., Bustamante, J., Houghton, J. A., and Matharu, A. S. (2017). “Valorisation of biowastes for the production of green materials using chemical methods,” Topics in Current Chemistry 375(2), Article Number 46. DOI: 10.1007/s41061-017-0133-8

Fernando, S., Adhikari, S., Chandrapal, C., and Murali, N. (2006). “Biorefineries: Current status, challenges, and future direction,” Energy & Fuels 20(4), 1727-1737. DOI: 10.1021/ef060097w

FitzPatrick, M., Champagne, P., Cunningham, M. F., and Whitney, R. A. (2010). “A biorefinery processing perspective: Treatment of lignocellulosic materials for the production of value-added products,” Bioresource Technology 101(23), 8915-8922. DOI: 10.1016/j.biortech.2010.06.125

Gu, H., Luo, X., Wang, H., Wu, L., Wu, J., and Xu, J. (2015). “The characteristics of phenanthrene biosorption by chemically modified biomass of Phanerochaete chrysosporium,” Environmental Science & Pollution Research 22(15), 11850-11861. DOI: 10.1007/s11356-015-4451-5

Gu, H., Foong, S. Y., Su, S. L., Yue, X., Yang, J., and Peng, W. (2021). “Characterization and potential utilization of extracts and pyrolyzates from Jasminum nudiflorum lindl. bark,” Journal of Analytical and Applied Pyrolysis 155, article ID 105092. DOI: 10.1016/j.jaap.2021.105092

Hadi, M. Y., Mohammed, G. J., and Hameed, I. H. (2016). “Analysis of bioactive chemical compounds of Nigella sativa using gas chromatography-mass spectrometry,” Journal of Pharmacognosy & Phytotherapy 8(2), 8-24. DOI: 10.5897/jpp2015.0364

Hameed, I. H., Hamza, L. F., and Kamal, S. A. (2015). “Analysis of bioactive chemical compounds of Aspergillus niger by using gas chromatography-mass spectrometry and Fourier-transform infrared spectroscopy,” Journal of Pharmacognosy and Phytotherapy 7(8), 132-163. DOI: 10.5897/JPP2015.0354

Huang, H., Liu, J., Liu, H., Evrendilek, F., and Buyukada, M. (2020). “Pyrolysis of water hyacinth biomass parts: Bioenergy, gas emissions, and by-products using TG-FTIR and Py-GC/MS analyses,” Energy Conversion and Management 207, article ID 112552. DOI: 10.1016/j.enconman.2020.112552

Hussein, H. J., Hadi, M. Y., and Hameed, I. H. (2016). “Study of chemical composition of Foeniculum vulgare using Fourier transform infrared spectrophotometer and gas chromatography-mass spectrometry,” Journal of Pharmacognosy and Phytotherapy 8(3), 60-89. DOI: 10.5897/JPP2015.0372

Idan, S. A., Al-Marzoqi, A. H., and Hameed, I. H. (2015). “Spectral analysis and anti-bacterial activity of methanolic fruit extract of Citrullus colocynthis using gas chromatography-mass spectrometry,” African Journal of Biotechnology 14(46), 3131-3158. DOI: 10.5897/AJB2015.14957

Jiang, S., Wei, Y., Liu, Z., Ni, C., Gu, H., and Peng, W. (2020). “Molecules and functions of rosewood: Pterocarpus santalinus,” Journal of King Saud University-Science 32(2), 1712-1717. DOI: 10.1016/j.jksus.2020.01.006

Keskin, D., Ceyhan, N., Ugur, A., and Dbeys, A. D. (2012). “Antimicrobial activity and chemical constitutions of West Anatolian olive (Olea europaea L.) leaves,” Journal of Food Agriculture & Environment 10(2), 99-102. DOI: 10.1016/j.jff.2012.03.002

Kitajima, N., Takemura, K., MoRo-okA, Y., Yoshikuni, T., Akada, M., Tomotaki, Y., and Taniguchi, M. (1988). “The selective liquid-phase oxidation of 3,4,5-trimethoxytoluene to 3,4,5-trimethoxybenzaldehyde,” Bulletin of the Chemical Society of Japan 61(3), 1035-1037. DOI: 10.1246/bcsj.61.1035

Koch, A., Krzton, A., Finqueneisel, G., Heintz, O., Weber, J. V., and Zimny, T. (1998). “A study of carbonaceous char oxidation in air by semi-quantitative FTIR spectroscopy,” Fuel 77(6), 563-569. DOI: 10.1016/S0016-2361(97)00157-9

Kranjčec, B., Papeš, D., and Altarac, S. (2014). “D-mannose powder for prophylaxis of recurrent urinary tract infections in women: A randomized clinical trial,” World Journal of Urology 32(1), 79-84. DOI: 10.1007/s00345-013-1091-6

Li, X., Deng, S., and Fu, H. (2010). “Inhibition by Jasminum nudiflorum Lindl. leaves extract of the corrosion of cold rolled steel in hydrochloric acid solution,” Journal of Applied Electrochemistry 40(9), 1641-1649. DOI: 10.1007/s10800-010-0151-5

Liu, B., Li, Y., Wu, S., Li, Y., Deng, S., and Xia, Z. (2013). “Pyrolysis characteristic of tobacco stem studied by Py-GC/MS, TG-FTIR, and TG-MS,” BioResources 8(1), 220-230. DOI: 10.15376/biores.8.1.220-230

Liu, L., Hu, C., Liu, L., Zhang, S., Chen, K., and He, D. (2017). “Rapid detection and separation of olive oil and Camellia oil based on ion mobility spectrometry fingerprints and chemometric models,” European Journal of Lipid Science & Technology 119(3), article ID 1500463. DOI: 10.1002/ejlt.201500463

Loewus, F. A., and Murthy, P. P. (2000). “myo-Inositol metabolism in plants,” Plant Science 150(1), 1-19. DOI: 10.1016/S0168-9452(99)00150-8

Lou, J., Chen, J., Ni, C., and Peng, W. (2018). “Molecules and functions of rosewood: Pterocarpus cambodianus,” Arabian Journal of Chemistry 11(6), 763-770. DOI: 10.1016/j.arabjc.2017.12.030

Lourençon, T. V., Hansel, F. A., Silva, T. A. D., Ramos, L. P., Muniz, G. I. B. D., and Magalhães, W. L. E. (2015). “Hardwood and softwood kraft lignins fractionation by simple sequential acid precipitation,” Separation and Purification Technology 154, 82-88. DOI: 10.1016/j.seppur.2015.09.015

Mishra, R. K., and Mohanty, K. (2018). “Pyrolysis kinetics and thermal behavior of waste sawdust biomass using thermogravimetric analysis,” Bioresource Technology 251, 63-74. DOI: 10.1016/j.biortech.2017.12.029

Movasaghi, Z., Rehman, S., and ur Rehman, D. I. (2008). “Fourier transform infrared (FTIR) spectroscopy of biological tissues,” Applied Spectroscopy Reviews 43(2), 134-179. DOI: 10.1080/05704920701829043

Narayanan, B. L., Kannappan, K., Subburaju, T., and Sajeeth, C. I. (2017). “GC-MS analysis and HPTLC fingerprinting profile of hydroalcoholic extract of Polygonum barbatum Linn. leaves,” Indian Journal of Pharmaceutical Sciences 79(3), 477-482. DOI: 10.4172/pharmaceutical-sciences.1000253

Panda, S., and Kar, A. (2006). “Evaluation of the antithyroid, antioxidative and antihyperglycemic activity of scopoletin from Aegle marmelos leaves in hyperthyroid rats,” Phytotherapy Research 20(12), 1103-1105. DOI: 10.1002/ptr.2014

Pourmorad, F., Hosseinimehr, S. J., and Shahabimajd, N. (2006). “Antioxidant activity, phenol and flavonoid contents of some selected Iranian medicinal plants,” African Journal of Biotechnology 5(11), 1142-1145. DOI: 10.4314/ajb.v5i11.42999

Revathi, S., Govindarajan, R. K., Rameshkumar, N., Hakkim, F. L., Mohammed, A. B., Krishnan, M., and Kayalvizhi, N. (2017). Anti-cancer, anti-microbial and anti-oxidant properties of Acacia nilotica and their chemical profiling,” Biocatalysis and Agricultural Biotechnology 11, 322-329. DOI: 10.1016/j.bcab.2017.08.005

Rosatella, A. A., Simeonov, S. P., Frade, R. F., and Afonso, C. A. (2011). “5-Hydroxymethylfurfural (HMF) as a building block platform: Biological properties, synthesis and synthetic applications,” Green Chemistry 13(4), 754-793. DOI: 10.1039/c0gc00401d

Rosdiana, N. A., Dumarçay, S., Gérardin, C., Chapuis, H., Santiago-Medina, F. J., Sari, R. K., Syafii,W., Gelhaye, E., Raharivelomanana, P., Mohammed, R., et al. (2017). “Characterization of bark extractives of different industrial Indonesian wood species for potential valorization,” Industrial Crops and Products 108, 121-127. DOI: 10.1016/j.indcrop.2017.06.034

Sait, H. H., Hussain, A., Salema, A. A., and Ani, F. N. (2012). “Pyrolysis and combustion kinetics of date palm biomass using thermogravimetric analysis,” Bioresource Technology 118(4), 382-389. DOI: 10.1016/j.biortech.2012.04.081

Sheela, D., and Uthayakumari, F. (2013). “GC-MS analysis of bioactive constituents from coastal sand dune taxon–Sesuvium portulacastrum (L.),” Bioscience Discovery 4(1), 47-53.

Sosa, A. A., Bagi, S. H., and Hameed, I. H. (2016). “Analysis of bioactive chemical compounds of Euphorbia lathyrus using gas chromatography-mass spectrometry and Fourier-transform infrared spectroscopy,” Journal of Pharmacognosy and Phytotherapy 8(5), 109-126. DOI: 10.5897/JPP2015.0371

Stevenson, J. M., Perera, I. Y., Heilmann, I., Persson, S., and Boss, W. F. (2000). “Inositol signaling and plant growth,” Trends in Plant Science 5(6), 252-258. DOI: 10.1016/S1360-1385(00)01652-6

Suksuwan, A., Lomlim, L., Dickert, F. L., and Suedee, R. (2015). “Tracking the chemical surface properties of racemic thalidomide and its enantiomers using a biomimetic functional surface on a quartz crystal microbalance,” Journal of Applied Polymer Science 132(30), article ID 42309. DOI: 10.1002/app.42309

Takenaka, Y., Tanahashi, T., Taguchi, H., Nagakura, N., and Nishi, T. (2002). “Nine new secoiridoid glucosides from Jasminum nudiflorum,” Chemical and Pharmaceutical Bulletin 50(3), 384-389. DOI: 10.1248/cpb.50.384

Tanahashi, T., Takenaka, Y., Nagakura, N., and Nishi, T. (1999). “Three secoiridoid glucosides from Jasminum nudiflorum,” Journal of Natural Products 62(9), 1311-1315. DOI: 10.1021/np9901175

Taylor, A. L., Ziesche, S., Yancy, C., Carson, P., D’Agostino, Jr., R., Ferdinand, K., Taylor, M., Adams, K., Sabolinski, M., Worcel, M., et al. (2004). “Combination of isosorbide dinitrate and hydralazine in blacks with heart failure,” New England Journal of Medicine 351(20), 2049-2057. DOI: 10.1056/NEJMoa042934

Vadivel, E., and Gopalakrishnan, S. (2011). “GC-MS analysis of some bioactive constituents of Mussaenda frondosa Linn.,” International Journal of Pharmacy 4(3), 226-234. DOI: 10.1080/01919510600985945

Xie, Y., Ge, S., Jiang, S., Liu, Z., Chen, L., Wang, L., Chen, J., Qin, L., and Peng, W. (2017). “Study on biomolecules in extractives of Camellia Oleifera, fruit shell by GC-MS,” Saudi Journal of Biological Sciences 25(2), 234-236. DOI: 10.1016/j.sjbs.2017.08.006

Xing, S., Yuan, H., Qi, Y., Lv, P., Yuan, Z., and Chen, Y. (2016). “Characterization of the decomposition behaviors of catalytic pyrolysis of wood using copper and potassium over thermogravimetric and Py-GC/MS analysis,” Energy 114, 634-646. DOI: 10.1016/j.energy.2016.07.154

Wang, Y., Liu, Q., Zheng, D., Zhao, Y., Wang, T., Yan, X., and Gu, H. (2020). “Active constituents of Koelreuteria paniculata root,” Thermal Science 24(3, Part A), 1737-1744. DOI: 10.2298/TSCI190517046W

Wen, S., Yan, Y., Liu, J., Buyukada, M., and Evrendilek, F. (2019). “Pyrolysis performance, kinetic, thermodynamic, product and joint optimization analyses of incense sticks in N2 and CO2 atmospheres,” Renewable Energy 141, 814-827. DOI: 10.1016/j.renene.2019.04.040

Article submitted: August 11, 2021; Peer review completed: November 14, 2021; Revised version received and accepted: February 12, 2022; Published: March 15, 2022.

DOI: 10.15376/biores.17.2.2525-2546

 

APPENDIX (Supplementary Information)

Table S1. Py-GC-MS Analysis of the JNL Wood