Next Article in Journal
Molecular Characterization of a Recombinant Isolate of Tomato Leaf Curl New Delhi Virus Associated with Severe Outbreaks in Zucchini Squash in Southern Italy
Next Article in Special Issue
The Wild Carrot (Daucus carota): A Phytochemical and Pharmacological Review
Previous Article in Journal
Confocal Microscopy Investigations of Biopolymeric PLGA Nanoparticle Uptake in Arabidopsis thaliana L. Cultured Cells and Plantlet Roots
Previous Article in Special Issue
Medicinal Uses of the Fabaceae Family in Zimbabwe: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phytochemical Composition and Antibacterial Activity of Barleria albostellata C.B. Clarke Leaf and Stem Extracts

by
Serisha Gangaram
1,*,
Yougasphree Naidoo
1,
Yaser Hassan Dewir
2,
Moganavelli Singh
1,
Johnson Lin
1 and
Hosakatte Niranjana Murthy
3
1
School of Life Sciences, Westville Campus, University of KwaZulu-Natal, Private Bag X54001, Durban 4000, South Africa
2
Plant Production Department, College of Food and Agriculture Sciences, King Saud University, P.O. Box 2460, Riyadh 11451, Saudi Arabia
3
Department of Horticultural Science, Chungbuk National University, Cheongju 28644, Republic of Korea
*
Author to whom correspondence should be addressed.
Plants 2023, 12(13), 2396; https://doi.org/10.3390/plants12132396
Submission received: 17 May 2023 / Revised: 2 June 2023 / Accepted: 14 June 2023 / Published: 21 June 2023
(This article belongs to the Special Issue Phytochemistry and Pharmacological Properties of Medicinal Plants)

Abstract

:
Barleria albostellata (Acanthaceae) is a shrub located in South Africa and is relatively understudied. However, plants within this genus are well known for their medicinal and ethnopharmacological properties. This study aimed to characterise the phytochemical compounds and antibacterial efficacies of B. albostellata. Phytochemical analysis, fluorescence microscopy and gas chromatography-mass spectrometry (GC-MS) analysis were performed to determine the composition of compounds that may be of medicinal importance. Crude leaf and stem extracts (hexane, chloroform and methanol) were subjected to an antibacterial analysis against several pathogenic microorganisms. The qualitative phytochemical screening of leaf and stem extracts revealed the presence various compounds. Fluorescence microscopy qualitatively assessed the leaf and stem powdered material, which displayed various colours under bright and UV light. GC-MS chromatograms represents 10–108 peaks of various compounds detected in the leaf and stem crude extracts. Major pharmacologically active compounds found in the extracts were alpha-amyrin, flavone, phenol, phytol, phytol acetate, squalene and stigmasterol. Crude extracts positively inhibited Gram-positive and Gram-negative bacteria. Significance was established at p < 0.05 for all concentrations and treatments. These results indicate that the leaves and stems of B. albostellata are rich in bioactive compounds, which could be a potential source of antibacterial agents for treating various diseases linked to the pathogenic bacteria studied. Future discoveries from this plant could advance the use of indigenous traditional medicine and provide novel drug leads.

1. Introduction

The dependence on plants as a source of medicine is prevalent in developing countries, especially where traditional medicine plays an important role in health care [1,2,3]. As defined by the World Health Organisation (WHO), traditional medicine is the knowledge, skill and practices based on the beliefs and experiences in various cultures [4]. The demand for herbal medicines worldwide is rapidly increasing due to their availability, low cost and higher safety margins [5,6,7]. The adverse side-effects of conventional medicine are related to certain pharmacological compounds; therefore, creating different therapies with greater effectiveness and bioavailability, with fewer side-effects, is essential [8,9]. Natural compounds isolated from plants have been assumed to remain an essential part of the exploration for new medicines against human diseases [10,11].
Africa is considered the cradle of humankind, comprising a rich cultural and biological diversity and with healing practices varying in regions [5,7,12]. Infectious diseases are a serious problem in Africa [13,14], with one of the leading causes of morbidity and mortality arising from bacterial infections (Gram-positive and Gram-negative bacteria) [15,16,17,18]. However, with the use of herbal medicine, certain bacterial infections have been reported to be under control, while others are resistant [19]. According to the WHO, by 2050, there will be approximately 10 million deaths arising from common diseases such as respiratory and urinary tract infections and drug-resistant pathogens, surpassing the number of deaths resulting from cancer [20,21]. Therefore, it is crucial to find alternative solutions, such as herbal extracts, to overcome future threats in the medical field [22].
Throughout the history of mankind, plant extracts have been used to treat various ailments through herbal preparations [23,24,25,26]. These preparations include concoctions, decoctions, infusions and teas [27,28]. Plants are rich in several naturally occurring phytochemicals such as alkaloids, flavonoids, tannins and terpenoids, which have been found to possess antimicrobial properties. These secondary metabolites are important components of a plant’s natural defence mechanisms and are products of primary metabolism [29,30,31,32,33,34,35,36].
In South Africa, about 3000 medicinal plants were reported to be used regularly, and from these plants, 38 indigenous species were commercialised [37,38]. These commercialised plant species are available as processed material in various forms, such as capsules, ointments, tablets or teas [39,40]. The verified record of natural products being used in drug discovery [41,42] has provided compelling evidence for increasing the exploration of nature to identify unique active compounds as promising leads for effective drug development [43,44,45]. There has been significant progress in pharmaceutical industries in search of important plant-based medicinal compounds; however, a significant amount of the plant biodiversity remains unexplored [46,47]. Screening plants for biologically active compounds has resulted in the development of new therapeutic drugs to treat several diseases [48].
Barleria albostellata C.B. Clarke (Acanthaceae) is a shrub located in South Africa. This shrub extends from Limpopo, Gauteng and Mpumalanga to KwaZulu-Natal [49]. Although B. albostellata has no recorded practice in traditional medicine, many species within the genus have been widely used in traditional medicine and were confirmed to contain various compounds possessing biological effects such as analgesic, anti-inflammatory, antileukaemic, antihyperglycemic, antitumour, anti-amoebic, antibiotic and virucidal activities [50,51,52,53,54,55]. Amoo et al. [56] examined the antioxidant potential of the methanolic extracts of the leaves and stems of B. albostellata and found the greatest flavonoid content in the stems of B. albostellata stems. Additionally, Amoo et al. [57] verified the antibacterial, antifungal and anti-inflammatory activity of B. albostellata against B. subtillis, S. aureus and E. coli. Thus, further investigation on B. albostellata is important as this study will provide baseline information on the potential usage of extracts from this plant. There is a scarcity of scientific data on the phytochemical compounds of the leaf and stem extracts of B. albostellata and its potential antibacterial activity against human pathogens. This study was therefore carried out to bridge these gaps.

2. Results and Discussion

2.1. Organoleptic Characteristics and Crude Extract Yield of B. albostellata

Organoleptic evaluation is a conventional, qualitative method whereby an individual uses their sight, smell, taste and touch to document the characteristic features of crude drugs. These assessments may serve as a baseline for preliminary phytochemical and pharmacological screening of a plant [58]. The organoleptic features of B. albostellata were evaluated by using sensory organs (Table 1). The following features were noted on both adaxial and abaxial surfaces: The leaves were grey-green in colour but lighter on the lower surface. The stems appeared as ‘yellow-buff’ on the upmost internodes and white/cream below. The odour of the leaves was slightly aromatic, whilst the stems were inodorous. The taste was acrid for both leaves and stems. Both surfaces of the leaves were velvety, whereas the stems were woody and glabrescent. According to Shaheen et al. [59], organoleptic studies are important taxonomic parameters, assisting in the verification of several medicinal plant species.
The highest percentage yield of the crude extracts from B. albostellata was obtained from the methanolic extract of the leaves (16.78%), followed by 9.38% from the methanolic stem extract (Table 2). The lowest percentage yield was observed in the hexane stem extract (1.94%). Overall, this suggests that the percentage yield of phyto-compounds in B. albostellata was greater in the leaf crude extract than from the stem. Furthermore, this implies that there may be more polar compounds in the leaf extracts and a smaller amount of non-polar compounds in the stem. Therefore, the yield obtained indicates the polarity of the different solvents is related to the plants’ pharmacological importance [60,61]. Each crude extract (hexane, chloroform and methanol) displayed distinct colours (Table 2). Hexane extracts were oily upon evaporation of the solvent, whereas chloroform and methanol dried to a hard-sticky solid.

2.2. Phytochemical Screening for Major Classes of Compounds in Extracts of B. albostellata Using Qualitative Colour Tests, TLC and Fluorescence Analysis

Major compounds identified in leaf and stem extracts of B. albostellata were alkaloids, amino acids, carbohydrates, flavonoids, mucilage and gums, phenols, saponins, terpenoids and sterols (Table 3). Fixed oils and fats were present in leaves and absent in the stem extracts. The intensity of compounds in the leaf extracts was greater in comparison to the stem. These compounds can act as defence mechanisms against various microorganisms, herbivores and insects [62,63,64].
Various phytochemicals have been known to contain diverse activities that may help protect against chronic diseases [65,66]. Amoo et al. [56] reported the presence of phenols, iridoids, gallotannins, flavonoids and condensed tannin in the leaves and stems of B. albostellata. These authors also found the total iridoid content to be the highest in the leaves of B. albostellata. Important pharmacological alkaloids can be found in iridoids, as this compound is known to be a precursor in the biosynthesis of alkaloids [67,68]. Similar compounds were also found in extracts of other Barleria species including B. acuminata [69], B. dintteri [10], B. cristata [70,71,72], B. longiflora [73] and B. prionitis [74].
Metabolites such as alkaloids and terpenoids (Table 3) were reported to contain antimicrobial, anticancer and anti-malarial properties [75,76,77,78,79,80]. Additionally, alkaloids have organic and natural constituents with sedative and analgesic roles [81]. Amino acids, carbohydrates and fixed oils and fats (Table 3) were reported to contain antioxidant properties [82,83,84]. Flavonoids and phenols (Table 3) possessed anti-inflammatory, anti-apoptosis, anti-carcinogen and anti-ageing properties [85,86,87,88].
Mucilage and gums (Table 3) are used in the treatment of gastric ulcers, for wound healing and as cytoprotective agents, and also contain antipyretic and antiseptic properties [89,90,91,92]. Saponins present in plant extracts (Table 3) are believed to contain anticancer, antioxidant, antiviral and anti-inflammatory properties [93,94]. Additionally, saponins display several hepatoprotective and antimicrobial activities [95]. Sterols were found to reduce cholesterol levels and contain anti-inflammatory and antioxidant properties (Table 3) [96,97,98,99,100].
A technique used in the qualitative assessment of natural products or crude drugs is fluorescence analysis, which is an important tool for pharmacognostic evaluation [101,102]. The powdered and fluorescence characteristics of the leaf and stem powder of B. albostellata are presented in Figure 1, Figure 2, Figure 3 and Figure 4. The powdered leaf and stem material treated with several reagents displayed various colours when observed under bright light, and this is compared to the colours observed under UV light. It should be noted that the colours indicated for the powdered leaf and stem material viewed under bright light were described according to the overall appearance. The purity and quality of crude drugs are occasionally authenticated using standard fluorescence characteristics, as certain natural products display no fluorescence in daylight but do so under UV light [103]. Natural products such as berberine alkaloids exhibit fluorescence under UV light but not in daylight [104,105]. As most crude drug materials do not fluoresce, these materials are converted either into fluorescent decomposition- or by-products with the aid of several reagents [101,102,106]. Furthermore, fluorescence analysis can be used to preserve the quality and effectiveness of crude drug materials by easily detecting adulterants and substituents [101,102,107,108].
The most prominent UV colour that stands out in both leaf and stem powder was blue (Figure 2 and Figure 4). This colour was observed in multiple plant samples where different reagents were used. According to Chase and Pratt [109], several drugs display duplication of colours, as there are sometimes more than four drugs found in a particular colour group. According to Sridharan and Gounder [110], powdered leaves of B. montana were separately exposed to 24 h of light with the addition of various reagents. These included powder + water, + ethanol, + ethyl acetate, + hexane, + chloroform and + acetone. Colours observed from the various regents after 24 h were orange, green, pale and light green, respectively. When these samples were exposed to UV light, colours observed were greenish-orange, light and dark green, pale yellow, yellowish-orange and pale red. Similar fluorescence results listed in this study were detected for certain reagents in the powdered leaf material for B. noctiflora [111] and B. gibsoni [112].

2.3. Phytochemical Screening for Major Classes of Compounds in B. albostellata Extracts Using GC-MS

The GC-MS chromatogram represents 10 peaks (Figure 5), 10 peaks (Figure 6) and 108 peaks (Figure 7) of various compounds detected in the leaf hexane, chloroform and methanolic extract, respectively. Additionally, 10 peaks (Figure 8), 10 peaks (Figure 9) and 104 peaks (Figure 10) were identified in the stem hexane, chloroform and methanolic extract. This analysis is used for the qualitative and quantitative examination of active compounds in plants [113,114]. Several overlapping peaks were observed in the middle stages of most chromatograms, Figure 5 (19–30 min), Figure 7 (14–33 min), Figure 8 (19–20 min), Figure 9 (19–25 min) and Figure 10 (19–33 min). Each peak in the chromatogram (Figure 5, Figure 6, Figure 7, Figure 8, Figure 9 and Figure 10) represents a signal produced when a compound is washed out with a solvent from the gas chromatography column into the detector [115,116].
Numerous small peaks were observed throughout the chromatogram. Major chemical compounds with high peaks and area percentages greater than one were selected and identified in leaf and stem extracts (hexane, chloroform and methanol) (Table 4, Table 5, Table 6, Table 7, Table 8 and Table 9). In certain circumstances, compounds with an area percentage less than one were only mentioned if they played an important role in the genus. The methanolic extracts for both leaf and stem revealed the highest number of compounds. Compounds with an area percentage greater than one were found in the leaf (16) and stem (15) methanolic extracts (Table 6 and Table 9).
For the leaf hexane chromatogram, the highest peak identified was tetratetracontane, which had the highest percentage area of 3.25% (Table 4). This compound displays antioxidant, cytoprotective and anti-inflammatory activities [113,117]. The lowest peak identified in the leaf hexane (Table 4), stem hexane (Table 7), and stem chloroform (Table 8) chromatogram was pentadecanoic acid, with an area percentage of 1.02%, 1.00% and 1.00%, respectively. It should be noted that pentadecanoic acid has not been reported in any species of Barleria; however, this compound is a fatty acid and is found in the milk fat of cows, regulates hormones, improves the immune system and boosts metabolism [118,119].
Octadecanoic acid, 2,3-dihydroxypropyl ester and tetratetracontane displayed the lowest peaks in the leaf chloroform chromatogram, with an area percentage of 1.00% (Table 5). These compounds have not been reported in any species of Barleria, though octadecanoic acid, 2,3-dihydroxypropyl ester displays anticancer and antimicrobial activities (Table 10). In the leaf methanolic chromatogram, flavone, 4′,5-dihydroxy-6,7-dimethoxy-exhibited the highest peak, with an area percentage of 11.69% (Table 6). The 13-docosenamide, (Z) displayed the lowest peak, with an average percentage of 2.46%, and has not been reported in any species of Barleria (Table 6) but was reported to exhibit antimicrobial properties in Ludwigia perennis [120].
The highest peak for the stem hexane chromatogram was 4,4,6a,6b,8a,11,11,14b-Octamethyl-1,4,4a,5,6,6a,6b,7,8,8a,9,10,11,12,12a,14,14a,14b-octadecahy-dro-2H-picen-3-one, which had an average percentage of 2.35 (Table 7). This compound was reported to exhibit antibacterial, antioxidant, antitumour and cancer preventives (Table 10). Additionally, the lowest peak on the stem methanol chromatogram was tributyl acetyl citrate (Table 7). Al-Rubaye et al. [121] examined the methanolic leaf extracts of Sinapis arvensis for its medicinal properties. These authors found tributyl acetyl citrate to display antioxidant and anti-inflammatory activities.
The identified compounds illustrated in Table 10, possessed various biological properties of medicinal importance. Several compounds found in the extracts of B. albostellata were also noted in other species of Barleria. Phyto-compounds such as phenol, 2,4-bis(1,1-dimethylethyl), found in B. albostellata (Table 10), were identified in B. prionitis [122], B. montana [123] and B. lupulina [124]. The 9,12,15-octadecatrienoic acid, (Z,Z,Z) (Table 10) was only prominent in B. buxifolia [125]. Kumari and Dubey [124] reported on octadecanoic acid (Table 10) in the extracts of B. lupulina, while Sriram and Sasikumar [123] found this compound in B. montana. Squalene found in B. albostellata (Table 10) was also identified in B. montana [126], B. longiflora [127], B. courtallica [128], B. lupulina [124] and B. grandiflora [129].
Eicosane, a solid n-alkane (Table 10), was found in extracts of B. courtallica [128], B. prionitis [130] and B. dinteri [10]. In the extracts of B. courtallica [128] and B. lupulina [124], phytol and acetate (Table 10) were identified. Furthermore, phytol (Table 10) was reported in B. montana [126], B. longiflora [127], B. courtallica [128], B. lupulina [124], B. strigosa [131], B. buxifolia [125] and B. prionitis [130]. Vitamin E (Table 10), a fat-soluble vitamin, was only noted in B. courtallica [128]. Flavones, a class of flavonoids, found in the extracts of B. albostellata (Table 10) were also reported in B. prionitis [132] and B. acanthoides [133].
Campesterol found in B. longiflora [134]; stigmasterol in B. courtallica [128], B. montana [123], B. longiflora [127], B. cristata, B. prionitis [135,136] and B. lupulina [137]; and beta-sitosterol identified in B. prionitis [130], B. courtallica [128], B. montana [126] and B. longiflora [127] are three characteristic phytosterols found in B. albostellata (Table 10). Stigmasta-3,5-dien-7-one has only been reported in B. albostellata (Table 10), while 13,14-seco-stigmasta-5,14-diene-3α-o was noted in B. prionitis [138]. Additionally, alpha-amyrin was noted in B. cristata [135] and B. prionitis [130]. Sujatha et al. [128] and Kumari and Dubey [124] reported the presence of 9,12-octdecadienoic acid (Z,Z) (Table 10) in the extracts of B. courtallica and B. lupulina, respectively.
To date, 1-heptacosanol; l-(+)-ascorbic acid 2,6-dihexadecanoate; tridecanoic acid; decanedioic acid, dibutyl ester; 1,2,3,5-cyclohexanetetrol; 1,2-15,16-diepoxyhexadecane; 1,4-benzenedicarboxylic acid, bis(2-ethylhexyl) ester; simiarenol; dichloroacetic acid, tridec-2-ynyl ester; 4,4,6a,6b,8a,11,11,14b-octamethyl-1,4,4a,5,6,6a,6b,7,8,8a,9,10,11,12,12a,14,14a,14b-octadecahydro-2H-picen-3-one; alpha amyrenone; acetic acid, 3-hydroxy-6-isopropenyl-4,8a-dimethyl-1,2,3,4,5,6,7,8; and cholest-4-en-3-one, found in B. albostellata, were not reported in any species of Barleria. Although GC-MS analysis identified the phytochemical constituents present in the hexane, chloroform and methanolic extracts, it should be noted that the most compounds were found in the leaf (Table 6) and stem methanolic (Table 9) extracts.
Table 10. Pharmacological activities of compounds found in B. albostellata.
Table 10. Pharmacological activities of compounds found in B. albostellata.
NoPhytochemical CompoundPharmacological ActionReferences
1Pentadecanoic acidFlavouring agent, lubricants, adhesive agents, ability to regulate hormones, improve the immune system, boost metabolism and inhibit production of uric acid[118,138,139,140,141]
29,12,15-Octadecatrienoic acid, (Z)Antioxidant, anti-inflammatory, antimicrobial, diuretic, anticancer, antitumour, chemo-preventive properties used in vaccine formulations and reduced complications in COVID-19 patients[142,143,144,145]
3Octadecanoic acidAntimicrobial activity[146,147,148]
413-Docosenamide, (Z)Antimicrobial activity[119,149]
5SqualeneCosmetics, skin ointments, antioxidant, antitumour, anticancer, chemo-preventive and sunscreen properties[150,151,152]
6EicosaneAntitumour, antifungal activity and bronchodilators[153,154,155]
71-HeptacosanolNematicidal, anticancer, antioxidant and antimicrobial properties[156,157,158]
8TetratetracontanePlant growth production, antioxidant, cytoprotective and anti-inflammatory activities[113,138,159,160]
9l-(+)-Ascorbic acid 2,6-dihexadecanoateAntioxidant food addictive, antimetastatic, anti-invasive, cancer, cardio protective and anti-infertility[161,162,163]
10Tridecanoic acidAntifungal, antibacterial and larvicidal[164,165,166]
11Decanedioic acid, dibutyl esterAntimicrobial, antispasmodic and anti-inflammatory effects[167]
12Octadecanoic acid, 2,3-dihydroxypropyl esterAnticancer, antimicrobial, acidifier, acidulant, arachidonic acid inhibitor and inhibits production of uric acid[139,141]
131,2,3,5-CyclohexanetetrolAntioxidant, antimicrobial and anti-inflammatory properties[168]
14Phytol, acetateAnti-inflammatory, antileishmanial, anti-trypanosomal, antimicrobial, anticancer and diuretic[32,169,170,171]
15n-Nonadecanol-1Antimicrobial and cytotoxic properties[172,173]
16PhytolAnticancer, antimicrobial, anti-inflammatory, antioxidant activity, diuretic, cosmetics and used in the fragrance industry[141,174]
171,2-15,16-DiepoxyhexadecaneAntitumour and anti-inflammatory properties[175]
181,4-Benzenedicarboxylic acid, bis(2-ethylhexyl) esterAnticancer properties[176,177]
19Vitamin ESkin repair, enhancing the immune system and has anticancer, antitumour and antioxidant properties[151,178,179]
20FlavoneAntibacterial, antimutagenic, antiviral and antioxidant activity[180,181,182]
21CampesterolAnti-inflammatory and anticancer activity[183,184]
22StigmasterolAnti-inflammatory, anti-asthma, anticancerous, anti-inflammatory, antiarthritic, hypoglycemic, antioxidant and thyroid-inhibiting properties, analgesic, antiosteoarthritic and antimutagenic activity[141,185,186]
23Beta-SitosterolReduces cholesterol levels, androgen blocker, anti-amyloid beta and anticancer properties[141,187]
24Alpha-AmyrinAlpha amylase and glucosidase inhibitor, antioxidant, antibacterial and anti-inflammatory properties[151,188]
25SimiarenolAntinociceptive activity[189]
269,12-Octdecadienoic acid (Z,Z)-Anti-inflammatory, antibacterial, antiarthritic, hepatoprotective, anti-histaminic, anticoronary and anticancer properties[139,190,191]
27Dichloroacetic acid, tridec-2-ynyl esterCosmetic treatments, anticancer, antimicrobial, antioxidant activity[192,193]
284,4,6a,6b,8a,11,11,14b-Octamethyl-1,4,4a,5,6,6a,6b,7,8,8a,9,10,11,12,12a,14,14a,14b-octadecahydro-2H-picen-3-oneAntibacterial, antioxidant, antitumour and cancer preventives[194,195]
29Phenol, 2,4-bis(1,1-dimethylethyl)-Antibacterial and anti-inflammatory activities[196]
30Tributyl acetylcitrateAnticancer and antimicrobial activities[197,198,199]
319-OctadecenamideAntimicrobial activity[148]
32Alpha. AmyrenoneAntibacterial and antimalarial activities[200,201]
33Acetic acid, 3-hydroxy-6-isopropenyl-4,8a-dimethyl-1,2,3,4,5,6,7,8Antimicrobial activity[202,203]
34Stigmasta-3,5-dien-7-oneAnti-diabetic and anticancer properties, free-radical scavenging activity[204,205,206]
35Cholest-4-en-3-oneAnti-obesity and an intestinal metabolite of cholesterol[207,208]

2.4. Antibacterial Activity of Leaf and Stem Extracts of B. albostellata

The current interest in herbal plants as therapeutic agents has increased in several parts of the world. This is due to the ever-increasing occurrence of drug-resistant bacteria and the influx of new pathogenic bacterial strains. Active phytochemicals found in hexane, chloroform and methanolic extracts of B. albostellata were subjected to antibacterial assays. Various concentrations (3.125, 6.25, 12.25, 25, 50 and 100 mg/mL) were tested against the Gram-positive (B. subtillus, methicillin-resistant S. aureus and S. aureus) and Gram-negative (E. coli and P. aeruginosa) bacteria. The zone of inhibition of the growth of bacteria was used to evaluate the antibacterial potential of the various extracts. Results presented in Table 11 of certain leaf and stem extracts showed significant inhibition compared to streptomycin and gentamicin (positive controls) (Table 11). The screening was done in triplicate with streptomycin (Gram-positive) and gentamicin (Gram-negative) used as the standard antibacterial positive controls, and 10% DMSO without plant extracts was used as the negative control. Clear zones of inhibition were observed in the leaf and stem crude extracts against the various strains. Significance was established at p < 0.05 for all concentrations and treatments. The various extracts and concentrations displayed a variable degree of bacterial growth against various bacterial strains.
As the concentration increased (3.125, 6.25, 12.25, 25, 50 and 100 mg/mL), the zone of inhibition against various bacterial strains also increased. Inhibition against the various bacterial strains for the various extracts was noted at concentrations > 25 mg/mL. The highest inhibitory activity was observed at 100 mg/mL for both leaf and stem extracts for B. subtillus and S. aureus. MRSA, E. coli and P. aeruginosa were resistant to the leaf hexane extracts, whilst the stem hexane extracts displayed no inhibition against E. coli and P. aeruginosa only (Table 11). Gram-positive and Gram–negative bacteria were resistant to all extracts at both 3.125 and 6.25 mg/mL concentrations. Amoo et al. [57] verified the antibacterial activity of B. albostellata against B. subtillis, S. aureus and E. coli. However, low activity was observed against Gram-negative bacteria [57]. Matu and Van Staden [209] suggested that a thick murein layer present in the structure of Gram-negative bacteria may prevent the entry of inhibitors. The differences in the bacterial inhibition varied for each crude extract. The leaf methanolic extracts at 100 mg/mL displayed the highest inhibition against all tested bacterial strains. However, the stem methanolic extracts had the highest inhibition against S. aureus and P. aeruginosa only.
Several notable bioactive compounds found in the leaf and stem extracts of B. albostellata using GC-MS analysis were reported to display antibacterial efficiency. The presence of phytol and flavone found in the leaf methanolic extracts of B. albostellata could be responsible for the antibacterial effects against the several tested strains. Phytol was reported to severely damage the deoxyribonucleic acid (DNA) of bacteria by inducing oxidative stress [210]. The presence of flavonoids blocks important enzymes that play a significant role in the reproduction, growth, cell rupture or functional modification in bacteria [211]. Stigmasterol, another compound found in the leaf methanolic extract, was reported to act as a lactamase inhibitor, which prevented antibacterial resistance by restoring the vulnerability of the antibiotic resistant bacteria to antibiotics [212]. The mode of action used by most bioactive compounds in treating microbial infections is by interacting with the microbial enzyme system, interfering with nucleic acids, the cell wall and the cell membrane [213,214,215].
Additionally, the antibacterial efficiency in the various extracts may be due to greater solubility of phyto-compounds in polar solvents than non-polar solvents [123]. It was recommended that the inability of plant extracts of other solvent systems to display antibacterial activity against the various bacterial strains could be due to these strains exhibiting some sort of resistance mechanism, e.g., alteration of target sites, enzyme inactivation, reduced drug accumulation or the amount of bioactive compounds present is very low [216]. Extracts of B. acuminata [69], B. cristata [217], B. greenii [57], B. prionitis [218] and B. montana [125] exhibited antibacterial activity against B. subtillis and S. aureus. However, B. cristata displayed low inhibition against E. coli [217], and B. montana [125] displayed moderate activity against E.coli and P. aeruginosa. According to Kumari and Dubey [123], ethanolic leaf extracts of B. lupulina inhibited the growth of E. coli, S. aureus and P. aeruginosa, whereas methanolic extracts displayed zones of inhibition against S. aureus and no inhibition against E. coli and P. aeruginosa [219]. Various medicinal plant extracts were reported to display greater activity against Gram-positive bacteria as opposed to Gram-negative bacteria [57,220,221]. The positive results in the present study could be attributed to the location of collected plant material, active constituents found in the different extracts, various extraction preparation, the broad range of treatment concentrations and the variety of bacterial strains. Therefore, the active constituents found in the different extracts of this plant were effective against both Gram-positive and -negative bacteria.
Qualitative phytochemical screening and GC-MS revealed various biologically active compounds which have been known to contain diverse activities that may help protect against chronic diseases [66,67]. Several compounds found in the extracts of B. albostellata were also noted in other species of Barleria [124,131,137,138,140,141]. Additionally, the phyto-constituents found in the leaf and stem crude extracts could inhibit the growth of various pathogenic strains. Various medicinal plant extracts were reported to display greater activity against Gram-positive bacteria as opposed to Gram-negative bacteria [57,220,221]. The antibacterial efficiency in the various extracts may be due to greater solubility of phyto-compounds in polar solvents than non-polar solvents [136].

3. Materials and Methods

3.1. Plant Materials

Leaves and stems of B. albostellata were collected from the University of KwaZulu-Natal, Westville campus (29°49′51.6″ S, 30°55′30″ E), Durban, South Africa. A voucher specimen (7973000) was deposited in the Ward Herbarium of the University of KwaZulu-Natal, Life Sciences, Westville campus.

3.2. Organoleptic Evaluation

The evaluation of crude leaf and stem material was completed with the aid of sensory organs following standard methods [221]. This protocol uses colour, odour, taste and texture. Organoleptic assessment is accomplished using organs of sense and describing specific features of the material. This assessment is regarded as a first step towards establishing the identity and degree of purity of the sample [222].

3.3. Preparation of Crude Extract

For the preparation of the crude extract, leaves and stems were oven-dried for 2 weeks at 35°. The dried materials were crushed to a fine powder with the aid of a mechanical blender (Russel Hobbs, model: RHB315). The powdered material underwent sequential extraction using various solvents (hexane, chloroform and methanol) in a Soxhlet apparatus. Approximately 10 g of powdered leaves were placed into a round-bottom flask containing 100 mL of hexane, the appropriate solvent, and boiled for 3 h at 40 °C. The extracted solution was filtered (Whatman® No. 1 filter paper) and retained. This procedure was conducted in replicates. Consecutive extractions of chloroform followed by methanol were achieved. Each solvent extraction followed the same process as mentioned above. Successive extractions were performed on the leaf and stem material.

Evaporation and Concentration

The concentration of each extract was left to evaporate in a dark fume-hood, at room temperature. The dried extracts were stored in airtight, labelled glass jars to prevent the material from reacting with the atmospheric humidity. The percentage yield of each extract was calculated using the following equation:
Extract   Yield   ( % ) = Weight   of   dried   extract   ( g ) Weight   of   plant   material   ( g ) × 100

3.4. Qualitative Phytochemical Analysis

Preliminary phytochemical screening was carried out on the powdered leaf and stem material and chemically tested for the presence of various constituents using standard protocols [223,224,225,226].

3.5. Fluorescence Analysis

Fluorescent analysis of the dried powdered plant material plays an important role in the determining the quality and purity of the tested drug. A small quantity of the dry plant powder (leaves and stems) was placed separately onto clean microscope slides. Two drops of each prepared reagent were dispensed, mixed gently by slanting the slide and allowed to stand for 3 min for the thorough absorption of the solution by the plant powder. The slides were then viewed using a Nikon Eclipse microscope, using bright field light and UV-2A (excitation 320/380) illumination. The colours attained by the application of various reagents were recorded. Fluorescence analysis of the leaf and stem powder was carried out using the standard method [227,228].

3.6. Gas Chromatography-Mass Spectrometry (GC-MS)

This analysis is used to examine liquid, gaseous or solid samples and produce several different peaks in the gas chromatogram. Each peak generates a specific mass spectrum which is used for compound identification. Leaf and stem methanolic extracts were analysed using the GC-MS (QP-2010 Ultra SE, Shimadzu, Kyoto, Japan) instrument, with an Rx_5Sil Ms capillary column (0.25 μm internal diameter and 0.25 μm film thickness) from Restek (Bellefonte, PA, USA). The carrier gas, helium, had a flow rate of 0.96 mL/min, a total flow of 4.9 mL/min and a linear velocity of 36.7 cm/sec at a purge flow of 3.0 mL/min. The injection temperature was set at 250 °C. The oven temperature was set at 50 °C and held for 1 min, increased to 310 °C and held for a further 10 min. Chemical compounds (analytes) were identified by relating their retention times with those of the polychlorinated biphenyl (PCB) standards found in the National Institute of Standards and Technology (NIST) library. This analysis was conducted at the Department of Chemistry at the University of KwaZulu-Natal, Westville campus.

3.7. Antibacterial Bioassay

Crude (hexane, chloroform and methanol) leaf and stem extracts were transferred to Eppendorf centrifuge tubes, dissolved in 10% dimethyl sulfoxide (DMSO) at various concentrations of 100, 50, 25, 12.5, 6.25 and 3.125 mg/mL and homogenised using a vortex. The prepared sample was stored at −4 °C until further use. The prepared crude extracts were subjected to antibacterial assays. Leaf and stem samples were tested against Gram-positive bacteria (Bacillus subtillus ATCC 6633, methicillin-resistant Staphylococcus aureus ATCC 43300 and Staphylococcus aureus ATCC 25923) and Gram-negative bacteria (Pseudomonas aeruginosa ATCC 25783 and Escherichia coli ATCC 35218). These bacterial strains were supplied by Professor Johnson Lin, School of Life Sciences (Microbiology Department), University of KwaZulu-Natal, and maintained in 75% glycerol at −80 °C before the experiment was conducted.
In vitro antibacterial screening of the prepared extracts was conducted using the agar disc diffusion technique as per the Clinical and Laboratory Standards Institute (CLSI) guidelines [229]. Both Gram-positive and Gram-negative bacteria from stock cultures were sub-cultured onto fresh agar plates and incubated overnight at 37 °C. Glass test tubes containing distilled water (15–20 mL) were autoclaved at 121 °C for 1 h. Colonies of bacteria from each Petri plate were harvested with a sterile loop and inoculated by transferring a loopful into glass test tubes containing 15 mL of sterile distilled water (0.5 McFarland scale). The absorbance of each bacterial culture was measured, adjusted and diluted to attain a viable cell count using the Cary 60 UV-Vis spectrophotometer.
Each bacterial strain was separately smeared uniformly over the surface of the Mueller–Hinton agar plates with a sterile cotton swab. Sterile Whatman filter paper No. 1 discs (diameter 6 mm) were impregnated with 20 μL of the respective extract concentrations (3.125, 6.25, 12.5, 25, 50, 100 mg/mL) and dried at room temperature for 1 h before use [230]. The prepared sterile discs containing extracts were placed carefully onto the agar using sterile forceps. Petri plates were sealed and incubated for 24 h at 37 °C. Zones of inhibition evident around the filter paper were taken as positive results. The diameters of inhibition were measured and photographed within 18–24 h after incubation to determine if the extract exhibited any antibacterial activity. Filter paper discs loaded with streptomycin and gentamycin were used as positive controls and 10% DMSO as the negative control [231]. The analyses were conducted in triplicate, and data were presented as mean ± standard deviation.

3.8. Statistical Analysis

All experiments conducted for the antibacterial assay were carried out in triplicate. Values were expressed as mean ± standard deviation (significant at p < 0.05 level). Antibacterial data were statistically analysed using the one-way analysis of variance (ANOVA).

4. Conclusions

It is evident from the present study that the qualitative colour tests, fluorescence and GC-MS analysis that the leaves and stems of B. albostellata possess biologically active compounds. Important compounds identified in leaf and stem extracts of B. albostellata were alkaloids, flavonoids and phenols. These compounds are known to display several diverse activities that may help protect against chronic diseases. Major pharmacologically active compounds found in the extracts were alpha-amyrin, flavone, phenol, phytol, phytol acetate, squalene and stigmasterol. Additionally, phyto-constituents found in the hexane, chloroform and methanol leaf and stem extracts of B. albostellata could inhibit the growth of various pathogenic strains (p < 0.05). Other solvents such as ethanol and acetone can be used in extracting phytochemical compounds from the leaves and stems. These extractions can be subjected to antibacterial assays in order to evaluate its potency against various pathogenic strains. Further studies should be conducted on the isolation, identification and characterisation of the bioactive compounds in B. albostellata that may be responsible for its bioactivity. This is important to further understand the mechanisms involved in the antibacterial activity. The bioactive compounds and pharmacological activities of B. albostellata will provide a basic understanding of the importance of this species as a medicinal plant and a potential source for novel and useful drugs. Additionally, other parts of the plant such as the flowers and roots should assessed for their safety and bioactivity and to identify any new therapeutic compounds or drug leads.

Author Contributions

Conceptualization and methodology, S.G.; investigation and data curation, S.G., Y.N., M.S. and J.L.; validation, S.G., Y.N., Y.H.D. and M.S.; writing—original draft preparation, S.G., Y.N. and M.S.; writing—review and editing, Y.H.D. and H.N.M.; supervision, Y.N., Y.H.D. and M.S. All authors have read and agreed to the published version of the manuscript.

Funding

Authors extend their appreciation to the National Research Foundation (Grant No. 118897) and the Researchers Supporting Project number (RSP2023R375), King Saud University, Riyadh, Saudi Arabia for funding this work.

Data Availability Statement

Not applicable.

Acknowledgments

Authors extend their appreciation to the National Research Foundation (Grant No. 118897), South Africa, and the staff at the Microscopy and Microanalysis Unit at the University of KwaZulu-Natal for their assistance with the microscopy components of the research. The authors gratefully acknowledge Researchers Supporting Project number (RSP2023R375), King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Farnsworth, N.O. The role of medicinal plants in drug development. In Natural Products and Drug Development; Krogsgaard-Larsen, P., Christensen, S.B., Kofod, H., Eds.; Balliere, Tindall and Cox: London, UK, 1984; pp. 8–98. [Google Scholar]
  2. Hosseinzadeh, S.; Jafarikukhdan, A.; Hosseini, A.; Armand, R. The application of medicinal plants in traditional and modern medicine: A review of Thymus vulgaris. Int. J. Clin. Med. 2015, 6, 635–642. [Google Scholar] [CrossRef] [Green Version]
  3. Bhat, B.B.; Udupa, N.; Sreedhar, D. Herbal products regulations in a few countries-a brief overview. Curr. Drug Discov. Technol. 2019, 16, 368–371. [Google Scholar] [CrossRef]
  4. WHO Traditional Medicine Strategy 2014–2023; WHO: Geneva, Switzerland. Available online: http://www.who.int/medicines/publications/traditional/trm_strategy14_23/en/ (accessed on 28 October 2020).
  5. Gurib-Fakim, A. Medicinal plants: Traditions of yesterday and drugs of tomorrow. Mol. Asp. Med. 2006, 27, 1–93. [Google Scholar] [CrossRef] [PubMed]
  6. Aiswarya, T.; Ravikumar, R. A comparative study on phytochemical analysis, antibacterial activity and antioxidant activity of Barleria prionitis leaves extract of petroleum ether and ethanol extract. Int. J. Chemtech Res. 2014, 6, 3025–3033. [Google Scholar]
  7. Eddouks, M.; Ajebli, M.; Hebi, M. Ethnopharmacological survey of medicinal plants used in Daraa-Tafilalet region (Province of Errachidia), Morocco. J. Ethnopharmacol. 2017, 198, 516–530. [Google Scholar] [CrossRef] [PubMed]
  8. Croteau, R.; Kutchan, T.M.; Lewis, N.G. Natural Products (Secondary Metabolites). In Biochemistry & Molecular Biology of Plants; Buchanan, B.B., Gruissem, W., Jones, R.L., Eds.; American Society of Plant Physiologists: Rockville, MD, USA, 2000. [Google Scholar]
  9. Salehi, B.; Albayrak, S.; Antolak, H.; Kregiel, D.; Pawlikowska, E.; Sharifi-Rad, M.; Uprety, Y.; Fokou, P.V.T.; Yousef, Z.; Zakaria, Z.A.; et al. Aloe genus plants: From farm to food applications and phytopharmacotherapy. Int. J. Mol. Sci. 2018, 19, 2843. [Google Scholar] [CrossRef] [Green Version]
  10. Semenya, C.; Maseko, R.; Gololo, S. Comparative qualitative phytochemical analysis of the different parts of Barleria dinteri (Oberm): A contribution to sustainable use of the plant species. J. Pharm. Chem. Biol. Sci. 2018, 6, 52–59. [Google Scholar]
  11. Alagawany, M.; Elnesr, S.S.; Farag, M.R.; Abd El-Hack, M.E.; Khafaga, A.F.; Taha, A.E.; Tiwari, R.; Yatoo, M.I.; Bhatt, P.; Marappan, G.; et al. Use of licorice (Glycyrrhiza glabra) herb as a feed additive in poultry: Current knowledge and prospects. Animals 2019, 9, 536. [Google Scholar] [CrossRef] [Green Version]
  12. Sadgrove, N.J. Southern Africa as a ‘cradle of incense’in wider African aromatherapy. Sci. Afr. 2020, 9, e00502. [Google Scholar]
  13. Van Vuuren, S.; Muhlarhi, T. Do South African medicinal plants used traditionally to treat infections respond differently to resistant microbial strains? S. Afr. J. Bot. 2017, 112, 186–192. [Google Scholar] [CrossRef]
  14. Nkengasong, J.N.; Sofonais, K.T. Africa needs a new public health order to tackle infectious disease threats. Cell 2020, 2, 296–300. [Google Scholar] [CrossRef] [PubMed]
  15. Elbashiti, T.A.; Elmanama, A.A.; Masad, A.A. The antibacterial and synergistic effects of some Palestinian plant extracts on Escherichia coli and Staphylococcus aureus. Funct. Plant Sci. Biotechnol. 2011, 5, 57–62. [Google Scholar]
  16. Ncube, B.; Finnie, J.; Van Staden, J. In vitro antimicrobial synergism within plant extract combinations from three South African medicinal bulbs. J. Ethnopharmacol. 2012, 139, 81–89. [Google Scholar] [CrossRef] [PubMed]
  17. Islam, R.; Rahman, M.S.; Rahman, S.M. GC-MS analysis and antibacterial activity of Cuscuta reflexa against bacterial pathogens. Asian Pac. J. Trop. Dis. 2015, 5, 399–403. [Google Scholar] [CrossRef]
  18. Prasad, P.; Gupta, A.; Sasmal, P.K. Aggregation-induced emission active metal complexes: A promising strategy to tackle bacterial infections. Chem. Commun. 2021, 57, 174–186. [Google Scholar] [CrossRef] [PubMed]
  19. Wise, R. The worldwide threat of antimicrobial resistance. Curr. Sci. 2008, 95, 181–187. [Google Scholar]
  20. Walsh, F. Superbugs to Kill “More Than Cancer” by 2050. 2014. Available online: https://www.bbc.com/news/health-30416844 (accessed on 6 December 2019).
  21. World Health Organization. New Report Calls for Urgent Action to Avert Antimicrobial Resistance Crisis. 2019. Available online: https://www.who.int/news-room/detail/29-04-2019-new-report-calls-for-urgent-action-toavert-antimicrobial-resistance-crisis (accessed on 12 September 2019).
  22. Yang, S.K.; Low, L.Y.; Yap, P.S.X.; Yuso, K.; Mai, C.W.; Lai, K.S.; Lim, S.H.E. Plant-Derived antimicrobials: Insights into mitigation of antimicrobial resistance. Rec. Nat. Prod. 2018, 12, 295–316. [Google Scholar] [CrossRef]
  23. Rabe, T.; Van Staden, J. Antibacterial activity of South African plants used for medicinal purposes. J. Ethnopharmacol. 1997, 56, 81–87. [Google Scholar] [CrossRef]
  24. Buwa, L.V.; Van Staden, J. Antibacterial and antifungal activity of traditional medicinal plants used against venereal diseases in South Africa. J. Ethnopharmacol. 2006, 103, 139–142. [Google Scholar] [CrossRef]
  25. Ekor, M. The growing use of herbal medicines: Issues relating to adverse reactions and challenges in monitoring safety. Front. Pharmacol. 2014, 4, 177. [Google Scholar] [CrossRef] [Green Version]
  26. Yuan, H.; Ma, Q.; Ye, L.; Piao, G. The traditional medicine and modern medicine from natural products. Molecules 2016, 21, 559. [Google Scholar] [CrossRef] [Green Version]
  27. Van Wyk, B.E.; Wink, M. Medicinal Plants of the World; Briza Publications: Pretoria, South Africa, 2004; p. 480. [Google Scholar]
  28. Nafiu, M.O.; Hamid, A.A.; Muritala, H.F.; Adeyemi, S.B. Preparation, standardization, and quality control of medicinal plants in Africa. In Medicinal Spices and Vegetables from Africa; Kuete, V., Ed.; Academic Press: Cambridge, MA, USA, 2017; pp. 171–204. [Google Scholar]
  29. Cowan, M.M. Plant products as antimicrobial agents. Clin. Microbiol. Rev. 1999, 12, 564–582. [Google Scholar] [CrossRef] [Green Version]
  30. Lewis, K.; Ausubel, F.M. Prospects for plant derived antibacterials. Nat. Biotechnol. 2006, 24, 1504–1507. [Google Scholar] [CrossRef] [PubMed]
  31. Al-Tameme, H.J.; Hadi, M.Y.; Hameed, I.H. Phytochemical analysis of Urtica dioica leaves by fourier-transform infrared spectroscopy and gas chromatography-mass spectrometry. J. Pharmacogn. Phytother. 2015, 7, 238–252. [Google Scholar] [CrossRef] [Green Version]
  32. Al-Marzoqi, A.H.; Hadi, M.Y.; Hameed, I.H. Determination of metabolites products by Cassia angustifolia and evaluate antimicrobial activity. J. Pharmacogn. Phytother. 2016, 8, 25–48. [Google Scholar] [CrossRef] [Green Version]
  33. Hadi, M.Y.; Mohammed, G.J.; Hameed, I.H. Analysis of bioactive chemical compounds of Nigella sativa using gas chromatography-mass spectrometry. J. Pharmacogn. Phytother. 2016, 8, 8–24. [Google Scholar]
  34. Umah, C.; Dorly; Sulistyaringsih, Y.C. Secretory structure, histochemistry and phytochemistry analyses of stimulant plant. IOP Conf. Ser. Earth Environ. Sci. 2017, 58, 012–048. [Google Scholar] [CrossRef]
  35. Davies, K.M.; Jibran, R.; Zhou, Y.; Albert, N.W.; Brummell, D.A.; Jordan, B.R.; Bowman, J.L.; Schwinn, K.E. The evolution of flavonoid biosynthesis: A bryophyte perspective. Front. Plant Sci. 2020, 11, 7. [Google Scholar] [CrossRef] [Green Version]
  36. Hatcher, C.R.; Ryves, D.B.; Millett, J. The function of secondary metabolites in plant carnivory. Ann. Bot. 2020, 125, 399–411. [Google Scholar] [CrossRef] [Green Version]
  37. Van Wyk, B.E.; Gericke, N. People’s Plants: A Guide to Useful Plants of Southern Africa; Briza Publications: Pretoria, South Africa, 2000; p. 351. [Google Scholar]
  38. Van Wyk, B.E. The potential of South African plants in the development of new medicinal products. S. Afr. J. Bot. 2011, 77, 812–829. [Google Scholar] [CrossRef] [Green Version]
  39. Van Wyk, B.E. A broad review of commercially important southern African medicinal plants. J. Ethnopharmacol. 2008, 119, 342–355. [Google Scholar] [CrossRef]
  40. Amoo, S.O.; Aremu, A.O.; Van Staden, J. Unraveling the medicinal potential of South African Aloe species. J. Ethnopharmacol. 2014, 153, 19–41. [Google Scholar] [CrossRef]
  41. Fabricant, D.S.; Farnsworth, N.R. The value of plants used in traditional medicine for drug discovery. Environ. Health Perspect. 2001, 109, 69–75. [Google Scholar]
  42. Shen, B. A new golden age of natural products drug discovery. Cell 2015, 163, 1297–1300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Cragg, G.M.; Newman, D.J.; Snader, M. Natural products in drug discovery and development. J. Nat. Prod. 1997, 60, 52–60. [Google Scholar] [CrossRef]
  44. Rybicki, E.P.; Chikwamba, R.; Koch, M.; Rhodes, J.I.; Groenewald, J.H. Plant-made therapeutics: An emerging platform in South Africa. Biotechnol. Adv. 2012, 30, 449–459. [Google Scholar] [CrossRef] [PubMed]
  45. Newman, D.J.; Cragg, G.M. Natural products as sources of new drugs from 1981 to 2014. J. Nat. Prod. 2016, 79, 629–661. [Google Scholar] [CrossRef] [Green Version]
  46. Cragg, G.M.; Newman, D.J. Drug from nature: Past achievement, future prospect. Adv. Phytomed. 2007, 1, 23–37. [Google Scholar]
  47. Wangchuk, P. Therapeutic applications of natural products in herbal medicines, biodiscovery programs, and biomedicine. J. Biol. Act. Prod. Nat. 2018, 8, 1–20. [Google Scholar] [CrossRef]
  48. Vidhya, R.; Udayakumar, R. Gas chromatography-Mass spectrometry (GC-MS) analysis of ethanolic extracts of Aerva lanata (L.). Int. J. Biochem. Res. Rev. 2015, 7, 192–203. [Google Scholar] [CrossRef]
  49. Froneman, W.; Le Roux, L.N. Barleria albostellata. 2007. Available online: http://pza.sanbi.org/barleria-albostellata (accessed on 2 February 2019).
  50. Yosook, C.; Panpisutchai, Y.; Chaichana, S.; Santisuk, T.; Reutrakul, V. Evaluation of anti-HSV-2 activities of Barleria lupulina and Clinacanthus nutans. J. Ethnopharmacol. 1999, 67, 179–187. [Google Scholar] [CrossRef] [PubMed]
  51. Wang, B.U.; Wu, M.; Perchellet, E.M.; Mcilvain, C.J.; Sperfslage, B.J.; Huang, X.; Tamura, M.; Stephany, H.A.; Hua, D.H.; Perchellet, J.P. Asynthetic triptycene bisquinone which blocks nucleoside transport and induces DNA fragmentation, retains its cytotoxic efficacy in daunorubicin-resistant HL-60 cell lines. Int. J. Oncol. 2001, 19, 1169–1178. [Google Scholar] [PubMed]
  52. Jassim, S.A.A.; Naji, A.M. Novel antiviral agents: A medicinal plant perspective. J. Appl. Microbiol. 2003, 95, 412–427. [Google Scholar] [CrossRef] [Green Version]
  53. Suba, V.; Murugesan, T.; Arunachalam, G.; Mandal, S.C.; Saha, B.P. Anti-diabetic potential of Barleria lupulina extract in rats. Phytomedicine 2004, 11, 202–205. [Google Scholar] [CrossRef] [PubMed]
  54. Suba, V.; Murugesan, T.; Kumaravelrajan, R.; Mandal, S.C.; Saha, B.P. Antiinflammatory, analgesic and antiperoxidative efficacy of Barleria lupulina Lindl. extract. Phytother. Res. 2005, 19, 695–699. [Google Scholar] [CrossRef]
  55. Chomnawang, M.T.; Surassmo, S.; Nukoolkarn, V.S.; Gritsanapan, W. Antimicrobial effects of Thai medicinal plants against acne-inducing bacteria. J. Ethnopharmacol. 2005, 101, 330–333. [Google Scholar] [CrossRef]
  56. Amoo, S.O.; Ndhlala, A.R.; Finnie, J.F.; Van Staden, J. Antifungal, acetylcholinesterase inhibition, antioxidant and phytochemical properties of three Barleria species. S. Afr. J. Bot. 2011, 77, 435–445. [Google Scholar] [CrossRef] [Green Version]
  57. Amoo, S.O.; Finnie, J.F.; Van Staden, J. In vitro pharmacological evaluation of three Barleria species. J. Ethnopharmacol. 2009, 121, 274–277. [Google Scholar] [CrossRef]
  58. Selvam, A.B.D. Standardization of organoleptic terminology with reference to description of vegetable crude drugs. Int. J. Pharm. Technol. 2015, 7, 3282–3289. [Google Scholar]
  59. Shaheen, S.; Ramzan, S.; Haroon, N.; Hussain, K. Ethnopharmacological and systematic studies of selected medicinal plants of Pakistan. Pak. J. Sci. 2014, 66, 175–180. [Google Scholar]
  60. Abubakar, E.M.; Misau, S.; Modibbo, S.; Bala, G.L. Percentage yield and acute toxicity of the plant extracts of Ceiba pentandra grown in Bauchi State, North Eastern Nigeria. J. Pharmacogn. Phytochem. 2017, 6, 1777–1779. [Google Scholar]
  61. Chintalapani, S.A.T.H.V.I.K.A.; Swathi, M.S.; Mangamoori, L.N. Phytochemical screening and in vitro antioxidant activity of whole plant extracts of Sesuvium portulacastrum L. Asian J. Pharm. Clin. Res. 2018, 11, 322–327. [Google Scholar] [CrossRef] [Green Version]
  62. Murugan, T.; Wins, J.A.; Murugan, M. Antimicrobial activity and phytochemical constituents of leaf extracts of Cassia auriculata. Indian J. Pharm. Sci. 2013, 75, 122–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Wink, M. Plant secondary metabolites modulate insect behavior-steps toward addiction? Front. Physiol. 2018, 9, 364. [Google Scholar] [CrossRef] [Green Version]
  64. Yactayo-Chang, J.P.; Tang, H.V.; Mendoza, J.; Christensen, S.A.; Block, A.K. Plant defense chemicals against insect pests. Agronomy 2020, 10, 1156. [Google Scholar] [CrossRef]
  65. Liu, R.H. Health benefits of fruit and vegetables are from additive and synergistic combinations of phytochemicals. Am. J. Clin. Nutr. 2003, 78, 517–520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Singh, N.; Sharma, B. Phytochemicals as therapeutics in heavy metal toxicity. In Advances in Pharmaceutical Biotechnology; Patra, J.K., Shukla, A.C., Das, G., Eds.; Springer: New York, NY, USA, 2020; pp. 91–100. [Google Scholar]
  67. Didna, B.; Debnath, S.; Harigaya, Y. Naturally occurring iridoids. A review, Part 1. Chem. Pharm. Bull. 2007, 55, 159–222. [Google Scholar]
  68. Tundis, R.; Loizzo, M.R.; Menichini, F.; Statti, G.A.; Menichini, F. Biological and pharmacological activities of iridoids: Recent developments. Mini-Rev. Med. Chem. 2008, 8, 399–420. [Google Scholar] [CrossRef]
  69. Bency, A.; Lohidas, J.; Murugan, M. Phytochemical studies and antibacterial activity of Barleria acuminata Nees (Acanthaceae). J. Pharmacogn. Hytochem. 2018, 7, 1909–1911. [Google Scholar]
  70. El-Mawla, A.M.A.A.; Ahmed, A.S.; Ibraheim, Z.Z.; Ernst, L. Phenylethanoid glycosides from Barleria cristata L. callus cultures. Bull. Pharm. Sci. 2005, 28, 199–204. [Google Scholar] [CrossRef]
  71. Hemalatha, K.; Hareeka, N.; Sunitha, D. Chemical constituents isolated from leaves of Barleria cristata Linn. Int. J. Pharm. Biol. Sci. 2012, 3, 609–615. [Google Scholar]
  72. Choudhary, M.; Kumar, V.; Gupta, P.K.; Singh, S. Anti-arthritic activity of Barleria prionitis Linn. leaves in acute and chronic models in Sprague Dawley rats. Bull. Fac. Pharm. 2014, 52, 199–209. [Google Scholar] [CrossRef] [Green Version]
  73. Baskaran, A.; Karthikeyan, V. Preliminary phytochemical analysis of Barleria longiflora L.f. using different solvent. World Sci. News 2019, 124, 319–325. [Google Scholar]
  74. Shukla, S.; Gunjegaokar, S.M. Pharmacognostical and pharmacological profiling of Barleria prionitis Linn. J. Biol. Sci. Med. 2018, 4, 41–50. [Google Scholar]
  75. Mahato, S.B.; Sen, S. Advances in triterpenoid research, 1990–1994. Phytochemistry 1997, 44, 1185–1236. [Google Scholar] [CrossRef] [PubMed]
  76. Dahanukar, S.A.; Kulkarni, R.A.; Rege, N.N. Pharmacology of medicinal plants and natural products. Indian J. Pharmacol. 2000, 32, 81–118. [Google Scholar]
  77. Russo, P.; Frustaci, A.; Del Bufalo, A.; Fini, M.; Cesario, A. Multitarget drugs of plants origin acting on Alzheimer’s disease. Curr. Med. Chem. 2013, 20, 1686–1693. [Google Scholar] [CrossRef]
  78. Usman, H.; Abdulrahman, F.I.; Ahmed, I.A.; Kaita, A.H.; Khan, I.Z. Antibacterial effects of cyanogenic glucoside isolated from the stem bark of Bauhinia rufescens Lam. Int. J. Biol. Chem. Sci. 2013, 7, 2139–2150. [Google Scholar] [CrossRef] [Green Version]
  79. Cushnie, T.P.; Cushnie, B.; Lamb, A.J. Alkaloids: An overview of their antibacterial, antibiotic-enhancing and antivirulence activities. Int. J. Antimicrob. Agents 2014, 44, 377–386. [Google Scholar] [CrossRef]
  80. Kaur, N.; Ahmed, T. Bioactive secondary metabolites of medicinal and aromatic plants and their disease-fighting properties. In Medicinal and Aromatic Plants: Healthcare and Industrial Applications; Aftab, T., Rehman, K.H., Eds.; Springer: Jeddah, Saudi Arabia, 2021; p. 113. [Google Scholar]
  81. Paul, R.; Gayathri, R.; Vishnu Priya, V. Preliminary phytochemical analysis and estimation of total phenol content in carrot extract. Int. J. Pharm. Sci. Rev. Res. 2017, 45, 34–36. [Google Scholar]
  82. Olsson, M.E.; Ekvall, J.; Gustavsson, K.E.; Nilsson, J.; Pillai, D.; Sjöholm, I.; Svensson, U.; Åkesson, B.; Nyman, M.G. Antioxidants, low molecular weight carbohydrates, and total antioxidant capacity in strawberries: Effects of cultivar, ripening, and storage. J. Agric. Food Chem. 2004, 52, 2490–2498. [Google Scholar] [CrossRef]
  83. Agoramoorthy, G.; Chandrasekaran, M.; Venkatesalu, H.M.J. Antibacterial and antifungal activities of fatty acid methyl esters of the blind-your-eye mangrove from India. Braz. J. Microbiol. 2007, 38, 739–742. [Google Scholar] [CrossRef] [Green Version]
  84. Dalle-Donne, I.; Rossi, R.; Colombo, G.; Giustarini, D.; Milzani, A. Protein S-glutathionylation: A regulatory device from bacteria to humans. Trends Biochem. Sci. 2009, 34, 85–96. [Google Scholar] [CrossRef]
  85. Han, X.; Shen, T.; Lou, H. Dietary polyphenols and their biological significance. Int. J. Mol. Sci. 2007, 8, 950–988. [Google Scholar] [CrossRef] [Green Version]
  86. Kunle, O.F.; Egharevba, H.O. Preliminary studies on Vernonia ambigua: Phytochemical and antimicrobial screening of the whole plant. Ethnobot. Leafl. 2009, 13, 1216–1221. [Google Scholar]
  87. Gopalakrishnan, K.; Udayakumar, R. Phytochemical content of leaf and stem of Marsilea quadrifolia (L.). J. Plant Sci. Phytopathol. 2017, 1, 26–37. [Google Scholar]
  88. Nguyen, T.H.; Nachtergael, A.; Nguyen, T.M.; Cornet, V.; Duez, P.; Muller, M.; Huong, D.T.L.; Kestemont, P. Anti-inflammatory properties of the ethanol extract from Clerodendrum cyrtophyllum Turcz based on in vitro and in vivo studies. J. Ethnopharmacol. 2020, 254, 112739. [Google Scholar] [CrossRef] [PubMed]
  89. Jani, G.K.; Shah, D.P.; Prajapati, V.D.; Jain, V.C. Gums and mucilages: Versatile excipients for pharmaceutical formulations. Asian J. Pharm. Sci. 2009, 4, 309–323. [Google Scholar]
  90. Wadhwa, J.; Nair, A.; Kumria, R. Potential of plant mucilages in pharmaceuticals and therapy. Curr. Drug Deliv. 2013, 10, 198–207. [Google Scholar] [CrossRef]
  91. Bhutada, S.A.; Muneer-Farhan, M.; Dahikar, S.B. Preliminary phytochemical screening and antibacterial activity of resins of Boswellia serrata Roxb. J. Pharmacogn. Phytochem. 2017, 6, 182–185. [Google Scholar]
  92. Dhingra, G.; Kamble, N.; Gaikwad, B.; Kamerkar, J.; Khedkar, P.; Madhavi, T. Isolation of mucilage from various plant sources and compare their disintegrant action in tablet formulation. Am. J. PharmTech Res. 2021, 11, 11–21. [Google Scholar] [CrossRef]
  93. Shi, J.; Arunasalam, K.; Yeung, D.; Kakuda, Y.; Mittal, G.; Jiang, Y. Saponins from edible legumes: Chemistry, processing, and health benefits. J. Med. Food 2004, 7, 67–78. [Google Scholar] [CrossRef] [PubMed]
  94. Qadir, U.; Paul, V.I.; Ganesh, P. Preliminary phytochemical screening and in vitro antibacterial activity of Anamirta cocculus (Linn.) seeds. J. King Saud Univ. Sci. 2015, 27, 97–104. [Google Scholar] [CrossRef] [Green Version]
  95. Prakash, J.; Vedanayaki, S. Organoleptic, fluorescence, qualitative and quantitative analysis of bulb extract of Zephyranthes citrina. J. Pharmacogn. Phytochem. 2019, 8, 2531–2536. [Google Scholar]
  96. Prieto, J.M.; Recio, M.C.; Giner, R.M. Anti-inflammatory activity of β-sitosterol in a model of oxazoloneinduced contact-delayed-type hypersensitivity. Lat. Am. Caribb. Bull. Med. Aromat. Plants 2006, 5, 57–62. [Google Scholar]
  97. Marangoni, F.; Poli, A. Phytosterols and cardiovascular health. Pharmacol. Res. 2010, 61, 193–199. [Google Scholar] [CrossRef]
  98. Baskar, A.A.; Al Numair, K.S.; Paulraj, M.G.; Alsaif, M.A.; Al Muamar, M.; Ignacimuthu, S. β-Sitosterol prevents lipid peroxidation and improves antioxidant status and histoarchitecture in rats with 1,2-dimethylhydrazine-induced colon cancer. J. Med. Food 2012, 15, 335–343. [Google Scholar] [CrossRef] [PubMed]
  99. Trautwein, E.; Vermeer, M.; Hiemstra, H.; Ras, R. LDL-cholesterol lowering of plant sterols and stanols—Which factors influence their efficacy? Nutrients 2018, 10, 1262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Martianto, D.; Bararah, A.; Andarwulan, N.; Średnicka-Tober, D. Cross-sectional study of plant sterols intake as a basis for designing appropriate plant sterol-enriched food in Indonesia. Nutrients 2021, 13, 452. [Google Scholar] [CrossRef]
  101. Zhao, Z.; Liang, Z.; Ping, G. Macroscopic identification of Chinese medicinal materials: Traditional experiences and modern understanding. J. Ethnopharmacol. 2011, 134, 556–564. [Google Scholar] [CrossRef]
  102. Andrews, S.; Azhagu Madhavan, S.; Ganesan, S.; Arjun, P.; Jeyaprakash, R.; Baskara Sanjeevi, S.; Ramasamy, M. Different bioactive constituents and biochemical composition of brown seaweed Spatoglossum marginatum. Waffen-Und Kostumkunde J. 2020, 14, 349–356. [Google Scholar]
  103. Pandavadra, M.; Chanda, S. Development of quality control parameters for the standardization of Limonia acidissima L. leaf and stem. Asian Pac. J. Trop. Med. 2014, 7, S244–S248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Kumar, M.; Mondal, P.; Borah, S.; Mahato, K. Physico-chemical evaluation, preliminary phytochemical investigation, fluorescence and TLC analysis of leaves of the plant Lasia spinosa (Lour) Thwaites. Int. J. Pharm. Pharm. Sci. 2013, 5, 306–310. [Google Scholar]
  105. Zheng, Y.H.; Li, W.H.; Chen, P.; Zhou, Y.; Lu, W.; Ma, Z.C. Determination of berberine in Rhizoma coptidis using a β-cyclodextrin-sensitized fluorescence method. R. Soc. Chem. Adv. 2020, 10, 40136–40141. [Google Scholar] [CrossRef]
  106. Chanda, S. Importance of pharmacognostic study of medicinal plants: An overview. J. Pharmacogn. Phytochem. 2014, 2, 69–73. [Google Scholar]
  107. Carvalho, A.C.B.; Perfeito, J.P.S.; Ramalho, L.S.; Marques, R.F.d.O.; Silveira, D. Regulation of herbal medicines in Brazil: Advances and perspectives. Braz. J. Pharm. Sci. 2011, 47, 467–473. [Google Scholar] [CrossRef] [Green Version]
  108. Folashade, O.; Omoregie, H.; Ochogu, P. Standardization of herbal medicines—A review. Int. J. Biodivers. Conserv. 2012, 4, 101–112. [Google Scholar]
  109. Chase, C.R., Jr.; Pratt, R. Fluorescence of powdered vegetable drugs with particular reference to development of a system of identification. J. Am. Pharm. Assoc. 1949, 38, 324–331. [Google Scholar] [CrossRef]
  110. Sridharan, S.; Gounder, S.C. Pharmacognostic standardization and physicochemical analysis of the leaves of Barleria montana Wight & Nees. Asian Pac. J. Trop. Dis. 2016, 6, 232–234. [Google Scholar]
  111. Arumugam, S.; Natesan, S.K. Pharmacognostical studies and phytochemical investigation of Barleria noctiflora Linn (Acantheceae). Int. J. Pharmacogn. Phytochem. 2015, 7, 450–456. [Google Scholar]
  112. Tamboli, F.; More, H. Pharmacognostic and Physicochemical analysis of Barleria gibsoni Dalz. Pharmacophore 2016, 7, 118–123. [Google Scholar]
  113. Agarwal, A.; Prajapati, R.; Raza, S.K.; Thakur, L.K. GC-MS analysis and antibacterial activity of aerial parts of Quisqualis indica plant extracts. Indian J. Pharm. Educ. 2017, 51, 329–336. [Google Scholar] [CrossRef] [Green Version]
  114. Yogashree, G.D.; Singh, P.; Shrishail, H.C. Phytochemical screening and GC-MS analysis of root extracts of Parkia biglandulosa (WIGHT&ARN.). Plant Arch. 2021, 21, 314–323. [Google Scholar]
  115. Hameed, I.H.; Salman, H.D.; Mohammed, G.J. Evaluation of antifungal and antibacterial activity and analysis of bioactive phytochemical compounds of Cinnamomum zeylanicum (Cinnamon bark) using gas chromatography-mass spectrometry. Orient. J. Chem. 2016, 32, 16–25. [Google Scholar] [CrossRef] [Green Version]
  116. Falaki, F. Sample preparation techniques for gas chromatography. In Gas Chromatography-Derivatization, Sample Preparation, Application; Kush, P., Ed.; IntechOpen: London, UK, 2019; pp. 1–30. [Google Scholar]
  117. Mallick, S.S.; Dighe, V.V. Detection and estimation of alpha-Amyrin, beta-Sitosterol, Lupeol, and n-Triacontane in two medicinal plants by high performance thin layer chromatography. Adv. Chem. 2014, 2014, 143948. [Google Scholar] [CrossRef] [Green Version]
  118. Hansen, R.P.; Shorland, F.B.; Cooke, N.J. The occurrence of n-pentadecanoic acid in hydrogenated mutton fat. Biochem. J. 1954, 58, 516–517. [Google Scholar] [CrossRef]
  119. Sharmila, M.; Rajeswari, M.; Jayashree, I. GC-MS Analysis of bioactive compounds in the whole plant of ethanolic extract of Ludwigia perennis L. Int. J. Pharm. Sci. Rev. Res. 2017, 46, 124–128. [Google Scholar]
  120. Al-Rubaye, A.F.; Kadhim, M.J.; Hameed, I.H. Determination of bioactive chemical composition of methanolic leaves extract of Sinapis arvensis using GC-MS technique. Int. J. Toxicol. Pharmacol. Res. 2017, 9, 163–178. [Google Scholar] [CrossRef]
  121. Ghosh, S.; Chacko, M.J.; Harke, A.N.; Gurav, S.P.; Joshi, K.A.; Dhepe, A.; Kulkarni, A.S.; Shinde, V.S.; Parihar, V.S.; Asok, A.; et al. Barleria prionitis leaf mediated synthesis of silver and gold nanocatalysts. J. Nanomed. Nanotechnol. 2016, 7, 394. [Google Scholar] [CrossRef]
  122. Sriram, S.; Sasikumar, C.G. Evaluation of antimicrobial activity and GC-MS profiling of Barleria montana. J. Pharm. Res. 2012, 5, 2921–2925. [Google Scholar]
  123. Kumari, R.; Dubey, R.C. HPTLC and GC-MS profile of Barleria lupulina Lindl extracts and their effect on enteric bacterial pathogens. J. Appl. Pharm. 2016, 8, 62–68. [Google Scholar] [CrossRef]
  124. Tamil, S.S.; Jamuna, S.; Thekan, S.; Paulsamy, S. Profiling of bioactive chemical entities in Barleria buxifolia L. using GC-MS analysis—A significant ethno medicinal plant. J. Ayurvedic Herb. Med. 2017, 3, 63–77. [Google Scholar]
  125. Natarajan, D.; Gomathi, M.; Yuvarajan, R. Phytochemical and antibacterial evaluation of Barleria montana Nees. (mountain Barleria). Asian J. Pharm. Clin. Res. 2012, 5, 44–46. [Google Scholar]
  126. Baskaran, A.; Karthikeyan, V.; Rajasekaran, C.S. Gas chromatography-mass spectrometry (GC-MS) analysis of ethanolic extracts of Barleria longiflora L.f. World J Pharm. Pharm. Sci. 2016, 5, 1233–1246. [Google Scholar]
  127. Sujatha, A.; Evanjaline, M.; Muthukumarasamy, S.; Mohan, V.; Evanjaline, M. Determination of bioactive components of Barleria courtallica nees (Acanthaceae) by gas chromatography-mass spectrometry analysis. Asian J. Pharm. Clin. Res. 2017, 10, 273–283. [Google Scholar]
  128. Kumari, S.; Jain, P.; Sharma, B.; Kadyan, P.; Dabur, R. In vitro antifungal activity and probable fungicidal mechanism of aqueous extract of Barleria grandiflora. Appl. Biochem. Biotechnol. 2015, 175, 3571–3584. [Google Scholar] [CrossRef] [PubMed]
  129. Pandey, K.; Gupta, H.; Kamble, B. Barleria prionitis: Journey from Ayurveda to modern medicine. Nat. Prod. J. 2018, 8, 109–130. [Google Scholar] [CrossRef]
  130. Manapradit, N.; Poeaim, S.; Charoenying, P. Cytotoxicity and antimicrobial activities of leaf extracts from Barleria strigosa. Int. J. Agric. Technol. 2015, 11, 551–561. [Google Scholar]
  131. Banerjee, S.; Banerjee, S.; Jha, G.K.; Bose, S. Barleria prionitis L.: An Illustrative traditional, phytochemical and pharmacological review. J. Nat. Prod. 2021, 11, 258–274. [Google Scholar] [CrossRef]
  132. Karim, A.; Noor, A.T.; Malik, A.; Qadir, M.I.; Choudhary, M.I. Barlerisides A and B, new potent superoxide scavenging phenolic glycosides from Barleria acanthoides. J. Enzym. Inhib. Med. Chem. 2009, 24, 1332–1335. [Google Scholar] [CrossRef]
  133. Rao, E.V.; Sridhar, P.; Kumar, J.R.; Lakshmi, T.V. Anthraquinones and arnidiol from Barleria longiflora Linn F. Indian J. Pharm. Sci. 1999, 61, 282. [Google Scholar]
  134. El-Emary, N.A.; Makboul, M.A.; Abdel-Hafiz, M.A.; Ahmed, A.S. Phytochemical study of Barleria cristata L. and Barleria prionitis L. cultivated in Egypt. Bull. Pharm. Sci. Assiut 1990, 13, 65–72. [Google Scholar] [CrossRef]
  135. Chowdhury, N.; Al Hasan, A.; Tareq, F.S.; Ahsan, M.; Azam, A.Z. 4-Hydroxy-trans-cinnamate derivatives and triterpene from Barleria cristata. Dhaka Univ. J. Pharm. Sci. 2014, 12, 143–145. [Google Scholar] [CrossRef]
  136. Wanikiat, P.; Panthong, A.; Sujayanon, P.; Yoosook, C.; Rossi, A.G.; Reutrakul, V. The anti-inflammatory effects and the inhibition of neutrophil responsiveness by Barleria lupulina and Clinacanthus nutans extracts. J. Ethnopharmacol. 2008, 116, 234–244. [Google Scholar] [CrossRef] [PubMed]
  137. Kosmulalage, K.S.; Zahid, S.; Udenigwe, C.C.; Akhtar, S.; Ata, A.; Samarasekera, R. Glutathione S-transferase, acetylcholinesterase inhibitory and antibacterial activities of chemical constituents of Barleria prionitis. Z. Nat. B 2007, 62, 580–586. [Google Scholar]
  138. Smedman, A.E.; Gustafsson, I.B.; Berglund, L.G.; Vessby, B.O. Pentadecanoic acid in serum as a marker for intake of milk fat: Relations between intake of milk fat and metabolic risk factors. Am. J. Clin. Nutr. 1999, 69, 22–29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Arora, S.; Kumar, G.; Meena, S. GC-MS analysis of bioactive compounds from the whole plant hexane extract of Cenchrus setigerus Vahl. Pharma Sci. Monit. 2017, 8, 137–146. [Google Scholar]
  140. Bekinbo, M.T.; Amah-Tariah, F.S.; Dapper, D.V. Comparative GC-MS determination of bioactive constituents of the methanolic extracts of Curcuma longa rhizome and Spondias mombin leaves. J. Med. Plant Res. 2020, 8, 1–6. [Google Scholar]
  141. Janaki, C.S.; Prabhu, K.; Rao, M.R.K.; Ramaiah, V.; Dinkar, S.; Vijayalakshmi, N.; Kalaivannan, J. The GC MS Analysis of ethyl acetate extract of Merremia emerginata Burm. F (Ipomoea reniformis). Indian J. Nat. Sci. 2021, 12, 33638–33646. [Google Scholar]
  142. Prabhadevi, V.; Sahaya, S.S.; Johnson, M.; Venkatramani, B.; Janakiraman, N. Phytochemical studies on Allamanda cathartica L. using GC–MS. Asian Pac. J. Trop. Biomed. 2012, 2, S550–S554. [Google Scholar] [CrossRef]
  143. Sen, A.N.T.A.R.A.; Batra, A.M.L.A. Chemical composition of methanol extract of the leaves of Melia azedarach L. Asian J. Pharm. Clin. Res. 2012, 5, 42–45. [Google Scholar]
  144. Rampilla, V.; Khasim, S.M. GC-MS analysis of organic extracts of Cymbidium aloifolium (L.) Sw. (Orchidaceae) leaves from Eastern ghats of India. In Orchid Biology: Recent Trends & Challenges; Khasim, S., Hegde, S., González-Arnao, M., Thammasiri, K., Eds.; Springer: Singapore, 2020; pp. 507–517. [Google Scholar]
  145. Weill, P.; Plissonneau, C.; Legrand, P.; Rioux, V.; Thibault, R. May omega-3 fatty acid dietary supplementation help reduce severe complications in COVID-19 patients? Biochimie 2020, 179, 275–280. [Google Scholar] [CrossRef] [PubMed]
  146. Kumaradevan, G.; Damodaran, R.; Mani, P.; Dineshkumar, G.; Jayaseelan, T. Phytochemical screening and GC-MS analysis of bioactive components of ethanol leaves extract of Clerodendrum phlomidis (L.). Am. J. Biol. Pharm. Res. 2015, 2, 142–148. [Google Scholar]
  147. Ganesh, M.; Mohankumar, M. Extraction and identification of bioactive components in Sida cordata (Burm. f.) using gas chromatography-mass spectrometry. J. Food Sci. Technol. 2017, 54, 3082–3091. [Google Scholar] [CrossRef]
  148. Pavani, P.; Naika, R. Evaluation of antibacterial activity and GC-MS analysis of Zanthoxylum ovalifolium fruit extracts. J. Pharm. Res. Int. 2021, 33, 7–17. [Google Scholar] [CrossRef]
  149. Olusola, A.O.; Olusola, A.O.; Ogidan, T.O.; Elekan, A.O.; Ekun, O.E.; Onoagbe, I.O. GC-MS analysis of alkaloid-rich fraction of Zanthoxylum zanthoxyloides leaf. Int. J. Pharm. Sci. Res. 2020, 5, 13–17. [Google Scholar]
  150. Ezhilan, B.P.; Neelamegam, R. GC-MS analysis of phytocomponents in the ethanol extract of Polygonum chinense L. Pharmacogn. Res. 2012, 4, 11–14. [Google Scholar]
  151. Rao, M.R.K.; Anisha, G. Preliminary phytochemical and GC-MS study of one medicinal plant Carissa spinarum. Indo Am. J. Pharamaceutical Res. 2018, 8, 414–421. [Google Scholar]
  152. Nainggolan, M.; Sinaga, A.G.S. Characteristics of fatty acid composition and minor constituents of red palm olein and palm kernel oil combination. J. Adv. Pharm. Technol. Res. 2021, 12, 22–26. [Google Scholar] [CrossRef]
  153. Sivasubramanian, R.; Brindha, P. In-vitro cytotoxic, antioxidant and GC-MS studies on Centratherum punctatum Cass. Int. J. Pharm. Pharm. Sci. 2013, 4, e8. [Google Scholar]
  154. Ahsan, T.; Chen, J.; Zhao, X.; Irfan, M.; Wu, Y. Extraction and identification of bioactive compounds (eicosane and dibutyl phthalate) produced by Streptomyces strain KX852460 for the biological control of Rhizoctonia solani AG-3 strain KX852461 to control target spot disease in tobacco leaf. AMB Express 2017, 7, 54. [Google Scholar] [CrossRef] [Green Version]
  155. Subramanian, S.; Dowlath, M.J.H.; Karuppannan, S.K.; Saravanan, M.; Arunachalam, K.D. Effect of solvent on the phytochemical extraction and GC-MS Analysis of Gymnema sylvestre. Pharmacogn. J. 2020, 12, 749–761. [Google Scholar] [CrossRef]
  156. Raman, B.V.; Samuel, L.A.; Pardha, S.M.; Narashimha, R.B.; Krishna, N.V.; Sudhakar, M.; Radhakrishnan, T.M. Antibacterial, antioxidant activity and GC-MS analysis of Eupatorium odoratum. Asian J. Pharm. Clin. Res. 2012, 5, 99–106. [Google Scholar]
  157. Chowdhary, K.; Kaushik, N. Biodiversity study and potential of fungal endophytes of peppermint and effect of their extract on chickpea rot pathogens. Arch. Phytopathol. Plant Prot. 2018, 51, 139–155. [Google Scholar] [CrossRef]
  158. Pei-Xia, L.; Wei-Yi, L.; Tian-Tian, J.; Dong-Hao, H.; Yi, H. Chemical constituents analysis of ethyl acetate extract from MSR-1707 by GC-MS. Asian J. Biol. 2020, 9, 26–33. [Google Scholar]
  159. Siddiquee, S.; Cheong, B.E.; Taslima, K.; Kausar, H.; Hasan, M.M. Separation and identification of volatile compounds from liquid cultures of Trichoderma harzianum by GC-MS using three different capillary columns. J. Chromatogr. Sci. 2012, 50, 358–367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Rajisha, K.; Fernandes, J. Identification of compounds from different fractions of Exacum biolor ROXB. by GC-MS analysis. Plant Arch. 2020, 20, 4531–4536. [Google Scholar]
  161. Al-Marzoqi, A.H.; Hameed, I.H.; Idan, S.A. Analysis of bioactive chemical components of two medicinal plants (Coriandrum sativum and Melia azedarach) leaves using gas chromatography-mass spectrometry (GC-MS). Afr. J. Biotechnol. 2015, 14, 2812–2830. [Google Scholar] [CrossRef] [Green Version]
  162. Kadam, D.; Lele, S.S. Extraction, characterization and bioactive properties of Nigella sativa seed cake. J. Food Sci. Technol. 2017, 54, 3936–3947. [Google Scholar] [CrossRef]
  163. Khan, I.H.; Javaid, A.; Ahmed, D.; Khan, U. Identification of volatile constituents of ethyl acetate fraction of Chenopodium quinoa roots extract by GC-MS. Int. J. Biol. Biotechnol. 2020, 17, 17–21. [Google Scholar]
  164. McGaw, L.J.; Jäger, A.K.; Van Staden, J. Isolation of antibacterial fatty acids from Schotia brachypetala. Fitoterapia 2002, 73, 431–433. [Google Scholar] [CrossRef]
  165. Sivakumar, R.; Jebanesan, A.; Govindarajan, M.; Rajasekar, P. Larvicidal and repellent activity of tetradecanoic acid against Aedes aegypti (Linn.) and Culex quinquefasciatus (Say.) (Diptera: Culicidae). Asian Pac. J. Trop. Med. 2011, 4, 706–710. [Google Scholar] [CrossRef] [Green Version]
  166. Kushwaha, P.; Yadav, S.S.; Singh, V.; Dwivedi, L.K. Phytochemical screening and GC-MS studies of the methanolic extract of Tridax procumbens. Int. J. Pharm. Sci. Res. 2019, 10, 2492–2496. [Google Scholar]
  167. Sahi, N.M. Evaluation of insecticidal activity of bioactive compounds from Eucalyptus citriodora against Tribolium castaneum. Int. J. Pharmacogn. Phytochem. Res. 2016, 8, 1256–1270. [Google Scholar]
  168. Sarumathy, K.; Dhana Rajan, M.S.; Vijay, T.; Jayakanthi, J. Evaluation of phytoconstituents, nephro-protective and antioxidant activities of Clitoria ternatea. J. Appl. Pharm. Sci. 2011, 1, 164–172. [Google Scholar]
  169. Kalaisezhiyen, P.; Sasikumar, V. GC-MS evaluation of chemical constituents from methanolic leaf extract of Kedrostis foetidissima (Jacq.) Cogn. Asian J. Pharm. Clin. Res. 2012, 5, 77–81. [Google Scholar]
  170. Nisha, S.N.; Jothi, B.A.; Geetha, B.; Hanira, S. Phytochemical analysis of Pistia stratiotes by GC-MS analysis. Phytochem. Anal. 2018, 3, 4–6. [Google Scholar]
  171. Vinoth, M.; Natarajan, B.; Sundaram, C.S. Characterization and evaluation ethyl acetate extract of Melochia corchorifolia leaf-anticancer antibiological and molecular docking studies on breast cancer estrogen receptor. Indian J. Chem. Technol. 2021, 28, 188–196. [Google Scholar]
  172. Hsouna, A.B.; Trigui, M.; Mansour, R.B.; Jarraya, R.M.; Damak, M.; Jaoua, S. Chemical composition, cytotoxicity effect and antimicrobial activity of Ceratonia siliqua essential oil with preservative effects against Listeria inoculated in minced beef meat. Int. J. Food Microbiol. 2011, 148, 66–72. [Google Scholar] [CrossRef]
  173. Kuppuswamy, K.M.; Jonnalagadda, B.; Arockiasamy, S. Gc-ms analysis of chloroform extract of Croton bonplandianum. Int. J. Pharma Bio Sci. 2013, 4, 613–617. [Google Scholar]
  174. Grover, N.; Patni, V. Phytochemical characterization using various solvent extracts and GC-MS analysis of methanolic extract of Woodfordia fruticosa Kurz. Int. J. Pharm. Pharm. Sci. 2013, 5, 291–295. [Google Scholar]
  175. Shareef, H.K.; Muhammed, H.J.; Hussein, H.M.; Hameed, I.H. Antibacterial effect of ginger (Zingiber officinale) roscoe and bioactive chemical analysis using gas chromatography mass spectrum. Orient. J. Chem. 2016, 32, 20–40. [Google Scholar] [CrossRef] [Green Version]
  176. Save, S.A.; Lokhande, R.S.; Chowdhary, A.S. Determination of 1,2-Benzenedicarboxylic acid, bis (2-ethylhexyl) ester from the twigs of Thevetia peruviana as a Colwell Biomarker. J. Innov. Pharm. Biol. Sci. 2015, 2, 349–362. [Google Scholar]
  177. Vijay, K.; Sree, K.K.; Devi, T.S.; Soundarapandian, S.; Ramasamy, V.; Thangavel, K. Computational biology approaches revealing novel target in vascular wilt pathogen Fusarium oxysporum f. sp. lycopersici for the ligands of marine actinobacterial origin. J. Pure Appl. Microbiol. 2020, 14, 363–373. [Google Scholar] [CrossRef] [Green Version]
  178. El Sohaimy, S.A.; Hamad, G.M.; Mohamed, S.E.; Amar, M.H.; Al-Hindi, R.R. Biochemical and functional properties of Moringa oleifera leaves and their potential as a functional food. Glob. Adv. Res. J. Agric. Sci. 2015, 4, 188–199. [Google Scholar]
  179. Abraham, A.; Kattoor, A.J.; Saldeen, T.; Mehta, J.L. Vitamin E and its anticancer effects. Crit. Rev. Food Sci. Nutr. 2019, 59, 2831–2838. [Google Scholar] [CrossRef] [PubMed]
  180. Buening, M.K.; Fortner, J.G.; Kappas, A.; Conney, A.H. 7,8-Benzoflavone stimulates the metabolic activation of aflatoxin B1 to mutagens by human liver. Biochem. Biophys. Res. Commun. 1978, 82, 348–355. [Google Scholar] [CrossRef]
  181. Buening, M.K.; Chang, R.L.; Huang, M.T.; Fortner, J.G.; Wood, A.W.; Conney, A.H. Activation and inhibition of benzo (a) pyrene and aflatoxin B1 metabolism in human liver microsomes by naturally occurring flavonoids. Cancer Res. 1981, 41, 67–72. [Google Scholar] [PubMed]
  182. Sivaraj, C.; Yamini, S.; Yahavi, A.; Kumar, R.P.; Arumugam, P.; Manimaaran, A. Antioxidant, antimicrobial activities and GC-MS analysis of fruit extract of Solanum nigrum L. J. Pharmacogn. Phytochem. 2020, 9, 1114–1121. [Google Scholar]
  183. Santhanamari, N.; Uthayakumari, F.; Maria Sumathi, B. Phytochemical investigation of whole plant extracts of Cyperus bulbosus Vahl by GC-MS analysis. World J. Pharm. Pharm. Sci. 2016, 5, 1671–1676. [Google Scholar]
  184. Beschi, D.A.; Appavoo, M.R.; Wilsy, J.I. GC-MS analysis, collected from Kavalkinaru area, Tirunelveli District, Tamil Nadu, India. Eur. J. Mol. Clin. Med. 2021, 7, 4287–4292. [Google Scholar]
  185. Jegajeevanram, P.; Alhaji, N.M.; Kumaravel, S. Identification of pesticide compounds of Cynodon doctylon by GC-MS analysis. Int. J. Pharm. Biol. Sci. 2014, 5, 604–608. [Google Scholar]
  186. Panda, S.; Jafri, M.; Kar, A.; Meheta, B.K. Thyroid inhibitory, antiperoxidative and hypoglycemic effects of stigmasterol isolated from Butea monosperma. Fitoterapia 2009, 80, 123–126. [Google Scholar] [CrossRef] [PubMed]
  187. Bharathy, V.; Sumathy, B.; Uthayakumari, F. Determination of phytocomponents by GC–MS in leaves of Jatropha gossypifolia L. Sci. Res. Rep. 2012, 2, 286–290. [Google Scholar]
  188. Okoye, N.N.; Ajaghaku, D.L.; Okeke, H.N.; Ilodigwe, E.E.; Nworu, C.S.; Okoye, F.B.C. Beta-amyrin and alpha-amyrin acetate isolated from the stem bark of Alstonia boonei display profound anti-inflammatory activity. Pharm. Biol. 2014, 52, 1478–1486. [Google Scholar] [CrossRef] [Green Version]
  189. Kuroshima, K.N.; Campos-Buzzi, F.; Yunes, R.A.; Delle Monache, F.; Cechinel Filho, V. Chemical composition and antinociceptive properties of Hyeronima alchorneoides leaves. Pharm. Biol. 2005, 43, 573–578. [Google Scholar] [CrossRef]
  190. Chinnadurai, V.; Viswanathan, P.; Kalimuthu, K.; Vanitha, A.; Ranjitha, V.; Pugazhendhi, A. Comparative studies of phytochemical analysis and pharmacological activities of wild and micro propagated plant ethanol extracts of Manihot esculenta. Biocatal. Agric. Biotechnol. 2019, 19, 101166. [Google Scholar] [CrossRef]
  191. Thirumalai, V.; Nirmala, P.; Venkatanarayanan, R. Phytochemical characterization of cold macerated methanolic leaf extract of Cadaba indica Lam. Using GC-MS. Int. J. Pharm. Sci. Res. 2021, 12, 3185–3192. [Google Scholar]
  192. Roy, S.J.; Baruah, P.S.; Lahkar, L.; Gurung, L.; Saikia, D.; Tanti, B. Phytochemical analysis and antioxidant activities of Homalomena aromatic Schott. J. Pharmacogn. Phytochem. 2019, 8, 1379–1385. [Google Scholar]
  193. Francis, S.; Gideon, V.A.; Britto, S.J. Antibacterial and GC-MS analysis of stem and leaf of Premna paucinervis (CB Clarke) gamble (Lamiaceae)—An endemic and rediscovered species. Int. J. Bot. Stud. 2021, 6, 282–292. [Google Scholar]
  194. Duan, D.D.; Bu, C.Y.; Cheng, J.; Wang, Y.N.; Shi, G.L. Isolation and identification of acaricidal compounds in Inula japonica (Asteraceae). J. Econ. Entomol. 2011, 104, 375–378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Durairaj, V.; Hoda, M.; Shakya, G.; Babu, S.P.P.; Rajagopalan, R. Phytochemical screening and analysis of antioxidant properties of aqueous extract of wheatgrass. Asian Pac. J. Trop. Med. 2014, 7, S398–S404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Amaral, A.C.; Gomes, L.A.; Silva, J.R.; Ferreira, J.L.; Ramos, A.R.; Rosa, M.S.S.; Vermelho, A.B.; Rodrigues, I.A. Liposomal formulation of turmerone-rich hexane fractions from Curcuma longa enhances their antileishmanial activity. Biomed Res. Int. 2014, 2014, 694934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Hussein, H.M.; Hameed, I.H.; Ibraheem, O.A. Antimicrobial Activity and spectral chemical analysis of methanolic leaves extract of Adiantum capillus-veneris using GC-MS and FT-IR spectroscopy. Int. J. Pharmacogn. Phytochem. Res. 2016, 8, 369–385. [Google Scholar]
  198. Hugar, A.L.; Kanjikar, A.P.; Londonkar, R.L. Cyclamen persicum: Methanolic extract using gas chromatography-mass spectrometry (GC-MS) technique. Int. J. Pharm. Clin. Res. 2017, 8, 200–213. [Google Scholar]
  199. Khan, S.; Richa, K.H.; Jhamta, R. Evaluation of antioxidant potential and phytochemical characterization using GC-MS analysis of bioactive compounds of Achillea filipendulina (L.) leaves. J. Pharmacogn. Phytochem. 2019, 8, 258–265. [Google Scholar]
  200. Manjari, M.S.; Karthi, S.; Ramkumar, G.; Muthusamy, R.; Natarajan, D.; Shivakumar, M.S. Chemical composition and larvicidal activity of plant extracts from Clausena dentata (Willd) (Rutaceae) against dengue, malaria, and filariasis vectors. Parasitol. Res. 2014, 113, 2475–2481. [Google Scholar] [CrossRef]
  201. Vashisht, S.; Singh, M.P.; Chawla, V. GC-MS Analysis of phytocomponents in the various extracts of Shorea robusta Gaertn F. Int. J. Pharmacogn. Phytochem. 2017, 9, 783–788. [Google Scholar]
  202. Shanmugavel, K.; Maria-Louis, P.L.; Soosaimicheal, M.J.K.; Veerabahu, R.M.; Arumugam, M. GC-MS analysis of ethanol extract of stem of Nothapodytes nimmoniana (Graham) Mabb. J. Pharm. Biol. Sci. 2015, 3, 1145–1150. [Google Scholar]
  203. Hase, G.J.; Deshmukh, K.K.; Pokharkar, R.D.; Gaje, T.R.; Phatanagre, N.D. Phytochemical Studies on Nerium oleander L. using GC-MS. Int. J. Pharmacogn. Phytochem. Res. 2017, 9, 885–891. [Google Scholar] [CrossRef]
  204. Delazar, A.; Nazifi, E.; Movafeghi, A.; Nazemiyey, H.; Hemmati, S.; Nahar, L.; Sarker, S.D. Analyses of phytosterols and free radical scavengers in the bulbs of Ornithogalum cuspidatum. Lat. Am. Caribb. Bull. Med. Aromat. Plants 2010, 9, 87–92. [Google Scholar]
  205. Balogun, O.S.; Oladosu, I.A.; Akiinnusi, A.; Zhiqiang, L. Fatty acid composition α-glucosidae inhibitory potential and cytotoxicity activity of Oncoba spinosa Forssk. Elix. Appl. Chem. 2013, 59, 15637–15641. [Google Scholar]
  206. Chouni, A.; Pal, A.; Gopal, P.K.; Paul, S. GC-MS Analysis and screening of anti-proliferative potential of methanolic extract of Garcinia cowa on different cancer cell lines. Pharmacogn. J. 2021, 13, 347–361. [Google Scholar] [CrossRef]
  207. Suzuki, K. Anti-obesity effect of cholest-4-en-3-one, an intestinal catabolite of cholesterol, on mice. J. Nutr. Sci. Vitaminol. 1993, 39, 537–543. [Google Scholar] [CrossRef] [PubMed]
  208. Kanjikar, A.; Arunal, H.; Ramesh, L. Novel investigation on in-vitro anti-diabetic and volatile profile of bioactive compounds present in methanolic extract of Ficus krishnae. Int. J. ChemTech Res. 2017, 10, 220–228. [Google Scholar]
  209. Matu, E.N.; Van Staden, J. Antibacterial and anti-inflammatory activities of some plants used for medicinal purposes in Kenya. J. Ethnopharmacol. 2003, 87, 35–41. [Google Scholar] [CrossRef]
  210. Lee, W.; Woo, E.R.; Lee, D.G. Phytol has antibacterial property by inducing oxidative stress response in Pseudomonas aeruginosa. Free. Radic. Res. 2016, 50, 1309–1318. [Google Scholar] [CrossRef]
  211. Awolola, G.V.; Koorbanally, N.A.; Chenia, H.; Shode, F.O.; Baijnath, H. Antibacterial and anti-biofilm activity of flavonoids and triterpenes isolated from the extracts of Ficus sansibarica Warb. Subsp. Sansibarica (Moraceae) extracts. Afr. J. Tradit. Complement. Altern Med. 2014, 11, 124–131. [Google Scholar] [CrossRef] [Green Version]
  212. Yenna, T.W.; Khanb, M.A.; Syuhadaa, N.A.; Ringa, L.C.; Ibrahimc, D.; Tan, W.N. Stigmasterol: An adjuvant for beta lactam antibiotics against beta-lactamase positive clinical isolates. Steroids 2017, 128, 68–71. [Google Scholar] [CrossRef] [PubMed]
  213. Neu, H.C. The crisis in antibiotic resistance. Science 1992, 257, 1064–1073. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Mukhopadhyay, A.; Peterson, R.T. Fishing for new antimicrobials. Curr. Opin. Chem. Biol. 2006, 10, 327–333. [Google Scholar] [CrossRef]
  215. Tenover, F.C. Mechanisms of antimicrobial resistance in bacteria. Am. J. Med. 2006, 119, S3–S10. [Google Scholar] [CrossRef]
  216. Schwarz, S.; Noble, W.C. Aspects of bacterial resistance to antimicrobials used in veterinary dermatological practice. Vet. Dermatol. 1999, 10, 163–176. [Google Scholar] [CrossRef]
  217. Kumar, H.; Agrawal, R.; Kumar, V. Barleria cristata: Perspective towards phytopharmacological aspects. J. Pharm. Pharmacol. 2018, 70, 475–487. [Google Scholar] [CrossRef] [Green Version]
  218. Chavan, C.B.; Shinde, U.V.; Hogade, M.; Bhinge, S. Screening of in vitro antibacterial assay of Barleria proinitis Linn. J. Herb. Med. Toxicol. 2010, 4, 197–200. [Google Scholar]
  219. Moin, S.; Babu, S.S.; Mahalakshmipriya, A. In vitro callus production and antibacterial activity of Barleria lupulina lindl. Asia-Pac. J. Mol. Biol. Biotechnol. 2012, 20, 59–64. [Google Scholar]
  220. Tekwu, E.M.; Pieme, A.C.; Beng, V.P. Investigations of antimicrobial activity of some Cameroonian medicinal plant extracts against bacteria and yeast with gastrointestinal relevance. Ethnopharmacology 2012, 142, 65–273. [Google Scholar] [CrossRef] [PubMed]
  221. Rubio-Moraga, Á.; Argandoña, J.; Mota, B.; Pérez, J.; Verde, A.; Fajardo, J.; Gómez-Navarro, J.; Castillo-López, R.; Ahrazem, O.; Gómez-Gómez, L. Screening for polyphenols, antioxidant and antimicrobial activities of extracts from eleven Helianthemum taxa (Cistaceae) used in folk medicine in south-eastern Spain. J. Ethnopharmacol. 2013, 148, 287–296. [Google Scholar] [CrossRef] [PubMed]
  222. Jarald, E.E.; Jarald, S.E. A Text Book of Pharmacognosy and Phytochemistry, 1st ed.; CBS Publishers and Distributors: New Delhi, India, 2007. [Google Scholar]
  223. Harborne, J.B. Photochemical Methods: A Guide to Modern Techniques of Plant Analysis; Chapman and Hall: London, UK, 1973. [Google Scholar]
  224. Trease, G.E.; Evans, W.C. Pharmacology, 11th ed.; Bailliere Tindall Ltd.: London, UK, 1978. [Google Scholar]
  225. Sofowora, A. Medicinal Plants and Traditional Medicines in Africa; Chichester John Wiley and Sons: New York, NY, USA, 1993; pp. 97–145. [Google Scholar]
  226. Tiwari, P.; Kumar, B.; Kaur, M.; Kaur, G.; Kaur, H. Phytochemical screening and extraction: A review. Int. Pharm. Sci. 2011, 1, 98–106. [Google Scholar]
  227. Kumar, D.; Gupta, J.; Kumar, S.; Arya, R.; Kumar, T.; Gupta, G. Pharmacognostic evaluation of Cayratia trifolia (Linn.) leaf. Asian Pac. J. Trop. Biomed. 2012, 2, 6–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Akwu, N.A.; Naidoo, Y.; Singh, M. A comparative study of the proximate, FTIR analysis and mineral elements of the leaves and stem bark of Grewia lasiocarpa E. Mey. ex Harv.: An indigenous southern African plant. S. Afr. J. Bot. 2019, 123, 9–19. [Google Scholar] [CrossRef]
  229. Clinical and Laboratory Standards Institute. Performance Standards for Antimicrobial Disk Susceptibility Tests, Approved Standard-Ninth Edition (M2-A9); CLSI: Philadelphia, PA, USA, 2006. [Google Scholar]
  230. Marathe, N.P.; Rasane, M.H.; Kumar, H.; Patwardhan, A.A.; Shouche, Y.S.; Diwanay, S.S. In vitro antibacterial activity of Tabernaemontana alternifolia (Roxb) stem bark aqueous extracts against clinical isolates of methicillin resistant Staphylococcus aureus. Ann. Clin. Microbiol. Antimicrob. 2013, 12, 26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Okla, M.K.; Alatar, A.A.; Al-Amri, S.S.; Soufan, W.H.; Ahmad, A.; Abdel-Maksoud, M.A. Antibacterial and antifungal activity of the extracts of different parts of Avicennia marina (Forssk.) Vierh. Plants 2021, 10, 252. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Leaf powder analysis (bright light) of B. albostellata. (A) Powder; (B) powder + water; (C) powder + H2SO4; (D) powder + acetic acid; (E) powder + aqueous NaOH; (F) powder + HCl; (G) powder + ethanol; (H) powder + ethyl acetate; (I) powder + hexane; (J) powder + chloroform; (K) powder + methanol; (L) powder + petroleum ether; (M) powder + diethyl ether; (N) powder + acetone.
Figure 1. Leaf powder analysis (bright light) of B. albostellata. (A) Powder; (B) powder + water; (C) powder + H2SO4; (D) powder + acetic acid; (E) powder + aqueous NaOH; (F) powder + HCl; (G) powder + ethanol; (H) powder + ethyl acetate; (I) powder + hexane; (J) powder + chloroform; (K) powder + methanol; (L) powder + petroleum ether; (M) powder + diethyl ether; (N) powder + acetone.
Plants 12 02396 g001
Figure 2. Fluorescence analysis (UV light) of leaf powder from B. albostellata. (A) Powder; (B) powder + water; (C) powder + H2SO4; (D) powder + acetic acid; (E) powder + aqueous NaOH; (F) powder + HCl; (G) powder + ethanol; (H) powder + ethyl acetate; (I) powder + hexane; (J) powder + chloroform; (K) powder + methanol; (L) powder + petroleum ether; (M) powder + diethyl ether; (N) powder + acetone.
Figure 2. Fluorescence analysis (UV light) of leaf powder from B. albostellata. (A) Powder; (B) powder + water; (C) powder + H2SO4; (D) powder + acetic acid; (E) powder + aqueous NaOH; (F) powder + HCl; (G) powder + ethanol; (H) powder + ethyl acetate; (I) powder + hexane; (J) powder + chloroform; (K) powder + methanol; (L) powder + petroleum ether; (M) powder + diethyl ether; (N) powder + acetone.
Plants 12 02396 g002
Figure 3. Stem powder analysis (bright light) from B. albostellata. (A) Powder; (B) powder + water; (C) powder + H2SO4; (D) powder + acetic acid; (E) powder + aqueous NaOH; (F) powder + HCl; (G) powder + ethanol; (H) powder + ethyl acetate; (I) powder + hexane; (J) powder + chloroform; (K) powder + methanol; (L) powder + petroleum ether; (M) powder + diethyl ether; (N) powder + acetone.
Figure 3. Stem powder analysis (bright light) from B. albostellata. (A) Powder; (B) powder + water; (C) powder + H2SO4; (D) powder + acetic acid; (E) powder + aqueous NaOH; (F) powder + HCl; (G) powder + ethanol; (H) powder + ethyl acetate; (I) powder + hexane; (J) powder + chloroform; (K) powder + methanol; (L) powder + petroleum ether; (M) powder + diethyl ether; (N) powder + acetone.
Plants 12 02396 g003
Figure 4. Fluorescence analysis (UV light) of stem powder from B. albostellata. (A) Powder; (B) powder + water; (C) powder + H2SO4; (D) powder + acetic acid; (E) powder + aqueous NaOH; (F) powder + HCl; (G) powder + ethanol; (H) powder + ethyl acetate; (I) powder + hexane; (J) powder + chloroform; (K) powder + methanol; (L) powder + petroleum ether; (M) powder + diethyl ether; (N) powder + acetone.
Figure 4. Fluorescence analysis (UV light) of stem powder from B. albostellata. (A) Powder; (B) powder + water; (C) powder + H2SO4; (D) powder + acetic acid; (E) powder + aqueous NaOH; (F) powder + HCl; (G) powder + ethanol; (H) powder + ethyl acetate; (I) powder + hexane; (J) powder + chloroform; (K) powder + methanol; (L) powder + petroleum ether; (M) powder + diethyl ether; (N) powder + acetone.
Plants 12 02396 g004
Figure 5. GC-MS chromatogram of leaf hexane extract of B. albostellata.
Figure 5. GC-MS chromatogram of leaf hexane extract of B. albostellata.
Plants 12 02396 g005
Figure 6. GC-MS chromatogram of leaf chloroform extract of B. albostellata.
Figure 6. GC-MS chromatogram of leaf chloroform extract of B. albostellata.
Plants 12 02396 g006
Figure 7. GC-MS chromatogram of leaf methanol extract of B. albostellata.
Figure 7. GC-MS chromatogram of leaf methanol extract of B. albostellata.
Plants 12 02396 g007
Figure 8. GC-MS chromatogram of stem hexane extract of B. albostellata.
Figure 8. GC-MS chromatogram of stem hexane extract of B. albostellata.
Plants 12 02396 g008
Figure 9. GC-MS chromatogram of stem chloroform extract of B. albostellata.
Figure 9. GC-MS chromatogram of stem chloroform extract of B. albostellata.
Plants 12 02396 g009
Figure 10. GC-MS chromatogram of stem methanol extract of B. albostellata.
Figure 10. GC-MS chromatogram of stem methanol extract of B. albostellata.
Plants 12 02396 g010
Table 1. Organoleptic features of different parts of B. albostellata.
Table 1. Organoleptic features of different parts of B. albostellata.
Organoleptic FeaturesLeafStem
ColourGrey-green on both surfaces but lighter on the lower sideYellow-buff on uppermost internodes, white or cream below
OdourSlightly aromaticInodorous
TasteAcridAcrid
TextureVelvetyWoody, glabrescent
Table 2. Percentage yield of the leaf and stem crude extracts of B. albostellata.
Table 2. Percentage yield of the leaf and stem crude extracts of B. albostellata.
Crude ExtractLeavesStemLeavesStemLeavesStem
Dried Extract Yield (g)Percentage Yield (%)Colour
Hexane0.1390.1941.391.94Dark yellowLight yellow
Chloroform0.2650.2192.652.19Dark greenLight green
Methanol1.6780.93816.789.38Dark brownLight yellow
Table 3. Preliminary phytochemical screening for major classes of compounds in hexane, chloroform and methanolic leaf and stem extracts of B. albostellata.
Table 3. Preliminary phytochemical screening for major classes of compounds in hexane, chloroform and methanolic leaf and stem extracts of B. albostellata.
Compound GroupPhytochemical TestLeavesStems
HexaneChloroformMethanolHexaneChloroformMethanol
AlkaloidsDragendorffs+++++++++++++++++
Mayers++++++−
Wagners+++++++−+−++++−
Amino acidsNinhydrin+−+−++
CarbohydratesBenedicts+++++
Fehlings++++−++++++++
Molisch+−+−+−
Fixed oils and fatsFilter paper+++−+−
FlavonoidsLead acetate+++++++−+++++++−
Mucilage and GumsRuthenium+++++++++++++++++
PhenolsFerric trichloride+−++++++−+−+−
SaponinsFroth+−+++−+−+−++
Foam+−+−+++++++++++
TerpenoidsChloroform+++++++−+−++
SterolsSalkowski’s+++−+−
Intensity of reaction: (−) no observed changes, (+−) low intensity, (++) medium intensity, (+++) high intensity.
Table 4. Phytochemical compounds identified in leaf hexane extracts of B. albostellata by GC-MS analysis.
Table 4. Phytochemical compounds identified in leaf hexane extracts of B. albostellata by GC-MS analysis.
PeakRetention TimePhytochemical CompoundMolecular FormulaMolecular WeightCAS No.Area %
117.738Pentadecanoic acidC15H30O22421002-84-21.02
219.4229,12,15-Octadecatrienoic acid, (Z)C18H30O2278463-40-11.25
319.628Octadecanoic acidC18H36O228457-11-41.09
424.59113-Docosenamide, (Z)C22H43NO337112-84-51.12
524.804SqualeneC30H50410111-02-41.06
626.742EicosaneC20H42282112-95-81.39
726.7841-HeptacosanolC27H56O3962004-39-91.21
828.136TetratetracontaneC44H906187098-22-83.25
929.965l-(+)-Ascorbic acid 2,6-dihexadecanoateC38H68O865228474-90-02.85
Table 5. Phytochemical compounds identified in leaf chloroform extracts of B. albostellata by GC-MS analysis.
Table 5. Phytochemical compounds identified in leaf chloroform extracts of B. albostellata by GC-MS analysis.
PeakRetention TimePhytochemical CompoundMolecular FormulaMolecular WeightCAS No.Area %
117.738Pentadecanoic acidC15H30O22421002-84-21.02
219.4189,12,15-Octadecatrienoic acid, (Z)C18H30O2278463-40-11.04
3.19.646Tridecanoic acidC13H26O2214638-53-91.01
419.674Decanedioic acid, dibutyl esterC18H34O314109-43-31.02
524.168Octadecanoic acid, 2,3-dihydroxypropyl esterC21H42O443123-94-41.00
624.60513-Docosenamide, (Z)C22H43NO338112-84-51.03
728.120TetratetracontaneC44H906197098-22-81.00
Table 6. Phytochemical compounds identified in leaf methanol extracts of B. albostellata by GC-MS analysis.
Table 6. Phytochemical compounds identified in leaf methanol extracts of B. albostellata by GC-MS analysis.
PeakRetention TimePhytochemical CompoundMolecular FormulaMolecular Weight (g/mol)CAS No.Area %
114.7431,2,3,5-CyclohexanetetrolC6H12O461953585-08-33.63
217.057Phytol, acetateC22H42O3390-00-07.29
319.514n-Nonadecanol-1C19H40O2851454-84-85.35
419.747PhytolC20H40O297150-86-74.66
521.1401,2-15,16-DiepoxyhexadecaneC16H30O22540-00-03.10
624.59113-Docosenamide, (Z)C22H43NO337112-84-52.46
724.8971,4-Benzenedicarboxylic acid, bis(2-ethylhexyl) esterC24H38O43916422-86-26.46
825.388SqualeneC30H5O410111-02-45.39
927.3401-HeptacosanolC27H56O3972004-39-94.27
1027.579Vitamin EC29H50O243159-02-93.67
1127.723Flavone, 4′,5-dihydroxy-6,7-dimethoxy-C17H14O63146601-62-311.69
1228.491CampesterolC28H48O401474-62-45.16
1328.705StigmasterolC29H48O41383-48-74.01
1429.257Beta-SitosterolC29H50O41583-46-56.70
1529.763Alpha-AmyrinC30H50O427638-95-93.22
1630.390SimiarenolC30H50O4271615-94-74.25
Table 7. Phytochemical compounds identified in stem hexane extracts of B. albostellata by GC-MS analysis.
Table 7. Phytochemical compounds identified in stem hexane extracts of B. albostellata by GC-MS analysis.
PeakRetention TimePhytochemical CompoundMolecular FormulaMolecular WeightCAS No.Area %
117.759Pentadecanoic acidC15H30O22421002-84-21.00
219.3739,12-Octdecadienoic acid (Z)C18H32O228060-33-32.33
319.433Dichloroacetic acid, tridec-2-ynyl esterC15H24Cl2O23060-00-01.52
419.646Octadecanoic acidC18H36O228457-11-41.52
519.673Decanedioic acid, dibutyl esterC18H34O4314109-43-31.39
624.60113-Docosenamide, (Z)C22H43NO337112-84-51.32
728.125TetratetracontaneC44H906187098-22-81.45
828.7964,4,6a,6b,8a,11,11,14b-Octamethyl-1,4,4a,5,6,6a,6b,7,8,8a,9,10,11,12,12a,14,14a,14b-octadecahydro-2H-picen-3-oneC30H48O4240-00-02.35
Table 8. Phytochemical compounds identified in stem chloroform extracts of B. albostellata by GC-MS analysis.
Table 8. Phytochemical compounds identified in stem chloroform extracts of B. albostellata by GC-MS analysis.
PeakRetention TimePhytochemical CompoundMolecular FormulaMolecular WeightCAS No.Area %
112.711Phenol, 2,4-bis(1,1-dimethylethyl)-C14H22O20696-76-41.20
217.750Pentadecanoic acidC15H30O22421002-84-21.00
319.641Octadecanoic acidC18H36O228457-11-41.06
419.675Decanedioic acid, dibutyl esterC18H34O4314109-43-31.07
524.168Octadecanoic acid, 2,3-dihydroxypropyl esterC21H42O4358123-94-41.03
624.61213-Docosenamide, (Z)C22H43NO337112-84-51.21
Table 9. Phytochemical compounds identified in stem methanol extracts of B. albostellata by GC-MS analysis.
Table 9. Phytochemical compounds identified in stem methanol extracts of B. albostellata by GC-MS analysis.
PeakRetention TimePhytochemical CompoundMolecular FormulaMolecular WeightCAS No.Area %
119.514n-Nonadecanol-1C19H40O2851454-84-85.84
220.942Tributyl acetylcitrateC20H34O840277-90-71.05
321.1391,2-15,16-DiepoxyhexadecaneC16H30O22540-00-01.23
421.9679-OctadecenamideC18H35NO281301-02-02.19
524.58813-Docosenamide, (Z)C22H43NO337112-84-51.14
624.9011,4-Benzenedicarboxylic acid, bis(2-ethylhexyl) esterC24H38O43916422-86-225.68
725.386SqualeneC30H50410111-02-41.39
827.3371-HeptacosanolC27H56O3972004-39-91.72
928.703StigmasterolC29H48O41383-48-71.89
1029.252Beta-SitosterolC29H50O41583-46-55.68
1129.453Alpha. AmyrenoneC30H48O4250-00-02.18
1229.755Alpha-AmyrinC30H50O427638-95-93.16
1329.928Acetic acid, 3-hydroxy-6-isopropenyl-4,8a-dimethyl-1,2,3,4,5,6,7,8C17H26O32780-00-02.01
1430.195Stigmasta-3,5-dien-7-oneC29H46O4112034-72-21.51
1530.604Cholest-4-en-3-oneC27H44O385601-57-01.13
Table 11. Antibacterial activity of crude extracts from leaves and stem of B. albostellata against human pathogenic strains (zone of inhibition, mm).
Table 11. Antibacterial activity of crude extracts from leaves and stem of B. albostellata against human pathogenic strains (zone of inhibition, mm).
StrainConcentration (mg/mL)ExtractsPositive Control
(mg/mL)
Leaf HexaneLeaf ChloroformLeaf MethanolStem HexaneStem ChloroformStem MethanolLeafStem
B. subtillus (ATCC 6633)3.125RRRRRR9.00 ± 1.0011.00 ± 1.00
6.25RRRRRR
12.5RRRRRR
25R8.33 ± 1.538.00 ± 1.00R7.33 ± 0.588.00 ± 0.00
507.67 ± 2.087.67 ± 0.589.33 ± 0.587.67 ± 2.088.00 ± 1.008.67 ± 0.58
1009.00 ± 3.467.00 ± 0.0010.00 ± 2.008.67 ± 1.5210.00 ± 3.619.67 ± 0.58
Methicillin-resistant S. aureus (ATCC 43300)3.125RRRRRR9.33 ± 0.589.00 ± 1.00
6.25RRRRRR
12.5RRRRRR
25R8.67 ± 0.58RRRR
50R9.00 ± 0.008.67 ± 2.08R7.33 ± 0.588.00 ± 1.00
100R10.00 ± 0.0011.00 ± 2.6510.67 ± 2.318.00 ± 1.009.00 ± 2.00
S. aureus (ATCC 25923)3.125RRRRRR9.67 ± 0.5810.00 ± 1.00
6.25RRRRRR
12.5RR7.33 ± 0.58RR7.00 ± 0.00
25R7.67 ± 0.588.00 ± 0.008.00 ± 1.007.33 ± 0.588.00 ± 1.00
507.33 ± 0.588.67 ± 0.588.67 ± 1.539.00 ± 0.008.33 ± 0.5810.00 ± 1.73
1009.33 ± 0.589.33 ± 0.5810.33 ± 1.5310.33 ± 1.539.00 ± 1.7311.00 ± 2.65
E. coli (ATCC 35218)3.125RRRRRR8.67 ± 0.589.33 ± 0.58
6.25RRRRRR
12.5RRRRRR
25R9.67 ± 0.58RR9.33 ± 0.58R
50R10.67 ± 1.159.00 ± 3.46R10.00 ± 1.009.67 ± 2.08
100R12.33 ± 2.0812.67 ± 0.58R11.33 ± 1.1511.33 ± 1.15
P. aeruginosa (ATCC 25783)3.125RRRRRR9.33 ± 0.588.67 ± 1.15
6.25RRRRRR
12.5RRRRRR
25R8.67 ± 1.537.33 ± 0.58R7.00 ± 0.009.33 ± 1.15
50R9.00 ± 0.008.67 ± 1.52R8.67 ± 0.5810.67 ± 2.87
100R10.00 ± 3.0014.33 ± 1.53R9.67 ± 1.1512.33 ± 0.58
R-resistant, positive controls (streptomycin 10 mg/mL (Gram-positive bacteria), gentamicin 10 mg/mL (Gram-negative bacteria)).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gangaram, S.; Naidoo, Y.; Dewir, Y.H.; Singh, M.; Lin, J.; Murthy, H.N. Phytochemical Composition and Antibacterial Activity of Barleria albostellata C.B. Clarke Leaf and Stem Extracts. Plants 2023, 12, 2396. https://doi.org/10.3390/plants12132396

AMA Style

Gangaram S, Naidoo Y, Dewir YH, Singh M, Lin J, Murthy HN. Phytochemical Composition and Antibacterial Activity of Barleria albostellata C.B. Clarke Leaf and Stem Extracts. Plants. 2023; 12(13):2396. https://doi.org/10.3390/plants12132396

Chicago/Turabian Style

Gangaram, Serisha, Yougasphree Naidoo, Yaser Hassan Dewir, Moganavelli Singh, Johnson Lin, and Hosakatte Niranjana Murthy. 2023. "Phytochemical Composition and Antibacterial Activity of Barleria albostellata C.B. Clarke Leaf and Stem Extracts" Plants 12, no. 13: 2396. https://doi.org/10.3390/plants12132396

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