Next Article in Journal
Resveratrol Treatment Induces Mito-miRNome Modification in Follicular Fluid from Aged Women with a Poor Prognosis for In Vitro Fertilization Cycles
Next Article in Special Issue
The Potential Role of Everlasting Flower (Helichrysum stoechas Moench) as an Antihypertensive Agent: Vasorelaxant Effects in the Rat Aorta
Previous Article in Journal
Oxygen Sensing: Physiology and Pathophysiology
Previous Article in Special Issue
Unveiling the Antioxidant Therapeutic Functionality of Sustainable Olive Pomace Active Ingredients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unveiling the Phytochemical Profile and Biological Potential of Five Artemisia Species

by
Adriana Trifan
1,†,
Gokhan Zengin
2,†,
Kouadio Ibrahime Sinan
2,
Elwira Sieniawska
3,*,
Rafal Sawicki
4,
Magdalena Maciejewska-Turska
5,
Krystyna Skalikca-Woźniak
3 and
Simon Vlad Luca
6,*
1
Department of Pharmacognosy, Grigore T. Popa University of Medicine and Pharmacy Iasi, 700115 Iasi, Romania
2
Physiology and Biochemistry Research Laboratory, Department of Biology, Science Faculty, Selcuk University, University Campus, 42130 Konya, Turkey
3
Department of Natural Products Chemistry, Medical University of Lublin, 20-093 Lublin, Poland
4
Department of Biochemistry and Biotechnology, Medical University of Lublin, 20-093 Lublin, Poland
5
Department of Pharmacognosy with the Medicinal Plant Garden, Medical University of Lublin, 20-093 Lublin, Poland
6
Biothermodynamics, TUM School of Life Sciences, Technical University of Munich, 85354 Freising, Germany
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 12 April 2022 / Revised: 17 May 2022 / Accepted: 19 May 2022 / Published: 21 May 2022

Abstract

:
The Artemisia L. genus comprises over 500 species with important medicinal and economic attributes. Our study aimed at providing a comprehensive metabolite profiling and bioactivity assessment of five Artemisia species collected from northeastern Romania (A. absinthium L., A. annua L., A. austriaca Jacq., A. pontica L. and A. vulgaris L.). Liquid chromatography–tandem high-resolution mass spectrometry (LC-HRMS/MS) analysis of methanol and chloroform extracts obtained from the roots and aerial parts of the plants led to the identification of 15 phenolic acids (mostly hydroxycinnamic acid derivatives), 26 flavonoids (poly-hydroxylated/poly-methoxylated flavone derivatives, present only in the aerial parts), 14 sesquiterpene lactones, 3 coumarins, 1 lignan and 7 fatty acids. Clustered image map (CIM) analysis of the phytochemical profiles revealed that A. annua was similar to A. absinthium and that A. pontica was similar to A. austriaca, whereas A. vulgaris represented a cluster of its own. Correlated with their total phenolic contents, the methanol extracts from both parts of the plants showed the highest antioxidant effects, as assessed by the DPPH and ABTS radical scavenging, CUPRAC, FRAP and total antioxidant capacity methods. Artemisia extracts proved to be promising sources of enzyme inhibitory agents, with the methanol aerial part extracts being the most active samples against acetylcholinesterase and glucosidase. All Artemisia samples displayed good antibacterial effects against Mycobacterium tuberculosis H37Ra, with MIC values of 64–256 mg/L. In conclusion, the investigated Artemisia species proved to be rich sources of bioactives endowed with antioxidant, enzyme inhibitory and anti-mycobacterial properties.

1. Introduction

Artemisia L. is a genus of small herbs and shrubs belonging to the Asteraceae family which inhabit the northern temperate regions of Asia, Europe and North America [1]. The Artemisia genus comprises over 500 species with significant medicinal and economic attributes due to their biological and chemical diversity [2]. Artemisia species are recognized for their characteristic strong aromas and bitter tastes, which are assigned to the presence of terpenes and sesquiterpene lactones [3]. Nonetheless, other classes of phenolic compounds, such as flavonoids, phenolic acids and coumarins, have been identified in various phytochemical studies [4,5]. Their aerial parts have a longstanding traditional use and are employed in the treatment of various ailments, including digestive disorders, inflammatory diseases, bronchitis, malaria, hepatitis and malignant diseases [5,6,7]. Over the past decades, the genus has attracted increasing attention in the field of drug discovery and development; many studies have unveiled its pleiotropic pharmacological profile, which includes anthelmintic, antimalarial, antitubercular, antiviral, antihyperlipidemic, antiemetic, antidepressant, anticancer, antiasthmatic, antihypertensive, antidiabetic, anxiolytic, hepatoprotective, gastroprotective and insecticidal effects [6,8,9,10].
The genus Artemisia is represented by 34 species in the flora of Romania [11], among which A. absinthium L., A. annua L., A. austriaca Jacq., A. pontica L. and A. vulgaris L. A. absinthium L., wormwood, are the best known species, with a wide distribution throughout Europe, North Africa, the Middle East and Asia. Wormwood is an ornamental and medicinal plant and has been used since antiquity as a bitter tonic, choleretic, anthelmintic and wound-healing agent [12]. The aerial parts contain essential oil (up to 1.5%, with marker compounds such as α- and β-thujone, thujyl alcohol, guaiazulene, (Z)-epoxyocimene, sabinyl acetate and chrysantenyl acetate), sesquiterpene lactones (absinthin and its isomers), flavonoids (apigenin, kaempferol, quercetin, artemethin and rutin), phenolic acids (caffeic, chlorogenic, ferulic, gallic, syringic and vanillic acids), coumarins (coumarin and herniarin), tannins, lignans, carotenoids, fatty acids and resins [1,10,12]. To date, numerous studies have unveiled other important bioactivities of A. absinthium aerial parts, e.g., antibacterial, antifungal, antiprotozoal, analgesic, anti-inflammatory, gastroprotective, hepatoprotective, neuroprotective, antidepressant and immunomodulatory properties [12,13,14]. Moreover, alongside its longstanding use in the alcoholic drinks industry (i.e., in the making of vermouth-type wines and absinthe), its current range of applications has rapidly expanded into the cosmetic and food industries [10].
A. annua L., sweet wormwood, is a herbaceous species that inhabits the temperate regions of Asia, Europe, Northern and Southern America, and Australia. The leaves and aerial parts have been traditionally used in Chinese and Hindu medicines as antipyretic agents in the treatment of malaria, tuberculosis and bacterial dysentery, and in the treatment of wounds, hemorrhoids and autoimmune diseases [8]. With the isolation in 1971 of the sesquiterpene lactone artemisinin as the active principle of A. annua against malaria, the species has received an increased interest from the scientific community [15]. Both artemisinin and its semi-synthetic derivatives artemether, arteether and artesunate have been employed clinically in the prophylaxis and treatment of malaria [16]. Alongside sesquiterpene lactones (e.g., artemisinin, arteannuins A–O, artemisitine and artemisinic acid), the aerial parts contain essential oil (up to 4%, with artemisia ketone, cadinene, camphene, camphor, β-caryophyllene and β-pinene as marker compounds), flavonoids (artemetin, casticin, derivatives of apigenin, kaempferol, isorhamnetin, luteolin and quercetin), phenolic acids (caffeic, rosmarinic, quinic and chlorogenic acids), coumarins (coumarin, esculetin, isofraxidine, melilotoside, tomentin, scopoletin and scopolin), polyalkenes, tannins, saponins, phytosterols and fatty acids [7,15,17,18]. Over the past decades, pharmacological studies have confirmed the known traditional uses of A. annua but have also unraveled novel bioactivities, such as anti-inflammatory, analgesic, anticancer, antihypertensive, antimicrobial, antioxidant and nephroprotective properties [17,19,20,21].
A. austriaca Jacq., Austrian wormwood, is a perennial herb found in the semi-arid lands of Central and Eastern Europe, Russia, Iran, Turkey and Northern China [22]. Phytochemical screening studies of aerial parts of A. austriaca have revealed the presence of essential oil (up to 1.1%, mainly comprising camphor, 1,8-cineole and camphene) [10,23], sesquiterpene lactones (arborescin, austricin, hydroxyachillin, artausin, matricarin and santonin) [22,24], flavonoids (cirsilineol, hesperidin, rutin, quercetin and luteolin-7-glycoside) [24,25,26] and phenolic acids (caffeic, chlorogenic and isochlorogenic acids) [27]. A. austriaca herbal extracts are traditionally used for their wound healing, choleretic, anthelmintic, anticonvulsant, anti-inflammatory and hemostatic activities [22,28]. Several studies have unveiled other important pharmacological properties, including antibacterial [29,30], antifungal [24,27,29,31] and antimalarial activities [24].
A. pontica L., Roman wormwood, is a perennial species distributed throughout Southeastern Europe, Siberia and Central Asia [8]. Compared to the allied Artemisia species, the phytochemical profile of A. pontica herbal extracts has not been comprehensively investigated, though several studies have reported the presence of sesquiterpene lactones (artemin, hydroxytaurin and hydroxyeudesmanolides) [8], essential oil (up to 2%, with 1,8-cineole, camphor, artemisia ketone and α-thujone as the main constituents) [1,32,33] and flavonoids (genkwanin and methyl esters of apigenin) [34]. A. pontica aerial parts are traditionally used as a bitter tonic and as sedatives and anthelmintics [33], but recent studies have proved additional beneficial effects, such as anti-inflammatory, analgesic [35], antioxidant [36] and insecticidal properties [37].
A. vulgaris L., common mugwort, is a species growing in the temperate and cold-temperature regions of Asia, Europe and North America, and has been employed as a culinary and medicinal herb [2]. The aerial parts have been traditionally used as a bitter tonic and anti-flatulent in treating gastrointestinal disorders and to alleviate gynecological ailments, such as amenorrhea or dysmenorrhea [38]. The main phytochemicals found in A. vulgaris include essential oil (up to 0.3%, comprising 1,8-cineole, sabinene, β-thujone and caryophyllene oxide as the main constituents), sesquiterpene lactones (vulgarin, psilostachyin and psilostachyin C), flavonoids (derivatives of kaempferol and quercetin), coumarins (coumarin, esculin, scopoletin and umbelliferone), phenolic acids (caffeic and chlorogenic acids), sterols, carotenoids and polyacetylenes [4,9,17]. To date, studies on A. vulgaris have confirmed its known traditional uses and revealed novel significant biological properties, e.g., antioxidant, spasmolytic, antibacterial, antifungal, antinociceptive, hepatoprotective, estrogenic and cytotoxic effects [2,17].
Our study aimed at promoting interest in the Romanian Artemisia species by providing novel insights into their metabolite profiles and bioactivities. To date, most studies have focused on the valorization of the aboveground parts with respect to their analgesic [35], anti-inflammatory [35,39], antimicrobial [40,41,42], antioxidant and cytotoxic properties [39,43,44]. To the best of our knowledge, we report herein for the first time a comprehensive phytochemical characterization of both the roots and aerial parts of five Artemisia spp. (A. absinthium, A. annua, A. austriaca, A. pontica and A. vulgaris) from the spontaneous flora of northeastern Romania by means of liquid chromatography–tandem high-resolution mass spectrometry (LC-HRMS/MS). The biological profile screening was achieved by in vitro testing of antioxidant (free radical scavenging, metal chelating and reducing power, and total antioxidant capacity), enzyme inhibitory (anti-cholinesterase, anti-tyrosinase, anti-amylase and anti-glucosidase) and anti-Mycobacterium activities.

2. Materials and Methods

2.1. Plant Materials and Preparation of Extracts

The aerial parts of the investigated Artemisia species were collected during August–September 2020, while the roots were collected during November–December 2020 from Neamt and Iasi counties, Romania, as follows: A. absinthium—Secuienii Noi (Neamt county, GPS coordinates: 46.843800, 26.887132), A. annua—Carlig (Iasi county, GPS coordinates: 47.195154, 27.567940), A. austriaca—Hadambu (Iasi county, GPS coordinates: 47.014004, 27.440596), A. pontica—Carlig (Iasi county, GPS coordinates: 47.200588, 27.562738), A. vulgaris—Secuienii Noi (Neamt county, GPS coordinates: 46.8408833, 26.8903426). The plant material was collected and authenticated by one of the authors (A.T.) and Dr. Constantin Mardari, Botanic Garden Anastasie Fatu, Iasi, Romania. Voucher specimens (AABH/2020, AANH/2020, AAUH/2020, APH/2020, AVH/2020, AABR/2020, AANR/2020, AAUR/2020, APR/2020 and AVR/2020) were deposited in the Department of Pharmacognosy, Grigore T. Popa University of Medicine and Pharmacy Iasi, Romania. The plant materials (aerial parts and roots collected from the investigated Artemisia spp.) were dried and ground and then 10 g was separately extracted with methanol and chloroform (100 mL) by ultrasonication (3 cycles of 30 min each, at room temperature). The obtained extracts were evaporated to dryness under vacuum (with the yields shown in Table 1) and kept at −20 °C until further analysis.

2.2. Total Phenolic and Flavonoid Content

Total phenolic content (TPC) and total flavonoid content (TFC) were determined as previously described [45,46], with the data provided as mg gallic acid equivalents (GAE)/g extract (TPC) and mg rutin equivalents (RE)/g extract (TFC), respectively.

2.3. LC-HRMS/MS Analysis

The liquid chromatography–tandem high-resolution mass spectrometry (LC-HRMS/MS) analysis of the methanol and chloroform extracts obtained from the roots and aerial parts of the five Artemisia species was carried out on an Agilent 1200 HPLC system (Agilent Technologies, Santa Clara, CA, USA) equipped with auto-sampler (G1329B), degasser (G1379B), binary pump (G1312C), thermostat (G1316A) and ESI-Q-TOF mass spectrometer (G6530B). The chromatographic separations were performed as follows: Phenomenex Gemini C18 column (2 mm × 100 mm, 3 μm); mobile phase 0.1% formic acid in water (A) and 0.1% formic acid in acetonitrile (B); gradient 5–60% B (0–45 min), 95% B (46–55 min); flow rate 0.2 mL/min; injection volume 10 μL. The following MS parameters were used: negative ionization mode; m/z range 100–1000; gas (N2) temperature 275 °C; N2 flow 10 L/min; nebulizer 35 psi; sheath gas temperature 325 °C; sheath gas flow rate 12 L/min; capillary voltage 4000 V; nozzle voltage 1000 V; skimmer 65 V; fragmentor 140 V; collision-induced dissociation energies 10 and 30 V.

2.4. Antioxidant and Enzyme Inhibitory Activity

The 1,1-diphenyl-2-picrylhydrazyl (DPPH) and 2,2′-azino-bis (3-ethylbenzothiazoline) 6-sulfonic acid (ABTS) radical scavenging, cupric ion reducing antioxidant capacity (CUPRAC), ferric ion reducing antioxidant power (FRAP), metal chelating ability (MCA) and phosphomolybdenum (PBD) assays were performed as previously detailed [45,46]. The results were expressed as mg Trolox equivalents (TE)/g for DPPH, ABTS, CUPRAC and FRAP assays, mg EDTA equivalents (EDTAE)/g for the MCA assay and mmol TE/g extract for the PBD assay. The inhibition assays against acetylcholinesterase (AChE), butyrylcholinesterase (BChE), tyrosinase, amylase and glucosidase were caried out as previously described [45,46]. The results were provided as mg galanthamine equivalents (GALAE)/g extract in the AChE and BChE assays, mg kojic acid equivalents [47]/g extract in the tyrosinase assay and mmol acarbose equivalents (ACAE)/g extract in the amylase and glucosidase assays.

2.5. Anti-Mycobacterium Activity

2.5.1. Inoculum Preparation

Mycobacterium tuberculosis H37Ra (ATCC 25177) was grown for two weeks on Löwenstein-Jensen slopes. The collected bacteria were transferred to 7H9-S medium (Middlebrook 7H9 broth supplemented with 10% ADC (albumin–dextrose–catalase) and 0.2% glycerol and vortexed with glass beads (1 mm diameter) for three minutes. After 30 min of room temperature incubation for larger clump sedimentation, the upper phase was transferred to a sterile tube and left for the second sedimentation for 15 min. Next, planktonic bacteria from above the sediment were placed in a fresh tube, with the turbidity adjusted to 0.5 McFarland standard with 7H9-S broth.

2.5.2. MIC Determination

Artemisia extracts were tested in a concentration range of 256 to 16 mg/L. Serial twofold dilutions were prepared in dimethyl sulfoxide (DMSO) using a 7H9-S medium as dilution. The final DMSO concentration did not exceed 1% (v/v) and did not influence the growth of the tested strain. Ethambutol, rifampicin and streptomycin were used as reference standards. Stock solutions were prepared according to the manufacturer’s instructions. Final twofold dilutions from 16 to 0.001 mg/L were prepared in 7H9-S broth. The round bottom micro-well plates were prepared as follows: 50 μL of inoculum and 50 μL of tested substances were added to each well. The sterility, growth and 1% DMSO controls were included. The final density of the inoculum in each well was approximately 5 × 105 CFU/mL. The plates were closed with sealing foil to prevent liquid evaporation and incubated for 8 days at 37 °C. Next, 10 µL of resazurin (Alamar Blue) solution was added to each well, followed by incubation for 48 h at 37 °C and assessment for color development. The minimum inhibitory concentration (MIC) was defined as the lowest drug concentration that prevented a blue to pink color change. The MIC determination was repeated twice. The obtained results were identical.

2.6. Data Analysis

A biological activities dataset was scaled, centered and submitted to principal component analysis (PCA) and hierarchical clustering analysis (HCA). For both PCA and HCA, “Ward’s rule” and “Euclidean distance” were employed for clustering. Afterwards, the biomolecules dataset was logarithm-transformed, scaled, centered and submitted to clustered image maps (CIMs). All multivariate analyses were performed using R v 4.1.2 software (R Foundation for Statistical Computing, Vienna, Austria). The Pearson correlation test was used to examine the relationship between phytoconstituents in tested extracts and biological activities. GraphPad. 9.0 (GraphPad Software, San Diego, CA, USA) was used for the correlation analysis.

3. Results and Discussion

3.1. Total Phenolic and Flavonoid Content

The TPCs and TFCs of the Artemisia extracts were determined using colorimetric methods. The results are given in Table 1. Apparently, methanol extracts contained more phenolics and flavonoids than chloroform in both plant parts. In addition, with the exception of A. pontica, the extracts of the aerial parts were richer than those of the roots in terms of total phenolics. The highest level of phenolics were determined in the methanol aerial part extracts of A. vulgaris, with 106.34 mg GAE/g. After that, the methanol aerial part extracts of A. pontica and A. annua contained significant levels of total phenolics (>50 mg GAE/g). Among the root extracts, the methanol extract from A. annua reached the highest value with 76.34 mg GAE/g, followed by the methanol extracts from A. pontica (65.65 mg GAE/g) and A. austriaca (41.68 mg GAE/g). With regard to the TFCs, the highest content was determined in the methanol extract of A. annua aerial parts with 47.74 mg RE/g, followed by the methanol extracts of A. austriaca (40.30 mg RE/g) and A. vulgaris (39.39 mg RE/g) aerial parts. The lowest level of total flavonoids was found in the chloroform extract of A. absinthium (0.37 mg RE/g). A number of studies have shown a comparable total content of phenolics and flavonoids in the Artemisia genus. For example, in a previous study, Ali et al. [48] found that the total phenolic content in A. absinthium extract was 3.61 mg GAE/g extract, which was lower than our values. In addition, Guo et al. [49] showed that the TPC and TFC in the aqueous extract of A. annua were 39.58 mg GAE/g and 7.04 mg RE/g, respectively. Our findings are also comparable to the results in the literature for other Artemisia species, such as A. copa (155.6 mg GAE/g dry plant in infused extract, reported by Larrazábal-Fuentes et al. [50]), A. vulgaris (117.14 mg GAE/g extract in methanol extract, reported by Jakovljevic et al. [51]), A. alba (110.20 mg GAE/g extract in methanol extract, reported by Jakovljevic et al. [51]) and A. argy (108.56 mg GAE/g extract in methanol extract, reported by Xiao et al. [52]). Although the spectrophotometric methods are widely used in phytochemical studies, recently, most phytochemists have been more concerned with colorimetric methods for assessing bioactive components [53]. This could be explained by the complex nature of phytochemicals and the fact that only specific compounds do not reduce the reagents used in the relevant assays. Given these facts, further chromatographic techniques are needed to evaluate the chemical profiles of plant extracts.

3.2. LC-HRMS/MS Analysis

The methanol and chloroform extracts obtained from the roots and aerial parts of the five Artemisia species were subsequently subjected to an in-depth LC-HRMS/MS analysis. The assignment of the peaks observed in the base peak chromatograms (BPCs) of the extract samples was performed by comparing the spectrometric data with the relevant literature [54,55,56,57,58,59,60,61] or online databases (KNApSacK; METLIN; NIST Chemistry WebBook). The metabolite profiling allowed the annotation of 73 compounds belonging to different phytochemical classes, such as phenolic acids, flavonoids, sesquiterpenes, organic acids, sugars, coumarins, triterpenes, lignans and fatty acids (Table 2, Table S1). In the following sub-sections, a brief description of these categories will be provided, whereas the intra- and interspecies differences will be thoroughly detailed in the Multivariate Analysis Section.
Out of the 15 phenolic acids identified in the Artemisia extracts, 2 were hydroxybenzoic acid derivatives (3 and 4), whereas the remaining were hydroxycinnamic acid derivatives. Of these, chlorogenic acid (9) was confirmed by comparing with the standard, whereas its isomers, neochlorogenic (5) and cryptochlorogenic (8) acids, were tentatively assigned based on their specific HRMS/MS fragments ions described in the literature [58,59]. Chlorogenic acid was identified in all five Artemisia species, especially in the aerial part extracts, whereas the other two isomers were present in all species, except for A. annua. In addition, several other quinic acid congeners were noticed, such as dicaffeoylquinic acids (18, 30 and 32), feruloylquinic acid (16), coumaroylquinic acid (19), coumaroylcaffeoylquinic acid (33) and feruloylcaffeoylquinic acids (35 and 38). Compounds 18, 19 and 33 were noticed only in A. vulgaris, whilst the ferulic acid derivatives (16, 35 and 38) were present in A. absinthium and A. annua. Lastly, two glycosides of caffeic acid (14) and coumaric acid (23) were identified as specific metabolites in A. vulgaris (Table 2, Table S1).
Flavonoids were the representative category of phytochemicals, with 26 different derivatives present exclusively in the extracts obtained from the aerial parts of Artemisia. Formally, they were grouped into free aglycones, O-glycosides and C-glycosides. Besides luteolin (41), identified based on standard injection, and eriodictyol (39), the other aglycones were tentatively identified as poly-hydroxylated/poly-methoxylated flavone derivatives. For instance, eupatolitin (42) was annotated only in A. annua; rhamnetin (43), diosmetin (53) and genkwanin (65) were characteristic of A. pontica; homoeriodictyol (48) was present only in A. austriaca; whereas eupalitin (59) and eupatilin (68) were spotted only in A. vulgaris. On the other hand, dihydroxytrimethoxyflavone (52) and casticin (61) were absent in A. austriaca and A. pontica, respectively. Cirsimaritin (56) and penduletin (57) were characteristic of both A. austriaca and A. pontica, while rhamnazin (50) was specific to A. austriaca, A. pontica and A. vulgaris. With respect to the O-glycosides, the following structures were tentatively proposed: mearnsetin-di-O-hexoside (12) in A. annua, quercetin-di-O-hexoside (22) in A. austriaca, luteolin-O-deoxyhexoside-O-hexoside (28) and rhamnetin-O-hexoside (37) in A. vulgaris, eupatolitin-O-deoxyhexoside-O-hexoside (29) in A. absinthium and eupatolitin-di-O-hexoside (34) and rhamnetin-di-O-hexoside (36) in A. pontica. On the other hand, quercetin-O-deoxyhexoside-O-hexoside (25) and mearnsetin-O-hexoside (26) were absent in A. austriaca, whilst quercetin-O-hexoside (27) was not present in A. annua. Lastly, the two apigenin-C-hexoside-C-pentosides (21 and 24) were specifically observed in A. annua (Table 2, Table S1).
A number of 14 sesquiterpenes have been identified in the aerial part extracts of Artemisia species; besides artemisinin (40), confirmed by standard injection, the other structures were proposed strictly in a tentative manner (Table 2, Table S1). Artemisinin (40) as well as deoxyartemisinins (45 and 47), pseudosantonin (54), artemisin C (58), arteannuin B (60), dihydroarteannuin B (64) and dihydrosantamarin (66) were specifically noticed in A. annua. Chrysartemins A (13) and B (15) were the only two sesquiterpenes in A. austriaca, whereas artabsinolide A (17) was the sole congener in A. absinthium. Artecanin hydrate (20) and cnicin (62) were characteristic of A. pontica, whilst santonin (46) was found exclusively in A. vulgaris.
Quinic acid (1) and sucrose (2) were two non-specific metabolites identified in all five Artemisia species. On the contrary, the five triterpenes, namely, absinthin (55), artenolide (63), isoabsinthin (70) and two absinthin derivatives (67 and 71), were characteristically noticed in A. absinthium. Three coumarins, such as esculetin (7) and two of its hexosides (6 and 11), were observed in A. vulgaris; additionally, one of the two esculetin-O-hexosides was also present in the roots of A. austriaca. Tracheloside, a glycosylated lignan, was putatively labeled in the aerial parts of A. annua, A. austriaca and A. vulgaris. Lastly, seven oxygenated fatty acids were assigned: trihydroxyoctadecenoic acid I (49), hydroxyoctadecatrienoic acid (72), hydroxyoctadecadienoic acid (73) and tuberonic acid-O-hexoside (10) in all Artemisia species; trihydroxyoctadecadienoic acid (44) in A. absinthium, A. austriaca and A. vulgaris; trihydroxyoctadecenoic acid II (51) in A. absinthium and A. vulgaris; and hydroperoxyoctadecadienoic acid (69) in A. annua and A. austriaca (Table 2, Table S1).

3.3. Antioxidant Activity

Antioxidant compounds are of increasing interest in the pharmaceutical and nutraceutical fields. These compounds provide powerful shields against free radicals, and a negative correlation between their consumption and the prevalence of chronic and degenerative diseases has been reported. In this sense, phytochemicals are considered a great treasure trove of antioxidants, and many compounds found in plants have been identified as natural and safe antioxidants. In the light of these facts, attempts were made to determine whether the tested Artemisia species are a source of natural antioxidants. Various chemical assays were performed, including radical quenching (ABTS and DPPH), reducing power (CUPRAC and FRAP), metal chelation and phosphomolybdenum assays. The results are shown in Table 3. Non-biological radicals, such as DPPH and ABTS, are commonly used in in vitro experiments to assess the abilities of plant extracts to scavenge radicals. From Table 3, the methanol extracts showed stronger radical scavenging abilities than the chloroform extracts in both parts. The best radical scavenging ability was found in the methanol extract of A. annua roots (DPPH: 237.03 mg TE/g; ABTS: 240.78 mg TE/g), followed by the methanol extracts of A. pontica roots (DPPH: 179.63 mg TE/g; ABTS: 176.12 mg TE/g) and A. vulgaris aerial parts (DPPH: 139.56 mg TE/g; ABTS: 173.86 mg TE/g) in both assays. The weakest radical scavenging ability was recorded in the chloroform extract of A. absinthium roots (DPPH: 5.11 mg TE/g; ABTS: 7.54 mg TE/g). The term “reducing power” refers to the ability of antioxidant compounds to donate electrons. For this purpose, CUPRAC and FRAP assays involving the conversion of Cu+2 to Cu+ and Fe+3 to Fe+2, respectively, were performed. In both plant parts, the methanol extracts had higher reducing potentials than the chloroform extracts. The methanol extracts of A. vulgaris aerial parts (498.32 mg TE/g), A. annua roots (438.43 mg TE/g) and A. pontica aerial parts (290.14 mg TE/g) exhibited the highest CUPRAC activities. With one small exception, the methanol extracts of A. annua roots (294.52 mg TE/g), A. vulgaris aerial parts (198.51 mg TE/g) and A. pontica roots (165.55 mg TE/g) demonstrated the highest levels of capability in the FRAP assay. The obtained results from the free radical scavenging and reducing power assays are almost consistent with the total phenolic results for the extracts. In this sense, the phenolic components in the extracts can be considered as the main contributors to the free radical scavenging and reducing abilities. Similar to our findings, several researches [62,63] reported a strong correlation between total phenolic content and antioxidant properties. However, we observed different results for metal chelation abilities. The chelating abilities of plant extracts may reflect the inhibition of hydroxyl radicals’ production in the Fenton reaction. The best metal chelating ability (MCA) was found in the methanol extract of A. pontica roots with 22.93 mg EDTAE/g extract. Intriguingly, in three Artemisia species tested, the chloroform extracts from the aerial parts had higher potentials than the methanol extracts (A. annua, A. pontica and A. vulgaris). In addition, two chloroform extracts of the roots (A. annua and A. vulgaris) showed no metal chelating ability. Non-phenolic chelators, such as polysaccharides, peptides or sulfides, may be responsible for the conflicting results. In support of our findings, several investigators reported a weak correlation between total phenolic content and metal chelating ability [64,65]. Some researchers also pointed out that the chelating ability of phenolics contributes only in a small extent to the antioxidant properties of plant extracts [66]. The phosphomolybdenum assay is related to the reduction of Mo (VI) to (Mo (V) by antioxidant compounds and is considered one of the total antioxidant capacity assays. As can be seen in Table 3, we observed different results for each species. For example, the aerial part extracts from two Artemisia species (A. absinthium and A. vulgaris) exhibited greater potentials than root extracts. In addition, the root extracts from two species (A. annua and A. pontica) showed stronger activity as compared to aerial parts. In the literature, several researchers have reported that the results from phosphomolybdenum assays exhibited weak correlations with total phenolic content [67,68]. This fact could be explained by the presence of non-phenolic antioxidants, such as tocopherol, ascorbic acid and terpenoids. Artemisia members have been previously found to possess interesting antioxidant properties. For example, in a recent study by Minda et al. [69], the DPPH radical scavenging abilities of three Artemisia species (A. absinthium, A. dracunculus and A. annua) were investigated, the materials exhibiting more than 90% scavenging ability at a concentration of 1000 µg/mL. In another study conducted by Kamarauskaite et al. [70], the fractions of A. absinthium and A. ludoviciana were assessed by ABTS and FRAP assays and their values were found to be 367–1693 µM TE/g and 5385–6952 µM TE/g, respectively. Ferrante et al. [71] also investigated the antioxidant properties of A. santonicum methanol extract (DPPH: 278.57 mg TE/g; ABTS: 217.60 mg TE/g; CUPRAC: 515.30 mg TE/g; FRAP: 255.35 mg TE/g; metal chelating: 21.96 mg EDTAE/g and phosphomolybdenum: 2.20 mmol TE/g). Other examples of Artemisia species whose antioxidant capacities have been determined in the literature include A. lactiflora [47], A. indica [72], A. santolinifolia [73] and A. monosperma [74]. Based on the solvents, plant parts, and species triangle, our results may provide new information on the antioxidant properties of Artemisia species.

3.4. Enzyme Inhibitory Activity

Enzyme inhibition is a concept that is currently gaining traction in the treatment of various global health problems, such as type 2 diabetes, obesity and Alzheimer’s disease. This phenomenon demonstrates that the inhibition of specific enzymes can be a highly effective therapeutic strategy to alleviate disease symptoms [75]. Amylase and glucosidase, for example, are thought to be important players in the management of blood glucose levels in diabetics [76]. Furthermore, acetylcholinesterase inhibitors may improve memory function in Alzheimer’s patients by increasing acetylcholine levels in the synapses [77]. Enzyme inhibitors, therefore, are being sought as a safe and effective way to treat the diseases listed above. In this sense, plants are considered excellent treasures [78]. Given these facts, we looked into the tested Artemisia species’ enzyme inhibitory properties. The results are summarized in Table 4. The best AChE inhibition was determined in the methanol extract of A. absinthium with 3.02 mg GALAE/g, followed by the chloroform extracts of A. absinthium (2.50 mg GALAE/g) and A. annua (2.36 mg GALAE/g) aerial parts. With regard to BChE inhibition, the chloroform root extracts (A. annua, A. austriaca and A. absinthium) were recorded as the strongest extracts. With the exception of A. vulgaris roots, the chloroform extracts were more active against BChE than the methanol extracts in all tested Artemisia species. Two methanol root extracts (A. austriaca and A. pontica) were not active on BChE. As can be seen from Table 4, tyrosinase inhibitory effects were higher in the methanol extracts as compared to the chloroform extracts, except for A. vulgaris aerial parts. The most active methanol extracts were A. annua (49.42 mg KAE/g), A. austriaca (47.27 mg KAE/g) and A. pontica (44.91 mg KAE/g). The weakest tyrosinase inhibition potential was found in the chloroform extract of A. austriaca roots with 13.16 mg KAE/g. For all parts and species, the chloroform extracts had stronger amylase inhibitory effects than the methanol extracts. The best amylase inhibitory effects were recorded in the chloroform extracts of A. austriaca parts (root: 0.57 mmol ACAE/g; aerial parts: 0.54 mmol ACAE/g). As for glucosidase inhibitory activity, the aerial parts of all Artemisia species showed stronger abilities than the root extracts, and the best ability was obtained by the methanol extract of A. vulgaris aerial parts (11.32 mmol ACAE/g). The methanol root extract of A. austriaca had the weakest glucosidase inhibitory effect (0.16 mmol ACAE/g). To the best of our knowledge, scientific information on enzyme inhibitory properties of the members of the genus Artemisia is scarce [56,71,79,80,81,82,83]. In this sense, the present work could provide further insights into the future application of Artemisia species as natural sources of enzyme inhibitory agents.

3.5. Anti-Mycobacterium Activity

The emergence of multidrug-resistant Mycobacterium tuberculosis strains represents a major barrier to tuberculosis eradication, leading to longer treatment regimens, higher toxicity and even treatment failure [84]. Thus, there is an urgent demand to explore novel drugs and combinations to improve tuberculosis therapy. Recent work has presented the antimalarial drug artemisinin as a promising antitubercular agent [85,86]. Moreover, Martini et al. found that dichloromethane extracts from leaves of A. annua L. and A. afra Jacq. ex Willd. displayed even higher anti-mycobacterial effects than sesquiterpene artemisinin [87]. In the present study, M. tuberculosis H37Ra was exposed to chloroform and methanol extracts obtained from the roots and aerial parts of five Artemisia species. Within the tested concentration range, most of the samples were active against the mycobacterial strain, except for the methanol root extracts of A. absinthium, A. austriaca and A. vulgaris (Table 5). The chloroform extract of A. austriaca aerial parts showed the highest anti-mycobacterial effect (MIC = 64 mg/L), while the other active extracts displayed similar degrees of potency, with MICs of 128–256 mg/L. Considering that plant extracts can be categorized as having strong activity when their MIC is within the range of 50–500 mg/L, moderate activity with an MIC of 500–1500 mg/L and weak activity with an MIC above 1500 mg/L [88], it can be stated that the Artemisia samples possess strong anti-Mycobacterium effects. To the best of our knowledge, previous studies referred only to the anti-mycobacterial potential of Artemisia herbal extracts. We report herein for the first time on the anti-mycobacterial activity of Artemisia root extracts and that their potency was found to be similar to that of the aerial part extracts. Our results are in agreement with the study of Bhowmick et al. [89] which showed the inhibitory effects of an hexane extract from A. annua aerial parts against Mycobacterium smegmatis (MIC range: 250–1000 mg/L). Further, the bioactivity-linked fractionation of the extract revealed that its inhibitory effects were due to the sesquiterpenes deoxyartemisinin and artemisinic acid [89]. Regarding our Artemisia species, the LC-MS/MS analysis showed that chrysartemin A and B were present only in the most active sample—A. austriaca herbal extract (Table S1). These two compounds, which were previously reported in other Artemisia species (A. mexicana and A. klotzchiana) [90], belong to the guaianolide-type sesquiterpenes known for their inhibitory effects against Mycobacterium strains [90,91]. Therefore, we can conclude that sesquiterpenes found in Artemisia extracts may contribute to their overall anti-mycobacterial activity; moreover, it cannot be excluded that the synergistic effects among different classes of constituents identified in our samples (e.g., sesquiterpenes, phenolic acids, flavonoids, coumarins, fatty acids) might explain the observed effects.

3.6. Multivariate Analysis

A Pearson correlation analysis of bioactive compounds and biological activities was performed. The correlation heatmap is shown in Figure 1. Clearly, total phenolic content was highly correlated (R > 0.8) with scavenging and reducing abilities. However, the metal chelation and phosphomolybdenum assays correlated moderately with total phenolic levels. This fact could be explained by the presence of non-phenolic chelators (polypeptides, sulfides, etc.) or antioxidants (vitamin C, tocopherols, etc.). As can be seen in Figure 1, the individual compounds exhibited different correlations with the biological activities. Dihydroxybenzoic acid hexoside (3) and feruloylcaffeoylquinic acid I (35) were strongly correlated with free radical scavenging and reducing power abilities. However, different compounds exhibited a linear correlation in the metal chelation and phosphomolybdenum assays. The main players were dihydroxytetramethoxyflavone (61) and chlorogenic acid (9) in the metal chelation assay. In the phosphomolybdenum assay, artabsinolide A (17) and absinthin derivative II (71) were found to be the main contributors. With regard to the enzyme inhibition assays, different compounds acted as inhibiting agents in each assay. Tuberonic acid-O-hexoside (10) was moderately correlated with cholinesterases (AChE and BChE). For tyrosinase inhibition, dicaffeoylquinic acid derivatives (30 and 32) had low correlation values (R < 0.4). Dihydroxytrimethoxyflavone (52) and dihydroxytetramethoxyflavone (61) were closely associated with glucosidase inhibition assay (R > 0.7). Two sesquiterpenes (chrysartemin A (13) and B (15)) showed stronger correlations with anti-Mycobacterium ability compared to other compounds. Taken together, the tested extracts have great potential as natural sources of bioactive agents and could therefore be considered as valuable raw materials with pharmaceutical and nutraceutical applications.
The results of the PCA for the antioxidant and enzyme inhibitory activities of Artemisia species are presented in Figure 2. Firstly, the screening of the eigenvalues suggested that the first three principal components (PCs) were sufficient to synthesize most of the data variation. Indeed, these components manifested a variance of 45%, 21% and 11%, respectively. The first PC represented the variation in the antioxidant activities, since it was predominantly and negatively linked with both radical scavenging (ABTS and DPPH) and reducing power (FRAP and CUPRAC) (Figure 2A). The second PC discriminated the samples based on their anti-glucosidase, phosphomolybdenum, anti-amylase and anti-BChE activities (Figure 2B), as it was significantly and positively bound to the mentioned bioactivities. The third PC separated the samples in terms of their metal chelating capacity and anti-tyrosinase activities (Figure 2C). Examination of the scatter plots (PC1 vs. PC2, PC1 vs. PC3 and PC2 vs. PC3) reported in Figure 2D–F evidenced considerable variability among the samples. Three groups seem to emerge in the first scatter plot (PC1 vs. PC2) (Figure 2D); the same trend is noticed in the second scatter plot (PC1 vs. PC3); however, the samples representing the three groups were different from those obtained previously (Figure 2E). In the third scatter plot (PC2 vs. PC3), no clear clusters were identified (Figure 2F). For a better identification of the different groups, HCA was applied by using the coordinates of the samples on the three dimensions of the PCA. By using Ward’s method and Euclidean distance, we obtained three main clusters (Figure 3). Among the clusters, the samples representing cluster 1 (i.e., the methanol root extracts of A. annua, A. pontica and aerial part extracts of A. vulgaris) were characterized by the highest radical scavenging (ABTS and DPPH) and reducing power (FRAP and CUPRAC) activities.
Next, to determine the phytochemical differences between the studied Artemisia species, CIM analysis with respect to the phytochemical compounds dataset was carried out. The extracts were separated into two large clusters, namely, the roots on the one hand and the aerial parts on the other (Figure 4). Overall, several compounds were more abundant in the extracts from the aerial parts than in the extracts from the roots. Furthermore, in cluster 2, which represented all the aerial part samples, the methanol and chloroform extracts of all species were very similar. On the other hand, in cluster 2, A. annua extracts were similar to A. absinthium extracts and A. pontica extracts were similar to A. austriaca extracts, whereas the A. vulgaris aerial part extracts represented a distinct cluster. Similarly, in cluster 1, both A. vulgaris root extracts were found to be clearly different as compared to the other extracts. These findings suggest that A. vulgaris distinguished itself from the other four Artemisia species investigated in the current work. Moreover, some compounds of A. vulgaris (i.e., caffeic acid-O-pentoside, esculetin-O-hexoside I, coumaroylquinic acid, coumaric acid-O-pentoside, luteolin-O-deoxyheoside-O-hexoside and coumaroylcaffeoylquinic acid) could be used as potential markers for this species, due to their abundance in the aerial parts of A. vulgaris.

4. Conclusions

In this work, five Artemisia species collected from the spontaneous flora of northeastern Romania, namely, A. absinthium, A. annua, A. austriaca, A. pontica and A. vulgaris, were comprehensively investigated with respect to their phytochemical profiles and multi-biological potential (antioxidant, enzyme inhibitory and anti-mycobacterial). The LC-HRMS/MS-based metabolite profiling allowed the annotation of 73 different compounds, of which 15 were phenolic acids (i.e., chlorogenic, neochlorogenic, dicaffeoylquinic, feruloylquinic, coumaroylquinic acids), 26 were flavonoids (i.e., as poly-hydroxylated/poly-methoxylated flavones) and 14 were sesquiterpenes (i.e., artemisinin, pseudosantonin, arteannuin B). CIM analysis of the phytochemical profile revealed three main clusters, the first comprising A. annua together with A. absinthium, the second A. pontica together with A. austriaca and the third A. vulgaris. The antioxidant activity analysis of the five species revealed the superior antioxidant activity of the aerial part extracts as compared to the root extracts, as well as the better antioxidant activity of the methanol extracts as compared to the chloroform extracts. Furthermore, PCA and HCA allowed us to differentiate the samples into three main clusters with respect to antioxidant and enzyme inhibitory potential, with one cluster (cluster 1—the methanol root extracts of A. annua and A. pontica and the aerial part extracts of A. vulgaris) being characterized by the highest radical scavenging (ABTS and DPPH) and reducing power (FRAP and CUPRAC) activities. In addition, the chloroform extract of A. austriaca aerial parts showed the highest antibacterial effects against M. tuberculosum H37Ra (MIC = 64 mg/L), while other extracts displayed MIC values of 128–256 mg/L. Aside from the chemotaxonomic importance, the current study makes significant contributions to knowledge of the chemical and versatile biological profile of the investigated Artemisia ssp. collected from Romanian flora. Overall, our research could open prospects for the large-scale exploitation of Artemisia species (both roots and aerial parts) as rich sources of bioactive metabolites endowed with interesting antioxidant, enzyme inhibitory and anti-mycobacterial properties.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/antiox11051017/s1, Table S1. Full spectro-chromatographic data of compounds identified in the Artemisia root and aerial parts extracts by LC-HRMS/MS.

Author Contributions

Conceptualization, A.T., S.V.L., E.S. and G.Z.; methodology, A.T., S.V.L., G.Z., K.S.-W., E.S. and R.S.; software, K.I.S.; validation, S.V.L. and K.S.-W.; formal analysis, A.T., S.V.L., G.Z. and K.I.S.; investigation, A.T., S.V.L., G.Z., K.I.S., E.S., M.M.-T. and R.S.; resources, A.T., S.V.L., G.Z. and E.S.; data curation, S.V.L. and K.I.S.; writing—original draft preparation, A.T., S.V.L., G.Z. and K.I.S..; writing—review and editing, E.S., M.M-T., R.S. and K.S.-W.; visualization, A.T., E.S. and S.V.L.; supervision, A.T., S.V.L., G.Z., K.S.-W. and E.S.; project administration, A.T. and S.V.L.; funding acquisition, A.T. and S.V.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We kindly acknowledge Constantin Mardari (Botanic Garden Anastasie Fatu Iasi, Romania) for help with the identification of Artemisia species. We thank Aleksandra Zaroda for help with the anti-mycobacterial testing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abad, M.J.; Bedoya, L.M.; Apaza, L.; Bermejo, P. The Artemisia L. genus: A review of bioactive essential oils. Molecules 2012, 17, 2542–2566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Nigam, M.; Atanassova, M.; Mishra, A.P.; Pezzani, R.; Devkota, H.P.; Plygun, S.; Salehi, B.; Setzer, W.N.; Sharifi-Rad, J. Bioactive compounds and health benefits of Artemisia species. Nat. Prod. Com. 2019, 14, 1934578X19850354. [Google Scholar]
  3. Pellicer, J.; Garcia, S.; Canela, M.; Garnatje, T.; Korobkov, A.A.; Twibell, J.; Vallès, J. Genome size dynamics in Artemisia L. (Asteraceae): Following the track of polyploidy. Plant Biol. 2010, 12, 820–830. [Google Scholar] [CrossRef] [PubMed]
  4. Ali, M.; Abbasi, B.H.; Ahmad, N.; Khan, H.; Ali, G.S. Strategies to enhance biologically active-secondary metabolites in cell cultures of Artemisia–current trends. Crit. Rev. Biotechnol. 2017, 37, 833–851. [Google Scholar] [CrossRef]
  5. Turi, C.E.; Shipley, P.R.; Murch, S.J. North American Artemisia species from the subgenus Tridentatae (Sagebrush): A phytochemical, botanical and pharmacological review. Phytochemistry 2014, 98, 9–26. [Google Scholar] [CrossRef]
  6. Bisht, D.; Kumar, D.; Kumar, D.; Dua, K.; Chellappan, D.K. Phytochemistry and pharmacological activity of the genus Artemisia. Arch. Pharm. Res. 2021, 44, 439–474. [Google Scholar] [CrossRef]
  7. Taleghani, A.; Emami, S.A.; Tayarani-Najaran, Z. Artemisia: A promising plant for the treatment of cancer. Bioorg. Med. Chem. 2020, 28, 115180. [Google Scholar] [CrossRef]
  8. Trendafilova, A.; Moujir, L.M.; Sousa, P.M.; Seca, A.M. Research advances on health effects of edible Artemisia species and some sesquiterpene lactones constituents. Foods 2020, 10, 65. [Google Scholar] [CrossRef]
  9. Bora, K.S.; Sharma, A. The genus Artemisia: A comprehensive review. Pharm. Biol. 2011, 49, 101–109. [Google Scholar] [CrossRef] [Green Version]
  10. Pandey, A.K.; Singh, P. The genus Artemisia: A 2012–2017 literature review on chemical composition, antimicrobial, insecticidal and antioxidant activities of essential oils. Medicines 2017, 4, 68. [Google Scholar] [CrossRef] [Green Version]
  11. Sârbu, I.; Ştefan, N.; Oprea, A. Plante Vasculare din România: Determinator Ilustrat de Teren; Editura Victor B Victor: Bucuresti, Romania, 2013; pp. 810–816. [Google Scholar]
  12. Szopa, A.; Pajor, J.; Klin, P.; Rzepiela, A.; Elansary, H.O.; Al-Mana, F.A.; Mattar, M.A.; Ekiert, H. Artemisia absinthium L.—Importance in the history of medicine, the latest advances in phytochemistry and therapeutical, cosmetological and culinary uses. Plants 2020, 9, 1063. [Google Scholar] [CrossRef] [PubMed]
  13. Bordean, M.-E.; Muste, S.; Marțiș, G.S.; Mureșan, V.; Buican, B.-C. Health effects of wormwood (Artemisia absinthium L.): From antioxidant to nutraceutical. J. Agroalim. Proc. Technol. 2021, 27, 211–218. [Google Scholar]
  14. Batiha, G.E.-S.; Olatunde, A.; El-Mleeh, A.; Hetta, H.F.; Al-Rejaie, S.; Alghamdi, S.; Zahoor, M.; Magdy Beshbishy, A.; Murata, T.; Zaragoza-Bastida, A. Bioactive compounds, pharmacological actions, and pharmacokinetics of wormwood (Artemisia absinthium). Antibiotics 2020, 9, 353. [Google Scholar] [CrossRef] [PubMed]
  15. Feng, X.; Cao, S.; Qiu, F.; Zhang, B. Traditional application and modern pharmacological research of Artemisia annua L. Pharmacol. Ther. 2020, 216, 107650. [Google Scholar] [CrossRef] [PubMed]
  16. Cui, L.; Su, X.-Z. Discovery, mechanisms of action and combination therapy of artemisinin. Expert Rev. Anti Infect. Ther. 2009, 7, 999–1013. [Google Scholar] [CrossRef]
  17. Ekiert, H.; Pajor, J.; Klin, P.; Rzepiela, A.; Ślesak, H.; Szopa, A. Significance of Artemisia vulgaris L. (Common Mugwort) in the history of medicine and its possible contemporary applications substantiated by phytochemical and pharmacological studies. Molecules 2020, 25, 4415. [Google Scholar] [CrossRef]
  18. Septembre-Malaterre, A.; Lalarizo Rakoto, M.; Marodon, C.; Bedoui, Y.; Nakab, J.; Simon, E.; Hoarau, L.; Savriama, S.; Strasberg, D.; Guiraud, P. Artemisia annua, a traditional plant brought to light. Int. J. Mol. Sci. 2020, 21, 4986. [Google Scholar] [CrossRef]
  19. Efferth, T. From ancient herb to modern drug: Artemisia annua and artemisinin for cancer therapy. In Seminars in Cancer Biology; Academic Press: Cambridge, MA, USA, 2017; pp. 65–83. [Google Scholar]
  20. Sadiq, A.; Hayat, M.Q.; Ashraf, M. Ethnopharmacology of Artemisia annua L.: A review. In Artemisia Annua-Pharmacology and Biotechnology; Aftab, T., Ferreira, J.F.S., Khan, M.M.A., Naeem, M., Eds.; Springer: Berlin/Heidelberg, Germany, 2014; pp. 9–25. [Google Scholar]
  21. Willcox, M. Artemisia species: From traditional medicines to modern antimalarials—And back again. J. Altern. Complement. Med. 2009, 15, 101–109. [Google Scholar] [CrossRef] [Green Version]
  22. Adekenov, S. Chemical study of Artemisia austriaca Jacq. Int. J. Biol. Chem. 2021, 14, 156–163. [Google Scholar] [CrossRef]
  23. Costa, R.; De Fina, M.R.; Valentino, M.R.; Rustaiyan, A.; Dugo, P.; Dugo, G.; Mondello, L. An investigation on the volatile composition of some Artemisia species from Iran. Flavour Fragr. J. 2009, 24, 75–82. [Google Scholar] [CrossRef]
  24. Kikhanova, Z.S.; Iskakova, Z.B.; Dzhalmakhanbetova, R.; Seilkhanov, T.; Ross, S.; Suleimen, E. Constituents of Artemisia austriaca and their biological activity. Chem. Nat. Compd. 2013, 49, 967–968. [Google Scholar] [CrossRef]
  25. Cubukcu, B.; Melikoğlu, G. Flavonoids of Artemisia austriaca. Planta Med. 1995, 61, 488. [Google Scholar] [CrossRef] [PubMed]
  26. Valant-Vetschera, K.M.; Wollenweber, E. Flavonoid aglycones from the leaf surfaces of some Artemisia spp. (Compositae-Anthemideae). Z. Naturforsch. C 1995, 50, 353–357. [Google Scholar] [CrossRef]
  27. Ivashchenko, I. Fungicidal properties of artemisia aromatic plants towards Fusarium oxysporum. Ukr. J. Ecol. 2015, 5, 44. [Google Scholar]
  28. Mercan, U.; Öner, A.C.; Öntürk, H.; Cengiz, N.; Erten, R.; Özgökç, F.; Özbek, H. Investigation of acute liver toxicity and anti-inflammatory effects of Artemisia austriaca J. Jacq. Pharmacologyonline 2008, 1, 131–138. [Google Scholar]
  29. Altunkaya, A.; Yıldırım, B.; Ekici, K.; Terzioğlu, Ö. Determining essential oil composition, antibacterial and antioxidant activity of water wormwood extracts. GIDA 2014, 39, 17–24. [Google Scholar]
  30. Smirnova, G.; Samoilova, Z.; Muzyka, N.; Oktyabrsky, O. Influence of plant polyphenols and medicinal plant extracts on antibiotic susceptibility of Escherichia coli. J. Appl. Microbiol. 2012, 113, 192–199. [Google Scholar] [CrossRef]
  31. Yiğit, D.; Yiğit, N.; Özgen, U. An investigation on the anticandidal activity of some traditional medicinal plants in Turkey. Mycoses 2009, 52, 135–140. [Google Scholar] [CrossRef]
  32. Kiplimo, J.J.; Chichira, D.; Kaiyonia, B.; Yugi, J.; Kebenei, S.J. Chemical composition and larvicidal activity of essential oil of Artemisia pontica. J. Afr. Res. Dev. 2016, 1, 1–7. [Google Scholar]
  33. Bos, R.; Stojanova, A.S.; Woerdenbag, H.J.; Koulman, A.; Quax, W.J. Volatile components of the aerial parts of Artemisia pontica L. grown in Bulgaria. Flav. Fragr. J. 2005, 20, 145–148. [Google Scholar] [CrossRef]
  34. Talzhanov, N.; Sadyrbekov, D.; Smagulova, F.; Mukanov, R.; Raldugin, V.; Shakirov, M.; Tkachev, A.; Atazhanova, G.; Tuleuov, B.; Adekenov, S. Components of Artemisia pontica. Chem. Nat. Compd. 2005, 41, 178–181. [Google Scholar] [CrossRef]
  35. Ivanescu, B.; Corciova, A.; Vlase, L.; Gheldiu, A.M.; Miron, A.; Ababei, D.C.; Bild, V. Analgesic and anti-inflammatory activity of Artemisia extracts on animal models of nociception. Balneo PRM Res. J. 2021, 12, 34–39. [Google Scholar] [CrossRef]
  36. Nikolova, M.; Gussev, C.; Nguyen, T. Evaluation of the Antioxidant action and flavonoid composition of Artemisia species extracts. Biotechnol. Biotechnol. Equip. 2010, 24, 101–103. [Google Scholar] [CrossRef] [Green Version]
  37. Derwich, E.; Benziane, Z.; Boukir, A. Chemical compositions and insectisidal activity of essential oils of three plants Artemisia sp: Artemisia herba-alba, Artemisia absinthium and Artemisia pontica (Morocco). Elec. J. Env. Agricult. Food Chem. 2009, 8, 1202–1211. [Google Scholar]
  38. Bisset, N.G.; Wichtl, M. Herbal Drugs and Phytopharmaceuticals: A Handbook for Practice on a Scientific Basis; Medpharm GmbH Scientific Publishers Stuttgart, CRC Press: Boca Raton, FL, USA, 1994; pp. 88–89. [Google Scholar]
  39. Moacă, E.-A.; Pavel, I.Z.; Danciu, C.; Crăiniceanu, Z.; Minda, D.; Ardelean, F.; Antal, D.S.; Ghiulai, R.; Cioca, A.; Derban, M. Romanian wormwood (Artemisia absinthium L.): Physicochemical and nutraceutical screening. Molecules 2019, 24, 3087. [Google Scholar] [CrossRef] [Green Version]
  40. Marinas, I.C.; Oprea, E.; Chifiriuc, M.C.; Badea, I.A.; Buleandra, M.; Lazar, V. Chemical composition and antipathogenic activity of Artemisia annua essential oil from Romania. Chem. Biodiv. 2015, 12, 1554–1564. [Google Scholar] [CrossRef] [PubMed]
  41. Badea, M.L.; Delian, E. In vitro antifungal activity of the essential oils from Artemisia spp. L. on Sclerotinia sclerotiorum. Rom. Biotechnol. Lett. 2014, 19, 9345–9352. [Google Scholar]
  42. Obistioiu, D.; Cristina, R.T.; Schmerold, I.; Chizzola, R.; Stolze, K.; Nichita, I.; Chiurciu, V. Chemical characterization by GC-MS and in vitro activity against Candida albicans of volatile fractions prepared from Artemisia dracunculus, Artemisia abrotanum, Artemisia absinthium and Artemisia vulgaris. Chem. Cent. J. 2014, 8, 6. [Google Scholar] [CrossRef] [Green Version]
  43. Craciunescu, O.; Constantin, D.; Gaspar, A.; Toma, L.; Utoiu, E.; Moldovan, L. Evaluation of antioxidant and cytoprotective activities of Arnica montana L. and Artemisia absinthium L. ethanolic extracts. Chem. Cent. J. 2012, 6, 97. [Google Scholar] [CrossRef] [Green Version]
  44. Stan, R.L.; Sevastre, B.; Ionescu, C.; Olah, N.K.; Vicaș, L.G.; Páll, E.; Moisa, C.; Hanganu, D.; Sevastre-Berghian, A.C.; Andrei, S. Artemisia annua L. extract: A new phytoproduct with sod-like and antitumour activity. Farmacia 2020, 68, 812–821. [Google Scholar] [CrossRef]
  45. Uysal, S.; Zengin, G.; Locatelli, M.; Bahadori, M.B.; Mocan, A.; Bellagamba, G.; De Luca, E.; Mollica, A.; Aktumsek, A. Cytotoxic and enzyme inhibitory potential of two Potentilla species (P. speciosa L. and P. reptans Willd.) and their chemical composition. Front. Pharmacol. 2017, 8, 290. [Google Scholar] [CrossRef] [PubMed]
  46. Grochowski, D.M.; Uysal, S.; Aktumsek, A.; Granica, S.; Zengin, G.; Ceylan, R.; Locatelli, M.; Tomczyk, M. In vitro enzyme inhibitory properties, antioxidant activities, and phytochemical profile of Potentilla thuringiaca. Phytochem. Lett. 2017, 20, 365–372. [Google Scholar] [CrossRef]
  47. Ali, M.; Iqbal, R.; Safdar, M.; Murtaza, S.; Mustafa, M.; Sajjad, M.; Bukhari, S.A.; Huma, T. Antioxidant and antibacterial activities of Artemisia absinthium and Citrus paradisi extracts repress viability of aggressive liver cancer cell line. Mol. Biol. Rep. 2021, 48, 7703–7710. [Google Scholar] [CrossRef] [PubMed]
  48. Guo, S.; Ma, J.; Xing, Y.; Xu, Y.; Jin, X.; Yan, S.; Shi, B. Artemisia annua L. aqueous extract as an alternative to antibiotics improving growth performance and antioxidant function in broilers. Ital. J. Anim. Sci. 2020, 19, 399–409. [Google Scholar] [CrossRef] [Green Version]
  49. Larrazábal-Fuentes, M.J.; Fernández-Galleguillos, C.; Palma-Ramírez, J.; Romero-Parra, J.; Sepúlveda, K.; Galetovic, A.; González, J.; Paredes, A.; Bórquez, J.; Simirgiotis, M.J. Chemical profiling, antioxidant, anticholinesterase, and antiprotozoal potentials of Artemisia copa Phil. (Asteraceae). Front. Pharmacol. 2020, 11, 1911. [Google Scholar] [CrossRef]
  50. Jakovljević, M.R.; Grujičić, D.; Vukajlović, J.T.; Marković, A.; Milutinović, M.; Stanković, M.; Vuković, N.; Vukić, M.; Milošević-Djordjević, O. In vitro study of genotoxic and cytotoxic activities of methanol extracts of Artemisia vulgaris L. and Artemisia alba Turra. S. Afr. J. Bot. 2020, 132, 117–126. [Google Scholar] [CrossRef]
  51. Xiao, J.-Q.; Liu, W.-Y.; Sun, H.-p.; Li, W.; Koike, K.; Kikuchi, T.; Yamada, T.; Li, D.; Feng, F.; Zhang, J. Bioactivity-based analysis and chemical characterization of hypoglycemic and antioxidant components from Artemisia argyi. Bioorg. Chem. 2019, 92, 103268. [Google Scholar] [CrossRef]
  52. Sánchez-Rangel, J.C.; Benavides, J.; Heredia, J.B.; Cisneros-Zevallos, L.; Jacobo-Velázquez, D.A. The Folin–Ciocalteu assay revisited: Improvement of its specificity for total phenolic content determination. Anal. Methods 2013, 5, 5990–5999. [Google Scholar] [CrossRef]
  53. Suberu, J.; Gromski, P.S.; Nordon, A.; Lapkin, A. Multivariate data analysis and metabolic profiling of artemisinin and related compounds in high yielding varieties of Artemisia annua field-grown in Madagascar. J. Pharm. Biomed. Anal. 2016, 117, 522–531. [Google Scholar] [CrossRef]
  54. Singh, P.; Bajpai, V.; Khandelwal, N.; Varshney, S.; Gaikwad, A.N.; Srivastava, M.; Singh, B.; Kumar, B. Determination of bioactive compounds of Artemisia Spp. plant extracts by LC–MS/MS technique and their in-vitro anti-adipogenic activity screening. J. Pharm. Biomed. Anal. 2021, 193, 113707. [Google Scholar] [CrossRef]
  55. Olennikov, D.N.; Chirikova, N.K.; Kashchenko, N.I.; Nikolaev, V.M.; Kim, S.-W.; Vennos, C. Bioactive phenolics of the genus Artemisia (Asteraceae): HPLC-DAD-ESI-TQ-MS/MS profile of the Siberian species and their inhibitory potential against α-amylase and α-glucosidase. Front. Pharmacol. 2018, 9, 756. [Google Scholar] [CrossRef] [PubMed]
  56. Mohamed, A.E.-H.H.; El-Sayed, M.; Hegazy, M.E.; Helaly, S.E.; Esmail, A.M.; Mohamed, N.S. Chemical constituents and biological activities of Artemisia herba-alba. Rec. Nat. Prod. 2010, 4, 1–25. [Google Scholar]
  57. Melguizo-Melguizo, D.; Diaz-de-Cerio, E.; Quirantes-Piné, R.; Švarc-Gajić, J.; Segura-Carretero, A. The potential of Artemisia vulgaris leaves as a source of antioxidant phenolic compounds. J. Funct. Foods 2014, 10, 192–200. [Google Scholar] [CrossRef]
  58. Han, J.; Ye, M.; Qiao, X.; Xu, M.; Wang, B.-r.; Guo, D.-A. Characterization of phenolic compounds in the Chinese herbal drug Artemisia annua by liquid chromatography coupled to electrospray ionization mass spectrometry. J. Pharm. Biomed. Anal. 2008, 47, 516–525. [Google Scholar] [CrossRef] [PubMed]
  59. Bourgou, S.; Rebey, I.B.; Mkadmini, K.; Isoda, H.; Ksouri, R.; Ksouri, W.M. LC-ESI-TOF-MS and GC-MS profiling of Artemisia herba-alba and evaluation of its bioactive properties. Food Res. Int. 2017, 99, 702–712. [Google Scholar] [CrossRef]
  60. Nagy, C.; Pesti, A.; Andrási, M.; Vasas, G.; Gáspár, A. Determination of artemisinin and its analogs in Artemisia annua extracts by capillary electrophoresis–mass spectrometry. J. Pharm. Biomed. Anal. 2021, 202, 114131. [Google Scholar] [CrossRef]
  61. Bajomo, E.M.; Aing, M.S.; Ford, L.S.; Niemeyer, E.D. Chemotyping of commercially available basil (Ocimum basilicum L.) varieties: Cultivar and morphotype influence phenolic acid composition and antioxidant properties. NFS J. 2022, 26, 1–9. [Google Scholar] [CrossRef]
  62. Zannou, O.; Pashazadeh, H.; Ibrahim, S.A.; Koca, I.; Galanakis, C.M. Green and highly extraction of phenolic compounds and antioxidant capacity from kinkeliba (Combretum micranthum G. Don) by natural deep eutectic solvents (NADESs) using maceration, ultrasound-assisted extraction and homogenate-assisted extraction. Arab. J. Chem. 2022, 15, 103752. [Google Scholar] [CrossRef]
  63. Dargham, M.B.; Boumosleh, J.M.; Farhat, A.; Abdelkhalek, S.; Bou-Maroun, E.; El Hosry, L. Antioxidant and anti-diabetic activities in commercial and homemade pomegranate molasses in Lebanon. Food Biosci. 2022, 46, 101540. [Google Scholar] [CrossRef]
  64. Chen, M.; He, X.; Sun, H.; Sun, Y.; Li, L.; Zhu, J.; Xia, G.; Guo, X.; Zang, H. Phytochemical analysis, UPLC-ESI-Orbitrap-MS analysis, biological activity, and toxicity of extracts from Tripleurospermum limosum (Maxim.) Pobed. Arab. J. Chem. 2022, 15, 103797. [Google Scholar] [CrossRef]
  65. Bibi Sadeer, N.; Montesano, D.; Albrizio, S.; Zengin, G.; Mahomoodally, M.F. The versatility of antioxidant assays in food science and safety—Chemistry, applications, strengths, and limitations. Antioxidants 2020, 9, 709. [Google Scholar] [CrossRef] [PubMed]
  66. Bouasla, I.; Hamel, T.; Barour, C.; Bouasla, A.; Hachouf, M.; Bouguerra, O.M.; Messarah, M. Evaluation of solvent influence on phytochemical content and antioxidant activities of two Algerian endemic taxa: Stachys marrubiifolia Viv. and Lamium flexuosum Ten. (Lamiaceae). Eur. J. Integr. Med. 2021, 42, 101267. [Google Scholar] [CrossRef]
  67. Rocchetti, G.; Miras-Moreno, M.B.; Zengin, G.; Senkardes, I.; Sadeer, N.B.; Mahomoodally, M.F.; Lucini, L. UHPLC-QTOF-MS phytochemical profiling and in vitro biological properties of Rhamnus petiolaris (Rhamnaceae). Ind. Crops Prod. 2019, 142, 111856. [Google Scholar] [CrossRef]
  68. Minda, D.; Ghiulai, R.; Banciu, C.D.; Pavel, I.Z.; Danciu, C.; Racoviceanu, R.; Soica, C.; Budu, O.D.; Muntean, D.; Diaconeasa, Z. Phytochemical Profile, Antioxidant and Wound Healing Potential of Three Artemisia Species: In Vitro and In Ovo Evaluation. Appl. Sci. 2022, 12, 1359. [Google Scholar] [CrossRef]
  69. Kamarauskaite, J.; Baniene, R.; Raudone, L.; Vilkickyte, G.; Vainoriene, R.; Motiekaityte, V.; Trumbeckaite, S. Antioxidant and Mitochondria-Targeted Activity of Caffeoylquinic-Acid-Rich Fractions of Wormwood (Artemisia absinthium L.) and Silver Wormwood (Artemisia ludoviciana Nutt.). Antioxidants 2021, 10, 1405. [Google Scholar] [CrossRef]
  70. Ferrante, C.; Zengin, G.; Menghini, L.; Diuzheva, A.; Jekő, J.; Cziáky, Z.; Recinella, L.; Chiavaroli, A.; Leone, S.; Brunetti, L. Qualitative Fingerprint Analysis and Multidirectional Assessment of Different Crude Extracts and Essential Oil from Wild Artemisia santonicum L. Processes 2019, 7, 522. [Google Scholar] [CrossRef] [Green Version]
  71. Kooltheat, N.; Chujit, K.; Nuangnong, K.; Nokkaew, N.; Bunluepuech, K.; Yamasaki, K.; Chatatikun, M. Artemisia lactiflora Extracts Prevent Inflammatory Responses of Human Macrophages Stimulated with Charcoal Pyrolysis Smoke. J. Evid. Based Integr. Med. 2021, 26, 2515690X211068837. [Google Scholar] [CrossRef]
  72. Dahal, P.; Bista, G.; Dahal, P. Phytochemical screening, antioxidant, antibacterial, and antidandruff activities of leaf extract of Artemisia indica. J. Rep. Pharm. Sci. 2021, 10, 231. [Google Scholar] [CrossRef]
  73. Nowsheen, T.; Hazrat, A.; Shah, S.W.A.; Bibi, S.; Begum, A.; Mukhtiar, M.; Khan, A. Phytochemical, antibacterial and antioxidant screening of Artemisia santolinifolia various parts. J. Biosci. 2021, 37, e37061. [Google Scholar] [CrossRef]
  74. Romeilah, R.M.; El-Beltagi, H.S.; Shalaby, E.A.; Younes, K.M.; Hani, E.; Rajendrasozhan, S.; Mohamed, H. Antioxidant and cytotoxic activities of Artemisia monosperma L. and Tamarix aphylla L. essential oils. Not. Bot. Horti Agrobot. Cluj Napoca 2021, 49, 12233. [Google Scholar] [CrossRef]
  75. Taqui, R.; Debnath, M.; Ahmed, S.; Ghosh, A. Advances on plant extracts and phytocompounds with acetylcholinesterase inhibition activity for possible treatment of Alzheimer’s disease. Phytomed. Plus 2022, 2, 100184. [Google Scholar] [CrossRef]
  76. Ayua, E.O.; Nkhata, S.G.; Namaumbo, S.J.; Kamau, E.H.; Ngoma, T.N.; Aduol, K.O. Polyphenolic inhibition of enterocytic starch digestion enzymes and glucose transporters for managing type 2 diabetes may be reduced in food systems. Heliyon 2021, 7, e06245. [Google Scholar] [CrossRef] [PubMed]
  77. Ahmed, S.; Khan, S.T.; Zargaham, M.K.; Khan, A.U.; Khan, S.; Hussain, A.; Uddin, J.; Khan, A.; Al-Harrasi, A. Potential therapeutic natural products against Alzheimer’s disease with Reference of Acetylcholinesterase. Biomed. Pharmacother. 2021, 139, 111609. [Google Scholar] [CrossRef]
  78. Rauf, A.; Jehan, N. Natural products as a potential enzyme inhibitors from medicinal plants. In Enzyme Inhibitors and Activators; Senturk, M., Ed.; IntechOpen Limited: London, UK, 2017; pp. 165–177. [Google Scholar]
  79. El-Askary, H.; Salem, H.H.; Abdel Motaal, A. Potential Mechanisms Involved in the Protective Effect of Dicaffeoylquinic Acids from Artemisia annua L. Leaves against Diabetes and Its Complications. Molecules 2022, 27, 857. [Google Scholar] [CrossRef]
  80. Bhat, S.H.; Ullah, M.F.; Abu-Duhier, F.M. Bioactive extract of Artemisia judaica causes in vitro inhibition of dipeptidyl peptidase IV and pancreatic/intestinal enzymes of the carbohydrate absorption cascade: Implication for anti-diabetic new molecular entities (NMEs). Orient. Pharm. Exp. Med. 2019, 19, 71–80. [Google Scholar] [CrossRef]
  81. Shoaib, M.; Shah, I.; Adikhari, A.; Ali, N.; Ayub, T.; Ul-Haq, Z.; Ali Shah, S.W. Isolation of flavonoides from Artemisia macrocephala anticholinesterase activity: Isolation, characterization and its in vitro anticholinesterse activity supported by molecular docking. Pak. J. Pharm. Sci. 2018, 31, 1347–1354. [Google Scholar] [PubMed]
  82. Zengin, G.; Mollica, A.; Aktumsek, A.; Picot, C.M.N.; Mahomoodally, M.F. In vitro and in silico insights of Cupressus sempervirens, Artemisia absinthium and Lippia triphylla: Bridging traditional knowledge and scientific validation. Eur. J. Integr. Med. 2017, 12, 135–141. [Google Scholar] [CrossRef]
  83. Raza, M.A.; Danish, M.; Mushtaq, M.; Sumrra, S.H.; Saqib, Z.; Rehman, S.U. Phenolic profiling and therapeutic potential of local flora of Azad Kashmir; In vitro enzyme inhibition and antioxidant. Open Chem. 2017, 15, 371–379. [Google Scholar] [CrossRef] [Green Version]
  84. Allué-Guardia, A.; García, J.I.; Torrelles, J.B. Evolution of drug-resistant Mycobacterium tuberculosis strains and their adaptation to the human lung environment. Front. Microbiol. 2021, 12, 612675. [Google Scholar] [CrossRef]
  85. Choi, W.H. Novel pharmacological activity of artesunate and artemisinin: Their potential as anti-tubercular agents. J. Clin. Med. 2017, 6, 30. [Google Scholar] [CrossRef] [Green Version]
  86. Zheng, H.; Colvin, C.J.; Johnson, B.K.; Kirchhoff, P.D.; Wilson, M.; Jorgensen-Muga, K.; Larsen, S.D.; Abramovitch, R.B. Inhibitors of Mycobacterium tuberculosis DosRST signaling and persistence. Nat. Chem. Biol. 2017, 13, 218–225. [Google Scholar] [CrossRef] [PubMed]
  87. Martini, M.C.; Zhang, T.; Williams, J.T.; Abramovitch, R.B.; Weathers, P.J.; Shell, S.S. Artemisia annua and Artemisia afra extracts exhibit strong bactericidal activity against Mycobacterium tuberculosis. J. Ethnopharmcol. 2020, 262, 113191. [Google Scholar] [CrossRef] [PubMed]
  88. de Oliveira Lima, M.; de Medeiros, A.A.; Silva, K.S.; Cardoso, G.; de Oliveira Lima, E.; de Oliveira Pereira, F. Investigation of the antifungal potential of linalool against clinical isolates of fluconazole resistant Trichophyton rubrum. J. Mycol. Med. 2017, 27, 195–202. [Google Scholar] [CrossRef] [PubMed]
  89. Bhowmick, S.; Baptista, R.; Fazakerley, D.; Whatley, K.E.; Hoffmann, K.F.; Shen, J.; Mur, L.A. The anti-mycobacterial activity of Artemisia annua L. is based on deoxyartemisinin and artemisinic acid. bioRxiv 2020. [Google Scholar] [CrossRef]
  90. Stavri, M.; Mathew, K.; Gordon, A.; Shnyder, S.D.; Falconer, R.A.; Gibbons, S. Guaianolide sesquiterpenes from Pulicaria crispa (Forssk.) oliv. Phytochemistry 2008, 69, 1915–1918. [Google Scholar] [CrossRef] [PubMed]
  91. Cantrell, C.L.; Franzblau, S.G.; Fischer, N.H. Antimycobacterial plant terpenoids. Planta Med. 2001, 67, 685–694. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Correlation analysis of the phytochemical composition and biological activities. ABTS, 2,2′-azino-bis (3-ethylbenzothiazoline) 6-sulfonic acid; AChE, acetylcholinesterase; BChE, butyrylcholinesterase; CUPRAC, cupric ion reducing antioxidant capacity; DPPH, 1,1-diphenyl-2-picrylhydrazyl; FRAP, ferric ion reducing antioxidant power; MCA, metal chelating activity; PDA, phosphomolybdenum activity; TPAC, total phenolic acid content; TPC, total phenolic content. Compounds numbered as in Table 2.
Figure 1. Correlation analysis of the phytochemical composition and biological activities. ABTS, 2,2′-azino-bis (3-ethylbenzothiazoline) 6-sulfonic acid; AChE, acetylcholinesterase; BChE, butyrylcholinesterase; CUPRAC, cupric ion reducing antioxidant capacity; DPPH, 1,1-diphenyl-2-picrylhydrazyl; FRAP, ferric ion reducing antioxidant power; MCA, metal chelating activity; PDA, phosphomolybdenum activity; TPAC, total phenolic acid content; TPC, total phenolic content. Compounds numbered as in Table 2.
Antioxidants 11 01017 g001
Figure 2. Exploratory principal component analysis. (AC) Contribution of biological activities to the principal components of the PCA. (DF) Scatter plot showing the distribution of the samples in the factorial plan derived from the three retained principal components.
Figure 2. Exploratory principal component analysis. (AC) Contribution of biological activities to the principal components of the PCA. (DF) Scatter plot showing the distribution of the samples in the factorial plan derived from the three retained principal components.
Antioxidants 11 01017 g002
Figure 3. Clustered image map (red color: high bioactivity; blue color: low bioactivity) based on the biological activities dataset.
Figure 3. Clustered image map (red color: high bioactivity; blue color: low bioactivity) based on the biological activities dataset.
Antioxidants 11 01017 g003
Figure 4. Clustered image map (red color: high bioactivity; blue color: low bioactivity) based on the chemical composition dataset.
Figure 4. Clustered image map (red color: high bioactivity; blue color: low bioactivity) based on the chemical composition dataset.
Antioxidants 11 01017 g004
Table 1. Extraction yields and total phenolic and flavonoid contents of Artemisia spp. extracts.
Table 1. Extraction yields and total phenolic and flavonoid contents of Artemisia spp. extracts.
Artemisia SpeciesPartExtraction
Solvent
Yield
(%)
TPC
(mg GAE/g)
TFC
(mg RE/g)
A. absinthium L.RootsMeOH15.6619.77 ± 0.20 f2.35 ± 0.04 e
CHCl35.795.78 ± 0.10 g0.37 ± 0.02 g
Aerial partsMeOH18.7553.38 ± 0.16 d28.74 ± 0.51 d
CHCl39.4718.28 ± 0.15 i24.25 ± 1.28 e
A. annua L.RootsMeOH2.1976.35 ± 0.75 a10.41 ± 0.27 a
CHCl30.6826.10 ± 0.10 d1.41 ± 0.07 f
Aerial partsMeOH17.5760.00 ± 0.24 c47.74 ± 0.79 a
CHCl310.6925.27 ± 0.20 g35.36 ± 0.30 c
A. austriaca Jacq.RootsMeOH11.7341.68 ± 0.25 c4.65 ± 0.20 c
CHCl31.5726.59 ± 0.23 d2.37 ± 0.01 e
Aerial partsMeOH13.8848.42 ± 0.49 e40.30 ± 0.94 b
CHCl38.1224.50 ± 0.14 g32.89 ± 1.58 c
A. pontica L.RootsMeOH6.5565.65 ± 0.46 b6.99 ± 0.17 b
CHCl31.1623.59 ± 0.58 e2.46 ± 0.03 e
Aerial partsMeOH22.9565.06 ± 0.59 b33.01 ± 0.43 c
CHCl39.4622.65 ± 0.18 h26.85 ± 1.02 de
A. vulgaris L.RootsMeOH12.4527.36 ± 0.99 d3.29 ± 0.12 d
CHCl31.0322.21 ± 0.82 e1.13 ± 0.17 f
Aerial partsMeOH15.67106.34 ± 0.61 a39.39 ± 0.86 b
CHCl35.8637.62 ± 0.09 f11.02 ± 0.78 f
Data are presented as mean ± standard deviation (SD) of three determinations; different superscript letters within columns indicate significant differences in the tested extracts for the same parts (p < 0.05). GAE, gallic acid equivalents; RE, rutin equivalents; TFC, total flavonoid content; TPC, total phenolic content.
Table 2. LC-HRMS/MS-based phytochemical profiling of Artemisia spp. extracts.
Table 2. LC-HRMS/MS-based phytochemical profiling of Artemisia spp. extracts.
No.Proposed IdentityClassTR (min)HRMSExp. (m/z)Calcd. (m/z)Δ (ppm)HRMS/MS (m/z)
1Quinic acid *Organic acid1.83[M − H]191.0557191.05612.14173.0381, 127.0340, 111.0384
2SucroseSugar1.86[M − H]341.1097341.1089−2.24179.0571, 119.0312
3Dihydroxybenzoic acid hexosidePhenolic acid7.58[M − H]315.0706315.07224.92153.0105, 109.0215
4Hydroxybenzoic acid *Phenolic acid9.98[M − H]137.0241137.02442.30109.0358
5Neochlorogenic acidPhenolic acid10.27[M − H]353.0893353.0878−4.22191.0484, 179.0252, 135.0370
6Esculetin-O-hexoside ICoumarin10.91[M − H]339.0715339.07221.93177.0233, 149.0157, 133.0217, 105.0327
7EsculetinCoumarin14.41[M − H]177.0207177.0193−7.68133.0227, 105.0266
8Cyrptochlorogenic acidPhenolic acid15.71[M − H]353.0880353.0878−0.55191.0488, 173.0429, 161.0239, 135.0412
9Chlorogenic acid *Phenolic acid16.69[M − H]353.0893353.0878−4.22191.0568, 173.0429, 135.0461
10Tuberonic acid-O-hexosideFatty acid17.48[M − H]387.1681387.1661−5.27207.1010, 163.1121, 119.0376
11Esculetin-O-hexoside IICoumarin18.19[M − H]339.0728339.0722−1.89177.0203, 149.0144, 133.0215
12Mearnsetin-di-O-hexosideFlavonoid18.57[M − H]655.1575655.1516−1.39493.1190, 331.0475, 315.0138
13Chrysartemin ASesquiterpene18.99[M − H]277.1072277.10813.41233.1193, 218.0969, 215.1098, 191.1061, 175.0763, 135.0835
14Caffeic acid-O-pentosidePhenolic acid19.50[M − H]311.0767311.07721.73179.0343, 149.0461, 135.0440
15Chrysartemin BSesquiterpene20.38[M − H]277.1066277.10815.56233.1165, 215.0981, 191.1034, 175.0705, 160.0463, 135.0839
16Feruloylquinic acidPhenolic acid20.52[M − H]367.1049367.1035−3.92191.0563, 173.0460, 134.0349
17Artabsinolide ASesquiterpene20.60[M − H]279.1237279.12370.35261.1027, 243.0906, 217.1121, 199.1051, 175.1082
18Dicaffeoylquinic acid IPhenolic acid20.65[M − H]515.1198515.1195−0.58353.0763, 191.0485, 179.0265, 135.0373
19Coumaroylquinic acidPhenolic acid20.67[M − H]337.0926337.09290.86191.0589, 173.0462, 145.0322, 109.0380
20Artecanin hydrateSesquiterpene21.15[M − H]295.1187296.1187−4.69251.1300, 207.1409, 189.1280, 151.0831
21Apigenin-C-hexoside-C-pentoside IFlavonoid21.33[M − H]563.1404563.14060.41503.1277, 383.0784, 353.0680, 325.0671, 297.0714
22Quercetin-di-O-hexosideFlavonoid21.51[M − H]625.1418625.1410−1.24463.0810, 300.0246, 271.0240, 151.0020
23Coumaric acid-O-pentosidePhenolic acid21.81[M − H]295.0819295.08231.44163.0416, 149.0463, 119.0494
24Apigenin-C-hexoside-C-pentoside IIFlavonoid22.12[M − H]563.1414563.1406−1.37443.1010, 383.0747, 353.0663, 325.0728, 297.0763
25Quercetin-O-deoxyhexoside-O-hexosideFlavonoid23.15[M − H]609.1476609.1461−2.44300.0167, 271.0154, 150.994
26Mearnsetin-O-hexosideFlavonoid23.91[M − H]493.0998493.0988−2.10331.0495, 315.0183, 287.0218, 271.0266
27Quercetin-O-hexosideFlavonoid24.10[M − H]463.0860463.08824.74300.0210, 255.0139, 150.9999
28Luteolin-O-deoxyhexoside-O-hexosideFlavonoid24.79[M − H]593.1540593.1512−4.72285.0443, 255.0292, 227.0355, 151.0042
29Eupatolitin-O-deoxyhexoside-O-hexosideFlavonoid25.01[M − H]653.1729653.1723−0.88345.0825, 330.0441, 301.0478, 287.0236
30Dicaffeoylquinic acid IIPhenolic acid25.92[M − H]515.1198515.1195−0.58353.0763, 191.0485, 179.0265, 135.0373
31TrachelosideLignan26.38[M − H]549.1985549.1978−1.36505.1054, 387.1727, 301.0335, 207.1026, 161.0258
32Dicaffeoylquinic acid IIIPhenolic acid26.78[M − H]515.1188515.11951.36353.0773, 191.0481, 179.0258, 173.0390
33Coumaroylcaffeoylquinic acidPhenolic acid27.81[M − H]499.1307499.1246−1.63353.0922, 337.0981, 191.0566, 163.0440
34Eupatolitin-di-O-hexosideFlavonoid27.98[M − H]669.1645669.16724.09345.0667, 330.0402, 301.0186, 179.0381, 161.0253
35Feruloylcaffeoylquinic acid IPhenolic acid28.31[M − H]529.1397529.13510.85367.1386, 353.1172, 191.0748, 179.0486, 161.0397
36Rhamnetin-di-O-hexosideFlavonoid28.95[M − H]639.1527639.15676.21413.1265, 315.0608, 300.0298, 284.0403, 271.0291, 255.0388
37Rhamnetin-O-hexosideFlavonoid29.02[M − H]477.1016477.10384.71433.1382, 315.0767, 161.0276, 153.0227, 109.0304
38Feruloylcaffeoylquinic acid IIPhenolic acid29.12[M − H]529.1353529.1351−0.28367.1035, 353.0930, 191.0589, 179.0320, 173.0484
39EriodictyolFlavonoid29.39[M − H]287.0567287.0561−2.04151.0046, 135.0479
40Artemisinin *Sesquiterpene30.44[M − H]281.1385281.13943.36263.1319, 237.1529, 193.1612
41Luteolin *Flavonoid31.04[M − H]285.0400285.04051.61175.0386, 133.0313
42Tetrahydroxydimethoxyflavone (e.g., eupatolitin)Flavonoid31.49[M − H]345.0602345.06164.02330.0402, 315.0188, 287.0296, 259.0301, 259.0301, 215.0351, 175.0091, 149.0308, 121.0326
43Tetrahydroxymethoxyflavone (e.g., rhamnetin)Flavonoid31.55[M − H]315.0509315.05100.40300.0327, 271.0269, 255.0312, 243.0322, 227.0356, 215.0350, 171.0409, 147.0202
44Trihydroxyoctadecadienoic acidFatty acid31.79[M − H]327.2181327.2177−1.43229.1442, 211.1319
45Deoxyartemisinin ISesquiterpene32.09[M − H]265.1435265.14453.88247.1335, 221.1582, 203.1459, 185.1346, 151.1148
46SantoninSesquiterpene32.42[M − H]245.1174245.11833.73201.1282, 186.1064, 161.0962, 147.0805, 135.0841
47Deoxyartemisinin IISesquiterpene32.70[M − H]265.1447265.1445−1.00247.1357, 221.1557, 203.1451, 151.1154
48Trihydroxymethoxyflavanone (e.g., homoeriodictyol)Flavonoid32.98[M − H]301.0724301.0718−2.11151.0049, 134.0413
49Trihydroxyoctadecenoic acid IFatty acid33.62[M − H]329.2331329.23330.75 229.1470, 211.1353, 199.1170
50Dihydroxydimethoxyflavone I (e.g., rhamnazin)Flavonoid33.69[M − H]329.0678329.0667−3.40314.0456, 299.0241, 271.0279, 271.0272, 243.0312, 227.0430, 215.0360, 199.0421, 185.0236, 161.0264, 151.0068, 133.0347
51Trihydroxyoctadecenoic acid IIFatty acid34.05[M − H]329.2338329.2333−1.37229.1433, 199.1155
52DihydroxytrimethoxyflavoneFlavonoid34.72[M − H]359.0772359.07720.11344.0575, 329.03351, 314.0086, 297.0051, 286.0162, 270.0287, 258.0184, 230.0225, 214.0302, 202.0280
53Trihydroxymethoxyflavone (e.g., diosmetin)Flavonoid36.93[M − H]299.0566299.0561−1.63284.0259, 255.0179, 239.0292, 227.0330, 151.0077, 133.0252
54PseudosantoninSesquiterpene36.96[M − H]263.1279263.12893.72245.1127, 219.1366, 201.1230, 159.1152
55AbsinthinTriterpene37.16[M + HCO2]541.2801541.28071.19351.6359, 275.5226
56Hydroxydimethoxyflavone (e.g., cirsimaritin)Flavonoid37.25[M − H]313.0711313.07182.11298.0722, 283.0375, 269.0628
57Hydroxytrimethoxyflavone I (e.g., penduletin)Flavonoid37.67[M − H]343.0813343.08232.98328.0382, 313.0382, 298.0133, 285.0421, 270.0199, 255.0318, 242.0284
58Artemisinin CSesquiterpene37.72[M − H]247.1329247.13404.30231.1403, 203.1469, 187.1442, 161.1372, 133.1030
59Dihydroxydimethoxyflavone II (e.g., eupalitin)Flavonoid37.99[M − H]329.0678329.0667−3.40314.0456, 299.0241, 271.0279, 271.0272, 243.0312, 227.0430, 215.0360, 199.0421, 185.0236, 161.0264, 151.0068, 133.0347
60Arteannuin BSesquiterpene38.36[M − H]247.1341247.1340−0.53203.1449, 133.1019
61Dihydroxytetramethoxyflavone (e.g., casticin)Flavonoid38.81[M − H]373.0939373.0929−2.70358.0729, 343.0494, 300.0407, 285.0054, 269.0079, 257.0103, 241.0132, 229.0140, 213.0161, 201.0202, 185.0220
62CnicinSesquiterpene39.38[M − H]377.1617377.1606−2.97295.1213, 251.1322, 189.1257, 151.07060
63ArtenolideTriterpene40.18[M + HCO2]573.2714573.2705−1.66527.2685, 325.1304, 263.1287, 185.1288
64Dihydroarteannuin BSesquiterpene40.28[M − H]249.1496249.14960.70231.1415, 207.1742, 187.1523
65Dihydroxymethoxyflavone (e.g., genkwanin)Flavonoid40.84[M − H]283.0601283.06123.86268.0423, 240.0392, 211.0419
66DihydrosantamarinSesquiterpene41.37[M − H]249.1508249.1496−4.72231.1471, 205.1599, 187.1494
67Absinthin derivative ITriterpene41.96[M + HCO2]555.2582555.26003.44509.2392, 491.2392, 447.2558, 265.1365, 243.1047, 229.1237, 199.1137
68Hydroxytrimethoxyflavone II (e.g., eupatilin)Flavonoid37.67[M − H]343.0813343.08232.98328.0382, 313.0382, 298.0133, 285.0421, 270.0199, 255.0318, 242.0284
69Hydroperoxyoctadecadienoic acidFatty acid44.54[M − H]311.2212311.22285.07293.2171, 211.1341, 171.0999
70IsoabsinthinTriterpene44.67[M + HCO2]541.2801541.28071.19495.2583, 351.6359, 275.5226
71Absinthin derivative IITriterpene46.17[M + HCO2]539.2672539.2650−4.37247.1212, 204.1637, 185.1479
72Hydroxyoctadecatrienoic acidFatty acid47.14[M − H]293.2118293.21221.42275.1973, 224.1359, 195.1381
73Hydroxyoctadecadienoic acidFatty acid48.69[M − H]295.2269295.22793.27277.2162, 195.1407, 171.1029
* Identified based on the standard.
Table 3. Antioxidant activity of Artemisia spp. extracts.
Table 3. Antioxidant activity of Artemisia spp. extracts.
Artemisia
Species
PartExtraction
Solvent
DPPH
(mg TE/g)
ABTS
(mg TE/g)
CUPRAC
(mg TE/g)
FRAP
(mg TE/g)
MCA
(mg EDTAE/g)
PBD
(mmol TE/g)
A. absinthium L.RootsMeOH43.59 ± 1.08 c47.67 ± 0.34 f82.69 ± 1.76 e52.80 ± 2.52 e7.25 ± 0.23 e1.20 ± 0.11 e
CHCl35.11 ± 0.22 f7.54 ± 0.21 g24.24 ± 0.26 g12.40 ± 0.09 h8.33 ± 0.14 d0.84 ± 0.08 f
Aerial partsMeOH67.57 ± 3.55 cd95.95 ± 3.61 c188.11 ± 5.6885.36 ± 1.20 c14.68 ± 0.91 cd2.10 ± 0.20 bc
CHCl310.52 ± 0.80 f27.35 ± 0.15 gh47.05 ± 0.9426.45 ± 0.21 f11.25 ± 0.99 de2.46 ± 0.14 a
A. annua L.RootsMeOH237.03 ± 5.93 a240.78 ± 1.27 a438.43 ± 10.59 a294.52 ± 8.32 a14.38 ± 0.60 c2.24 ± 0.08 a
CHCl329.98 ± 0.55 d60.61 ± 0.62 d87.39 ± 4.62 e54.97 ± 0.08 en.a.2.37 ± 0.06 a
Aerial partsMeOH102.66 ± 2.15 b134.36 ± 2.28 b156.62 ± 4.1558.67 ± 1.45 d17.46 ± 3.03 bc1.55 ± 0.02 f
CHCl313.04 ± 0.70 f32.49 ± 0.34 fg58.26 ± 1.3324.79 ± 1.22 f20.91 ± 1.10 ab1.85 ± 0.06 de
A. austriaca Jacq.RootsMeOH48.99 ± 0.07 c77.19 ± 0.06 c168.90 ± 2.70 c105.77 ± 3.02 c15.84 ± 0.19 b1.59 ± 0.05 c
CHCl319.68 ± 0.22 e52.04 ± 2.09 e82.69 ± 1.76 e40.68 ± 1.77 f3.76 ± 0.46 g1.70 ± 0.05 c
Aerial partsMeOH64.85 ± 0.09 d75.19 ± 0.42 d143.59 ± 2.2159.60 ± 0.61 d22.16 ± 0.88 a1.66 ± 0.17 ef
CHCl311.05 ± 0.07 f37.64 ± 0.55 f52.72 ± 0.5525.57 ± 2.21 f12.76 ± 1.19 de1.56 ± 0.07 f
A. pontica L.RootsMeOH179.63 ± 2.60 b176.12 ± 2.64 b263.94 ± 1.87 b165.55 ± 3.83 b22.93 ± 0.32 a1.97 ± 0.00 b
CHCl331.34 ± 0.41 d51.69 ± 1.19 e80.33 ± 1.19 e45.77 ± 0.74 ef6.48 ± 0.28 ef1.53 ± 0.09 cd
Aerial partsMeOH71.65 ± 3.52 c98.45 ± 3.20 c290.14 ± 8.95113.33 ± 1.15 b9.89 ± 0.99 e1.55 ± 0.07 f
CHCl310.55 ± 1.18 f25.86 ± 0.50 h49.99 ± 1.0022.18 ± 1.39 f17.84 ± 0.44 bc1.52 ± 0.10 f
A. vulgaris L.RootsMeOH48.83 ± 0.04 c49.36 ± 0.40 ef113.97 ± 2.77 d66.51 ± 2.80 d5.78 ± 0.13 f1.23 ± 0.02 e
CHCl326.16 ± 0.50 d63.81 ± 1.19 d46.26 ± 4.81 f24.87 ± 2.41 gn.a.1.35 ± 0.09 de
Aerial partsMeOH139.56 ± 3.19 a173.86 ± 3.66 a498.32 ± 4.02198.51 ± 5.00 a12.35 ± 1.15 de2.33 ± 0.11 ab
CHCl333.74 ± 0.49 e56.54 ± 0.50 e111.48 ± 2.0147.73 ± 1.66 e20.94 ± 1.65 ab1.89 ± 0.04 cd
Data are presented as mean ± standard deviation (SD) of three determinations; different superscript letters within columns indicate significant differences in the tested extracts for the same parts (p < 0.05). ABTS, 2,2′-azino-bis (3-ethylbenzothiazoline) 6-sulfonic acid; CUPRAC, cupric ion reducing antioxidant capacity; DPPH, 1,1-diphenyl-2-picrylhydrazyl; EDTAE, EDTA equivalents; FRAP, ferric ion reducing antioxidant power; MCA, metal chelating activity; n.a., not active; PBD, phosphomolybdenum assay; TE, trolox equivalents.
Table 4. Enzyme inhibitory activity of Artemisia spp. extracts.
Table 4. Enzyme inhibitory activity of Artemisia spp. extracts.
Artemisia SpeciesPartExtraction
Solvent
AChE
(mg GALAE/g)
BChE
(mg GALAE/g)
Tyrosinase
(mg KAE/g)
Amylase
(mmol ACAE/g)
Glucosidase
(mmol ACAE/g)
A. absinthium L.RootsMeOH3.02 ± 0.05 a1.19 ± 0.11 d41.20 ± 0.73 ab0.30 ± 0.01 e0.88 ± 0.01 a
CHCl32.18 ± 0.14 b2.91 ± 0.21 c19.44 ± 0.67 c0.32 ± 0.01 e0.87 ± 0.01 ab
Aerial partsMeOH2.33 ± 0.02 ab2.32 ± 0.16 bc37.09 ± 1.31 bcde0.40 ± 0.01 e11.18 ± 0.20 a
CHCl32.50 ± 0.00 a2.67 ± 0.43 ab35.78 ± 1.53 cde0.44 ± 0.00 cd10.85 ± 0.13 ab
A. annua L.RootsMeOH2.00 ± 0.06 bcn.a.49.42 ± 4.51 a0.31 ± 0.01 e0.81 ± 0.01 abc
CHCl32.20 ± 0.02 b4.52 ± 0.18 a45.74 ± 0.56 ab0.50 ± 0.02 b0.87 ± 0.02 ab
Aerial partsMeOH1.97 ± 0.14 c2.14 ± 0.08 bc35.71 ± 2.04 cde0.41 ± 0.02 e5.93 ± 0.93 d
CHCl32.36 ± 0.08 ab3.11 ± 0.10 a36.39 ± 2.36 cde0.54 ± 0.01 a8.84 ± 1.08 bc
A. austriaca Jacq.RootsMeOH1.86 ± 0.07 cdn.a.47.27 ± 5.68 ab0.31 ± 0.00 e0.16 ± 0.03 f
CHCl32.16 ± 0.05 b3.45 ± 0.39 b13.16 ± 2.73 c0.57 ± 0.03 a0.79 ± 0.01 bc
Aerial partsMeOH2.00 ± 0.08 c1.94 ± 0.55 bc39.37 ± 0.77 abc0.42 ± 0.00 de6.07 ± 0.40 d
CHCl32.16 ± 0.02 bc2.55 ± 0.22 abc39.37 ± 0.94 abcd0.54 ± 0.01 a9.84 ± 0.71 abc
A. pontica L.RootsMeOH2.05 ± 0.12 cdn.a.44.91 ± 5.05 ab0.30 ± 0.00 e0.65 ± 0.05 d
CHCl31.82 ± 0.05 bc0.93 ± 0.06 d38.30 ± 1.69 b0.38 ± 0.01 d0.77 ± 0.01 c
Aerial partsMeOH1.92 ± 0.05 c1.82 ± 0.05 c44.64 ± 0.40 a0.46 ± 0.02 c6.21 ± 0.79 d
CHCl32.15 ± 0.13 bc2.34 ± 0.02 bc42.82 ± 2.30 ab0.50 ± 0.01 b8.54 ± 0.58 c
A. vulgaris L.RootsMeOH1.72 ± 0.08 bc0.46 ± 0.08 e41.08 ± 0.68 ab0.31 ± 0.01 e0.30 ± 0.07 e
CHCl31.98 ± 0.16 bc0.12 ± 0.01 ef13.79 ± 2.78 c0.44 ± 0.02 c0.87 ± 0.01 abc
Aerial partsMeOH2.04 ± 1.14 c1.86 ± 0.31 c31.38 ± 2.74 e0.40 ± 0.01 e11.32 ± 0.38 a
CHCl32.10 ± 0.12 bc2.14 ± 0.12 bc33.23 ± 4.15 de0.51 ± 0.01 ab10.01 ± 1.30 abc
Data are presented as mean ± standard deviation (SD) of three determinations; different superscript letters within columns indicate significant differences in the tested extracts for the same parts (p < 0.05). ACAE, acarbose equivalents; AChE, acetylcholinesterase; BChE, butyrylcholinesterase; GALAE, galanthamine equivalents; KAE, kojic acid equivalents; n.a., not active.
Table 5. Anti-Mycobacterium tuberculosis H37Ra activity of Artemisia spp. extracts.
Table 5. Anti-Mycobacterium tuberculosis H37Ra activity of Artemisia spp. extracts.
Artemisia SpeciesPartExtraction
Solvent
MIC
(mg/L)
A. absinthium L.RootsMeOH>256
CHCl3256
Aerial partsMeOH256
CHCl3128
A. annua L.RootsMeOH256
CHCl3128
Aerial partsMeOH256
CHCl3128
A. austriaca Jacq.RootsMeOH>256
CHCl3256
Aerial partsMeOH128
CHCl364
A. pontica L.RootsMeOH256
CHCl3128
Aerial partsMeOH256
CHCl3256
A. vulgaris L.RootsMeOH>256
CHCl3128
Aerial partsMeOH256
CHCl3128
Etambutol2
Streptomycin0.5
Rifampicin0.002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Trifan, A.; Zengin, G.; Sinan, K.I.; Sieniawska, E.; Sawicki, R.; Maciejewska-Turska, M.; Skalikca-Woźniak, K.; Luca, S.V. Unveiling the Phytochemical Profile and Biological Potential of Five Artemisia Species. Antioxidants 2022, 11, 1017. https://0-doi-org.brum.beds.ac.uk/10.3390/antiox11051017

AMA Style

Trifan A, Zengin G, Sinan KI, Sieniawska E, Sawicki R, Maciejewska-Turska M, Skalikca-Woźniak K, Luca SV. Unveiling the Phytochemical Profile and Biological Potential of Five Artemisia Species. Antioxidants. 2022; 11(5):1017. https://0-doi-org.brum.beds.ac.uk/10.3390/antiox11051017

Chicago/Turabian Style

Trifan, Adriana, Gokhan Zengin, Kouadio Ibrahime Sinan, Elwira Sieniawska, Rafal Sawicki, Magdalena Maciejewska-Turska, Krystyna Skalikca-Woźniak, and Simon Vlad Luca. 2022. "Unveiling the Phytochemical Profile and Biological Potential of Five Artemisia Species" Antioxidants 11, no. 5: 1017. https://0-doi-org.brum.beds.ac.uk/10.3390/antiox11051017

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