Next Article in Journal
Unveiling the Neem (Azadirachta indica) Effects on Biofilm Formation of Food-Borne Bacteria and the Potential Mechanism Using a Molecular Docking Approach
Previous Article in Journal
Genetic Diversity of Tulipa alberti and T. greigii Populations from Kazakhstan Based on Application of Expressed Sequence Tag Simple Sequence Repeat Markers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Insights into Chemical Diversity and Potential Health-Promoting Effects of Ferns

1
Department of Pharmacognosy, Faculty of Pharmacy, Ain Shams University, Abbassia, Cairo 11566, Egypt
2
Food Science and Technology Program, Department of Life Sciences, BNU-HKBU United International College, 2000 Jintong Road, Tangjiawan, Zhuhai 519087, China
*
Author to whom correspondence should be addressed.
Plants 2024, 13(18), 2668; https://doi.org/10.3390/plants13182668
Submission received: 22 August 2024 / Revised: 18 September 2024 / Accepted: 20 September 2024 / Published: 23 September 2024
(This article belongs to the Section Phytochemistry)

Abstract

:
The scientific community is focusing on how to enhance human health and immunity through functional foods, and dietary supplements are proven to have a positive as well as a protective effect against infectious and chronic diseases. Ferns act as a taxonomical linkage between higher and lower plants and are endowed with a wide chemical diversity not subjected to sufficient scrutinization before. Even though a wealth of traditional medicinal fern uses were recorded in Chinese medicine, robust phytochemical and biological investigations of these plants are lacking. Herein, an extensive search was conducted using the keywords ferns and compounds, ferns and NMR, ferns and toxicity, and the terms ferns and chemistry, lignans, Polypodiaceae, NMR, isolation, bioactive compounds, terpenes, phenolics, phloroglucinols, monoterpenes, alkaloids, phenolics, and fatty acids were utilized with the Boolean operators AND, OR, and NOT. Databases such as PubMed, Web of Science, Science Direct, Scopus, Google Scholar, and Reaxys were utilized to reveal a wealth of information regarding fern chemistry and their health-promoting effects. Terpenes followed by phenolics represented the largest number of isolated active compounds. Regarding the neuroprotective effects, Psilotium, Polypodium, and Dryopteris species possessed as their major phenolics component unique chemical moieties including catechins, procyanidins, and bioflavonoids. In this updated chemical review, the pharmacological and chemical aspects of ferns are compiled manifesting their chemical diversity in the last seven years (2017–2024) together with a special focus on their nutritive and potential health-promoting effects.

1. Introduction

The history of vascular cryptogams, known as pteridophytes, comprises a wealth of traditional uses and applications practiced by native ethnic people all around the world. Among cryptogams or free-sporing plants, including around 12,000 species in tropical and subtropical regions, ferns rather than lycophytes are more closely related to gymnosperms. Among vascular plants, ferns are the second species-rich clade after angiosperms. Ferns have exhibited many biological effects such as antihyperglycemic [1,2], trypanocidal [3], hepatoprotective, and anti-inflammatory activities. Phytochemical components reported in ferns encompass fatty acids, carotenoids, terpenoids, and polyphenols and are particularly implicated in current age-related diseases due to their marked antioxidant and free radical scavenging power. This was noted in populations largely consuming polyphenols in their diet [4].
Fern’s distribution differs from other plants, as angiosperms are restricted and lower in number, with research groups having limited access; thus, reported information about them is far from enough. A common feature noticed in the production of natural products from ferns was their medicinal uses and ethnobotanical benefits to people where they were located; therefore, their ecological adaptation and chemical diversity are noteworthy. Based on their botanical classification, pteridophytes were unique phylogenetically in their biosynthetic genes from other plants, which made their compounds distinct from other natural sources both structurally and functionally [5].
One example of an imminent world problem is antibiotic resistance, which is more common than ever before and is on the rise, with more than 2.8 million cases in the US resulting in about 35 thousand deaths, per the Centers for Disease Control and Prevention (CDC) annual report of 2019 [6,7]. Ferns were historically established as antibiotic agents dating back to 1980 when the water and alcoholic extracts of 114 pteridophytes were screened for their antimicrobial effects, and about 64% of the samples showed bactericidal activity against Gram positive, Gram negative, and fungal strains; to mention just a few, Dryopteris cochleate (D.Don) C.Chr.; Leptochillus decurrens Blume.; Phymatodes ebenipes (Hook.), Polypodium irioides Poir.; Pyrrosia mannii (Giesenh.) Ching.; and Microsorium alternifolium (Willd.) Copel. had activity [8]. Additionally, ferns have inherited the unique ability to adapt to extreme environments, from the Rocky Mountains and deserts to misty tropical and subtropical regions, owing to their protected spores and modified leaf shapes, which have honed their chemical diversity and provided us with an arsenal of promising bioactive compounds. Thus, fern research could contribute effectively to solving, among other problems, the antibiotic resistance dilemma. This has motivated researchers to embark on more and more studies to unveil the unexplored nature of ferns as a promising solution for the world’s problems [9]. Ferns are evolutionary forerunners compared to gymnosperms and angiosperms, with a genetic machinery that favors chemical diversity and unusual skeletons [9].
In this study, we introduce a chemical review highlighting the most recent isolated phytochemical compounds from ferns between 2017 and 2023 with a critical discussion of their novel aspects and biological activities. Previous reports on ferns were few, in which some discussed taxonomy [10] or the culinary uses of edible ferns in China [11]; moreover, Soeder studied the phytochemical components of known ferns [12]. Cao et al. prepared a compilation of ferns’ compounds isolated from 1980 to 2017 [5], yet incorrectly included lycophytes as Selaginella species under the fern classification. However, scientific interest in fern research has risen dramatically, and more than 300 compounds are reviewed here, most of which are unprecedented.
The main aim of this study is to review the isolation and characterization of more than 300 bioactive compounds from ferns from seven different chemical classes and their reported medicinal effects in the previous six years, thus pointing out the significance of fern research and paving the way for more funding for drug discovery from ferns. Furthermore, the health and nutritive benefits presented by ferns as well as their toxicity concerns are summarized.

2. Methodology

Paper Selection Criteria

In this survey, an extensive search was conducted using the keywords ferns and compounds, ferns and NMR, and ferns and toxicity, and the terms ferns and chemistry, lignans, Polypodiaceae, NMR, isolation, bioactive compounds, terpenes, phenolics, phloroglucinols, monoterpenes, alkaloids, phenolics, and fatty acids were utilized with the Boolean operators AND, OR, and NOT. Databases such as PubMed, Web of Science, Science Direct, Scopus, Google Scholar, and Reaxys were used. The initial results for keywords such as ferns and phenolics yielded 13,000 papers, ferns and terpenes yielded 2690 papers, and ferns and phloroglucinols yielded 520 papers. These were manually filtered to select only reports discussing chemistry work, which resulted in 2000, 556, and 43 articles, respectively. Subsequently, only reports that discovered new natural products with significant bioactivity were included to yield a reference number of 167 papers. Regarding the biological activities of ferns, 4070 papers were initially obtained and filtered by removing 3007 articles focusing on chemical synthesis and 850 articles about lycophytes to give a final total of 50 papers.

3. Chemical Classes of Compounds Isolated from Ferns

Phloroglucinols are subdivided according to their number of monomers into mono, di, tri, tetra and phlorotanins. As one of the monomeric phloroglucinols, polyprenylated acylphloroglucinols (PPAPs) are endowed with fascinating chemical structures and biological activities as a source of lead compounds because they are unique, and many are still untapped. The search results in the PPAPs databases demonstrated the presence of more than 850 molecules with disparate bioactivities. To mention just a few, antiadipogenic, antidepressant, antiparasitic, antimicrobial, antiviral, leishmanicidal, molluscicide and immunomodulatory effects were noted [13].

3.1. Polyketide-Terpenoid Hybrids

Dryopteridaceae Family

The two genera Elaphoglossum and Dryopteris contained all the phloroglucinol derivatives isolated before and reported; for instance, there were C-glycosides as dryopteroside from Dryopteris crassirhizoma Nakai [14], trisflavaspidic acid ABB and BBB, flavaspidic acid AB and PB, filixcic acid ABA and PBP, dryocrassin, and albaspidin AA [14]. Others included elaphopilosins A-E and elaphogayanin A-C [15] and lindbergins A-D and E-I [16]
PPAPs serve as chemotaxonomic and phylogenetic markers of the genera Dryopteris and Elaphoglossum, being of significant abundance in this type of molecule [17]. Two new acyl phloroglucinol dimer derivatives, namely, paleacenins A 1 and B 2 with filicinic acid geranylated moieties, were isolated from the nonpolar fractions of the fern Elaphoglossum paleaceum (Hook. & Grev.) Sledge. Paleacenin A was a combined filicinic acid and phloroglucinol ring separated by a central methylene. The prenyl side chain attached to C-4 of the filicinic acid was related to the known yungesins. Furthermore, the acetate group at C-6 was verified through long-range coupling between OH-5 and C-19 as well as the carbonyl at C-1, thus establishing the β-triketone system with its keto enol tautomerisms, which explained the chelation and extra deshielding of the acetate group nuclei. While paleacenin A was active as an inhibitor of MAO-A and MAO-B with IC50 values of 31 and 4.7 μM, respectively, and more selectivity towards MAO-A, paleacenin B revealed IC50 values of 1.3 and 4.4 μM with selectivity towards MAO-B [18].
These compounds were unique in possessing the geranylated filicinic acid moiety and were only preceded by the yungensins obtained from Elaphoglossum yungense de la Sota [15]. It is notable that antimicrobial assays and biofilm inhibition might yield promising results, and researchers are encouraged to carry out these investigations based on the reported activities of yungensins and their structural similarity with the paleacenin compounds.
Dryocrassoids (A-J) 3-28 (Figure 1), unique acylphloroglucinol meroterpenoid enantiomeric pairs based on nerolidol-type sesquiterpenes, were isolated from Dryopteris crassirhizoma Nakai with six other known phenolic compounds and recorded moderate to low antiviral activity against HSV-1. The structures were assigned based on 1D, 2D, HRMS, and ECD and compared to (±)-hyperjaponol A [19,20] and nerolidol. Dryocrassoid A was identified as a racemic keto nol tautomer, and its signal showed two fragments I and II marking an unconventional link between the filicinic acid and a nerolidol sesquiterpene, respectively, which warrants further biosynthetic pathway identification work, proposedly with similarity to hyperjaponol A [19]. Fragment I only differed from filicinic acid in the presence of an acetyl group instead of the isobutyl and was located at C-2. The linkage was established between the two fragments based on Correlation Spectroscopy (COSY) and Heteronuclear Multiple Bond Correlation (HMBC) that verified the C-7–C-10′ bond. Furthermore, C-5 and C-11′ were shifted down field, revealing the presence of an oxo-bridge that formed a pyran ring. The amounts of the discovered isomers were minute, between 6 and 32 mg, which further encourages biotechnological interventions to enhance the production of these compounds [20].
The unique meroterpenoidal acylphloroglucinols, with the parent name dryoptol, were isolated from Dryopteris crassirhizoma Nakai as either mono or dimeric skeletons fused at C-10″/C11″ (type-I) or C-6″/C-7″ (type-II) to the nerolidiol moiety 67-90 (Figure 1). (±)-Dryoptols B-E and their (±)-3″-epi-dryoptols B, C, D, and E 31-38; (±)-3″-epi-dryoptol G; (±)-dryoptol G 41-42; and (±)-dryoptols J-L and their 3″epimers 47-52 were isolated as racemes, and the crystallographic data were deposited for (±)-dryoptol A 29 and its 3″-epi isomer 30 and its NMR compared to dryocrassoid E as a reference compound. Dryoptol isomers featured the first structurally mixed molecules between mono acylphloroglucinols and dehydro-theonelline with a pronounced antifungal activity of 1.61 µg/mL for dryoptol D and its epimer compared to fluconazole, whose MIC value is 3.41 µg/mL [21]. Comparing the assignments on the left of C-3″ manifested that dryoptols G, H, and I were shielded compared to the respective signals in their epi-3″ dryoptols G, H, and I, and assignments on the right of C-3″ in dryoptols G, H, and I were all deshielded compared to their C-3″ epimers epi-3″ dryoptols G, H, and I, suggesting the structural correlation of the six compounds [21].
The meroterpenoids dryoptins and 11″ epi dryoptins A-E 53-62 were isolated from the same fern, revealing a new possibility of the combination of dimeric and trimeric phloroglucinols with dehydrotheonelline with a marked antifungal activity against standard Candida albicans with an MIC value of 1.61 µg/mL compared to fluconazole (FLC): 3.41 µg/mL [22].
The genus Dryopteris was characterized mainly by phloroglucinol derivatives [23,24,25] with few reports on phenolic constituents. Novel tocopherol skeletons were decorated and identified in Dryopteris crassirhizoma Nakai. For example, drycrasspherol A 63 featured a 1,2,4 tri oxosubstituted benzene ring combined with a 13′-carboxy-α-tocopherol side chain with RRS absolute configurations in the 9, 13, 17-positions, similar to previously isolated tocopherols. The C-22 methoxy isomer was isolated and named drycrasspherol B 64 (Figure 1). Drycrasspherol C 65 (Figure 1) was composed partially of dryocrasspherol A with two more degrees of unsaturation to form rings C and D; moreover, it showed a resemblance to the known compound 3,4-dihydroxy-4-methylcyclohexanone. Compounds 65-67 (Figure 1) represented the new 6/6/6/6 (A/B/C/D) ring system with the same stereochemistry at C-9, C-13, and C-17 based on their common biosynthetic pathway. Drycrasspherol E 67 signified a unique dimer of pentacyclic structure. Correlations between the ring E methylene protons were asserted with C-6, C-5, and C-1 or C-2′, C-3′, and C-4′ in each of the similar monomers of ring B, and the typical Diels–Alder reaction was suggested to be the mechanism of dimerization. Only drycrasspherol A exhibited a promising antiviral effect against respiratory syncytial virus with an IC50 value of 6.50 μM compared to ribavarin [26].

3.2. Polyketides

3.2.1. Dryopteridaceae Family

Five propionyl and butyryl acylated monomeric phloroglucinol derivatives 68-72 were isolated from the dichloromethane fraction of Dryopteris crassirhizoma Nakai [27]. Upon studying the antiplatelet activity, structure–activity relationships concluded that three key features in the phloroglucinol skeleton were necessary for the antiplatelet effect: the C-1 butanonyl group, the C-4 hydroxy group, and the C-3 methyl group [27]. Compounds 68 and 69 (Figure 2) were reported to have antimicrobial activity (2 µg/mL) against MRSA that was very comparable to vancomycin both in vitro and in vivo, where virulence factor suppression and ribosomal inhibition played major roles [28].
Dryopteris crassirhizoma Nakai rhizome methanolic extract was the source of about 20 phloroglucinol derivatives 73-92 (Figure 2), with monomeric, dimeric, trimeric, and tetrameric types whose protein tyrosine phosphatase 1B enzyme (PTP1B) inhibition was measured from the dimeric flavaspidic acids AB and PB and norflavaspidic acid PB to the trimeric antiviral filixic acid ABA or the tetrameric types as dryocrassin ABBA [29]. The new molecules were 76 and 87. While dryopidin PB 76 (Figure 2) was identified by comparison to the known methylene-bis-phlorobutyrophenon, as the former manifested a propyl group instead of a methyl group in the latter, dryopcrassirine AB 87 (Figure 2) revealed a unique phloroglucinol structure with a pyrrolidine-1-carboximidamide group confirmed by the MS/MS analysis together with the HMQC and HMBC correlations to the quaternary carbons in the butyrylated phloroglucinol ring. Signals of the pyrrolidine ring were further correlated to the imine functionality by HMBC. Both compounds had a moderate to low PTP1B inhibitory effect compared to ursolic acid [29]. The former was compared to the monomeric known phlorobutyrophenone whose carbon number and obtained molecular formula did not match but suggested the possibility of its dimeric structure. The weakest enzyme inhibition, IC50 from 45 to >100 μM, was recorded for monomeric phloroglucinols 73 and 74, followed by the dimeric compounds with IC50 values from 6.0 to 41 μM with a strong emphasis on the nature of the side chain chemical structures, as it could alter the enzyme sensitivity as in 75, whose IC50 value was 14.3 μM despite being a monomer. Propyl, acetyl, and butyryl side chains resulted in significant activity as in flavaspidic acid PB 86, flavaspidic acid AB 85, and norflavaspidic acid PB 81, respectively [29].
Disaspidin BB 93, a pseudoaspidinol derivative attached to a phloroglucinol ring [30], was isolated from Dryopteris fragrans (L.) Schott and revealed promising topical antimicrobial effects against S. haemolyticus (SHA), Staphylococcus epidermidis (SEP), and methicillin-resistant S. aureus (MRSA); moreover, various ceftazidime-resistant strains were sensitive to 93 with MIC values of 1.67–2.71 μg/mL. Furthermore, it showed a biofilm scavenging effect on SEP [31,32]. The study is a continuation of Teng’s work [30], which outlined a SAR correlation for pseudo aspidinols, particularly the C-4 butyryl, C-6, and the phloroglucinol hydroxyl groups.
Six known phenolic compounds 94-99 from the fronds’ exudate of Dryopteris crassirhizoma Nakai exhibited moderate to low antiviral activity against HSV-1 [20,33]. The compounds represented the albaspidin and araspidin nuclei with only acyl group variations in the dimeric filicinic acid units linked with a methylene group.
More acylphloroglucinols with a structural similarity to phloropyron BB were obtained from Dryopteris championii (Benth.) C. Chr. in small amounts of between 3 and 11 mg. Prominent features that aided in the identification were the length of the aliphatic chain next to the carbonyl C-3, which varied from methyl, ethyl, or propyl in phloropyrons B and A 100-101 (Figure 2) compared to phloropyron BB as manifested by the H-H COSY between H-9 and H-10 and the HMBC between H–C10 and C8. Moreover, the pyronone attached chain was isobutyl in phloropyron C 102 (Figure 2). Another two compounds showed analogy to margaspidin BB and were characterized as the C-4 methoxy derivative with a methyl group instead of the carbonyl attached propyl group, named margaspidin A 103 (Figure 2), and margaspidin B 104 (Figure 2) with two additional methylenes and the absence of the methyl group. Only phloropyron A and pseudoaspidinol A 105 (Figure 2) showed moderate bactericidal activity against Staphylococcus aureus, Escherichia coli, Bacillus subtilis, and Dickeya zeae; this was particularly the case for Bacillus, where the two compounds displayed MIC values of 16 μg/mL [34].
Dryofragone 106 and dryofracoumarin B 107 (Figure 2) were isolated and identified from the petroleum ether fraction of the fragrant woodfern Dracaena fragrans (L.) Schott methanolic extract; furthermore, six known compounds 108-113 were isolated by means of cytotoxicity-guided separation using the MTT and CCK-8 assays [35]. Dryofragone was an acylphloroglucinol enantiomeric pair with a cyclohexadiene ring hypothesized based on the demonstrated long-range coupling between H-4/C-2, C-3, C-5, and C-6 as well as between CH3-11/C-1, C-2, and C-3. A butyryl group was attached to C-6 as manifested by the HMBC correlations from C-8 protons to carbons 6, 7, 8, and 9 [35]. A strong in vitro antiviral effect against H5N1 neuraminidase was revealed by filixic acid ABA 114 and dryocrassin ABBA with IC50 values of 29.5 and 18.5 μM, respectively [23]. Similarly, 114 demonstrated an anti-influenza effect against the A/Puerto Rico/8/1934 (H1N1) with an IC50 of 38 μg/mL compared to ribavirin, whose IC50 was 40 μg/mL [21].
Methoxylated flavonoids 115-124 were obtained from Asplenium anceps and demonstrated an antitumor effect against A549, MCF-7, and HL-7702 with a moderate IC50 range between 16 and 85 μM [36].

3.2.2. Pteridaceae Family

Three new lignans 125-127 (Figure 3) were identified from Pteris laeta Wall. as analogs of 8′-hydroxy-(+)-isolariciresinol-9-O-β-D-xylopyranoside. While 125 was denoted as having a hydroxy group in the 8 position instead of 8′, 126 and 127 were syringaresinol derivatives revealing a xylose sugar moiety instead of glucose in the former and a 7R, 8S, 7′S, 8′R stereochemistry in 127 as manifested by comparing the experimental ECD to the calculated one [37]. Lignans 128-130 were not tested for their biological activity.

3.2.3. Aspleniaceae Family

Interestingly, the same fern was reported to contain iridoid glycosides such as camptoside 131 (Figure 3), which was detected with a unique C-C bond connecting a 4-hydroxystyryl group to its C-6 position [38]. Although it differed from the known methyl ester of E-6-O-p-coumaroyl scandoside by only m/z 44, possibly interpreted as liberated CO2 during the MS fragmentation, camptoside was not detected upon heating the E-6-O-p-coumaroyl scandoside and re-running the same experiment again, which proved its natural origin. This unusual iridoid with other known compounds 131-133 (Figure 3) revealed a potent NO production inhibitory effect in RAW 264.7 macrophage cells of 11.2 μM compared to the curcumin standard of 10.1 μM.

3.2.4. Onocleaceae Family

The genus Matteuccia is composed of five species, and most of the phytochemical work was carried out on three of them: Matteuccia orientalis (HOOK.), Matteuccia Intermedia C. Chr., and Matteuccia struthiopteris (L.) Tod. [39]. The alcoholic rhizome extract of Matteuccia orientalis (HOOK.) (Onocleaceae) yielded two new phenolic compounds, matteucens I 135 and J 136 (Figure 3). Matteucen I was a C-β-D-glucoside connected to a phenyl chromone aglycone moiety. Matteucen II was a bibenzyl phenolic compound analogous to the known molecule 5-β-D-glucosyloxy-3-hydroxy-trans-stilbene-2-carboxylic acid, yet it is lacking the trans-olefins and instead possesses two methylenes between the two benzenes. Although these two compounds did not have any reports of their biological activities, many pharmacological effects were recorded for Matteuccia sp.; for instance, antidiabetic, antioxidative, and anti-inflammatory effects were reported [40]. The methyl group attached to ring A was inserted in the fern’s biosynthetic reaction pattern; thus, several C-methylated flavanones have been reported from the genus Matteuccia with different ring B oxygenation combinations [41].
The C-methylated ring A flavanonol demethyl matteucinol 137 was obtained from Matteuccia Intermedia C. Chr. [42] together with the C-methylated flavanone glucosides 137-145 (Figure 3) and twenty known compounds 146-167 all presenting C-methylation in ring A at the C-8 position. The SAR correlations suggested the importance of the 7-OH group as a crucial substitution for α-glucosidase inhibitory activity, which agreed with the previous results of Cao et al. [43]. Moreover, the combined C-methylation in the 6 and 8 positions in ring A with either or both the hydroxylation or methoxylation of ring B added more to the α-glucosidase inhibitory effect [39]. Compounds 149-160 and 162-170 were either inactive as α-glucosidase inhibitors or were not assayed at all due to insufficient amounts [44].
The extracts from two ferns, Matteuccia orientalis (Hook.) Trevis and M. struthiopteris (L.) Tod., which are widely used in Chinese cuisine, exhibited reduced IL-1β gene expression in a dose-dependent manner in RAW264.7 cells compared to LPS [45], possibly based on their unique phenolic content, which was found to be increased under the influence of light and stress responses [46].

3.2.5. Tectariaceae Family

The phenolics detected from Tectaria coadunata (J.Sm.) C.Chr., a widely traditionally used ethnobotanical pteridophyte in the family Tectariaceae, were A-type procyanidins with different polymerization rates, namely, eriodictyol-7-O-glucuronide 171 and luteolin-7-O-glucoronide 172, which demonstrated inhibitory bioactivity against several enzymes such as tyrosinases, α-amylase, α-glucosidase, and acetyl or butyryl cholinesterase. A moderate activity against the BxPC3 cancer cell line was also reported [47].

3.2.6. Dryopteridaceae Family

Eight phenolics with a novel phenyl propanoid fragranoside B 173 (Figure 4) were isolated from Dracaena fragrans (L.) Schott, with known phenolic acids 174-179 and only 173 and 179 demonstrating a remarkable anticancer effect against MCF7 cells [48].

3.2.7. Polypodiaceae Family

A few studies were conducted on the Polish traditional fern Polypodium vulgare L. revealing its phytochemical components as mainly ecdysteroids, flavonoids, and flavan-3ol derivatives [49]. The rhizomes yielded (+)-afzelechin-7-O-α-L-arabinofuranoside 180 upon water extraction, Moreover, (+)-afzelechin-7-O-β-D-apiofuranoside 181 and the aglycone (+)-afzelechin 182 with the known flavan-3-ol glycosides (+)-catechin-7-O-α-L-arabinofuranoside 183 and (+)-catechin-7-O-β-D-apiofuranoside 184 were also isolated. Very few studies investigated the biological applications of afzelechin/epiafzelechin; for instance, they were reported to have antioxidant and anti-fresh cut browning effects [1]. The polypody rhizome was also a powerful substance in folk medicine to treat kidney-related ailments. Afzelchins or flavan-3-ols were rarely isolated from ferns; for instance, monomers, dimers, and trimers of the afzelchin glucopyranoside and allopyranosides were isolated from the rhizome of Drynaria fortune (Kunze) J. Sm. [50], while monomers and trimers of afzelchin/epiafzechin stereoisomers linked by A- and B-type linkages were isolated from Selliguea feei (Bory.) [51].
The rhizomes of the epiphytic fern Drynaria roosii Nakaike were extracted to yield five liglaurates A-E 185-189 (Figure 4), the novel bis (lauric acid-12-yl) lignanoates, and the methyl and glyceryl 12-caffeoyloxylaurates. While liglaurates A and B were obtained as pure enantiomers after oxidative coupling reaction, liglaurates C and D were racemic mixtures, and the anticancer effect measured against Hela cells was found to have IC50 values of 0.11, 0.24, 0.02, 0.13, 0.34, and 0.17 μM for liglaurates (+)-A, (-)-A, (+)-B, (-)-B, (±)-C, and (±)-D, respectively [52].
The isolated liglaurates were highly analogous to the known epiphyllic acid. These were composed of a lignan skeleton and two methyl 12-hydroxyllaurate, which form a diester between the C9, C9′, and 12″ carboxylic groups. Due to the conformational instability of liglaurate A, determining the absolute configuration of C-7 and C-8 as 7S,8R, was achieved via a synthetic route where the tetra-O-methyl derivatives were prepared by reaction with iodomethane followed by sodium hydroxide, which was further confirmed by HRMS [52]. A novel protocatechuic methyl ester 192 (Figure 4) was isolated from Phymatopteris hastata (Thunb.) Pic. Serm. comprising two protocatechuic units linked via an ether link from the C-4 to the anomeric carbon in a central glucose unit and between the C-6 of glucose and the C-7′ of the second protocatechuate [53].
The new naringenin glycoside 193 (Figure 4) was confirmed by 1D NMR to be dihydro-flavone, which gave zero optical rotation indicating the possibility of having a racemic mixture. The α-glucosidase inhibition activity was assessed for 192 and 194-203, and they showed potent effects with IC50 values ranging between 99 and 697 μg/mL, compared to the standard acarbose [53].

3.2.8. Marattiaceae Family

The rarely reported compounds from ferns, (-)-epi-osmundalactone 204 and angiopteroside 205, were isolated from the medium polarity fractions of Angiopteris helferiana C. Presl of the family Marattiaceae; their content was measured and quantified by UPLC to be 1.54% and 3.2% in the fleshy fern extract, respectively, with promising anti-obesity and anti-inflammatory effects [54]. Osmundalactone was previously isolated from the fern Angiopteris caudatiformis Hieron. and showed insect antifeeding activity [55]. Angiopteroside was isolated exclusively from the genus Angiopteris and manifested an activity against HIV-1 reverse transcriptase with an IC50 of 0.9 mM compared to the didanosine/ddl control with an IC50 of 0.87 mM [56].

3.2.9. Metaxyaceae Family

New methylated xanthones were detected in the fern Metaxya rostrata (Kunth) C. Presl. This fern has been largely under investigated, from which nearly only xanthones were separated and characterized [57,58,59]. Compounds 206-208 were isolated from the nonpolar extract of the rootlets of Metaxya rostrata (Kunth) C. Presl., and structure–activity relationships were suggested based on their assay results as inhibitors of FOxM1 or the induction of cell cycle arrest in G2/M phase [60]. While a single methylation exerted an effect to induce apoptosis, the unmethylated counterparts were of higher efficacy. Furthermore, the di-methylated 2-deprenyl-rheediaxanthone B (XB) and 2-deprenyl-7-hydroxy-rheediaxanthone B were safe with no cytotoxicity. The lack of methylation in the OH-group at position 7 was favored to inhibit topoisomerase I [60].

3.2.10. Cibotiaceae Family

Substituted glycosides and glycoses were obtained from Cibotium barometz (L.) J. Sm. “Gou-ji”, the fern traditionally used for liver and kidney damage. Three new glycosidic compounds, namely, 4-O-coumaroyl-β-D-allose 209 (Figure 5), 4-O-caffeoyl-β-D-allose 210, and ethyl (4-O-caffeoyl)-β-D-glucopyranoside 211, were isolated. Phenolic alloses were seldom found in nature, with few reports documenting their chemistry or biological activity. Furthermore, compound 209 significantly exceeded the bicyclol positive control in the hepatoprotective assay against APAP-induced HepG2 cell damage, as it diminished the serum-deprived PC12 cell death and improved viability from 76.4% to 82.9% [42].

3.2.11. Thelypteridaceae Family

Red flavonoids of variant nature were isolated in low yield and characterized from Pronephrium penangianum (Hook.) Holttum rhizomes, namely, jixueqisus A-F 220-225 (Figure 5). Jixueqisus A and B possessed the rare 6,8-dimethyl-2-phenyl-7H-1-benzopyran-7-one extended conjugated system, which explained their rich red color. Jixueqisus B was the acetylated analog of Jixueqisus A. Jixueqisus C was comparable to the known abacopterin L except for the lack of the acetyl group. Jixueqisus D was a typical chalcone with four spin systems, the ABX system in the B ring of chalcone, the meta coupled aromatic protons, and the saccharide unit protons; additionally, it had trans-olefinic protons with large J-values. Jixueqisus E shared many features with the D analog, such as the ABX spin protons and the meta coupled protons in ring A, yet the HMBC correlations manifested the presence of an aurone moiety. Cytotoxicity was assayed against the cell lines MCF-7, BGC-823, HepG2, and HCT-116, yet the jixueqisus congeners had negligible activity using paclitaxel as a control [61]. Jixueqisus red pigments were proposed to be generated biosynthetically from 220 through acidification, hydration, tautomerism, hydrogenation, and acetylation processes; furthermore, the heterocyclic ring found in all jixueqisus compounds was anticipated to originate from Michel nucleophilic attack between an unsaturated ketone and a phenol [62]. Flavonoids isolated from the same fern were mostly inactive except for 226 and 228-229, which were reported to have hepatoprotective or cytotoxic effects [61] (Figure 5).

3.3. Steroids and Fatty Acids

Athyriaceae Family

For the first time from ferns, (2R)-3-(4′-hydroxyphenyl) lactic acid 236 was isolated and identified in the fronds of Diplazium esculentum (Retz.) Sw. together with ecdysteroids as ponastrone A, amarasterone A1, and makisterone C, 235-238. Diplazium esculentum (Retz.) Sw. fronds are eaten as vegetables in many countries and are used in folk medicine as a treatment for diabetes and topically for scabies and boils [63]. The stereochemistry of 236 was based on the optical rotation values measured and by comparing them to published values of the R and S enantiomers [64]. Ecdysteroids, known as insect hormones, are polyhydroxy steroids new to the genus Diplazium and deserve additional study regarding their chemo taxonomical significance. The Amarasterone A1 stereochemistry was assigned as the (24R, 25S)-isomer based on the superimposition of its NMR data with that previously isolated and confirmed by the synthesis of the molecule from Leuzea carthamoides Willd. (Asteraceae) [65] (Figure 6).
β-Ecdysterone is the typical 20-hydroxyecdysone, which was revealed to have a promising bone proliferative effect via the BMP2/Smad/Runx2 signaling pathway [36]. Despite the utility of forming the 7,8-dihydro analog based on their wide biological activity, this was sterically hindered in ecdysteroids with their γ-hydroxy-α, β-enone group.
The number of fatty acids with incorporated cyclopropane rings is scarce in nature. A few examples mentioned here from plants, sponges, and bacteria include sterculinic acid isolated from the kernel of Sterculia joetida L. seeds, whose anticancer effect and interaction with body fat composition were reported [66]. Similarly, malvalic acid was isolated from the cluster mallow Malva verticillate L. family (Malvaceae) and other Malva species. Amphimic acids A and B from the sponge Amphimedon sp. growing in Australia demonstrated effective inhibitory action on DNA topoisomerase I, and majusculoic acid obtained from the marine cyanobacteria depicted potent antifungal and anticandidal activities (MIC value of 8 µM). From ferns, the novel unsaturated FA associated with a cyclopropane ring, adiantic acid, was obtained from the fern Adiantum flabellulatum L. and characterized as (S, E)-7-(2-octylcyclopropylidene) heptanoic acid 240, along with three known compounds: stigmast-4-en-6-β-ol-3-one 241, β-sitosterol 243, and isoadiantol B 242. Adiantic acid showed promising anti-protein tyrosine phosphatase enzyme activity (anti-PTP1B) with an IC50 value of 6.99 μM [67].

3.4. Terpenes Family

Unlike seed plants, which contain terpene synthases, ferns possess microbial terpene synthase-like (MTPSL) enzymes for terpene production; likewise, fungi, liverworts, hornworts, mosses, and lycophytes contain these enzymes [68] (Figure 7).

3.4.1. Pteridaceae Family

The genus Adiantum is composed of 17 species, called the maidenhair fern with several medicinal uses, and about 135 isolated compounds, among them triterpenes, were reported as migrated hopane-, oleanane- [69], and lanostane-type triterpenes [70]. Three hopane-type triterpenes were isolated from the aerial parts of Adiantum capillus-veneris Linn. 245-247 (Figure 3 and Figure 4) with significant antifungal effects against Micrococcus lysodeikticus, Alternaria alternata, and Helminthosporium maydis, and their MIC was down into the range of 12.5–3.125 μg/mL [71]. The botanical insecticide 22-hydroxyhopane 248 (Figure 7) was isolated from the fronds of Adiantum latifolium Lam., and its trehalase inhibitory and larvicidal activity were recorded since trehalose is a main carbohydrate source in these organisms and its loss contributes to insect mortality [72].
Fern sesterterpenes were reported to assume either a tricyclic cheilanthane or a tetracyclic 18-episcalarane skeleton. The genus Aleuritopteris produced sesterterpenes, which were hypothesized to be produced by triterpene synthases in ferns [73]. Although relatively rare in plants, the sesterterpene group of compounds was unveiled in Aleuritopteris anceps (Blanf.) Panigrahi, a genus from which only six sesterterpene compounds were detected before with a (17Z)-13, 19-dihydroxylcheilanth-17-ene] moiety: one from Aleuritopteris Khunii [74], another from Aleuritopteris mexicana Fée [75], and five from Aleuritopteris krameria (Franch. & Sav.) Ching [76]. Ancepsone A 249 (Figure 7) was a novel 13, 19-epoxycheilanthane skeleton that was unambiguously characterized with its absolute configuration using single crystal X-ray crystallography. Ancepsone A 249 exhibited a moderate anticancer effect against MCF-7 cell line cells with an IC50 value of 12.63 μM as compared to its milder effect against normal HL-7702 cells [36]. Cheilanthane sesterterpenes were known to be isolated from a few plants such as Aletris farinose as well as insects and marine sponges [77].
While rarely described from ferns and limited to basidiomycetes and fungi [78], cyathane diterpenes were isolated from Aleuritopteris albofusca (Baker) Pic. Serm.; specifically, 14-oxy-7α,20-dihydroxycyath-12,18-diene 250 (Figure 7) was isolated, and its structure was comparable to onychiol B, whose abundance and high yield in this fern suggested the possible production of cyathanes by ferns rather than their fungal counterparts. Moreover, ent-8[14],15-pimaradiene-2β,19-diol 251 (Figure 7) was reported and characterized by single-crystal X-ray crystallography [79]. Even though pimaradienes have been reported from plants, they mainly had 3,19-dihydroxy groups; the 2β,19-diol pimaradienes were not common [79]. Ptercresions A 255, B 256, and C 257 (Figure 7) were isolated from Pteris cretica L., and the latter two compounds showed significant liver-protective effects in paracetamol-induced L02 cells in vitro with EC50 values between 107 and 213 μM compared to silymarin [80].
The genus Pteris includes about 300 species, and most of them were used to treat common illnesses in Chinese medicine as Pteris multifida Poir. and Pteris spinosa (L.) Desv. [80,81]. Illudoid sesquiterpenes, known as pterosins, were recognized as the chemo taxonomical markers of this genus and the family Pteridaceae. Pterosins were classified according to their carbon skeleton number into C13, C14, and C15 compounds, which possess several biological activities such as antitumor, anti-inflammatory, and antimicrobial effects [37].
The 1-indanone type sesquiterpenes, which are pterosin compounds, belonging to the illudoid C14 and C15 molecules, were first isolated and characterized from Bracken ferns Pteridium aquilinum (L.) Kuhn back in 1978. For example, pterosins A, D, K, L, and Z together with their esters and glucosides, were named pterosides. The nor-sesquiterpenes were also reported by the same group as pterosins B, C, E, F, G, I, J, O, and N with their esters and glucosides [82]. Several later derivatives such as (2R)-pterosin P and dehydropterosin B were reported from Pteris multifida Poir. [83]. These molecules possessed two chiral centers at the carbon 2 and 3 positions, which formed their cotton effect CD (Circular Dichroism) at 310–370 nm caused by the n-π* transitions of their conjugated ketones [84].
Recently, more pterosin glucosides of (2S,3S) pterosin C were isolated from the roots of Pteris multifida Poir (Pteridaceae) as two new derivatives possessing either caffeoyl or coumaroyl moieties attached to the glucose C-4 or C-6 position, respectively. The first C14 sesquiterpene was (2S,3S)-pterosin C3-O-β-D-[4′-(E)]-caffeoyl]-glucopyranoside 261 (Figure 7), and the second was (2S,3S)-pterosin C3-O-β-D-[6′-(E)-p-coumaroyl]-glucopyranoside 262 (Figure 7). Both compounds showed positive cotton effects at 320 nm indicating that the C-3 hydroxyl group was pseudo axial. Additionally, the anti-configuration of C-3 was assured from a small coupling constant. It is notable that Kim et al. incorrectly referred to the C-3 configuration as trans, which should be anti, as the system was a single bond with free rotation. (2S,3S)-3-Methoxypterosin was reported from nature for the first time after its synthesis by [85]. Furthermore, ten known compounds 263-272 were detected (Table 1). Only compounds 261 and 262 (Figure 7) displayed a significant anticancer effect against the HCT116 human colorectal cancer cell line with IC50 values of 8 and 17 μM by increasing the levels of caspase 9 and procaspase 9 [86].
The roots of Pteris laeta Wall were the source of about twenty sesquiterpenes 273-288 (Figure 7 and Figure 8) with a powerful neuroprotective activity in Alzheimer disease (AD), as evidenced by in vitro studies involving H2O2-induced N2a cell damage where the survival rate in all sesquiterpenes ranged from 86 to 91%, with pterosin B exhibiting a neuroprotective effect 200 times that of ginsenoside Rg1 [37]. Although the Pteris species are producers of pterosin toxins, edible pterosin types were identified in this genus as well [87,88,89]. Pterosinsides A, B, and C 273-275 (Figure 7) were isolated and characterized by an olefinic C-2 and C-3 trans glycol system, compared to pterosin E methylene and methylidyne carbons. Zhang et al. inadequately named the sesquiterpene pterosin E compound as peterosin E [84] and the COSY analysis as COZY, which should be pointed out critically here to avoid confusion.
The indanone-skeleton-based sesquiterpenes were isolated from Pteris obtusiloba Ching & S.H. Wu. Among the reported bioactivities were antidiabetic, anti-tumor, and antihyperlipidemic effects [90]. Nevertheless, in a unique study considered to be the only one recorded from Pteris obtusiloba Ching & S.H. Wu, Peng et al. reported eleven 1H-inden-1-one-stype sesquiterpenes HSs, 288-298 including three novel dimers 288 and 290 (Figure 8) and an isomeric pair 288 and 289. These skeletons were verified by NMR, 2D NMR, ESI-MS, CD, and X-ray. 1H-Inden-1-one-type sesquiterpenes HSs were also isolated from fungi previously such as Basoidiomycota, yet none of them significantly inhibited the α-glucosidase enzyme. These results, possibly attributed to experimental errors, warrant more investigation since pterosins and their glycosides were mentioned by more than one study to exhibit promising antidiabetic activity [90,91]. The three obtupterosin dimers A 288-290 revealed IC50 values of 27.5 μM, 30.6 μM, and 12.8 μM against the human colon cancer cell line HCT-116 [90,92].
Pteris cretica L. aerial parts were the source of the new pterosins, creticolacton A 299 (Figure 8), 13-hydroxy-2(R),3(R)-pterosin L 300, creticoside A 301 with a 1, 2 glycol side chain attached to C-6 of 300, and the glycosidic derivative spelosin 3-O-β-D-glucopyranoside 302 (Figure 6). Creticolacton A 299 and 300 (Figure 8) manifested a cytotoxic activity against different human cell lines such as the neuroblastoma cell line SH-SY5Y, gastric cancer cell line SGC-7901, colon cancer cell line HCT-116, and colorectal cancer cell line Lovo, with IC50 values of 22.4 μM and 15.8 μM, respectively [93]. It is worth mentioning that the yield was very low from Pteris cretica L., as the authors ended up with a few milligrams of each compound although they started with 5 kg of the dried plant and a weight of over one-hundred grams of extract by liquid extraction of n-butanol, petroleum ether, dichloromethane, and ethyl acetate.
Diterpenes of the family Pteridaceae were classified into ent-kaurane or ent-kaurenes with hydroxylation mainly encountered in any of the 2, 16, 11, 7, 9, 15, 14, 19, and 12 positions as mono-, bi-, tri-, or tetra-hydroxylation. Epoxidations were seen usually in positions 9, 11, or 16, and oxidation was shown mostly in position 15. Sugar attachments were commonly formed by the hydroxy groups in position 2 or 19, and the type of sugar was primarily glucose or allose [5]. Kaurenes were reported in abundance from plants and fungi [94], and in ferns, the family Pteridaceae produced almost all the ent-kaurene or ent-kaurane moieties. Several kaurene and ent-kaurene compounds were isolated in the 95% ethanolic extract of the spike-tooth fern, Pteris dispar Kunze. whole plant, among which the 18-acetoxy derivatives 306 and 307 were isolated for the first time and hydroxylations were seen in carbons 17 and 7 [95]; moreover, the isodon 7,20-epoxy-ent-kuarenoid neolaxiflorin L was identified, which was previously isolated from the plant Isodon excisoides (Y.Z.Sun) H.Hara, studied for its SAR, and many of its analogs were synthesized to optimize its promising anticancer effects [96]. Neolaxiflorin L 310 manifested a potent cytotoxicity against the four cell lines HCT-116, HepG2, NCI-H1650, and A2780 with an IC50 range between 1 and 5 µg/mL [95]. The ent-kaurane-6β,16α-diol-3-one 315 (Figure 8) was identified from Pteris ensiformis Burm. as the first ent-kaurane with an oxidized C-3 position in the Pteris genus and Pterideaceae family [97].

3.4.2. Aspleniaceae Family

Another class of triterpenes was the cycloartane type glycosides characterized by the rare 9α,10α-cyclopropyl ring; these were isolated from Asplenium ruprechtii Sa. Kurata. While 321 and 322 possessed the typical three-membered ring, 323-325 (Figure 8) comprised a new Asplenium moiety of 9,19-seco-cycloartane-9,11-en triterpene aglycone in a highly 27 oxidized skeleton of methylene, methylene, or quaternary carbons in the order of 3, 7 (or 23), 24, 25, 30 [98]. 321 and 322 possessed an antitumor activity within 18–60 μM (Table 1).
The cycloartane glycoside aspleniumside A 321 (Figure 9) was isolated from Asplenium ruprechtii Sa.Kurata. The relative configuration was assigned using NOESY knowing that CH2-19 in this plant genus is found in the β-orientation. Aspleniumside B was comparably assigned to aspleniumside A, being that it only lacked one hexosyl moiety, while aspleniumside C-E belonged to the 9,11-en, 9,19-seco-cycloartane triterpene skeleton. The in vitro assays against the HL-60 and HepG2 cell lines showed significant cytotoxic effects of these compounds with IC50 values between 18 and 60 μM [98]. Both the 23, 25-dihydroxycycloartane, aspleniumside F 326 (Figure 9) [99] and the 9, 19-cycloartanes, aspleniumside G 327 [100], as well as aspleniumside H 328 and I 329 were isolated from Asplenium ruprechtii Sa.Kurata (Figure 9) but with no appreciable bioactivity. It was notable that these compounds were isolated from the same fern by the same group, and thus, it was better to publish them in the same study. He et al. revealed the isolation of cycloartane triterpenoidal saponins 330-332 (Figure 9) with a triglycosidic nature, which was further confirmed by acid hydrolysis. 333 and 334 were closely related to 330 except for an extra double bond in C-25 or a change in the sugar moieties, respectively. The compounds were obtained in minute yields and may benefit from biotechnological applications to optimize their amounts [101].
Table 1. Compilation of isolated compounds, their sources, and bioactivities from ferns.
Table 1. Compilation of isolated compounds, their sources, and bioactivities from ferns.
No.Chemical NameFern NameBiological ActivityReferences
1Paleacenins A Elaphoglossum paleaceum (Hook. & Grev.) Sledge MAO inhibitor[18]
2Paleacenins B Elaphoglossum paleaceum (Hook. & Grev.) Sledge MAO inhibitor[18]
3Aspidinol PDryopteris crassirhizoma Nakai antimicrobial[27]
4Aspidinol BDryopteris crassirhizoma Nakaiantimicrobial[27]
5Propionyl-3-methylphloroglucinolDryopteris crassirhizoma Nakai [27]
6Butyryl-3-methylphloroglucinol Dryopteris crassirhizoma Nakai [27]
7Butyrylphloroglucinol Dryopteris crassirhizoma Nakai [27]
81,3-(2,4,6-trihydroxyphenyl)dibutanoneDryopteris crassirhizoma NakaiPTP1B inhibitor[29]
9Tripropionylphloroglucinol Dryopteris crassirhizoma NakaiPTP1B inhibitor[29]
10Tributyrylphloroglucinol Dryopteris crassirhizoma NakaiPTP1B inhibitor[29]
11Dryopidin PBDryopteris crassirhizoma NakaiPTP1B inhibitor[29]
12Methylene-bis-phlorobutyrophenoneDryopteris crassirhizoma NakaiPTP1B inhibitor[29]
13Araspidin BB Dryopteris crassirhizoma Nakai [29]
14Norflavaspidic acid APDryopteris crassirhizoma Nakai [29]
15Norflavaspidic acid AB Dryopteris crassirhizoma NakaiPTP1B inhibitor[29]
16Norflavaspidic acid PB Dryopteris crassirhizoma NakaiPTP1B inhibitor[29]
17Norflavaspidic acid BB Dryopteris crassirhizoma Nakai [29]
18Flavaspidic acid AADryopteris crassirhizoma Nakai [29]
19Flavaspidic acid AP Dryopteris crassirhizoma Nakai [29]
20Flavaspidic acid AB Dryopteris crassirhizoma Nakai [29]
21Flavaspidic acid PBDryopteris crassirhizoma Nakai [29]
22Dryopcrassirine AB Dryopteris crassirhizoma Nakai [29]
23Nortrisflavaspidic acid ABBDryopteris crassirhizoma NakaiPTP1B inhibitor[29]
24Trisflavaspidic acid ABBDryopteris crassirhizoma Nakai [29]
25Trisflavaspidic acid BBB Dryopteris crassirhizoma Nakai [29]
26Filixic acid ABADryopteris crassirhizoma NakaiAntiviral H1N1[29]
27Dryocrassin ABBADryopteris crassirhizoma Nakai [29]
28Disaspidin BB Dryopteris fragrans (L.) Schottantimicrobial [30]
29(+)-Dryocrassoid ADryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
30(-)-Dryocrassoid ADryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
31Enol-(+)-Dryocrassoid ADryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
32Enol-(-)-Dryocrassoid ADryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
33(+)-Dryocrassoid BDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
34(-)-Dryocrassoid BDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
35Enol(+)-Dryocrassoid BDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
36Enol(-)-Dryocrassoid BDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
37(+)-Dryocrassoid CDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
38(-)-Dryocrassoid CDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
39(+)-Dryocrassoid DDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
40(-)-Dryocrassoid DDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
41Enol(+)-Dryocrassoid DDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
42Enol(-)-Dryocrassoid DDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
43(+)-Dryocrassoid EDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
44(-)-Dryocrassoid EDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
45(+)-Dryocrassoid FDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
46(-)-Dryocrassoid FDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
47(+)-Dryocrassoid GDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
48(-)-Dryocrassoid GDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
49(+)-Dryocrassoid HDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
50(-)-Dryocrassoid HDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
51(+)-Dryocrassoid IDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
52(-)-Dryocrassoid IDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
53(+)-Dryocrassoid JDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
54(-)-Dryocrassoid JDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
55Albaspidin AP Dryopteris crassirhizoma Nakai [20]
56Albaspidin AA Dryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
57Albaspidin ABDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
58Araspindin BBDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
591-[3-[(3-acetyl-2,4,6-trihydroxyphenyl)methyl]-2,4,6-trihydroxyphenyl] butanoneDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
60Methylene-bis-phlorobutyrophenoneDryopteris crassirhizoma Nakaianti-HSV 1 and anti-RSV[20]
61Phloropyron A Dryopteris championii (Benth.) C. Chr. antibacterial[34]
62Phloropyron BDryopteris championii (Benth.) C. Chr. bactericidal[34]
63Phloropyron CDryopteris championii (Benth.) C. Chr. [34]
64Margaspidin ADryopteris championii (Benth.) C. Chr. [34]
65Margaspidin BDryopteris championii (Benth.) C. Chr. [34]
66Pseudoaspidinol ADryopteris championii (Benth.) C. Chr. bactericidal[34]
67Dryoptol ADryopteris crassirhizoma Nakai [21]
683″-Epi-dryoptol A Dryopteris crassirhizoma Nakaianti-inflammatory[21]
69Dryoptol BDryopteris crassirhizoma Nakaiantifungal+FLC[21]
703″-epi-dryoptol B Dryopteris crassirhizoma Nakaiantifungal+FLC[21]
71Dryoptol CDryopteris crassirhizoma Nakai [21]
723″-Epi-dryoptol CDryopteris crassirhizoma Nakai [21]
73Dryoptol DDryopteris crassirhizoma Nakaiantifungal+FLC[21]
743″-epi-dryoptol DDryopteris crassirhizoma Nakai [21]
75Dryoptol EDryopteris crassirhizoma Nakaiantifungal+FLC[21]
763″-Epi-dryoptol EDryopteris crassirhizoma Nakaiantifungal+FLC[21]
77Dryoptol FDryopteris crassirhizoma Nakai [21]
783”-Epi-dryoptol FDryopteris crassirhizoma Nakai [21]
79Dryoptol GDryopteris crassirhizoma Nakaianti-inflammatory[21]
803″-Epi-dryoptol GDryopteris crassirhizoma Nakaianti-inflammatory[21]
81Dryoptol HDryopteris crassirhizoma Nakai [21]
823″-Epi-dryoptol HDryopteris crassirhizoma Nakai [21]
83Dryoptol IDryopteris crassirhizoma Nakaiantifungal+FLC[21]
843″-Epi-dryoptol IDryopteris crassirhizoma Nakaiantifungal+FLC[21]
85Dryoptol JDryopteris crassirhizoma Nakai [21]
863″-Epi-dryoptol JDryopteris crassirhizoma Nakai [21]
87Dryoptol KDryopteris crassirhizoma Nakaianti-inflammatory y, antifungal+FLC[21]
883″-Epi-dryoptol KDryopteris crassirhizoma Nakaianti-inflammatory, antifungal+FLC[21]
89Dryoptol LDryopteris crassirhizoma Nakaiantifungal+FLC[21]
903″-Epi-dryoptol LDryopteris crassirhizoma Nakaiantifungal+FLC[21]
91Dryoptin ADryopteris crassirhizoma Nakai [23]
92Epi-dryoptin ADryopteris crassirhizoma Nakaiantifungal[23]
93Dryoptin BDryopteris crassirhizoma Nakai [23]
94Epi-dryoptin BDryopteris crassirhizoma Nakai [23]
95Dryoptin CDryopteris crassirhizoma Nakaiantifungal[23]
96Epi-dryoptin CDryopteris crassirhizoma Nakaiantifungal[23]
97Dryoptin DDryopteris crassirhizoma Nakaiantifungal[23]
98Epi-dryoptin DDryopteris crassirhizoma Nakaiantifungal[23]
99Dryoptin EDryopteris crassirhizoma Nakai [23]
100Epi-dryoptin EDryopteris crassirhizoma Nakai [23]
101Dryofragone Dracaena fragrans (L.) SchottCytotoxic[35]
102Dryofracoumarin B Dracaena fragrans (L.) SchottCytotoxic[35]
103NorflavesoneDracaena fragrans (L.) SchottCytotoxic[35]
104Aspidinol Dracaena fragrans (L.) Schottcytotoxic and immunomodulator[35]
105Dryofracoumarin A Dracaena fragrans (L.) Schottcytotoxic and immunomodulator[35]
106Vitamin E quinone Dracaena fragrans (L.) Schottcytotoxic[35]
107AlbicanolDracaena fragrans (L.) Schottcytotoxic[35]
1082′,4′-dihydroxy-6′-methoxy-3′,5′-dimethylchalcone Dracaena fragrans (L.) Schottcytotoxic[35]
109Fern-7(8)-en-19α, 28-diol Adiantum capillus-veneris Linn.antifungal[71]
1103β,4α,25-trihydroxyfilican Adiantum capillus-veneris Linn.antifungal[71]
111Pteron-14-ene-7α,19α,28-triol Adiantum capillus-veneris Linn.antifungal[71]
11222-hydroxyhopane Adiantum latifolium Lam.trehalase inhibitory and larvicidal activity[72]
113Ancepsone AAleuritopteris anceps (Blanf.) Panigrahianticancer[36]
1145-hydroxy-3,7,4′-trimethoxyflavoneAleuritopteris anceps (Blanf.) Panigrahianticancer[36]
115β-sitosterolAleuritopteris anceps (Blanf.) Panigrahianticancer[36]
1165-hydroxy-7,4′-dimethoxyflavoneAleuritopteris anceps (Blanf.) Panigrahianticancer[36]
1173,7,3′,4′-tetramethyl-quercetinAleuritopteris anceps (Blanf.) Panigrahianticancer[36]
1185,4′-dihydroxy-3,7-dimethoxyflavoneAleuritopteris anceps (Blanf.) Panigrahianticancer[36]
119KaempferolAleuritopteris anceps (Blanf.) Panigrahianticancer[36]
120Daucosterol Aleuritopteris anceps (Blanf.) Panigrahianticancer[36]
121Quercetin-3-glucopyranoside Aleuritopteris anceps (Blanf.) Panigrahianticancer[36]
12214-oxy-7α,20-dihydroxycyath-12,18-dieneAleuritopteris albofusca (Baker) Pic.Serm [79]
123Ent-8[14],15-pimaradiene-2β,19-diolAleuritopteris albofusca (Baker) Pic.Serm [79]
124Myricetin 3,7,3′,5′-tetramethyl ether Aleuritopteris albofusca (Baker) Pic.Serm [79]
125Ethyl caffeateAleuritopteris albofusca (Baker) Pic.Serm [79]
126Ent-16-oxo-17-norkauran-19-oic acidAleuritopteris albofusca (Baker) Pic.Serm [79]
127Onychiol B Aleuritopteris albofusca (Baker) Pic.Serm [79]
128Aleuritopsis BAleuritopteris albofusca (Baker) Pic.Serm [79]
129Astragalin Aleuritopteris albofusca (Baker) Pic.Serm [79]
130HyperinAleuritopteris albofusca (Baker) Pic.Serm [79]
131Ptercresion A Pteris cretica L. [80]
132Ptercresion BPteris cretica L.hepatoprotective[80]
133Ptercresion CPteris cretica L.hepatoprotective[80]
134Callisalignene D Pteris cretica L.hepatoprotective[80]
135Berberiside A Pteris cretica L. [80]
136Creoside I Pteris cretica L. [80]
137(2S,3S)-pterosin C 3-O-β-D-[4′-(E)-caffeoyl]-glucopyranosidePteris multifida Poir.anticancer[86]
138(2S,3S)-pterosin C 3-O-β-D-[6′-(E)-P-coumaroyl]-glucopyranosidePteris multifida Poir.anticancer[86]
139(2S,3S)-pterosin C3-O-β-D-glucopyranosidePteris multifida Poir. [86]
140(2S,3S)-3-methoxypterosinPteris multifida Poir. [86]
141(2S,3S)-pterosin CPteris multifida Poir. [86]
142(2S,3S)-pterosin CPteris multifida Poir. [86]
143(2R)-pterosin BPteris multifida Poir. [86]
144(2R)-pterosin B 4-O-β-D-glucopyranosidePteris multifida Poir. [86]
145(3R)-pterosin DPteris multifida Poir. [86]
146(3R)-pterosin WPteris multifida Poir. [86]
147(3R)-pteroside W Pteris multifida Poir. [86]
148(2R,3R)-pterosin LPteris multifida Poir. [86]
149Pterosinside A Pteris laeta Walls neuroprotective[37]
150Pterosinside BPteris laeta Walls neuroprotective[37]
151Pterosinside CPteris laeta Walls neuroprotective[37]
1528-Hydroxy-(-)-isolariciresinol-9-O-β-D-xylopyranoside.Pteris laeta Walls [37]
153(+)-(7S, 8R, 7′S, 8′S)-8-hydroxysyringaresinol-8-β-D-xylocoside]Pteris laeta Walls [37]
154(+)-(7S, 8R, 7′S, 8′S)-8-hydroxysyringaresinol-8-β-D-xylocosidePteris laeta Walls [37]
155Pteroside CPteris laeta Walls neuroprotective[37]
156Pteroside B Pteris laeta Walls neuroprotective[37]
157Pterosin E (unidentified chirality)Pteris laeta Walls neuroprotective[37]
158Pterosin Z (unidentified chirality)Pteris laeta Walls neuroprotective[37]
159Pterosin D (unidentified chirality)Pteris laeta Walls neuroprotective[37]
160Pterosin X (unidentified chirality)Pteris laeta Walls neuroprotective[37]
161Pterosin B (unidentified chirality)Pteris laeta Walls neuroprotective 200 times Ginsenoside Rg1[37]
162Dehydrated pterosin S Pteris laeta Walls neuroprotective[37]
163Pterosin S (unidentified chirality)Pteris laeta Walls neuroprotective[37]
164Pterosin C (unidentified chirality)Pteris laeta Walls neuroprotective[37]
1652-hydroxypterosin CPteris laeta Walls neuroprotective[37]
166Pterosin A (unidentified chirality)Pteris laeta Walls neuroprotective[37]
167Pterosin T (unidentified chirality)Pteris laeta Walls neuroprotective[37]
1681-hydroxy-2-epipinoresinol-1-β-D-glucosidePteris laeta Walls neuroprotective[37]
169Schilignan F Pteris laeta Walls neuroprotective[37]
1701-hydroxypinoresinol-1-β-D-glucosidePteris laeta Walls neuroprotective[37]
171Obtupterosin A Pteris obtusiloba Ching & S.H. Wucytotoxic[92]
172Obtupterosin BPteris obtusiloba Ching & S.H. Wucytotoxic[92]
173Obtupterosin CPteris obtusiloba Ching & S.H. Wucytotoxic[92]
174Dehydropterosin B Pteris obtusiloba Ching & S.H. Wu [92]
175Dehydropterosin Q Pteris obtusiloba Ching & S.H. Wu [92]
176(2S)-pterosin P Pteris obtusiloba Ching & S.H. Wu [92]
177(2S,3S)-pterosin S Pteris obtusiloba Ching & S.H. Wu [92]
178(2R,3S)-pterosin S Pteris obtusiloba Ching & S.H. Wu [92]
179(2S)-pterosin B 14-O-β-D-glucopyranoside Pteris obtusiloba Ching & S.H. Wu [92]
180(3R)-pteroside WPteris obtusiloba Ching & S.H. Wu [92]
181Pterosin S 14-O-β-D-glucopyranosidePteris obtusiloba Ching & S.H. Wu [92]
182Creticolacton A Pteris cretica L.cytotoxic[93]
18313-hydroxy-2(R),3(R)-pterosin L Pteris cretica L.cytotoxic[93]
184Creticoside APteris cretica L. [93]
185Spelosin 3-O-β-D-glucopyranoside Pteris cretica L. [93]
186Pterosin DPteris cretica L. [93]
187(3R)-pterosin W Pteris cretica L. [93]
188(3R)-pterosin D 3-O-β-D-glucopyranoside Pteris cretica L. [93]
189(3R)-pteroside WPteris cretica L. [93]
1902R,3R-pterosin L 3-O-β-D-glucopyranosidePteris cretica L. [93]
191Geopyxin BPteris dispar Kunzecytotoxic[23]
192Geopyxin EPteris dispar Kunzecytotoxic[23]
193Ent-11α-hydroxy-18-acetoxykaur-16-enePteris dispar Kunze [23]
194Ent-14β-hydroxy-18-acetoxykaur-16-enePteris dispar Kunze [23]
195 Neolaxiflorin LPteris dispar Kunze [23]
196Ent-3β, 19-dihydroxy-kaur-16-enePteris dispar Kunze [23]
197Ent-3β-hydroxy-kaur-16-enePteris dispar Kunze [23]
1987β, 17-dihydroxy-16α-ent-kauran-19-oic acid 19-O-β-D-glucopyranoside esterPteris dispar Kunze [23]
200Crotonkinin CPteris dispar Kunze [23]
201Ent-kaurane-6β,16α-diol-3-onePteris ensiformis Burm. [97]
202Ent-7α,9α-dihydroxy-15-oxokaur-16-en-196β-olidePteris ensiformis Burm.cytotoxic[97]
203Ent-11α-hydroxy-15-oxokaur-16-en-19-oic acid Pteris ensiformis Burm.cytotoxic[97]
2049, 15β-dihydroxy-ent-kaur-16-en-19-oic acidPteris ensiformis Burm. [97]
205Foliol Pteris ensiformis Burm. [97]
206(16R)-11β-hydroxy-15-oxo-ent-kauran-19-oic acidPteris ensiformis Burm. [97]
207Pterosin C (unidentified chirality)Pteris ensiformis Burm. [97]
208Pterosin DPteris ensiformis Burm. [97]
209Pterosin LPteris ensiformis Burm. [97]
210Aspleniumside AAsplenium ruprechtii Sa. Kuratacytotoxic[98]
211Aspleniumside BAsplenium ruprechtii Sa. Kuratacytotoxic[98]
212Aspleniumside CAsplenium ruprechtii Sa. Kuratacytotoxic[98]
213Aspleniumside DAsplenium ruprechtii Sa. Kurata [98]
214Aspleniumside EAsplenium ruprechtii Sa. Kurata [98]
215Aspleniumside F Asplenium ruprechtii Sa. Kurata [100]
216Aspleniumside GAsplenium ruprechtii Sa. Kurata [100]
217Aspleniumside HAsplenium ruprechtii Sa. Kurata [100]
218Aspleniumside IAsplenium ruprechtii Sa. Kurata [100]
2193β,23, 24, 25, 30-pentahydroxycycloartane-3, 25, 30-tri-O-β-D-glucopyranosideCamptosorus sibiricus Rupr. [101]
2203β,23,24,30-tetrahydroxycycloartane-25-en-3-O-β-D-glucopyranosyl-(1→4)-β-D-glucopyranosyl-(1→2)-β-D-glucopyranosideCamptosorus sibiricus Rupr. [101]
2213β,23,24,30-tetrahydroxycycloartane-25-en-3-O-β-D-Glucopyranosyl-(1→2)-β-D-glucopyranosyl-(1→2)-β-D-glucopyranosyl–(1→4)-β-D-glucopyranoside Camptosorus sibiricus Rupr. [101]
222Kaempferol-3-O-β-D-glucopyranosideCamptosorus sibiricus Rupr. [101]
223Kaempferol-3, 7-O-β-D-diglucopyranoside Camptosorus sibiricus Rupr. [101]
224Kaempferol-3-O-β-D-glucoside-7-O-α-L-rhamnosideCamptosorus sibiricus Rupr. [101]
225Kaempferol-3-O-β-D-glucopyranosyl-(1→2)-β-D-glucopyranoside Camptosorus sibiricus Rupr. [101]
226Camsibriside BCamptosorus sibiricus Rupr. [101]
227Camsibriside ACamptosorus sibiricus Rupr. [101]
228CamptosideCamptosorus sibiricus Rupr.NO inhibitor[38]
229Dehydrodiconiferyl alcohol-9′-O-β-D-glucopyranoside Camptosorus sibiricus Rupr.NO inhibitor[38]
230AesculetinCamptosorus sibiricus Rupr.NO inhibitor[38]
231VajicosideCamptosorus sibiricus Rupr. [38]
232(5R,6S,8R,9S,10R)-6-O-[β-D-glucopyranosyl-(1→4)-α-l-rhamnopyranosyl] cleroda-3,13(16),14-dieneDicranopteris pedata (Houtt) Nakaikeweak cytotoxicity[102]
233(5R,6S,8R,9S,10R,13S)-6-O-[β-D-glucopyranosyl-(1→4)-α-l-rhamnopyranosyl]-2-ox-oneocleroda-3,13-dien-15-ol.Dicranopteris pedata (Houtt) Nakaikeweak cytotoxicity[102]
234(5R,6S,8R,9S,10R)-6-O-[β D-glucopyranosyl-(1→4)-α-l-rhamnopyranosyl]-(13E)-2-oxoneocleroda-3,14-dien-13-olDicranopteris pedata (Houtt) Nakaikeweak cytotoxicity[102]
235Dryofraterpene ADracaena fragrans (L.) Schottcytotoxic[39]
236YomoginDracaena fragrans (L.) Schott [39]
237Pinoresinol Dracaena fragrans (L.) Schott [39]
238 (2R)-3-(4′-Hydroxyphenyl) lactic acid Diplazium esculentum (Retz.) Sw. [63]
239Ponastrone ADiplazium esculentum (Retz.) Sw. [63]
240Amarasterone A1Diplazium esculentum (Retz.) Sw. [63]
241Makisterone CDiplazium esculentum (Retz.) Sw. [63]
242Matteucens IMatteuccia orientalis (Hook.) Trevis. [40]
243Matteucens JMatteuccia orientalis (Hook.) Trevis. [40]
244Demethylmatteucinol Matteuccia intermedia C.Chr. [44]
245Matteflavosides HMatteuccia intermedia C.Chr. [44]
246Matteflavosides IMatteuccia intermedia C.Chr. [44]
247Matteflavoside JMatteuccia intermedia C.Chr. [44]
248Matteuinterate AMatteuccia intermedia C.Chr. [44]
249Matteuinterate BMatteuccia intermedia C.Chr. [44]
250Matteuinterate CMatteuccia intermedia C.Chr. [44]
251Matteuinterate DMatteuccia intermedia C.Chr. [44]
252Matteuinterate EMatteuccia intermedia C.Chr. [44]
253Matteuinterate FMatteuccia intermedia C.Chr. [44]
254DemethoxymatteucinolMatteuccia intermedia C.Chr. [44]
255Farrerol Matteuccia intermedia C.Chr.α-Glucosidase inhibitory[44]
256MatteucinMatteuccia intermedia C.Chr.α-Glucosidase inhibitory[44]
257Matteucinol Matteuccia intermedia C.Chr.α-Glucosidase inhibitory[44]
258MethoxymatteucinMatteuccia intermedia C.Chr.α-Glucosidase inhibitory[44]
2593′-hydroxy-matteucinolMatteuccia intermedia C.Chr.α-Glucosidase inhibitory[44]
260Cyrtominetin Matteuccia intermedia C.Chr.α-Glucosidase inhibitory[44]
2615,7-dihydroxy-4′-methoxy-6-methyl-flavanoneMatteuccia intermedia C.Chr. [44]
262Demethoxymateucinol 7-O-glucosideMatteuccia intermedia C.Chr. [44]
263Farrerol 7-O-glucoside Matteuccia intermedia C.Chr. [44]
264Matteucinol 7-O-glucoside Matteuccia intermedia C.Chr. [44]
265Myrciacitrin IIMatteuccia intermedia C.Chr. [44]
266Matteflavoside GMatteuccia intermedia C.Chr. [44]
267Matteuorienate AMatteuccia intermedia C.Chr. [44]
268Matteuorienate BMatteuccia intermedia C.Chr. [44]
269Matteuorienate DMatteuccia intermedia C.Chr. [44]
270Matteuorienate FMatteuccia intermedia C.Chr. [44]
271Matteuorienate H Matteuccia intermedia C.Chr. [44]
272Matteuorienate IMatteuccia intermedia C.Chr. [44]
273Matteuorienate JMatteuccia intermedia C.Chr. [44]
274Matteuorienate KMatteuccia intermedia C.Chr. [44]
275Matteuinterin AMatteuccia intermedia C.Chr. [39]
276Matteuinterin BMatteuccia intermedia C.Chr. [39]
277Matteuinterin CMatteuccia intermedia C.Chr. [39]
278Kankanoside PMatteuccia intermedia C.Chr. [39]
279(3S,6S)-6,7-dihydroxy-6,7-dihydrolinalool 3-O-β-D-glucopyranosideMatteuccia intermedia C.Chr.anti-inflammatory[39]
280(6R,7E,9R)-9-hydroxy-4,7-megastigmadien-3-one 9-O-β-D-glucopyranosideMatteuccia intermedia C.Chr. [39]
281(6S,7E,9R)-9-hydroxy-4,7-megastigmadien-3-one 9-O-β-D-gluco-pyranosideMatteuccia intermedia C.Chr. [39]
282Byzantionoside BMatteuccia intermedia C.Chr. [39]
283Isodonmegastigmane IMatteuccia intermedia C.Chr. [39]
2849-O-β-D-glucopyranosyloxy-5-megastigmen-4-oneMatteuccia intermedia C.Chr.anti-inflammatory[39]
285Eriodictyol-7-O-glucuronide Tectaria coadunata (Wall. ex Hook. & Grev.) C.Chr [47]
286Luteolin-7-O-glucoronide Tectaria coadunata (Wall. ex Hook. & Grev.) C.Chr [47]
287Dryocrasspherol ADryopteris crassirhizoma Nakai anti-respiratory syncytial virus activity[26]
288Dryocrasspherol BDryopteris crassirhizoma Nakai [26]
289Dryocrasspherol CDryopteris crassirhizoma Nakai [26]
290Dryocrasspherol DDryopteris crassirhizoma Nakai anti-inflammatory[26]
291Dryocrasspherol EDryopteris crassirhizoma Nakai anti-inflammatory[26]
292Fragranoside B Dracaena fragrans (L.) Schott [48]
293Dihydroconiferyl alcohol Dracaena fragrans (L.) Schott [48]
2941,3-dihydroxy-5-propylbenzeneDracaena fragrans (L.) Schottanticancer[48]
2954-hydroxyacetophenone Dracaena fragrans (L.) Schott [48]
2963,4-dihydroxybenzaldehyde Dracaena fragrans (L.) Schott [48]
2972-ethyl-6-hydroxybenzoic acid Dracaena fragrans (L.) Schott [48]
2983,4-dihydroxyacetophenone Dracaena fragrans (L.) Schottanticancer[48]
299(+)-afzelechin-7-O-α-L-arabinofuranoside Polypodium vulgare L. [1]
300(+)-afzelechin-7-O-β-D-apiofuranoside Polypodium vulgare L. [1]
301(+)-afzelechin Polypodium vulgare L. [1]
302(+)-catechin-7-O-α-L-arabinofuranoside Polypodium vulgare L. [1]
303(+)-catechin-7-O-β-D-apiofuranoside Polypodium vulgare L. [1]
304Liglaurate ADrynaria roosii Nakaikeantioxidant, anti-inflammatory, and cytotoxic[52]
305Liglaurate BDrynaria roosii Nakaikeantioxidant, anti-inflammatory, and cytotoxic[52]
306Liglaurate CDrynaria roosii Nakaikeantioxidant, anti-inflammatory, and cytotoxic[52]
307Liglaurate DDrynaria roosii Nakaikeantioxidant, anti-inflammatory, and cytotoxic[52]
308Liglaurate EDrynaria roosii Nakaike [52]
309Methyl 12-caffeoyloxylaurateDrynaria roosii Nakaike [52]
310Glyceryl 12-caffeoyloxylaurateDrynaria roosii Nakaikeantioxidant, anti-inflammatory[52]
311Protocatechuic acid methyl ester-4-O-(6′-O-proto-catEchuoyl)-β-D-glucopyranosidePhymatopteris hastata (Thunb.) Pic. Serm. α-glucosidase inhibitory[53]
312(-)/(+)-naringenin-7-O-α-L-rhamnopyranoside Phymatopteris hastata (Thunb.) Pic. Serm. [53]
313CoumarinPhymatopteris hastata (Thunb.) Pic. Serm. α-glucosidase inhibitory[53]
314Trans-O-coumaric acidPhymatopteris hastata (Thunb.) Pic. Serm. [53]
315Trans-caffeic acidPhymatopteris hastata (Thunb.) Pic. Serm. [53]
316JuglaninPhymatopteris hastata (Thunb.) Pic. Serm. [53]
317Kaempferol 3-O-α-L-rhamnopyranosidePhymatopteris hastata (Thunb.) Pic. Serm. [53]
318Kaempferol-7-O-α-L-rhamnopyranosidePhymatopteris hastata (Thunb.) Pic. Serm. [53]
319Avicularin Phymatopteris hastata (Thunb.) Pic. Serm. [53]
320ZeaxanthinPhymatopteris hastata (Thunb.) Pic. Serm. α-glucosidase inhibitory[53]
321Kaempferol-3-O-α-L-arabinofuranosyl-7-O-α-L-rhamnopyranosidePhymatopteris hastata (Thunb.) Pic. Serm. α-glucosidase inhibitory[53]
322Kaempferol 3-O-[α-D-apiofuranosyl-(1–2)-α-L-arabinofuranoside]-7-O-α-L-rhamnopyranosidePhymatopteris hastata (Thunb.) Pic. Serm. [53]
323β-ecdysterone Phymatopteris hastata (Thunb.) Pic. Serm. [53]
324Trans-melilotoside Phymatopteris hastata (Thunb.) Pic. Serm. [53]
325(-)-epi-osmundalactone Angiopteris helferiana C.Presl [54]
326AngiopterosideAngiopteris helferiana C.Presl [54]
3272-deprenyl-5-O-methyl-7-methoxy-rheediaxanthone B Metaxya rostrata (Kunth) C.Preslanticancer[60]
3282-deprenyl-6-O-methyl-7-hydroxy-rheediaxanthone B Metaxya rostrata (Kunth) C.Preslanticancer[60]
3292-deprenyl-5-O-methyl-7-hydroxy-rheediaxanthone BMetaxya rostrata (Kunth) C.Preslanticancer[60]
3304-O-coumaroyl-β-D-allose Cibotium barometz (L.) J. Sm. hepatoprotective, anti-inflammatory[42]
3314-O-caffeoyl-β-D-allose Cibotium barometz (L.) J. Sm. [42]
332Ethyl (4-O-caffeoyl)-β-D-glucopyranoside Cibotium barometz (L.) J. Sm. [42]
333Cibotiumbaroside DCibotium barometz (L.) J. Sm.hepatoprotective[42]
3344-O-coumaroylglucose Cibotium barometz (L.) J. Sm. [42]
3356-O-coumaroylglucose Cibotium barometz (L.) J. Sm. [42]
3364-O-caffeoylglucoseCibotium barometz (L.) J. Sm.hepatoprotective[42]
3376-O-caffeoylglucoseCibotium barometz (L.) J. Sm. [42]
338Caffeic acid 4-O-β-D-glucopyranoside Cibotium barometz (L.) J. Sm.anti-inflammatory[42]
339Ferulic acid 4-O-β-D-glucopyranoside Cibotium barometz (L.) J. Sm.hepatoprotective[42]
340Vanillic acid 4-O-β-D-glucopyranosideCibotium barometz (L.) J. Sm. [42]
341Cyathenosin ACibotium barometz (L.) J. Sm. [42]
342Jixueqisus A Pronephrium penangianum (Hook.) Holtt [61]
343Jixueqisus B Pronephrium penangianum (Hook.) Holtt [61]
344Jixueqisus CPronephrium penangianum (Hook.) Holtt [61]
345Jixueqisus D Pronephrium penangianum (Hook.) Holtt [61]
346Jixueqisus E Pronephrium penangianum (Hook.) Holtt [61]
347Jixueqisus F Pronephrium penangianum (Hook.) Holtt [61]
348(2S)-5,2′,5′-trihydroxy-7-methoxyflavanonePronephrium penangianum (Hook.) Holtt [61]
3495,2′, 5′-trihydroxy-7-methoxyflavonePronephrium penangianum (Hook.) Holtt [61]
350Abacopterin C Pronephrium penangianum (Hook.) Holtt cytotoxic[61]
351Abacopterin APronephrium penangianum (Hook.) Holtt cytotoxic[61]
352Eruberin BPronephrium penangianum (Hook.) Holtt cytotoxic[61]
353Triphyllin A Pronephrium penangianum (Hook.) Holtt cytotoxic[61]
354Eruberin APronephrium penangianum (Hook.) Holtt [61]
355Abacopterin I Pronephrium penangianum (Hook.) Holtt [61]
356Abacopterin KPronephrium penangianum (Hook.) Holtt [61]
357(S, E)-7-(2-octylcyclopropylidene) heptanoic acid
(Adiantic acid)
Adiantum flabellulatum L. [67]
358Stigmast-4-en-6-β-ol-3-one Adiantum flabellulatum L. [67]
359β-sitosterol Adiantum flabellulatum L. [67]
360Isoadiantol AAdiantum flabellulatum L. [67]

3.4.3. Gleicheniaceae Family

Despite its widespread distribution in marine organisms and the plant kingdom, clerodane diterpenes were reported only in ferns from five species of the genus Dicranopteris, which varied according to the junction position between the sugar and the decalin ring or the C-9 moiety [102]. Dicranopteris pedata (Houtt.) Nakaike introduced three new clerodane molecules 333-335 (Figure 9) with weak cytotoxic activity. All the structures were of high resemblance to the known compound (6S,13S)-6-{[β-D-glucopyranosyl-(1-4)-α-L-rhamnopyranosyl] oxy} cleroda 3,14-dien-13-ol except in the attachment of the sugars and the position of the hydroxyl or methyl groups in one or two positions [103].

3.4.4. Dryopteridaceae Family

From Dracaena fragrans (L.) Schott, the new sesquiterpene dryofraterpene A 336 (Figure 9) or the (7S,10S)-2,3-dihydroxy-calamenene-15-carboxylic acid methyl ester was isolated and characterized. Its antitumor activity was revealed against the cell lines A549, MCF7, HepG2, HeLa, and PC-3 with an IC50 below 10 μM compared to taxol as a positive control [104].

3.4.5. Athyriaceae Family

Gymnomitrane sesquiterpenes were newly reported from ferns and included matteuinterin A 339 (Figure 9). Interestingly, 340 showed a potent anti-inflammatory effect by reducing PGE2 in lipopolysaccharide (LPS)-induced RAW 264.7 murine macrophages with an IC50 value of 17 μM, which was even better than the minocycline control [44]. Gymnomitrane sesquiterpenes were detected from plants, liverworts, and fungi, namely, Codonopsis pilosula (Franch.) Nannf., Ganoderma lucida (Lingzhi or Reishi) [105], basidiomycete Antrodiella albocinnamomea [106], Cylindrocolea recurvifolia [Steph.] Inoue [107], and Jungermannia truncata NEES [108]. While matteuinterin A was identified by analogy of its aglycone part to 3-gymnomitren-15-ol, matteuinterin B displayed a resemblance to leptorumol 7-O-β-D-glucopyranoside, yet the attachment of the 3-hydroxy-3-methylglutaryl (HMG) fragment to the C-6′ position was not seen before and was confirmed by HMBC [44].

3.5. Polysaccharides

The widespread use of calorie- and simple sugar-rich foods, which are usually devoid of any dietary fibers, vitamins, or complex carbohydrates, throughout the previous decades was implicated in the obesity epidemic and its chronic consequences such as stroke, cardiovascular diseases, and cancer. Accordingly, scientific attention was drawn towards using functional foods and their phytochemical components for their undeniable various health-promoting aspects. Among other components are polysaccharides from different sources such as algae and plants, which are reported to have prebiotic, anticancer, antidiabetic, antiviral, immunomodulatory, wound-healing, and antibacterial effects [109]. Probably the most promising are the water-soluble polysaccharides based on their safety profile and variable bioactivities, for instance, antihyperglycemic, immunomodulatory, and antitumor activities [110,111,112]. The complex carbohydrate characterization depends on the chemical composition of monomers, the position of branches, molecular weight, and the type of glycosidic links [66].
BTP1 polysaccharide was obtained from Botrychium ternatum (Thunb.) Sw.; family Ophiolossaceaus. This water-soluble polysaccharide with a molecular weight of 11,638 Da according to the MALDI matrix was predicted to be a good immunomodulatory agent that could elevate the medicinal and nutritious value of the traditional Chinese fern Botrychium ternatum (Thunb.) Sw. Its chemical structure was determined putatively both by MALDI-TOF and NMR to be (1→)-linked α-L-Araf at O-2, O-3, and adjacent O-2 positions lying over a backbone of linear (1→5)-Araf. In the RAW 254.7 cell assays, BTP1 proved to enhance cell viability and promote NO release [66].
The polysaccharide DCP-3 was isolated from the fern Dryopteris crassirhizoma Nakai, which is used as a traditional antiviral agent in China. DCP-3 is composed of monomer units of mannose (9.22%), galactose (36.65%), arabinose (17.07%), and xylose (34.75%) arranged in an unprecedented triple helical structure with a molecular weight of 273.2 kDa. The ferric reducing power, hydroxyl radical, and DPPH scavenging radical activity were measured and revealed the potency of DCP-3 as an antioxidant; moreover, DCP-3 stimulated RAW264.7 macrophages and enhanced NO production. These results were encouraging enough to support the plausible use of Dryopteris crassirhizoma Nakai fern in the food industry and nutraceuticals [113].
Athyrium multidentatum Ching (AMC) polysaccharides were investigated by Ching et al., who revealed six main monomers, namely, arabinose, galactose, glucose, glucuronic acid, rhamnose, mannose, and fucose, with a total molecular weight of 33,203 Da isolated from the rhizomes. The chemical characterization was performed by NMR, FT-IR, monosaccharide analysis, and molecular weight determination; furthermore, the compounds’ antiaging properties were studied in the mouse model of D-galactose-induced aging. The results showed a significant reversal of the changes induced by d-galactose, particularly an elevated Bcl-2/Bax ratio, lowered caspase-3 level, Akt phosphorylation, enhanced Akt mRNA expression levels, and heme oxygenase-1 (HO-1) together with nuclear factor-erythroid 2-related factor 2 (Nrf2) mRNA expression. These outcomes supported the hypothesis that AMC polysaccharides could be implicated in antiaging or antiwrinkle preparations by affecting the PI3K/Akt/Nrf2 and FOXO3a pathways and their downstream antioxidant factors [114].

4. Health-Promoting Effects of Ferns

4.1. Anti-Inflammatory and Antioxidant Effects

Four fern extracts from Osmunda japonica, Matteuccia species struthiopteris and orientalis, and Pteris aquilinum were studied for their antioxidant and anti-inflammatory activities in two inflammatory pathways: gene expression inhibition of IL6 and IL1β, and iNOS gene expression by LPS-induced macrophages. All fern extracts reduced the IL1β gene expression, especially the Osmunda japonica roots and Malus orientalis (Hook.) Trev. Fronds, where their IC50 values were reported to be 17.8 and 50.0 µg/mL, respectively. The extract of Malus orientalis (Hook.) Trev fronds demonstrated a reduction in the iNOS pathway at a relatively low concentration of 20 µg/mL [115]. In vitro assays conducted using the Athyrium multidentatum (Doll.) Ching (AM) aerial parts in the lipopolysaccharide (LPS)-induced inflammatory model indicated that the AM extract lowered the gene expression of the two synthetases (iNOS and COX-2), leading to the downregulation of NO and PGE2 [116] (Table S1).
Stenochlaena palustris J. Sm. was highlighted as a favorable vegetable functional food that is rich year-round in polyphenols, anthocyanins, and flavonoids; moreover, antioxidant assays such as the ferric reducing and radical scavenging activities revealed that the total polyphenolic content of S. palustris extracts was 51.69 mg/g [117]. Dryopteris erythrosora (D.C. Eaton) Kuntze was one of the fern plants whose total flavonoid content was reciprocal to its antioxidant power as compared to their rutin equivalents, even though the ABTS, DPPH, superoxide anion scavenging, and FRAP were comparable to rutin. This was possibly because of the cytotoxic action of the flavonoids gliricidin 7-O-hexoside, quercetin 7-O-rutinoside, quercetin 7-O-galactoside, apigenin 7-O-glucoside keampferol 7-O-gentiobioside, myricetin 3-O-rhamnoside, and keampferol-3-O-rutinoside. The Dryopteris erythrosora (D.C. Eaton) Kuntze flavonoids exerted an acetylcholinesterase effect leading to their anti-inflammatory and/or antioxidant properties [118]. Among the highly nutritious ferns were the three potherbs Osmunda cinnamomea Linn, Pteridium aquilinum (L.) Kuhn, and Athyrium multidentatum (Doll.). A strong antioxidant power was attributed to them, especially Athyrium multidentatum (Doll.). [119]. Diplazium maximum [D. Don] displayed richness in phenolics, with procatechuic acid, myricetin, epicatechin, and catechin constituting noteworthy quantities, which bestows Diplazium maximum with the potential to be consumed widely as an instant functional food [120].
Asparagus officinalis L. Spears was rich in phenolics, giving it a superior antioxidant potential over other vegetables. Upon studying the various antioxidant assays such as the cupric reducing antioxidant capacity (CUPRAC) and ferric reducing/antioxidant power (FRAP), the extracts of Asparagus showed different contents of bioactive compounds, as the phenolics reached up to 16.9 mg GAE/g. Asparagus was favored as an important dietary component [121] (Table S1). Similarly, the Stenoloma chusanum (L.) Ching subterranean parts total flavonoid content (TFC) and total phenolic content (TPC) were quantified to be 24.63 ± 1.34% and 9.58 ± 0.41%, respectively [122]. In a recent study, more than 30 edible ferns from 13 families grown in Europe were examined for their antioxidant activity and nutraceutical application. The genera were noncytotoxic against ovine hepatocytes and included the following: Asplenium, Athyrium, Blechnum, Davallia, Pteridium, Dryopteris, Polystichum, Marsilea, Regnellidium, Matteuccia, Osmunda, Polypodium, Adiantum, Lastrea, Phegopteris, Thelypteris, and Cystopteris [123]. Compared to rocket and spinach vegetables, most of the species were revealed as good sources of carotenoids, for example, lutein (205 µg·g−1 dry weight) and β-carotene (161 µg·g−1 dry weight). The high antioxidant effect of these ferns exceeded 0.5 g Trolox eq/gm dry weight in the ORAC assay, with an IC50 value of less than 30 µg·mL−1 in the DPPH assays [124].
Isoetes sinensis Palmer is an endangered fern with a relatively high total flavonoid content of 10.74 ± 0.25 mg/g. Compounds such as flavones; apigenin-7-glucuronide, acacetin-7-O-glcopyranoside, and homoplantageninisoetin; one prodelphinidin (procyanidins) and one nothofagin (dihydrochalcone); and four flavonols, namely, isoetin, kaempferol-3-O-glucoside, quercetin-3-O-[6″-O-(3-hydroxy-3-methyl glutaryl)-β-D-glucopyranoside], and limocitrin-Neo, were detected in the mass spectrometry-DAD at 254 nm [125] (Table S1). Traditionally, ferns such as Diplazium and Pteridium were included as pickles in the Himalayans’ diet [126].
Even though Dryopteris crassirhizoma (DC) was used in folk medicine as anti-inflammatory and antibacterial agent, in vivo studies manifested its effectiveness in the treatment of allergic rhinitis through the suppression of Th2 cytokine overproduction and mast cell infiltration and the reduction of nasal fluid Treg cytokines [127]. Polypodium vulgare L. (Polypodiaceae) showed an antioxidant effect, possibly due to the high polyphenolic content, cellular repair activity in 3T3 fibroblast cells, as well as cytoprotective potential with no appreciable cytotoxicity in physiological conditions [128]. Adiantum capillus-veneris Linn. (ACVL) revealed a protective effect against inflammatory reactions induced by carbendazim pesticide (CBZ) in Sprague–Dawley female rats. Reduced CPZ toxicity and symptom reversal were recorded after the administration of the fern due to its anti-inflammatory effect and inhibition of NF-κB-P65 synthesis as the major inflammatory marker [129]. Ferns such as Adiantum, Dryopteris, Blechnum, Pteris, and Azolla species were reported as rich sources of 3-deoxyanthocyanidins, rare anthocyanin-type compounds with unique antioxidant, anticancer, and favorable food coloration effects [130]. Pteris ensiformis Burm. extracts were proposed as a food supplement due to their bioactivities and higher antioxidant effect in the superoxide anion and hydroxyl radical assays [131].
From the Sumatrian fern Trichomanes javanicum Blume (Hymenophyllaceae), mangifern revealed an activity twice as high as that of the standard ascorbic acid with an IC50 value of 3.89 μM. Another sumatrian fern, Oleandra pistillaris (Sw.) C. Chr. (Oleandraceae), is rich in gallic acid and phenolic acids and was significantly active against Salmonella thypimurium [132]. The leaves of Adiantum capillus-veneris Linn. were extracted, and thirteen phenolics were characterized putatively by reversed phase HPLC-DAD profiling; the phenolics included caftaric acid, 5-caffeoylquinic acid, hydroxybenzoic acid, chlorogenic acid, rosmarinic acid, kaempferol glycosides, p-coumaric, and quercetin glycosides [133]. (-)-epi-Osmundalactone and angiopteroside were studied using RAW cells, and anti-inflammatory bioactivity was revealed, especially for (-)-epi-osmundalactone [54].

4.2. Anti-Adipogenic Effect

The WHO has declared obesity an epidemic disease worldwide associated with many metabolic disorder diseases. Not only is obesity linked to cardiovascular aliments and hypertension but also to chronic diseases such as osteoarthritis, cancers, inflammatory pathologies, stroke, and sleep apnea. A large number and size of adipocytes is considered the main manifestation of obesity; accordingly, research interest is evolving in this area to target factors affecting adipogenic cells and inhibitors of lipid and carbohydrate digestion or absorption [54] (Table S1). Fern functional foods are among the promising ingredients that could be consumed regularly, thus promoting human health through its high-nutrient and low-fat content. The combined type 2 diabetes and obesity syndrome, denoted as diabesity, has a high occurrence rate at present. 3T3-L1 cells were used as an in vitro model of cells to assess the lipid production rate and the anti-adipogenic effect where both (-)-epi-osmundalactone and angiopteroside showed significant results with 30% and 20% inhibition rates, respectively. As far as the obesity metabolic pattern is concerned, inflammation is regarded as a key player affecting TNF-α and various immune cells to overproduce or stimulate the proliferation of adipogenic cells; therefore, it makes sense to consider the anti-inflammatory activity for promising antiadipogenic molecules. A structure correlation was deduced that linked the biological activity of both lactones with the glycosidic linkage. The absence of the glucose linkage was hypothesized to reduce the biological effectiveness as manifested by angiopteroside, which was weaker than (-)-epi-osmundalactone. Likewise, previous studies on natural flavonoids supported the hypothesis that aglycones and glycosides do have different bioavailabilities and activities [134].
Adiantum capillus-veneris Linn. was studied for its influence on the triglyceride profile of diabesity patients and showed in vitro efficiency against α-amylase/α-glucosidase comparable to acarbose, the standard antidiabetic drug. Surprisingly, Adiantum capillus-veneris Linn. was not able to hinder glucose diffusion, but it inhibited the pancreatic lipase (PL) enzyme together with its pure components ferulic, ellagic, and chlorogenic acids [135] (Table S1). The methanolic and aqueous extracts of Tectaria coadunata were investigated for their activity against five metabolic and degenerative α-glucosidase and α-amylase enzymes and showed promising activity. The reason was based on the rich procyanidin-type A content in the fern’s extract, mainly luteolin-7-O-glucoronide and eriodictyol-7-O-glucuronide [47]. The medicinal fern Stenochlaena palustris (Burm. f.) was reported as an inhibitor of α-glucosidase and amylase enzymes, contributing to carbohydrate digestion. These activities were shown by different fractions, particularly the polar fractions [136].
The butanol, water, and methanolic extracts of the Angiopteris helferiana C. Presl ferns’ rhizome were investigated in the DPPH assay to show their significant antioxidant and α-glucosidase activities; moreover, in 3T3-L1 cells, the three fractions revealed their promising antiadipogenic potential by inhibiting adipogenic markers (PPARγ, C/EBPα) [54].

4.3. Anticancer Effect

The use of nutraceuticals either to treat or protect from carcinogenic agents has been on the rise, in particular, due to their natural herbal origin, which patients or users deliberately trust as natural. Many phytochemicals such as flavonoids, carotenoids, or stilbenes have been proven to significantly enhance apoptosis through various mechanisms; thus, they are encouraged to be used in combination therapies [137,138]. Geopyxins B and E as well as other ent-kaurene derivatives manifested a moderate cytotoxic effect against the Bel-7402 and HepG2 cell lines with IC50 values between 6.6 and 10.6 μmol/L [95]. The water extract of Camptosorus sibiricus Rupr. was found to suppress ROS by activating HO-1 and NQO-1 Nrf2-mediated reductases. Furthermore, DNA damage caused by induced mutations and oxidative stress was attenuated in BEAS.2B cells and B(a)P-induced lung adenocarcinoma cells. Tumor size, volume and growth were diminished by C. sibiricus in in-vivo experiments [101]. Nine cell lines were targeted including MG63, HepG2, MB231, U251, A549, HeLa, HOS, U2OS, and SKBR-3, and the effectiveness ranged from IC50 1.24 µg/mL for pinoresinol in Hela cells to 48.3 µg/mL for yomogin in HepG2 cells, representing the highest and lowest potencies, respectively [104].
Beyond the antitumor effect of seed plants and their essential oils [139], fern herbs were remarkable in their anticancer effect. Among the highly nutritious ferns are three potherbs, Osmunda cinnamomea Linn, Pteridium aquilinum, Filicium decipiens, and Athyrium multidentatum (Doll.). A significant anticancer power was attributed to the three ferns, especially Athyrium multidentatum (Doll.) as a biomolecule protectant and as a cellular antioxidant. On the molecular level, extracts of Athyrium multidentatum (Doll.) led to inhibition of wound-healing and mitochondrial membrane potential (MMP) reduction in HepG2 cells [119].

4.4. Neuroprotective Effect

The edible fern Diplazium esculentum [Retz.] was studied both in vitro and in vivo to examine its effects as an acetylcholine esterase inhibitor as well as an anti-amyloid B (Aβ). The Drosophila models of α-amyloid toxicity were affected positively by the ethanolic extracts of Diplazium esculentum (Retz.), which exhibited high phenolic and flavonoid contents and reduced the BACE-1 and amyloid beta 42 (Aβ42) peptides, thus ameliorating the locomotor functioning in Alzheimer disease and other dementia aliments [140] (Figure 10). The methanolic and aqueous extracts of Tectaria coadunata (Wall. ex-Hook. & Grev.) C. Chr. were investigated for their activity against the AChE and BChE enzymes and showed promising activity. The reason was based on the rich procyanidin-type A content in the fern’s extract, mainly luteolin-7-O-glucoronide and eriodictyol-7-O-glucuronide [47]. The structural features of flavonoids such as the number and position of OH groups and the oxidation of the C-ring were associated with anti-Alzheimer activity, particularly the phenyl chroman backbone. Catechin and epicatechin glucuronide metabolites presented the highest concentration in the blood after the consumption of polyphenol-rich food. It is worth stating that these metabolites are associated with better synaptic functioning and transmission.
Polyphenols were not only implicated in preventing AD causative factors but also type 2 diabetes since both diseases share the common Aβ aggregation in the brain and human islets of the pancreas, respectively. The antioxidant power of polyphenols contributed effectively to their beneficial roles here. While the amyloid fiber topological structures would be impeded by the aromaticity and the hydrophobic groups encountered in many phenolic compounds, AB fibrils are destabilized in their aggregation by the antioxidant effect [141]. Catechols possessed disaggregation and anti-aggregation properties, and their SARs referred to hydrogen bonding formation with Glu22 as well as the OH group. For instance, hispidin derivatives, caffeoylquinic acid, schizotenuin A, kukoamines A and B, lycopic acids, clovamide, and rosmarinic acid successfully inhibited amyloid protein aggregation [142]. The number of catechol groups was proportional to the anti-amyloid activity, as it oxidized to O-benzoquinone covalently bound to the β-sheet structure of the amyloid proteins.
Diplazium esculentum (Retz.) ethanolic extracts manifested promising results when tested both in vitro and in Aβ-mediated toxicity Drosophila models. A high content of phenolics and flavonoids was shown to possibly contribute to the anti-AD and antioxidant effects. The locomotor activities were improved, with a notable inhibition of the BACE-1 and amyloid beta 42 (Aβ42) peptide decline [139]. By far, Pteris sesquiterpenes showed antioxidant potential, which gives them the potential to be efficient molecules for reducing the progression of and prohibiting AD symptoms [37].

4.5. Skin-Protective Effect

Asplenium australasicum (J. Sm.) Hook, the edible bird’s nest fern, revealed effectiveness as a melanogenesis inhibitor in skin tissues based on its phytoconstituents; it is mainly rich in polyphenols, fucose-rich polysaccharides, and p-coumaric acid. The study highlighted the effect of these natural products on higher collagen production, fibroblast growth, and skin elasticity [142] (Table S1). Polypodium leucotomos fern (PLF) was one of the components of the Over-the-Counter (OTC) dietary dermatological preparations for skin, hair, and nails. As a dietary supplement prescribed by 66% of dermatologists, the preparation is used by a large population in the USA. Other components were biotin, vitamin D, zinc, and nicotinamide [143]. Despite the insufficient clinical trials, PLF was reported to have in vitro antioxidant and photoprotective effects that reduce UV-induced skin erythema. Various evidence supports the use of PLF in actinic keratosis and idiopathic photodermatoses [143]. On the other hand, flavonoids with a 3-hydroxy, 4-keto group were effective tyrosinase inhibitors, which is possibly due to the analogy with the di-hydroxyphenyl group in l-DOPA, and the Tectaria coadunata (Wall. ex-Hook. & Grev.) C. Chr. alcoholic extracts depicted a promising tyrosinase inhibitory activity that reached around 149.41 mg kojic acid equivalent (KAE)/g [47]. The methanolic and aqueous extracts of Tectaria coadunata (Wall. ex-Hook. & Grev.) C. Chr. were investigated for their activities against tyrosinases and showed promising activity based on the rich procyanidin-type A content in the fern’s extract, mainly luteolin-7-O-glucoronide and eriodictyol-7-O-glucuronide [47]. The dimeric phloroglucinols from Dryopteris crassirhizoma Nakai roots revealed their suppressing effect against melanogenesis in B16F10 murine melanoma cells with IC50 values between 181.3 and 35.7 μM. Among the important targets to block the melanogenesis pathway is the tyrosinase enzyme. As a tyrosinase inhibitor, norflavaspidic acid PB manifested potencies with an IC50 value of 5.9 μM in B16F10 cells [21], and the study recorded the vital role of the ring substituents and their functional groups in determining the degree of tyrosinase inhibition.

4.6. Bone Mineralization Effect

The Davallia formosana Hayata (DFH) water and ethanolic extracts showed effectiveness in MC3T3E1 osteoblasts. The water extract enhanced cell survival and increased the bone morphogenetic and maturation markers BMP-2, RUNX-2, ALP, and CoL-1 and, hence, cell differentiation. Moreover, late mineralization was stimulated by both extracts. Upon studying the metabolic constituents of the water extract of DFH, (-)-epicatechin-3-O-D-allipyranoside was isolated and revealed superior results in promoting mineralization (218.7%) and cell survival (118.9%) [123] (Table S1).

4.7. Immunomodulatory Effect

Pteridium aquilinum (L.) Kuhn was the source of the PAP1-A polysaccharide composed of the monomers L-fucose, D-galactose, L-rhamnose, D-xylose, D-mannose, L-arabinose, and D-glucose. Upon measuring RAW264.7 cell proliferation and NO production, PAP1-A manifested a potent immunomodulatory effect (12.5–100 μg/mL) with negligible cytotoxicity and was regarded as a promising ingredient in food supplementation [144]. The effect of (Adiantum capillus-veneris Linn.) leaf powder was investigated in a fish feeding study in which it was added to the diet and exhibited a pronounced effect on the immune system [145].

4.8. Antimicrobial and Antiviral Effects

In the quest for new antimicrobial agents to combat antimicrobial resistance, ferns have been largely unexplored. The skin lysosomal activity and superoxide dismutase activity were raised compared to the control in the groups receiving the dietary fern Adiantum capillus-veneris Linn. Additionally, a skin bactericidal effect was reported against Escherichia coli (EHEC ATCC 43895), Klebsiella, Aeromonas hydrophila, S. aureus, Pseudomonas aeruginosa, and Escherichia coli (CI) [145]. The sugar compositions in the nectaries produced by 101 fern species from the family Polypodiaceae were appraised as a valuable food source since it attracted scale insects, ants, and snails and other opportunistic animals [146] (Figure 10) (Table S1). One study delt with the essential oils from six Pyrrosia species and their in vitro antimicrobial activity with 2,4-pentadienal, phytol, and nonanal as the major components detected by GC-MS. From this, the Pyrrosia lingua extract was reported to have the highest potency against S. aureus (ATCC 25923) at 2.5 μL/mL [147].

4.9. Nutritional Importance of Ferns

New sources of human diet have never been more in demand, and edible ferns in regions such as East Asia and North America have been in use for a long period of time. Compared to leafy vegetables such as spinach and rockets, the carotenoids content estimated in ferns belonging to 13 families was about 1.2 to 4 times the former and equivalent to 4.77 mg⋅g−1 DW, with particularly high levels of lutein and β-carotene. These latter pigments are of vital importance in retinal tissue. In Dryopteridaceae species, β-carotene contributed the highest percentage to the total carotenoid content, up to 14 times that in cabbage. β-Carotene is a powerful antioxidant and provitamin A that serves to protect the skin [148].
In China, fern use as a traditional food date back to more than 3000 years, with Pteridium aquilinum var. latiusculum and Osmunda japonica recorded as the most widely eaten ferns by Chinese people together with Athyrium multidentatum (Doll.) [119]. Many parts of these edible ferns are used, for instance, the fronds, stems, rhizomes, and young leaves, where their starch content could be extracted, processed, and used for cooking. Fern cakes are popular in China from this starch, and many plant parts are stir-fried with chicken or meat [146].
In the Russian Far East diet, the ostrich fern (Matteuccia struthiopteris (L.) Tod.) and bracken fern (Pteridium aquilinum (L.) Kuhn) were the most popular with a valuable long-chain fatty acid content up to 5.5 and 0.5 mg/g dry weight for ARA and EPA, respectively. Ferns are known for their long-chain polyunsaturated fatty acid content such as arachidonic acid, eicosapentaenoic acid, sciadonic acid, juniperonic acid, and others; thus, they were postulated as a possible ingredient in formulating vegetarian diets [149]. Athyrium multidentatum (Doll.), Pteridium aquilinum (L.) Kuhn, and Osmunda cinnamomea Linn were assessed as widely ingested potherbs in China to show that their nutritive values are rich in protein, carbohydrates, minerals, and fats; additionally, they possess effective antioxidant effects and biomolecule-protective actions based on their total phenolic and flavonoid contents [119]. Just like many mushrooms and plants, Diplazium maximum (D. Don) was used as a food by Western Himalayas tribes [11]. The dried young fronds exhibit a high dietary fiber content, essential amino acids, up to 50% PUFAs, crude proteins, and phenolics. Further studies on its applicability as a food product revealed the favorable dispersibility and swelling properties, which prepare DM to be a part of instant food preparations [120].
The nutritive content of European ferns belonging to eight families, namely, Aspleniaceae, Athyriaceae, Dennstaedtiaceae, Dryopteridaceae, Onocleaceae, Osmundaceae, Polypodiaceae, and Thelypteridaceae, was investigated with respect to fatty acids, phenolics, ascorbates, and carotenoids and revealed that these plants’ young fronds were comparable in their nourishment profile to commonly used vegetables as spinach and others sold in the market with a desirable n-6/n-3 ratio between 2 and 6.4. The antioxidant effects of most of these fiddleheads in the ORAC assay were above 1 g Trolox equivalent per gram/dry weight [149].
The hazardous consequences of the use of some ferns were elaborated despite their use as a cultural edible herb. Among which, Diplazium esculentum (Koenig ex Retz.) consumed by the Rajbanshi population after cooking displayed both immunotoxic and hemolytic activities in a dose-dependent manner [150]. The Matteuccia struthiopteris (L.) Tod. plant was the cause of acute poisoning and GIT symptoms upon its raw ingestion despite its documented health benefits, yet little beyond being destroyed by boiling was known about the toxic component [150].
Beyond the risk of ptaquiloside inclusion in ferns, most of the toxicities seem minor and reversible, although we recommend further in-depth studies to confirm this observation. Polypodium leucomotos was marketed as Fernblock® with no genotoxicity, oral toxicity, or mutagenicity effects. Regarding the genus Dryopteris, several acylphloroglucinols from its rhizomes were reported as anthelminthics and caused blindness in cattle, yet rat in vivo studies revealed no death associated with these extracts up to a dose of 5000 mg·kg−1. Even the potential nephron- and hepatoxicities caused at subchronic doses were reversible [151]. Similarly, Dryopteris crassirhizoma Nakai caused no genotoxicities up to 2000 mg·kg−1 in bacterial chromosomal aberration and bone marrow tests. Macrothelypteris torresiana (Gaudich) Ching was safe under the same dose and showed no hematological or biochemical changes. It is worth mentioning that appropriate cooking and frequent non-daily consumption could lower any toxicity risk as along as toxic ptaquilosides and analogs are absent [152].

5. Fern Use in Cosmetics

Many fern extracts demonstrated promising effects as cosmetic agents, for example, the volatile components of Matteuccia struthiopteris roots, trophophyll, and petioles [153]. Ophioglossum vulgatum L. was characterized by its rich flavonoids and polyols and formulated as a facial cleanser in the form of pickering emulsion stabilized by carbon dots (CDs-Arg) [154]. Ag nanoparticles (AgNPs) of the vegetable fern (named Gosari) revealed strong in vitro antioxidant potential [155]. Ferns such as bird’s nest extracts are composed of a substantial amount of p-coumaric acid and fucose-rich mucilage, which contribute to their antiaging effect via oxidative stress reduction. Several reports documented the role of polysaccharides in retaining the moisture and elasticity of skin [156]. The genus Asplenium produced a volatile profile with a predominant shikimic, terpene, and lipoidal nature, especially Asplenium onopteris in which 4-hydroxybenzoic acid represented 8.7% with a potential antiaging action [157,158]. Microsorum grossum (Polypodiaceae) extract rich in phytoecdysteroids such as 20-hydroxyecdysone was tested for its skin-protective effect using a human fibroblast model, gene expression and modulation study, and SIPS test, which indicated its efficacy and regulation of heme oxygenase 1 (HO1) to reduce oxidative stress that contributes to skin aging [159]. Polypodium leucotomos extract was suggested to be included in daily cosmetic preparations due to its antioxidant effect based on its caffeic acid and ferulic acid, which cured ultraviolet-induced erythema. Furthermore, the extract improved cell membrane integrity and elastin expression. Its future applications might be in photoprotection or its combined administration with sunscreens [160].

6. Application of Modern Extraction Methods to Ferns

Extraction methods and their efficiency are the foundation of good natural products research. Several novel techniques have started to gain scientific attention during the last few years, and their impact might significantly advance fern plant extractions and yield such as microwave-assisted extraction and supercritical fluid extraction. Dryopteris fragrans was extracted for its volatile oils using 1-Ethyl-3-methylimidazolium acetate as a digesting agent and microwave extraction at 79 degrees C for 3.6 min with a plant mass of 0.9 g to produce the same yield reported before but in a faster and more solvent- and sample-efficient way [161]. Phloroglucinol extraction was enhanced from Dryopteris fragrans via microwave-assisted extraction combined with ionic liquid-based surfactant, and 1-octyl-3-methylimidazolium bromide presented the highest relative extraction efficiency at 50 °C, time 7 min, and irradiation power 600 W [162]. Pteris multifida essential oil was extracted using CO2 supercritical fluid extraction and steam distillation to provide a total of twenty-seven and forty-five components, respectively [163]. The Cibotium barometz phospholipid content was measured to check the quality of the extract. The content of phosphatidylcholine was 0.198% and was utilized as the benchmark level. The method involved applying supercritical fluid extraction with a temperature of 46.2 °C, pressure at 31.5 MPa, and a 0.37 L/kg modifier [164].

7. Ethnopharmacological Uses of Ferns

Ferns have been widely used ethnopharmacologically in different regions (Table 2). Dryopteris crassirhizoma was added to the pharmacopeial preparation Fufang Qingdai Wan in 40 g amounts based on its soothing and cooling effect to treat and resolve dermal macules. This formulation was prescribed for rashes, itching, erythema, psoriasis, and sores. Moreover, the rhizomes killed worms and helped cure abdominal pain caused by worm infection. The genus Microsorum, known in the Pacific Islands as “Metuapua’a”, was used in the treatment of several ailments, particularly renal diseases, as a diuretic and anti-inflammatory agent [165]. Since only a few pharmacological studies have been conducted so far on ferns, many of these bioactivities were not correlated to their chemical active ingredients except for a few such as exdysteroids, which are widely recognized as anti-inflammatory agents. Another example is the Chinese remedy called “Gusuibu” formed of a unique combination of Pseudodrynaria coronans, Drynaria fortunei, Davallia divaricata, Davallia solida, Humata griffithiana, and Davallia mariesii [165] that was given to collagen carriers to increase bone formation in skull defects with more than 90% new bone formation compared to the active control [166]. “Anapsos” was registered in Spain after clinical trials as “Regender” and “Armaya fuerte” containing the extracts of Polypodium leucotomos extracts to treat dermatitis and psoriasis [167]. Fernblock could play an important role as a nutraceutical agent as it was shown to act as an immunomodulator, photoprotectant, and antioxidant [168]. Similarly, Fernblock was prepared from the aqueous extract of Polypodium leucotomos [169].

8. Future Aspects and Long-Term Avenues

Ferns are among the most unexploited species on earth despite their richness in unusual chemical skeletons and bioactivities. The world is facing several challenging problems that require every effort to solve these global issues, which include but are not limited to the antibiotic resistance problem. The WHO announced that the discovery of natural products is a strategic approach to limit antibiotic resistance. Moreover, ferns are traditionally known for their antibacterial effect and represent a rich repository of drug leads.
This review highlighted the intense discoveries made so far in fern research during the previous six years and the outcomes. The focus on ferns’ rare molecules of chemical classes such as acylphloroglucinols, terpenes, and phenols with superior antimicrobial, antiviral, and antifungal activities paves the way for the further development of antibiotic resistance solutions. Currently, little literature data are available about ferns, which contradicts the actual chemical and biological potential of these vascular plants. Another perspective is the nutraceutical use of ferns in the diet. The various fern species available and those used by locals will be collected, extracted, and analyzed for their chemical components. Their antimicrobial as well as their cytotoxic effects will be measured to select safe and effective species that could contribute based on their in vivo and in vitro experimental results to a healthy diet for people, particularly those suffering from diabetes, cardiovascular ailments, and obesity.

9. Conclusions

In this review, more than 25 acylphloroglucinols were compiled from fern data, and their antidepressant activity was shown. More than 40 terpene compounds were described here, and the majority are reported to have an acceptable safety profile, with only the kaurene diterpenes in Pteris dispar demonstrating a moderate cytotoxicity between 6 and 10 µmol/L. Around 50 phenolics were documented with promising activity in the binding sites of many enzymes such as tyrosinases, α-amylase, α-glucosidase, and acetyl cholinesterases. The inherent antioxidant potential of polyphenolic compounds predisposes them to combat cellular oxidative stress, which forms the pathological principle of many diseases such as stress-induced liver damage, cardiovascular problems, atherosclerosis, diabesity, Alzheimer, and immune disorders.
Terpenes, followed by phenolics, exhibited the largest number of isolated active compounds. Regarding the neuroprotective effect, the Psilotium, Polypodium, and Dryopteris species possessed as their major phenolic components unique chemical moieties including catechins, procyanidins, and bioflavonoids. Likewise, sesquiterpenes from Pteris species such as pterosin B exhibited a neuroprotective effect 200 times that of ginsenoside Rg1, yet the mechanism of this activity was unrevealed until now.
While the highest antiproliferative activity was manifested by liglaurates at 0.17 μM, an antiadipogenic effect was seen for osmundolactone and angiopterosides with 30–20% inhibition rates of TG in 3T3-L1 cells. Unmatching was observed between traditional uses and experimental study outcomes; for instance, Polypodium vulgare was used to treat kidney disorders, yet it was reported as an antioxidant. Dryopteris crassirhizoma was used as an antibacterial agent and recently proved to be antiallergic.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/plants13182668/s1, Table S1. Biological activities and potencies of compounds and/or extracts of different fern species.

Author Contributions

Conceptualization, A.Y.M. and B.X.; methodology, A.Y.M.; software, A.Y.M. and J.L.; validation, A.Y.M. and J.L.; formal analysis, A.Y.M. and J.L.; investigation, B.X.; resources, B.X.; data curation, A.Y.M.; writing—original draft preparation, A.Y.M.; writing—review and editing, A.Y.M., J.L. and B.X.; visualization, B.X.; supervision, B.X.; project administration, B.X.; funding acquisition, B.X. All authors have read and agreed to the published version of the manuscript.

Funding

This study is supported by two research grants (UICR0400015-24 and UICR0400016-24) from BNU-HKBU United International College, China.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Gleńsk, M.; Dudek, M.K.; Ciach, M.; Włodarczyk, M. Isolation and structural determination of flavan-3-ol derivatives from the Polypodium vulgare L. rhizomes water extract. Nat. Prod. Res. 2021, 35, 1474–1483. [Google Scholar] [CrossRef]
  2. Zheng, X.K.; Li, Y.J.; Zhang, L.; Feng, W.S.; Zhang, X. Antihyperglycemic activity of Selaginella tamariscina Beauv. Spring J. Ethnopharmacol. 2011, 133, 531–537. [Google Scholar] [CrossRef]
  3. Morais-Braga, M.F.B.; Souza, T.M.; Santos, K.K.A.; Andrade, J.C.; Guedes, G.M.M.; Tintino, S.R.; Coutinho, H.D.M. Citotocixidade atividade antiparasitária de Lygodium venustum SW. Acta Toxicol. Argent. 2013, 21, 50–56. [Google Scholar]
  4. De Araújo, F.F.; De Paulo Farias, D.; Neri-Numa, I.A.; Pastore, G.M. Polyphenols and their applications: An approach in food chemistry and innovation potential. Food Chem. 2021, 338, 127535. [Google Scholar] [CrossRef]
  5. Cao, H.; Chai, T.T.; Wang, X.; Morais-Braga, M.F.B.; Yang, J.H.; Wong, F.C.; Wang, R.; Yao, H.; Cao, J.; Cornara, L.; et al. Phytochemicals from fern species: Potential for medicine applications. Phytochem. Rev. 2017, 16, 379–440. [Google Scholar] [CrossRef]
  6. Moussa, A.Y.; Fayez, S.; Xiao, H.; Xu, B. New insights into antimicrobial and antibiofilm effects of edible mushrooms. Food Res. Int. 2022, 162, 111982. [Google Scholar] [CrossRef]
  7. Moussa, A.Y.; Xu, B. A narrative review on inhibitory effects of edible mushrooms against malaria and tuberculosis-the world’s deadliest diseases. Food Sci. Hum. Wellness 2023, 12, 942–958. [Google Scholar] [CrossRef]
  8. Banerjee, R.D.; Sen, S.P. Antibiotic activity of pteridophytes. Econ. Bot. 1980, 34, 284–298. [Google Scholar] [CrossRef]
  9. Marchant, D.B.; Chen, G.; Cai, S.; Chen, F.; Schafran, P.; Jenkins, J.; Shu, S.; Plott, C.; Webber, J.; Lovell, J.T.; et al. Dynamic genome evolution in a model fern. Nat. Plants 2022, 8, 1038–1051. [Google Scholar] [CrossRef]
  10. Christenhusz, M.J.; Chase, M.W. Trends and concepts in fern classification. Ann. Bot. 2014, 113, 571–594. [Google Scholar] [CrossRef]
  11. Liu, Y. Food uses of ferns in China: A review. Acta Soc. Bot. Pol. 2012, 81, 263–270. [Google Scholar] [CrossRef]
  12. Soeder, R.W. Fern constituents: Including occurrence, chemotaxonomy and physiological activity. Bot. Rev. 1985, 51, 442–536. [Google Scholar] [CrossRef]
  13. Bailly, C.; Vergoten, G. Anticancer properties and mechanism of action of oblongifolin C, guttiferone K and related polyprenylated acylphloroglucinols. Nat. Prod. Bioprospecting 2021, 11, 629–641. [Google Scholar] [CrossRef]
  14. Hang, X.C.; Wei, L.I.; Oike, K.K.; Lijun, W.U.; Ikaido, T.N. Phenolic constituents from the rhizomes of Dryopteris crassirhizoma. Chem. Pharm. Bull. 2006, 54, 748–750. [Google Scholar]
  15. Socolsky, C.; Borkosky, S.; Bardón, A. Structure-molluscicidal activity relationships of acylphloroglucinols from ferns. Nat. Prod. Commun. 2011, 6, 4–8. [Google Scholar] [CrossRef]
  16. Socolsky, C.; Salamanca, E.; Gime, A.; Borkosky, S.A.; Bardo, A. Prenylated acylphloroglucinols with leishmanicidal activity from the fern Elaphoglossum lindbergii. J. Nat. Prod. 2016, 79, 98–105. [Google Scholar] [CrossRef] [PubMed]
  17. Bridi, H.; de Carvalho Meirelles, G.; Von Poser, G.L. Structural diversity and biological activities of phloroglucinol derivatives from Hypericum species. Phytochemistry 2018, 155, 203–232. [Google Scholar] [CrossRef]
  18. Arvizu-Espinosa, M.G.; Von Poser, G.L.; Henriques, A.T.; Mendoza-Ruiz, A.; Cardador-Martínez, A.; Gesto-Borroto, R.; Cardoso-Taketa, A. Bioactive dimeric acylphloroglucinols from the Mexican fern Elaphoglossum paleaceum. J. Nat. Prod. 2019, 82, 785–791. [Google Scholar] [CrossRef]
  19. Lam, H.C.; Spence, J.T.; George, J.H. Biomimetic total synthesis of hyperjapones A-E and hyperjaponols A and C. Angew. Chem. Int. Ed. Engl. 2016, 35, 10368–10371. [Google Scholar] [CrossRef]
  20. Chen, N.; Wu, Z.; Li, W.; Li, Y.; Luo, D.; Chen, L.; Zhang, X.; Zhang, Y.; Wang, G.; Li, Y. Acylphloroglucinols-based meroterpenoid enantiomers with antiviral activities from Dryopteris crassirhizoma. Ind. Crops Prod. 2020, 150, 112415. [Google Scholar] [CrossRef]
  21. Hai, P.; Rao, K.; Jiang, N.; Liu, D.; Wang, R.; Gao, Y.; Liu, X.; Deng, S.; Zhou, Y.; Chen, X.; et al. Structure elucidation, biogenesis, and bioactivities of acylphloroglucinol-derived meroterpenoid enantiomers from Dryopteris crassirhizoma. Bioorg. Chem. 2022, 119, 105567. [Google Scholar] [CrossRef]
  22. Hai, P.; He, Y.; Wang, R.; Yang, J.; Gao, Y.; Wu, X.; Chen, N.; Ye, L.; Li, R. Antimicrobial acylphloroglucinol meroterpenoids and acylphloroglucinols from Dryopteris crassirhizoma. Planta Med. 2023, 89, 295–307. [Google Scholar] [CrossRef]
  23. Wang, J.; Yan, Y.T.; Fu, S.Z.; Peng, B.; Bao, L.L.; Zhang, Y.L.; Gao, Z.P. Anti-influenza virus H5N1 activity screening on the phloroglucinols from rhizomes of Dryopteris crassirhizoma. Molecules 2017, 22, 431. [Google Scholar] [CrossRef]
  24. Pharm, V.C.; Kim, O.; Lee, J.H.; Min, B.S.; Kim, J.A. Inhibitory effects of phloroglucinols from the roots of Dryopteris crassirhizoma on melanogenesis. Phytochem. Lett. 2017, 21, 51–56. [Google Scholar] [CrossRef]
  25. Yuk, H.J.; Kim, J.Y.; Sung, Y.Y.; Kim, D.S. Phloroglucinol derivatives from Dryopteris crassirhizoma as potent xanthine oxidase inhibitors. Molecules 2021, 26, 122. [Google Scholar] [CrossRef]
  26. Chen, N.; Li, W.; Zhang, J.; Xu, W.; Wu, Z.; Li, Y.; Zhang, Y.; Zhang, Q.; Wang, G. Isolation and structural elucidation of β-tocopherol derivatives from Dryopteris crassirhizoma. Ind. Crop. Prod. 2021, 172, 114010. [Google Scholar] [CrossRef]
  27. Yim, N.H.; Lee, J.J.; Lee, B.H.; Li, W.; Ma, J.Y. Antiplatelet activity of acylphloroglucinol derivatives isolated from Dryopteris crassirhizoma. Molecules 2019, 24, 2212. [Google Scholar] [CrossRef]
  28. Hua, X.; Yang, Q.; Zhang, W.; Dong, Z.; Yu, S.; Schwarz, S.; Liu, S. Antibacterial activity and mechanism of action of aspidinol against multi-drug-resistant methicillin-resistant Staphylococcus aureus. Front. Pharmacol. 2018, 9, 619. [Google Scholar] [CrossRef]
  29. Phong, N.V.; Oanh, V.T.; Yang, S.Y.; Choi, J.S.; Min, B.S.; Kim, J.A. PTP1B inhibition studies of biological active phloroglucinols from the rhizomes of Dryopteris crassirhizoma: Kinetic properties and molecular docking simulation. Int. J. Biol. Macromol. 2021, 188, 719–728. [Google Scholar] [CrossRef]
  30. Teng, X.; Wang, Y.; Gu, J.; Shi, P.; Shen, Z.; Ye, L. Antifungal agents: Design, synthesis, antifungal activity and molecular docking of phloroglucinol derivatives. Molecules 2018, 28, 3116. [Google Scholar] [CrossRef] [PubMed]
  31. Zheng, S.Q.; Song, G.Q.; Yin, C.P.; Chen, Y.F.; Wang, S.S. A new phloroglucinol from Dryopteris fragrans and its antibacterial activity in-vitro. China J. Chin. Mater. Med. 2022, 47, 2474–2479. [Google Scholar] [CrossRef]
  32. Lan, S.; Chen, X.; Yin, C.; Xie, S.; Wang, S.; Deng, R.; Shen, Z. Antibacterial and anti-biofilm activities of disaspidin BB against Staphylococcus epidermidis. Front. Microbiol. 2023, 19, 999449. [Google Scholar] [CrossRef]
  33. Wollenweber, E.; Stevens, J.F.; Ivanic, M.; Deinzer, M.L. Acylphloroglucinols and flavonoid aglycones produced by external glands on the leaves of two dryopteris ferns and currania robertiana. Phytochemistry 1998, 48, 931–939. [Google Scholar] [CrossRef]
  34. Chen, N.; Qian, Y.; Li, W.; Zhang, Y.; Zhou, Y.; Li, G. Six new acylphloroglucinols from Dryopteris championii. Chem Biodivers. 2017, 9, e1700001. [Google Scholar] [CrossRef]
  35. Zhang, T.; Wang, L.; Duan, D.H.; Zhang, Y.H.; Huang, S.X.; Chang, Y. Cytotoxicity-guided isolation of two new phenolic derivatives from Dryopteris fragrans (L.) Schott. Molecules 2018, 23, 1652. [Google Scholar] [CrossRef]
  36. Jiang, T.; Dai, X.; Gao, T.; Wang, L.; Yang, F.; Zhang, Y.; Wang, N.; Huang, G.; Cao, J. Ancepsone A, a new cheilanthane sesterterpene from Aleuritopteris anceps. Tetrahedron Lett. 2022, 100, 153869. [Google Scholar] [CrossRef]
  37. Zhang, Y.; Li, C.B.; Xu, H.S.; Lin, B.H.; Chen, J.C.; Zhao, Y.M.; Shi, Y.S. Pterosin sesquiterpenes and lignans from Pteris laeta Wall. and their neuroprotective bioactivity. J. Mol. Struct. 2022, 1261, 132890. [Google Scholar] [CrossRef]
  38. Wang, F.; Jia, Q.W.; Yuan, Z.H.; Lv, L.Y.; Li, M.; Jiang, Z.B.; Liang, D.L.; Zhang, D.Z. An anti-inflammatory C-stiryl iridoid from Camptosorus sibiricus Rupr. Fitoterapia 2019, 134, 378–381. [Google Scholar] [CrossRef]
  39. Li, X.; Zhu, L.; Chen, J.; Shi, C.; Niu, L.; Zhang, X. C-Methylated flavanones from the rhizomes of Matteuccia intermedia and their α -glucosidase inhibitory activity. Fitoterapia 2019, 136, 104147. [Google Scholar] [CrossRef]
  40. Zhu, L.J.; Song, Y.; Shao, P.; Zhang, X.; Yao, X.S. Matteucens I-J, phenolics from the rhizomes of Matteuccia orientalis. J. Asian Nat. Prod. Res. 2018, 20, 62–66. [Google Scholar] [CrossRef]
  41. Schroder, J.; Raiber, S.; Berger, T.; Schmidt, A.; Schmidt, J.; Soares-Sello, A.M.; Bardshiri, E.; Strack, D.; Simpson, T.J.; Veit, M.; et al. Plant polyketide synthases: A chalcone synthase-type enzyme which performs a condensation reaction with methylmalonyl-CoA in the biosynthesis of C-methylated chalcones. Biochemistry 1998, 37, 8417–8425. [Google Scholar] [CrossRef] [PubMed]
  42. Li, L.; Xie, M.P.; Sun, H.; Lu, A.Q.; Zhang, B.; Zhang, D.; Wang, S.J. Bioactive phenolic acid-substituted glycoses and glycosides from rhizomes of Cibotium barometz. J. Asian Nat. Prod. Res. 2019, 21, 947–953. [Google Scholar] [CrossRef] [PubMed]
  43. Cao, H.; Chen, X. Structures required of flavonoids for inhibiting digestive enzymes. Anticancer Agents Med. Chem. 2012, 12, 929–939. [Google Scholar] [CrossRef] [PubMed]
  44. Li, X.; Niu, L.; Meng, W.; Zhu, L.; Zhang, X. Matteuinterins A–C, three new glycosides from the rhizomes of Matteuccia intermedia. J. Asian Nat. Prod. Res. 2019, 22, 225–232. [Google Scholar] [CrossRef]
  45. Deligiannidou, G.; Gougoula, V.; Bezirtzoglou, E.; Kontogiorgis, C.; Constantinides, T.K. The role of natural products in rheumatoid arthritis: Current knowledge of basic in-vitro and in-vivo research. Antioxidants 2021, 10, 599. [Google Scholar] [CrossRef]
  46. Pietrak, A.; Salachna, P. Changes in growth, ionic status, metabolites content and antioxidant activity of two ferns exposed to shade, full sunlight, and salinity. Int. J. Mol. Sci. 2023, 24, 296. [Google Scholar] [CrossRef]
  47. Shrestha, S.S.; Sut, S.; Di Marco, S.B.; Zengin, G.; Gandin, V.; De Franco, M.; Rajbhandary, S. Phytochemical Fingerprinting and In Vitro Bioassays of the Ethnomedicinal Fern Tectaria coadunata (J. Smith) C. Christensen from Central Nepal. Molecules 2019, 24, 4457. [Google Scholar] [CrossRef]
  48. Liu, Z.D.; Zhao, D.D.; Jiang, S.; Xue, B.; Zhang, Y.L.; Yan, X.F. Anticancer phenolics from Dryopteris fragrans (L.) schott. Molecules 2018, 23, 680. [Google Scholar] [CrossRef]
  49. Karl, C.; Müller, G.; Pedersen, P.A. A new catechin glycoside from Polypodium vulgare. Z. Für Naturforschung C 1982, 37, 148–151. [Google Scholar] [CrossRef]
  50. Liang, Y.H.; Ye, M.; Yang, W.Z.; Qiao, X.; Wang, Q.; Yang, H.J.; Wang, X.L.; Guo, D.A. Flavan-3-ols from the rhizomes of Drynaria fortunei. Phytochemistry 2011, 72, 1876–1882. [Google Scholar] [CrossRef] [PubMed]
  51. Fu, C.; Wang, H.; Ng, W.L.; Song, L.; Huang, D. Antioxidant activity and proanthocyanidin profile of Selliguea feei rhizomes. Molecules 2013, 18, 4282–4292. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  52. Wufuer, H.; Xu, Y.; Wu, D.; He, W.; Wang, D.; Zhu, W.; Wang, L. Liglaurates A-E, cytotoxic bis(lauric acid-12yl)lignanoates from the rhizomes of Drynaria roosii Nakaike. Phytochemistry 2022, 198, 113143. [Google Scholar] [CrossRef] [PubMed]
  53. Chen, J.; Dai, X.; Jiang, C.; Fu, Y.; Jiang, T.; Tang, L.; Wang, L.; Wang, Q.; Huang, G.; Cao, J. One new protocatechuic acid methyl ester and one enantiomeric pair of dihydroflavones isolated from Phymatopteris hastata. Phytochemistry 2021, 43, 130–134. [Google Scholar] [CrossRef]
  54. Lamichhane, R.; Pandeya, P.R.; Lee, K.H.; Kim, S.G.; Devkota, H.P.; Jung, H.J. Anti-adipogenic and anti-inflammatory activities of (−)-epi-osmundalactone and angiopteroside from Angiopteris helferiana C. Presl. Molecules 2020, 25, 1337. [Google Scholar] [CrossRef] [PubMed]
  55. Yu, Y.M.; Yang, J.S.; Peng, C.Z.; Caer, V.; Cong, P.Z.; Zou, Z.M.; Lu, Y.; Yang, S.Y.; Gu, Y.C. Lactones from Angiopteris caudatiformis. J. Nat. Prod. 2009, 72, 921–924. [Google Scholar] [CrossRef] [PubMed]
  56. Taveepanich, S.; Kamthong, N.; Sawasdipuksa, N.; Roengsumran, S. Chemical constituents and biological activity of Angiopteris evecta. J. Sci. Res. 2005, 30, 187–192. [Google Scholar]
  57. Virtbauer, J.; Krenn, L.; Kählig, H.; Hüfner, A.; Donath, O.; Marian, B. Chemical and pharmacological investigations of Metaxya rostrata. Z. Naturforsch C. J. Biosci. 2008, 63, 469–475. [Google Scholar] [CrossRef] [PubMed]
  58. Kainz, K.P.; Virtbauer, J.; Kählig, H.; Arion, V.; Donath, O.; Reznicek, G.; Huber, W.; Marian, B.; Krenn, L. Two unusual methylidene cyclopropane glucosides from Metaxya rostrata C. Presl. Helv. Chim. Acta. 2012, 95, 1531–1537. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  59. Kainz, K.P.; Krenn, L.; Erdem, Z.; Kaehlig, H.; Zehl, M.; Bursch, W.; Berger, W.; Marian, B. 2-deprenyl-rheediaxanthone B isolated from Metaxya rostrata induces active cell death in colorectal tumor cells. PLoS ONE 2013, 8, e65745. [Google Scholar] [CrossRef]
  60. Mittermair, E.; Kählig, H.; Tahir, A.; Rindler, S.; Hudec, X.; Schueff, H.; Heffeter, P.; Marian, B.; Krenn, L. Methylated xanthones from the rootlets of Metaxya rostrata display cytotoxic activity in colorectal cancer cells. Molecules 2020, 25, 4449. [Google Scholar] [CrossRef]
  61. Shen, B.; Chen, S.; Zhou, Q.; Jian, Y.; Daniyal, M.; Sheng, W.; Gong, L.; Luo, D.; Liu, B.; Xu, G.; et al. Flavonoid glycosides from the rhizomes of Pronephrium penangianum. Phytochemistry 2020, 179, 112500. [Google Scholar] [CrossRef]
  62. Veitch, N.C.; Grayer, R.J. Chalcones, Dihydrochalcones, and Aurones; Flavonoids; Taylor & Francis Group, LLC: Abingdon, UK, 2006; pp. 1003–1100. [Google Scholar]
  63. Watanabe, M.; Miyashita, T.; Prasad, H. Phenolic compounds and ecdysteroids of Diplazium esculentum Retz Sw. Athyriaceae from Japan and their chemotaxonomic significance. Biochem. Syst. Ecol. 2021, 94, 104211. [Google Scholar] [CrossRef]
  64. Ke, Y.Y.; Tsai, C.H.; Yu, H.M.; Jao, Y.C.; Fang, J.M.; Wong, C.H. Latifolicinin A from a fermented soymilk product and the structure-activity relationship of synthetic analogues as inhibitors of breast cancer cell growth. J. Agric. Food Chem. 2015, 63, 9715–9721. [Google Scholar] [CrossRef] [PubMed]
  65. Budesínský, M.; Vok’ac, K.; Harmatha, J.; Cvacka, J. Additional minor ecdysteroid components of Leuzea carthamoides. Steroids 2008, 73, 502–514. [Google Scholar] [CrossRef]
  66. Yu, X.H.; Prakash, R.R.; Sweet, M.; Shanklin, J. Coexpressing Escherichia coli cyclopropane synthase with Sterculia foetida lysophosphatidic acid acyltransferase enhances cyclopropane fatty acid accumulation. Plant Physiol. 2014, 164, 455–465. [Google Scholar] [CrossRef]
  67. Chen, J.; Li, Y.; Cao, J.; Huang, J.; Jiang, C.; Dai, X.; Huang, G. Adiantic acid, a new unsaturated fatty acid with a cyclopropane moiety from Adiantum flabellulatum L. Nat. Prod. Res. 2022, 36, 2386–2392. [Google Scholar] [CrossRef] [PubMed]
  68. Jia, Q.; Li, G.; Köllner, T.G.; Fu, J.; Chen, X.; Xiong, W.; Chen, F. Microbial-type terpene synthase genes occur widely in nonseed land plants, but not in seed plants. Proc. Natl. Acad. Sci. USA 2016, 113, 12328–12333. [Google Scholar] [CrossRef]
  69. Nakane, T.; Maeda, Y.; Ebihara, H.; Arai, Y.; Masuda, K.; Takano, A.; Ageta, H.; Shiojima, K.; Cai, S.Q.; Abdel-Halim, O.B. Fern constituents: Triterpenoids from Adiantum capillus-veneris. Chem. Pharm. Bull. 2002, 50, 1273–1275. [Google Scholar] [CrossRef]
  70. Alam, M.S.; Chopra, N.; Ali, M.; Niwa, M. Normethyl pentacyclic and lanostane-type triterpenes from Adiantum venustum. Phytochemistry 2000, 54, 215–220. [Google Scholar] [CrossRef] [PubMed]
  71. Zhang, X.; Chen, H.L.; Hong, L.; Xu, L.L.; Gong, X.W.; Zhu, D.L.; Yang, X.L. Three new hopane-type triterpenoids from the aerial part of Adiantum capillus-veneris and their antimicrobial activities. Fitoterapia 2019, 133, 146–149. [Google Scholar] [CrossRef]
  72. Kumar, R.P.; Babu, K.V.D.; Evans, D.A. Isolation, characterization and mode of action of a larvicidal compound, 22-hydroxyhopane from Adiantum latifolium Lam. against Oryctes rhinoceros. Pestic. Biochem. Phys. 2017, 153, 161–170. [Google Scholar] [CrossRef] [PubMed]
  73. Shinozaki, J.; Shibuya, M.; Ebizuka, Y.; Masuda, K. Cyclization of all-E- and 2Z-geranylfarnesols by a bacterial triterpene synthase: Insight into sesterterpene biosynthesis in Aleuritopteris ferns. Biosci. Biotechnol. Biochem. 2013, 77, 2278–2282. [Google Scholar] [CrossRef] [PubMed]
  74. Kamaya, R.; Ageta, H. Fern constituents: Cheilanthenetriol and cheilanthenediol. sesterterpenoids isolated from the leaves of Aleuritopteris Khunii. Chem. Pharm. Bull. 1990, 38, 342–346. [Google Scholar] [CrossRef]
  75. Kamaya, R.; Masuda, K.; Suzuki, K.; Ageta, H.; Hsu, H.Y. Fern constituents: Sesterterpenoids isolated from fronds of Aleuritopteris mexicana. Chem. Pharm. 1996, 44, 690–694. [Google Scholar] [CrossRef]
  76. Kamaya, R.; Masuda, K.; Ageta, H.; Chang, H.C. Fern constituents: Sesterterpenoids isolated from fronds of Aleuritopteris agetae. Chem. Pharm. Bull. 1996, 44, 695–698. [Google Scholar] [CrossRef]
  77. Challinor, V.L.; Chap, S.; Lehmann, R.P.; Bernhardt, P.V.; De Voss, J.J. Structure and absolute configuration of methyl (3R)-malonyl-(13S)-hydroxycheilanth-17-en-19-oate, a sesterterpene derivative from the roots of Aletris farinosa. J. Nat. Prod. 2013, 26, 485–488. [Google Scholar] [CrossRef] [PubMed]
  78. Tang, H.-Y.; Yin, X.; Zhang, C.C.; Jia, Q.; Gao, J.-M. Structure diversity, synthesis, and biological activity of cyathane diterpenoids in higher fungi. Curr. Med. Chem. 2015, 22, 2375–2391. [Google Scholar] [CrossRef]
  79. Jiang, C.; Chen, J.; Dai, X.; Wang, Q.; Cao, J.; Huang, G. Ent-Pimaradiene and cyathane diterpenes from Aleuritopteris albofusca. Phytochem. Lett. 2020, 40, 10–14. [Google Scholar] [CrossRef]
  80. Hu, J.; Liu, S.; Long, Y.; Tao, N.; Yang, B.; Zhang, L. Ptercresions A–C: Three new terpenes with hepatoprotective activity from Pteris cretica L. Nat. Prod. Res. 2022, 36, 6252–6258. [Google Scholar] [CrossRef] [PubMed]
  81. Xiong, Z.; Cui, Y.; Wu, J.; Shi, L.; Quan, W.; Yang, S.; Feng, Y. Luteolin-7-O-rutinoside from Pteris cretica L. var. nervosa attenuates LPS/D-gal-induced acute liver injury by inhibiting PI3K/AKT/AMPK/NF-κB signaling pathway. Naunyn Schmiedebergs Arch. Pharmacol. 2022, 395, 1283–1295. [Google Scholar] [CrossRef]
  82. Fukuoka, M.; Kuroyanagi, M.; Yoshihira, K.; Natori, S. Chemical and toxicological studies on Bracken fern, Pteridium aquilinum var. latiusculum II. Structures of pterosins, sesquiterpenes having 1-indanone skeleton. Chem. Pharm. Bull. 1978, 26, 2365–2385. [Google Scholar] [CrossRef]
  83. Ouyang, D.; Ni, X.; Xu, H.; Chen, J. Pterosins from Pteris multifida. Planta Med. 2010, 76, 1896–1900. [Google Scholar] [CrossRef]
  84. Kuroyanagi, M.; Fukuoka, M.; Yoshihira, K.; Natori, S. Circular dichroism and conformations of pterosins, 1-indanone derivatives from Bracken. Chem. Pharm. Bull. 1979, 27, 731–741. [Google Scholar] [CrossRef]
  85. Attya, M.; Nardi, M.; Tagarelli, A.; Sindona, G. A new facile synthesis of D4-pterosin B and D4-bromopterosin, deuterated analogues of ptaquiloside. Molecules 2012, 17, 5795–5802. [Google Scholar] [CrossRef]
  86. Kim, J.W.; Kim, H.P.; Sung, S.H. Cytotoxic pterosins from Pteris multifida roots against HCT116 human colon cancer cells. Bioorganic Med. Chem. Lett. 2017, 27, 3144–3147. [Google Scholar] [CrossRef]
  87. Ge, X.; Ye, G.; Li, P.; Tang, W.J.; Gao, J.L.; Zhao, W.M. Cytotoxic diterpenoids and sesquiterpenoids from Pteris multifida. J. Nat. Prod. 2008, 71, 227–231. [Google Scholar] [CrossRef]
  88. Chen, J.J.; Wang, T.C.; Yang, C.K.; Liao, H.R.; Sung, P.J.; Chen, I.S.; Cheng, M.J.; Peng, C.F.; Chen, J.F. New pterosin sesquiterpenes and antitubercular constituents from Pteris ensiformis. Chem. Biodivers. 2013, 10, 1903–1908. [Google Scholar] [CrossRef]
  89. Luo, X.K.; Li, C.J.; Luo, P.; Lin, X.; Ma, H.; Seeram, N.P.; Song, C.; Xu, J.; Gu, Q. Pterosin sesquiterpenoids from Pteris cretica as hypolipidemic agents via activating liver X receptors. J. Nat. Prod. 2016, 79, 3014–3021. [Google Scholar] [CrossRef]
  90. Hsu, F.L.; Huang, C.F.; Chen, Y.W.; Yen, Y.P.; Wu, C.T.; Uang, B.J.; Liu, S.H. Antidiabetic effects of pterosin A, a small-molecular-weight natural product, on diabetic mouse models. Diabetes 2013, 62, 628–638. [Google Scholar] [CrossRef]
  91. Chen, C.; Chiu, F.; Lin, Y.; Huang, W.; Hsieh, P. Chemical constituents’ analysis and antidiabetic activity validation of four fern species from Taiwan. Int. J. Mol. Sci. 2015, 16, 2497–2516. [Google Scholar] [CrossRef]
  92. Peng, C.; Lu, J.; Liu, J.; Huang, H.; Zhu, Y.; Shu, J. Three novel pterosin dimers form Pteris obtusiloba. Fitoterapia 2020, 146, 104713. [Google Scholar] [CrossRef]
  93. Lu, J.; Peng, C.; Cheng, S.; Liu, J.; Ma, Q.; Shu, J. Four new pterosins from Pteris cretica and their cytotoxic activities. Molecules 2019, 24, 2767. [Google Scholar] [CrossRef]
  94. Moussa, A.Y.; Sobhy, H.A.; Eldahshan, O.A.; Singab, A.N.B. Caspicaiene: A new kaurene diterpene with anti-tubercular activity from an Aspergillus endophytic isolate in Gleditsia caspia desf. Nat Prod. Res. 2021, 35, 5653–5664. [Google Scholar] [CrossRef]
  95. Wang, J.N.; Zhang, Y.; Song, L.Y.; Peng, Y.B.; Li, W.J. Chemical constituents from Pteris dispar and their anti-tumor activity in-vitro. China J. Chin. Mater. Med. 2017, 42, 4159–4164. [Google Scholar] [CrossRef] [PubMed]
  96. Zhang, M.; Lee, M.M.; Ye, W.; Wong, W.; Chan, D.; Chen, S.; Zhu, L.; Tai, W.C.; Lee, C. Total synthesis enabled systematic SAR study for development of a bioactive alkyne-tagged derivative of Neolaxiflorin L. J. Org. Chem. 2019, 84, 7007–7016. [Google Scholar] [CrossRef]
  97. Shi, Y.S.; Zhang, Y.; Hu, W.Z.; Zhang, L.H.; Chen, X.; Zhang, N.; Li, G.; Tan, L.Y. Cytotoxic diterpenoids from Pteris ensiformis. J. Asian Nat. Prod. Res. 2017, 19, 188–193. [Google Scholar] [CrossRef]
  98. Wang, F.; Jiang, Z.B.; Wu, X.L.; Liang, D.L.; Zhang, N.; Li, M.; Zhang, D.Z. Structural determination and in-vitro tumor cytotoxicity evaluation of five new cycloartane glycosides from Asplenium ruprechtii Sa. Kurata. Bioorg. Chem. 2020, 102, 104085. [Google Scholar] [CrossRef]
  99. Jiang, Z.B.; Liu, X.X.; Chen, J.Z.; Guo, H.H.; Hu, Y.Q.; Guo, X.; Ma, X.L.; Wang, F.; Zhang, D.Z.; Wu, X.L. Structural determination of a new cycloartane glycoside from Asplenium ruprechtii. Chem. Biodivers. 2020, 17, e2000500. [Google Scholar] [CrossRef] [PubMed]
  100. Jiang, Z.B.; Chen, J.Z.; Guo, H.H.; Hu, Y.Q.; Guo, X.; Ma, X.L.; Yang, J.L.; Wang, F.; Zhang, D.Z.; Wu, X.L. A new 9,19-cycloartane glycoside from Asplenium ruprechtii. China J. Chin. Mater. Med. 2021, 46, 1155–1159. [Google Scholar] [CrossRef] [PubMed]
  101. He, S.; Ou, R.; Wang, W.; Ji, L.; Gao, H.; Zhu, Y.; Liu, X.; Zheng, H.; Liu, Z.; Wu, P.; et al. Camptosorus sibiricus rupr aqueous extract prevents lung tumorigenesis via dual effects against ROS and DNA damage. J. Ethnopharmacol. 2018, 220, 44–56. [Google Scholar] [CrossRef] [PubMed]
  102. Li, R.; Morris-natschke, S.L.; Lee, K. Clerodane diterpenes: Sources, structures, and biological activities. Nat. Prod. Rep. 2017, 33, 1166–1226. [Google Scholar] [CrossRef]
  103. Bei, B.; Yu, G.; Ou, F.; Feng, Q.; Zhi, Z.; Zhou, P.; Deng, Z.T.; Li, M.; Zhao, Q.S. Clerodane- type diterpene glycosides from Dicranopteris pedata. Nat. Prod. Bioprospecting. 2021, 11, 557–564. [Google Scholar] [CrossRef]
  104. Zhong, Z.C.; Zhao, D.D.; Liu, Z.D.; Jiang, S.; Zhang, Y.L. A new human cancer cell proliferation inhibition sesquiterpene, dryofraterpene A, from medicinal plant Dryopteris fragrans (L.) Schott. Molecules 2017, 22, 180. [Google Scholar] [CrossRef]
  105. Binh, P.T.; Descoutures, D.; Dang, N.H.; Nguyen, N.P.; Dat, N.T. A new cytotoxic gymnomitrane Sesquiterpene from Ganoderma lucidum fruiting bodies. Nat. Prod. Commun. 2015, 10, 1911–1912. [Google Scholar] [CrossRef] [PubMed]
  106. Chen, Z.M.; Chen, H.P.; Wang, F.; Li, Z.H.; Feng, T.; Liu, J.K. New triquinane and gymnomitrane sesquiterpenes from fermentation of the basidiomycete Antrodiella albocinnamomea. Fitoterapia 2015, 102, 61–66. [Google Scholar] [CrossRef] [PubMed]
  107. Wu, C.L.; Kao, T.L. A new gymnomitrane-type sesquiterpenoid from the liverwort Cylindrocolea recurvifolia. J. Asian Nat. Prod. Res. 2002, 4, 281–285. [Google Scholar] [CrossRef]
  108. Nagashima, F.; Kondoh, M.; Uematsu, T.; Nishiyama, A.; Saito, S.; Sato, M.; Asakawa, Y. Cytotoxic and apoptosis-inducing ent-kaurane-type diterpenoids from the Japanese liverwort Jungermannia truncata NEES. Chem. Pharm. Bull. 2002, 50, 808–813. [Google Scholar] [CrossRef] [PubMed]
  109. Zhao, X.; Li, J.; Liu, Y.; Wu, D.; Cai, P.; Pan, Y. Structural characterization and immunomodulatory activity of a water-soluble polysaccharide isolated from Botrychium ternatum. Carbohydr. Polym. 2017, 171, 136–142. [Google Scholar] [CrossRef]
  110. Chien, R.C.; Yen, M.T.; Tseng, Y.H.; Mau, J.L. Chemical characteristics and anti-proliferation activities of Ganoderma tsugae polysaccharides. Carbohydr. Polym. 2015, 128, 90–98. [Google Scholar] [CrossRef]
  111. Yu, Q.; Nie, S.P.; Wang, J.Q.; Huang, D.F.; Li, W.J.; Xie, M.Y. Signaling pathway involved in the immunomodulatory effect of Ganoderma atrum polysaccharide in spleen lymphocytes. J. Agric. Food Chem. 2015, 6, 2734–2740. [Google Scholar] [CrossRef]
  112. Ganesan, K.; Xu, B. Anti-diabetic effects and mechanisms of dietary polysaccharides. Molecules 2019, 24, 2556. [Google Scholar] [CrossRef]
  113. Zhao, Y.; Hu, W.; Zhan, H.; Ding, C.; Huang, Y.; Liao, J.; Zhang, Z.; Yuan, S.; Chen, Y.; Yuan, M. Antioxidant and immunomodulatory activities of polysaccharides from the rhizome of Dryopteris crassirhizoma Nakai. Int. J. Biol. Macromol. 2019, 130, 238–244. [Google Scholar] [CrossRef]
  114. Jing, L.; Jiang, J.; Liu, D.; Sheng, J.; Zhang, W.; Li, Z.; Wei, L. Structural characterization and antioxidant activityof polysaccharides from Athyrium multidentatum (Doll.) Ching in D-galactose-induced aging mice via PI3K/AKT pathway. Molecules 2019, 24, 3364. [Google Scholar] [CrossRef]
  115. Dion, C.; Haug, C.; Guan, H.; Ripoll, C.; Spiteller, P.; Coussaert, A.; Boulet, E.; Schmidt, D.; Wei, J.; Zhou, Y.L.K. Evaluation of the anti-inflammatory and antioxidative potential of four fern species from China intended for use as food supplements. Nat. Prod. Commun. 2015, 10, 597–603. [Google Scholar] [CrossRef]
  116. Han, X.Z.; Ma, R.; Chen, Q.; Jin, X.; Jin, Y.Z.; An, R.B.; Jiang, J. Anti-inflammatory action of Athyrium multidentatum extract suppresses the LPS-induced TLR4 signaling pathway. J. Ethnopharmacol. 2018, 217, 220–227. [Google Scholar] [CrossRef]
  117. Chai, T.T.; Panirchellvum, E.; Ong, H.C.; Wong, F.C. Phenolic contents and antioxidant properties of Stenochlaena palustris, an edible medicinal fern. Bot. Stud. 2012, 53, 439–446. [Google Scholar]
  118. Cao, J.; Xia, X.; Chen, X.; Xiao, J.; Wang, Q. Characterization of flavonoids from Dryopteris erythrosora and evaluation of their antioxidant, anticancer and acetylcholinesterase inhibition activities. Food Chem. Toxicol. 2013, 51, 242–250. [Google Scholar] [CrossRef]
  119. Qi, G.; Yang, L.; Xiao, C.; Shi, J.; Mi, Y.; Liu, X. Nutrient values and bioactivities of the extracts from three fern species in China: A comparative assessment. Food Funct. 2015, 6, 2918–2929. [Google Scholar] [CrossRef]
  120. Sareen, B.; Bhattacharya, A.; Srivatsan, V. Nutritional characterization and chemical composition of Diplazium maximum D. Don C. Chr. J. Food Sci. Technol. 2021, 58, 844–854. [Google Scholar] [CrossRef]
  121. Lee, B.; Choi, S.; Kim, H.; Jung, S.; Hwang, S.; Pyo, M.; Nah, S. 2016. Differential effects of quercetin and quercetin glycosides on human α 7 nicotinic acetylcholine receptor-mediated ion currents. Biomol. Ther. 2016, 24, 410–417. [Google Scholar] [CrossRef]
  122. Lee, J.W.; Lee, J.H.; Yu, I.H.; Gorinstein, S.; Bae, J.H.; Ku, Y.G. Bioactive compounds, antioxidant and binding activities and spear yield of Asparagus officinalis L. Plant Foods Hum. Nutr. 2014, 69, 175–181. [Google Scholar] [CrossRef]
  123. Wu, C.F.; Lin, Y.S.; Lee, S.C.; Chen, C.Y.; Wu, M.C.; Lin, J.S. Effects of Davallia formosana Hayata water and alcohol extracts on osteoblastic MC3T3-E1 cells. Phytother. Res. 2017, 31, 1349–1356. [Google Scholar] [CrossRef]
  124. Langhansova, L.; Pumprova, K.; Haisel, D.; Ekrt, L.; Pavicic, A.; Zajíčková, M.; Dvorakova, M. European ferns as rich sources of antioxidants in the human diet. Food Chem. 2021, 356, 129637. [Google Scholar] [CrossRef]
  125. Wang, X.; Ding, G.; Liu, B.; Wang, Q. Flavonoids and antioxidant activity of rare and endangered fern: Isoetes sinensis. PLoS ONE 2020, 15, e232185. [Google Scholar] [CrossRef]
  126. Chakraborty, R.; Roy, S. Exploration of the diversity and associated health benefits of traditional pickles from the Himalayan and adjacent hilly regions of Indian subcontinent. J. Food Sci. Technol. 2018, 55, 1599–1613. [Google Scholar] [CrossRef]
  127. Piao, C.H.; Kim, T.G.; Bui, T.T.; Song, C.H.; Shin, D.U.; Eom, J.E.; Chai, O.H. Ethanol extract of Dryopteris crassirhizoma alleviates allergic inflammation via inhibition of Th2 response and mast cell activation in a murine model of allergic rhinitis. J. Ethnopharmacol. 2019, 232, 21–29. [Google Scholar] [CrossRef]
  128. Farràs, A.; Mitjans, M.; Maggi, F.; Caprioli, G.; Vinardell, M.P.; López, V. Polypodium vulgare L. (Polypodiaceae) as a source of bioactive compounds: Polyphenolic profile, cytotoxicity and cytoprotective properties in different cell lines. Front. Pharmacol. 2021, 12, 727528. [Google Scholar] [CrossRef]
  129. Madboli, A.E.N.A.; Seif, M.M. Adiantum capillus-veneris Linn protects female reproductive system against carbendazim toxicity in rats: Immunohistochemical, histopathological, and pathophysiological studies. Environ. Sci. Pollut. Res. Int. 2021, 28, 19768–19782. [Google Scholar] [CrossRef]
  130. Xiong, Y.; Zhang, P.; Warner, R.D.; Fang, Z. 3-Deoxyanthocyanidin colorant: Nature, health, synthesis, and food applications. Compr. Rev. Food Sci. 2019, 18, 1533–1549. [Google Scholar] [CrossRef]
  131. Hou, M.; Hu, W.; Xiu, Z.; Shi, Y.; Hao, K.; Cao, D.; Yin, H. Efficient enrichment of total flavonoids from Pteris ensiformis Burm. extracts by macroporous adsorption resins and in-vitro evaluation of antioxidant and antiproliferative activities. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2020, 1138, 121960. [Google Scholar] [CrossRef]
  132. Putra, D.P.; Arbain, D. Antioxidant and antibacterial constituents from two Sumatran. Nat. Prod. Commun. 2017, 12, 1263–1264. [Google Scholar] [CrossRef]
  133. Zeb, A.; Ullah, F. Reversed phase HPLC-DAD profiling of carotenoids, chlorophylls and phenolic compounds in Adiantum capillus-veneris leaves. Front. Chem. 2017, 5, 29. [Google Scholar] [CrossRef]
  134. Zhang, H.; Hassan, Y.I.; Liu, R.; Mats, L.; Yang, C.; Liu, C.; Tsao, R. Molecular mechanisms underlying the absorption of aglycone and glycosidic flavonoids in a Caco-2 BBe1 cell model. ACS Omega 2020, 5, 10782–10793. [Google Scholar] [CrossRef]
  135. Kasabri, V.; Al-Hallaq, E.K.; Bustanji, Y.K.; Abdul-Razzak, K.K.; Abaza, I.F.; Afifi, F.U. Antiobesity and antihyperglycaemic effects of Adiantum capillus-veneris extracts: In vitro and in vivo evaluations. Pharm. Biol. 2016, 55, 164–172. [Google Scholar] [CrossRef]
  136. Chai, T.T.; Kwek, M.T.; Ong, H.C.; Wong, F.C. Water fraction of edible medicinal fern Stenochlaena palustris is a potent α-glucosidase inhibitor with concurrent antioxidant activity. Food Chem. 2015, 186, 26–31. [Google Scholar] [CrossRef]
  137. Torky, Z.A.; Moussa, A.Y.; Abdelghffar, E.A.; Abdel-Hameed, U.K.; Eldahshan, O.A. Chemical profiling, antiviral and antiproliferative activities of the essential oil of Phlomis aurea Decne grown in Egypt. Food Funct. 2021, 10, 4630–4643. [Google Scholar] [CrossRef]
  138. Fayez, S.; Ayoub, I.M.; Mostafa, N.M.; Moussa, A.Y.; Gamal ElDin, M.I.; El-Shazly, M. Nutraceuticals in cancer therapy, challenges, and opportunities. In Handbook of Oxidative Stress in Cancer: Therapeutic Aspects; Springer-Nature: Berlin/Heidelberg, Germany, 2022. [Google Scholar] [CrossRef]
  139. Kunkeaw, T.; Suttisansanee, U.; Trachootham, D.; Karinchai, J.; Chantong, B.; Potikanond, S.; Temviriyanukul, P. Diplazium esculentum (Retz.) Sw. reduces BACE-1 activities and amyloid peptides accumulation in Drosophila models of Alzheimer’s disease. Sci. Rep. 2021, 11, 23796. [Google Scholar] [CrossRef]
  140. Miyamae, Y.; Kurisu, M.; Murakami, K.; Han, J.; Isoda, H.; Irie, K.; Shigemori, H. Protective effects of caffeoylquinic acids on the aggregation and neurotoxicity of the 42-residue amyloid β-protein. Bioorg. Med. Chem. 2012, 20, 5844–5849. [Google Scholar] [CrossRef]
  141. Tanaka, T.; Betkekar, V.V.; Ohmori, K.; Suzuki, K.; Shigemori, H. Evaluation of amyloid polypeptide aggregation inhibition and disaggregation activity of a-type procyanidins. Pharmaceuticals 2021, 14, 1118. [Google Scholar] [CrossRef]
  142. Zeng, W.W.; Lai, L.S. Anti-melanization effects and inhibitory kinetics of tyrosinase of bird’s nest fern Asplenium australasicum frond extracts on melanoma and human skin. J. Biosci. Bioeng. 2019, 127, 738–743. [Google Scholar] [CrossRef]
  143. Thompson, K.G.; Kim, N. Dietary supplements in dermatology: A review of the evidence for zinc, biotin, vitamin D, nicotinamide, and Polypodium. J. Am. Acad. Dermatol. 2021, 84, 1042–1050. [Google Scholar] [CrossRef]
  144. Song, G.; Wang, K.; Zhang, H.; Sun, H.; Wu, B.; Ju, X. Structural characterization and immunomodulatory activity of a novel polysaccharide from Pteridium aquilinum. Int. J. Biol. Macromol. 2017, 102, 599–604. [Google Scholar] [CrossRef]
  145. Hoseinifar, S.H.; Jahazi, M.A.; Mohseni, R.; Raeisi, M.; Bayani, M.; Mazandarani, M.; Torfi Mozanzadeh, M. Effects of dietary fern Adiantum capillus-veneris leaves powder on serum and mucus antioxidant defence, immunological responses, antimicrobial activity and growth performance of common carp Cyprinus carpio juveniles. Fish Shellfish Immunol. 2020, 106, 959–966. [Google Scholar] [CrossRef]
  146. Mehltreter, K.; Tenhaken, R.; Jansen, S. Nectaries in ferns: Their taxonomic distribution, structure, function, and sugar composition. Am. J. Bot. 2022, 109, 46–57. [Google Scholar] [CrossRef]
  147. Fan, Y.; Feng, H.; Liu, L.; Zhang, Y.; Xin, X.; Gao, D. Chemical components and antibacterial activity of the essential oil of six Pyrrosia species. Chem. Biodivers. 2020, 17, e2000526. [Google Scholar] [CrossRef]
  148. Dvorakova, M.; Pumprova, K.; Antonínová, Z.; Rezek, J.; Haisel, D.; Ekrt, L.; Vanek, T.; Langhansova, L. Nutritional and antioxidant potential of fiddleheads from European ferns. Foods 2017, 10, 460. [Google Scholar] [CrossRef]
  149. Nekrasov, E.V.; Svetashev, V.I. Edible far Eastern ferns as a dietary source of long-chain polyunsaturated fatty acids. Foods 2021, 10, 1220. [Google Scholar] [CrossRef]
  150. Roy, S.; Tamang, S.; Dey, P.; Chaudhuri, T.K. Assessment of the immunosuppressive and hemolytic activities of an edible fern, Diplazium esculentum. Immunopharm. Immunot. 2013, 35, 365–372. [Google Scholar] [CrossRef]
  151. Dhir, S.B. Fiddlehead fern poisoning: A case report. Wilderness Environ. Med. 2020, 31, 226–229. [Google Scholar] [CrossRef]
  152. Erhirhie, E.O.; Ilodigwe, E.E. Sub-chronic toxicity evaluation of Dryopteris filix-mas L. schott. leaf extract in albino rats. Braz. J. Pharm. Sci. 2016, 55, e18107. [Google Scholar] [CrossRef]
  153. Wang, X.; Guo, J.; Zang, S.; Liu, B.; Wu, Y. Comparison of Flavonoid Content, Antioxidant Potential, Acetylcholinesterase Inhibition Activity and Volatile Components Based on HS-SPME-GC-MS of Different Parts from Matteuccia struthiopteris (L.) Todaro. Molecules 2024, 4, 1142. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  154. Xu, G.; Chen, S.; Shi, Q.; Wang, H.; Wu, L.; Pan, P.; Ying, H.; Xie, H. Properties of Ophioglossum vulgatum L. extract Pickering emulsion stabilized by carbon dots and its potential use in cosmetics. RSC Adv. 2024, 2, 390–396. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  155. Das, G.; Shin, H.S.; Patra, J.K. Comparative Bio-Potential Effects of Fresh and Boiled Mountain Vegetable (Fern) Extract Mediated Silver Nanoparticles. Plants 2022, 11, 3575. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  156. Zeng, W.W.; Lai, L.S. Multiple-physiological benefits of bird’s nest fern (Asplenium australasicum) frond extract for dermatological applications. Nat. Prod. Res. 2019, 33, 736–741. [Google Scholar] [CrossRef] [PubMed]
  157. Froissard, D.; Rapior, S.; Bessière, J.M.; Buatois, B.; Fruchier, A.; Sol, V.; Fons, F. Asplenioideae Species as a Reservoir of Volatile Organic Compounds with Potential Therapeutic Properties. Nat. Prod. Commun. 2015, 10, 1079–1083. [Google Scholar] [CrossRef] [PubMed]
  158. Kim, D.K.; Jeon, H.; Cha, D.S. 4-Hydroxybenzoic acid-mediated lifespan extension in Caenorhabditis elegans. J. Func. Foods. 2014, 7, 630–640. [Google Scholar] [CrossRef]
  159. Ho, R.; Teai, T.; Meybeck, A.; Raharivelomanana, P. UV-protective effects of phytoecdysteroids from Microsorum grossum extracts on human dermal fibroblasts. Nat. Prod. Commun. 2015, 10, 33–36. [Google Scholar] [CrossRef] [PubMed]
  160. Bhatia, N. Polypodium leucotomos: A potential new photoprotective agent. Am. J. Clin. Dermatol. 2015, 16, 73–79. [Google Scholar] [CrossRef] [PubMed]
  161. Jiao, J.; Gai, Q.Y.; Wang, W.; Luo, M.; Zhao, C.J.; Fu, Y.J.; Ma, W. Ionic-liquid-assisted microwave distillation coupled with headspace single-drop microextraction followed by GC-MS for the rapid analysis of essential oil in Dryopteris fragrans. J. Sep. Sci. 2013, 36, 3799–3806. [Google Scholar] [CrossRef] [PubMed]
  162. Li, X.J.; Yu, H.M.; Gao, C.; Zu, Y.G.; Wang, W.; Luo, M.; Gu, C.B.; Zhao, C.J.; Fu, Y.J. Application of ionic liquid-based surfactants in the microwave-assisted extraction for the determination of four main phloroglucinols from Dryopteris fragrans. J. Sep. Sci. 2012, 35, 3600–3608. [Google Scholar] [CrossRef] [PubMed]
  163. Chen, F.; Chen, H.; Li, S.J.; Lin, J.M.; Wang, Z.Y. Comparison of the chemical components of essential oil extracted by supercritical CO2 fluid and steam distillation from Pteris multifida. J. Chin. Med. Mater. 2013, 36, 1270–1274. (In Chinese) [Google Scholar] [PubMed]
  164. Xu, Z.; Chen, Z.; Chen, Z.; Ge, F.; Zhang, K. Determination of phospholipid in Cibotium barometz by supercritical fluid extraction and RP-HPLC. J. Chin. Med. Mater. 2001, 24, 174–175. (In Chinese) [Google Scholar] [PubMed]
  165. Chang, H.-C.; Huang, G.-J.; Agrawal, D.C.; Kuo, C.-L.; Wu, C.-R.; Tsay, H.-S. Antioxidant activities and polyphenol contents of six folk medicinal ferns used as Gusuibu. Bot. Stud. 2007, 48, 397–406. [Google Scholar]
  166. Wong, R.W.K.; Rabie, A.B.M. Effect of Gusuibu Graft on Bone Formation. J. Oral Maxillofac. Surg. 2006, 64, 770–777. [Google Scholar] [CrossRef]
  167. Sempere-Ortells, J.M.; Campos, A.; Velasco, I.; Marco, F.; Ramirez-Bosca, A.; Diaz, J.; Pardo, J. Anapsos (Polypodium leucotomos) modulates lymphoid cells and the expression of adhesion molecules. Pharm. Res. 2002, 46, 185–190. [Google Scholar] [CrossRef]
  168. Gombau, L.; Garcia, F.; Lahoz, A.; Fabre, M.; Roda-Navarro, P.; Majano, P.; Alonso-Lebrero, J.L.; Pivel, J.P.; Castell, J.V.; Gomez-Lechon, M.J. Polypodium leucotomos extract: Antioxidant activity and disposition. Toxicol. In Vitro 2006, 20, 464–471. [Google Scholar] [CrossRef]
  169. Garcia, F.; Pivel, J.P.; Guerrero, A.; Brieva, A.; Martinez-Alcazar, M.P.; Caamano-Somoza, M. Phenolic components and antioxidant activity of fernblock, an aqueous extract of the aerial parts of the fern Polypodium leucotomos. Methods Find. Exp. Clin. Pharmacol. 2006, 28, 157–160. [Google Scholar] [CrossRef]
Figure 1. Polyketide-terpene compounds 1-67 isolated from ferns (2017–2023). The two chiral centers have the same configuration, denoted as R*. The two chiral centers have opposite configurations, denoted as S*.
Figure 1. Polyketide-terpene compounds 1-67 isolated from ferns (2017–2023). The two chiral centers have the same configuration, denoted as R*. The two chiral centers have opposite configurations, denoted as S*.
Plants 13 02668 g001
Figure 2. Polyketide compounds 68-116 isolated from ferns (2017–2023).
Figure 2. Polyketide compounds 68-116 isolated from ferns (2017–2023).
Plants 13 02668 g002
Figure 3. Polyketide compounds 117-159 isolated from ferns (2017–2023).
Figure 3. Polyketide compounds 117-159 isolated from ferns (2017–2023).
Plants 13 02668 g003
Figure 4. Polyketide compounds 160-207 isolated from ferns (2017–2023).
Figure 4. Polyketide compounds 160-207 isolated from ferns (2017–2023).
Plants 13 02668 g004
Figure 5. Polyketide compounds 208-234 isolated from ferns (2017–2023).
Figure 5. Polyketide compounds 208-234 isolated from ferns (2017–2023).
Plants 13 02668 g005
Figure 6. Fatty acid and steroidal compounds 236-244 isolated from ferns (2017–2023).
Figure 6. Fatty acid and steroidal compounds 236-244 isolated from ferns (2017–2023).
Plants 13 02668 g006
Figure 7. Terpene compounds 245-279 isolated from ferns (2017–2023).
Figure 7. Terpene compounds 245-279 isolated from ferns (2017–2023).
Plants 13 02668 g007
Figure 8. Terpene compounds 280-320 isolated from ferns (2017–2023).
Figure 8. Terpene compounds 280-320 isolated from ferns (2017–2023).
Plants 13 02668 g008
Figure 9. Terpene compounds isolated from ferns (2017–2023).
Figure 9. Terpene compounds isolated from ferns (2017–2023).
Plants 13 02668 g009
Figure 10. Illustration of ferns’ chemical diversity and biological effects.
Figure 10. Illustration of ferns’ chemical diversity and biological effects.
Plants 13 02668 g010
Table 2. The ethnopharmacological uses of ferns in different regions.
Table 2. The ethnopharmacological uses of ferns in different regions.
FernsUsageCountry/RegionSymptoms TreatedReference
MicrosorumTraditional Pacific Island medicineThe PacificTreats various ailments, particularly renal diseases, as a diuretic and anti-inflammatory agent[165]
Dryopteris crassirhizomaPharmacopeial preparation component
(Fufang Qingdai Wan)
ChinaTreats dermal macules, rashes, itching, erythema, psoriasis, and sores[165]
Pseudodrynaria coronans, Drynaria fortunei, Davallia divaricata, Davallia solida, Humata griffithiana, Davallia mariesiiChinese formula (Gusuibu)ChinaPromotes bone formation in skull defects[165,166]
Polypodium leucotomosClinical trial drug (Anapsos)SpainTreats dermatitis and psoriasis[167]
Polypodium leucotomosNutraceutical preparation (Fernblock)SpainUsed as an immunomodulator, photoprotectant, and antioxidant[168,169]
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

Moussa, A.Y.; Luo, J.; Xu, B. Insights into Chemical Diversity and Potential Health-Promoting Effects of Ferns. Plants 2024, 13, 2668. https://doi.org/10.3390/plants13182668

AMA Style

Moussa AY, Luo J, Xu B. Insights into Chemical Diversity and Potential Health-Promoting Effects of Ferns. Plants. 2024; 13(18):2668. https://doi.org/10.3390/plants13182668

Chicago/Turabian Style

Moussa, Ashaimaa Y., Jinhai Luo, and Baojun Xu. 2024. "Insights into Chemical Diversity and Potential Health-Promoting Effects of Ferns" Plants 13, no. 18: 2668. https://doi.org/10.3390/plants13182668

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