[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (226)

Search Parameters:
Keywords = Euphorbia

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 4953 KiB  
Article
ECPUB5 Polyubiquitin Gene in Euphorbia characias: Molecular Characterization and Seasonal Expression Analysis
by Faustina Barbara Cannea, Daniela Diana, Rossano Rossino and Alessandra Padiglia
Genes 2024, 15(7), 957; https://doi.org/10.3390/genes15070957 - 21 Jul 2024
Viewed by 528
Abstract
The spurge Euphorbia characias is known for its latex, which is rich in antioxidant enzymes and anti-phytopathogen molecules. In this study, we identified a novel polyubiquitin protein in the latex and leaves, leading to the first molecular characterization of its coding gene and [...] Read more.
The spurge Euphorbia characias is known for its latex, which is rich in antioxidant enzymes and anti-phytopathogen molecules. In this study, we identified a novel polyubiquitin protein in the latex and leaves, leading to the first molecular characterization of its coding gene and expressed protein in E. characias. Using consensus-degenerate hybrid oligonucleotide primers (CODEHOP) and rapid amplification of cDNA ends (5′/3′-RACE), we reconstructed the entire open reading frame (ORF) and noncoding regions. Our analysis revealed that the polyubiquitin gene encodes five tandemly repeated sequences, each coding for a ubiquitin monomer with amino acid variations in four of the five repeats. In silico studies have suggested functional differences among monomers. Gene expression peaked during the summer, correlating with high temperatures and suggesting a role in heat stress response. Western blotting confirmed the presence of polyubiquitin in the latex and leaf tissues, indicating active ubiquitination processes. These findings enhance our understanding of polyubiquitin’s regulatory mechanisms and functions in E. characias, highlighting its unique structural and functional features. Full article
(This article belongs to the Special Issue Abiotic Stress in Land Plants: Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>This satellite view was obtained through Google Earth of the Mediterranean Sea, where Sardinia occupies a central position. The location of Dolianova is indicated with a red point.</p>
Full article ">Figure 2
<p>Nucleotide and amino acid sequence of <span class="html-italic">E. characias</span> polyubiquitin (ECPUB5). The ORF (1143 bp) consists of 5 tandem repeats, represented in parentheses. The 3′-UTR (48 bp) and 5′-UTR (191 bp) are indicated by a solid line above the block of nucleotides. Arrows (↓) indicate cleavage sites caused by deubiquitinase enzymes. The 3′-UTR includes a putative polyadenylation signal “aataaa” (in bold underlined), located 25 bp upstream of the poly(A) tail. The position and orientation of the specific sense (SP4, SP6) and antisense (SP5, SP7) primers relative to the gene are shown with arrows above or below the bold sequence. The numbering refers to the nucleotide coding sequence.</p>
Full article ">Figure 3
<p>Amino acid alignment of each repetition (ECUBI 1-5) of ECPUB5. The conserved amino acid residues forming the Ile-44 patch core (Leu 8, Ile 44, His 68, Val 70) are highlighted in gray. Lysine residues important for the formation of polyubiquitin chains are indicated with arrows (↓). Non-identical amino acids are shown in boxes. Asterisks (*) under the aligned sequences indicate fully conserved residues across all repetitions.</p>
Full article ">Figure 4
<p>Comparison of the secondary structure predictions of ubiquitin monomers derived from polyubiquitin genes with five tandem repeats in <span class="html-italic">Euphorbiaceae</span> species. Panel (<b>A</b>) shows the common secondary structure (S1) shared by <span class="html-italic">Euphorbia peplus</span>, <span class="html-italic">HHevea brasiliensis</span>, <span class="html-italic">Morchella esculenta</span>, <span class="html-italic">Ricinus communis</span>, and monomers 2, 3, and 4 of <span class="html-italic">E. characias</span>. Panel (<b>B</b>) shows the unique secondary structures (S2 and S3) of monomers 1 and 5 of <span class="html-italic">E. characias</span>, respectively. The letter “S” stands for the secondary structure. The differences are highlighted in boxes. The asterisks (*) above the ECUBI1 and ECUBI5 sequences indicate amino acid residues that are not present in the other ubiquitins. These differences led to the in silico prediction of the S2 and S3 structures. The S2 region of ECUBI1 and the S3 region of ECUBI5, which have different secondary structures compared to S1, are shown in boxes. The letters used to identify the particular types of secondary structure are C (coil), E (extended strand), H (helix), and T (turn).</p>
Full article ">Figure 5
<p>Northern blot and qRT-PCR analyses of the expression of mRNA encoding ECPUB5 over the course of the year (bimonthly analysis from February to December). (<b>A</b>) Northern blot analyses performed on RNA extracted from latex and (<b>B</b>) on RNA extracted from leaves. (<b>C</b>) ECPUB5 expression is presented as the threshold PCR cycle (Ct) mean value in the different samples. The average Ct values of the three replicates and the standard deviation (SD) are reported. The results reveal a prominent band in the RNA samples collected in August, which is the hottest and driest period of the year, with average temperatures reaching approximately 32 °C.</p>
Full article ">Figure 6
<p>Western blot analysis results. Lanes 1 and 2 contain samples from latex and leaf tissue, respectively, under reducing conditions. Lanes 3 and 4 contain samples from latex and leaf tissue, respectively, under non-reducing conditions. The samples were collected in August. Solid arrows indicate an immunoreactive band at approximately 5–6 kDa, likely corresponding to the ubiquitin monomer, as well as several immunoreactive bands above 50 kDa, likely representing ubiquitinated proteins. Dashed arrows highlight immunoreactive bands of about 35 kDa present only in samples prepared under reducing conditions, suggesting that these bands correspond to protein complexes or aggregates that dissociate in the presence of reducing agents.</p>
Full article ">
27 pages, 8937 KiB  
Article
Using Multiscale Molecular Modeling to Analyze Possible NS2b-NS3 Protease Inhibitors from Philippine Medicinal Plants
by Allen Mathew Fortuno Cordero and Arthur A. Gonzales
Curr. Issues Mol. Biol. 2024, 46(7), 7592-7618; https://doi.org/10.3390/cimb46070451 - 18 Jul 2024
Viewed by 551
Abstract
Within the field of Philippine folkloric medicine, the utilization of indigenous plants like Euphorbia hirta (tawa-tawa), Carica papaya (papaya), and Psidium guajava (guava) as potential dengue remedies has gained attention. Yet, limited research exists on their comprehensive [...] Read more.
Within the field of Philippine folkloric medicine, the utilization of indigenous plants like Euphorbia hirta (tawa-tawa), Carica papaya (papaya), and Psidium guajava (guava) as potential dengue remedies has gained attention. Yet, limited research exists on their comprehensive effects, particularly their anti-dengue activity. This study screened 2944 phytochemicals from various Philippine plants for anti-dengue activity. Absorption, distribution, metabolism, excretion, and toxicity (ADMET) profiling provided 1265 compounds demonstrating pharmacokinetic profiles suitable for human use. Molecular docking targeting the dengue virus NS2b-NS3 protease’s catalytic triad (Asp 75, Ser 135, and His 51) identified ten ligands with higher docking scores than reference compounds idelalisib and nintedanib. Molecular dynamics simulations confirmed the stability of eight of these ligand–protease complexes. Molecular Mechanics/Poisson–Boltzmann Surface Area (MM/PBSA) analysis highlighted six ligands, including veramiline (−80.682 kJ/mol), cyclobranol (−70.943 kJ/mol), chlorogenin (−63.279 kJ/mol), 25beta-Hydroxyverazine (−61.951 kJ/mol), etiolin (−59.923 kJ/mol), and ecliptalbine (−56.932 kJ/mol) with favorable binding energies, high oral bioavailability, and drug-like properties. This integration of traditional medical knowledge with advanced computational drug discovery methods paves new pathways for the development of treatments for dengue. Full article
(This article belongs to the Special Issue New Insight: Enzymes as Targets for Drug Development, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Cartoon representation of the NS2b-NS3 protease (PDB ID: 2FOM) showing the active site grid (orange box) and the catalytic triad residues (ASP 75, HIS 51, and SER 135). Illustration was generated using Schrodinger Maestro [<a href="#B12-cimb-46-00451" class="html-bibr">12</a>].</p>
Full article ">Figure 2
<p>Overlay of the docked glycerol (GOL) structure on 2FOM using Autodock 4.2 and the original crystallized GOL structure from 2FOM. Illustration was generated using Schrodinger Maestro [<a href="#B12-cimb-46-00451" class="html-bibr">12</a>].</p>
Full article ">Figure 3
<p>Best docking conformations of top ten ligand–protein complex based on binding energy. Reference ligands: (<b>A</b>) IDE; (<b>B</b>) NIN. Candidate ligands: (<b>C</b>) VER; (<b>D</b>) ISO; (<b>E</b>) ETI; (<b>F</b>) TOM; (<b>G</b>) CAR; (<b>H</b>) HYD; (<b>I</b>) CHL; (<b>J</b>) ECL; (<b>K</b>) CYC; (<b>L</b>) HON. Illustrations were generated using UCSF Chimera 1.16 [<a href="#B34-cimb-46-00451" class="html-bibr">34</a>].</p>
Full article ">Figure 4
<p>Complex-RMSD of the top ten complexes and two references at 100 ns. Reference ligands: (<b>A</b>) IDE; (<b>B</b>) NIN. Candidate ligands: (<b>C</b>) VER; (<b>D</b>) ISO; (<b>E</b>) ETI; (<b>F</b>) TOM; (<b>G</b>) CAR; (<b>H</b>) HYD; (<b>I</b>) CHL; (<b>J</b>) ECL; (<b>K</b>) CYC; (<b>L</b>) HON. Graphs were plotted using GraphPad Prism 8.2.1 (GraphPad Software, San Diego, CA, USA, <a href="http://www.graphpad.com" target="_blank">www.graphpad.com</a>).</p>
Full article ">Figure 5
<p>RMSD graphs for the reference and candidate ligands using ligands only as the reference point (Ligand-RMSD). Reference ligands: (<b>A</b>) IDE; (<b>B</b>) NIN. Candidate ligands: (<b>C</b>) VER; (<b>D</b>) ISO; (<b>E</b>) ETI; (<b>F</b>) TOM; (<b>G</b>) CAR; (<b>H</b>) HYD; (<b>I</b>) CHL; (<b>J</b>) ECL; (<b>K</b>) CYC; (<b>L</b>) HON. Graphs were plotted using GraphPad Prism 8.2.1 (GraphPad Software, San Diego, CA, USA, <a href="http://www.graphpad.com" target="_blank">www.graphpad.com</a>).</p>
Full article ">Figure 6
<p>Active site-RMSD of the top ten complexes and two references at 100 ns. Reference ligands: (<b>A</b>) IDE; (<b>B</b>) NIN. Candidate ligands: (<b>C</b>) VER; (<b>D</b>) ISO; (<b>E</b>) ETI; (<b>F</b>) TOM; (<b>G</b>) CAR; (<b>H</b>) HYD; (<b>I</b>) CHL; (<b>J</b>) ECL; (<b>K</b>) CYC; (<b>L</b>) HON. Graphs were plotted using GraphPad Prism 8.2.1 (GraphPad Software, San Diego, CA, USA, <a href="http://www.graphpad.com" target="_blank">www.graphpad.com</a>).</p>
Full article ">Figure 7
<p>RMSF graphs of NS2b-NS3 Protease docked with the reference and best candidate ligands during the last 10 ns of MD simulation. Residues with interactions were denoted by a purple vertical line (|) on the <span class="html-italic">x</span>-axis. Reference ligands: (<b>A</b>) IDE; (<b>B</b>) NIN. Best candidate ligands: (<b>C</b>) VER; (<b>D</b>) ETI; (<b>E</b>) HYD; (<b>F</b>) CHL; (<b>G</b>) ECL; (<b>H</b>) CYC. Graphs were plotted using GraphPad Prism 8.2.1 (GraphPad Software, San Diego, CA, USA, <a href="http://www.graphpad.com" target="_blank">www.graphpad.com</a>).</p>
Full article ">Figure 8
<p>Illustration of interacting residues from 2FOM during the last 10 ns of MD simulation for the top six ligands and references. Reference ligands: (<b>A</b>) IDE; (<b>B</b>) NIN. Candidate ligands: (<b>C</b>) HYD; (<b>D</b>) VER; (<b>E</b>) ETI; (<b>F</b>) CHL; (<b>G</b>) CYC; (<b>H</b>) ECL. Illustrations were generated using Schrodinger Maestro [<a href="#B12-cimb-46-00451" class="html-bibr">12</a>].</p>
Full article ">Figure 9
<p>Binding energy decomposition on complexes with the highest MM/PBSA energies. Reference ligands: (<b>A</b>) IDE; (<b>B</b>) NIN. Candidate ligands: (<b>C</b>) VER; (<b>D</b>) ETI; (<b>E</b>) CAR; (<b>F</b>) HYD; (<b>G</b>) CHL; (<b>H</b>) ECL; (<b>I</b>) CYC; (<b>J</b>) HON. Graphs were plotted using GraphPad Prism 8.2.1 (GraphPad Software, San Diego, CA, USA, <a href="http://www.graphpad.com" target="_blank">www.graphpad.com</a>).</p>
Full article ">Figure 10
<p>Best conformations of ligand-NS2b-NS3 protein complexes after MD simulation at 100 ns. Reference ligands: (<b>A</b>) IDE; (<b>B</b>) NIN. Best candidate ligands: (<b>C</b>) VER; (<b>D</b>) ETI; (<b>E</b>) HYD; (<b>F</b>) CHL; (<b>G</b>) ECL; (<b>H</b>) CYC. Illustrations were generated using Schrodinger Maestro [<a href="#B12-cimb-46-00451" class="html-bibr">12</a>].</p>
Full article ">Figure 11
<p>3D structures of ligands for both unbound and bound states. Reference ligands: (<b>A</b>) IDE; (<b>B</b>) NIN. Candidate ligands: (<b>C</b>) VER; (<b>D</b>) ETI; (<b>E</b>) CAR; (<b>F</b>) HYD; (<b>G</b>) CHL; (<b>H</b>) ECL; (<b>I</b>) CYC; (<b>J</b>) HON. Illustrations were generated using PyMol 2.5.4 [<a href="#B46-cimb-46-00451" class="html-bibr">46</a>].</p>
Full article ">
19 pages, 2854 KiB  
Article
Effects of Ni and Cu Stresses on Morphological and Physiological Characteristics of Euphorbia marginata Pursh Seedlings
by Xudan Zhou, Yue An, Tongbao Qu, Tian Jin, Lei Zhao, Hongliang Guo, Wei Wang and Chunli Zhao
Agronomy 2024, 14(6), 1223; https://doi.org/10.3390/agronomy14061223 - 5 Jun 2024
Viewed by 474
Abstract
Increasing soil contamination with nickel (Ni) and copper (Cu) is a growing environmental concern, adversely affecting ecosystems and the survival of both plants and animals. This study investigated the morphological and physiological responses of Euphorbia marginata Pursh seedlings to varying concentrations of Ni [...] Read more.
Increasing soil contamination with nickel (Ni) and copper (Cu) is a growing environmental concern, adversely affecting ecosystems and the survival of both plants and animals. This study investigated the morphological and physiological responses of Euphorbia marginata Pursh seedlings to varying concentrations of Ni and Cu over a 45-day period. The findings revealed that low concentrations of Ni and Cu enhanced morphological indexes, root indexes, biomass, and photosynthetic pigment content of E. marginata, while high concentrations inhibited these parameters. Compared to the control, Ni and Cu stresses induced membrane peroxidation, increased cell membrane permeability, and inhibited the synthesis of soluble proteins and proline in the leaves. The seedlings demonstrated an ability to mitigate Ni and Cu toxicity by increasing soluble sugar content and enhancing the activities of peroxidase (POD), superoxide dismutase (SOD), and catalase (CAT). Notably, E. marginata exhibited a higher capacity for Cu2+ enrichment and translocation compared to Ni2+. Combined Ni and Cu treatments reduced the maximum enrichment and translocation levels of both metals in E. marginata. This study highlights the superior tolerance of E. marginata to Ni and Cu stresses and elucidates the mechanisms underlying its response, providing a theoretical basis for the use of landscape plants in the remediation of heavy-metal-contaminated soils. Full article
(This article belongs to the Section Grassland and Pasture Science)
Show Figures

Figure 1

Figure 1
<p>Schematic Diagram of the Main Steps of the Experimental Procedure.</p>
Full article ">Figure 2
<p>Response of <span class="html-italic">E. marginata</span> Seedlings to Exposure to 45 day Ni and Cu Stresses.</p>
Full article ">Figure 3
<p>Effects of Ni and Cu Stress on <span class="html-italic">E. marginata</span> Biomass: (<b>a</b>) aboveground fresh weight; (<b>b</b>) belowground fresh weight; (<b>c</b>) aboveground dry weight; (<b>d</b>) belowground dry weight. Note: Different lowercase letters indicate significant differences between treatments (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 4
<p>Effects of Ni and Cu Stresses on <span class="html-italic">E. marginata</span> Antioxidant Activity: (<b>a</b>) Chlorophyll a content; (<b>b</b>) chlorophyll b content; (<b>c</b>) total chlorophyll (a + b) content; (<b>d</b>) chlorophyll a/b ratio; (<b>e</b>) carotenoid content. Note: Different lowercase letters indicate significant differences between treatments (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 5
<p>Effects of Ni and Cu Stresses on <span class="html-italic">E. marginata</span> Antioxidant Activity: (<b>a</b>) POD activity; (<b>b</b>) SOD activity; (<b>c</b>) CAT activity. Note: Different lowercase letters indicate significant differences between treatments (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 6
<p>Effects of Ni and Cu Stresses on <span class="html-italic">E. marginata</span> Osmoregulatory Substance Content: (<b>a</b>) soluble sugar content; (<b>b</b>) soluble protein content; (<b>c</b>) proline content. Note: Different lowercase letters indicate significant differences between treatments (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 7
<p>Effects of Ni and Cu Stresses on <span class="html-italic">E. marginata</span> Cell Membrane Permeability: (<b>a</b>) MDA content; (<b>b</b>) relative conductivity. Note: Different lowercase letters indicate significant differences between treatments (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">
8 pages, 251 KiB  
Review
Vertebrate Pollination of Angiosperms in the Mediterranean Area: A Review
by Benito Valdés
Plants 2024, 13(6), 895; https://doi.org/10.3390/plants13060895 - 20 Mar 2024
Viewed by 850
Abstract
For a long time, it was considered that entomogamy was the only pollination mechanism in the Mediterranean area. However, data recorded in this review prove that ornithogamy and saurogamy also take place. With the exception of the nectarivorous Cinnyris osea (Nectariniidae) which pollinates [...] Read more.
For a long time, it was considered that entomogamy was the only pollination mechanism in the Mediterranean area. However, data recorded in this review prove that ornithogamy and saurogamy also take place. With the exception of the nectarivorous Cinnyris osea (Nectariniidae) which pollinates the mistletoe Picosepalus acaciae in Israel, all birds responsible for the pollination of several plant species in this area are primarily insectivorous, sedentary, or migrating passerine birds, particularly Sylvia atricapilla, S. melanocephala, Phylloscopus collibita and Parus caeruleus. They contribute, together with insects, to the pollination of Anagyris foetida, three species of Scrophularia with big flowers, Rhamnus alaternus, Brassica oleracea, and some other plants. The lacertid lizard Podarcis lilfordi acts as a pollinating agent on several W Mediterranean islands, where it effectively pollinates Euphorbia dendroides, Cneorum tricocum, and presumably Rosmarinus officinalis and Chrithmum maritimum. The flowers of some other plant species are visited by birds or by Podarcis species in the Mediterranean area, where they could also contribute to their pollination. Full article
11 pages, 1745 KiB  
Article
A Biomimetic Approach to Premyrsinane-Type Diterpenoids: Exploring Microbial Transformation to Enhance Their Chemical Diversity
by Felipe Escobar-Montaño, Antonio J. Macías-Sánchez, José M. Botubol-Ares, Rosa Durán-Patrón and Rosario Hernández-Galán
Plants 2024, 13(6), 842; https://doi.org/10.3390/plants13060842 - 14 Mar 2024
Viewed by 814
Abstract
Premyrsinane-type diterpenoids have been considered to originate from the cyclization of a suitable 5,6- or 6,17-epoxylathyrane precursor. Their biological activities have not been sufficiently explored to date, so the development of synthetic or microbial approaches for the preparation of new derivatives would be [...] Read more.
Premyrsinane-type diterpenoids have been considered to originate from the cyclization of a suitable 5,6- or 6,17-epoxylathyrane precursor. Their biological activities have not been sufficiently explored to date, so the development of synthetic or microbial approaches for the preparation of new derivatives would be desirable. Epoxyboetirane A (4) is an 6,17-epoxylathyrane isolated from Euphorbia boetica in a large enough amount to be used in semi-synthesis. Transannular cyclization of 4 mediated by Cp2TiIIICl afforded premyrsinane 5 in good yield as an only diasteroisomer. To enhance the structural diversity of premyrsinanes so their potential use in neurodegenerative disorders could be explored, compound 5 was biotransformed by Mucor circinelloides NRRL3631 to give rise to hydroxylated derivatives at non-activated carbons (67), all of which were reported here for the first time. The structures and absolute configurations of all compounds were determined through extensive NMR and HRESIMS spectroscopic studies. Full article
(This article belongs to the Section Phytochemistry)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Key COSY (red bond) and HMBC (blue arrows) correlations for compound <b>5</b>. (<b>b</b>) Selected 1D and 2D NOESY correlations for compound <b>5</b>. <span class="html-italic">β</span>-Face correlations in red. <span class="html-italic">α</span>-Face correlations in blue. Correlations involving nuclei in α-face and <span class="html-italic">β</span>-face in black.</p>
Full article ">Figure 2
<p>Key COSY (red bond) and HMBC (blue arrows) correlations for compounds <b>6</b> and <b>7</b>.</p>
Full article ">Scheme 1
<p>Biogenetic relationship between lathyranes and premyrsinanes.</p>
Full article ">Scheme 2
<p>Chemical conversion of lathyradiene <b>1</b> into premyrsinanes <b>2a</b> and <b>2b</b> mediated by Fe.</p>
Full article ">Scheme 3
<p>Preparation of premyrsinane <b>5</b>.</p>
Full article ">Scheme 4
<p>Catalytic cycle for the Ti(III)-mediated radical transannular cyclization.</p>
Full article ">Scheme 5
<p>Biotransformation of <b>5</b> by <span class="html-italic">Mucor circinelloides</span> NRRL3631.</p>
Full article ">
29 pages, 3499 KiB  
Article
Folk Knowledge and Perceptions about the Use of Wild Fruits and Vegetables–Cross-Cultural Knowledge in the Pipli Pahar Reserved Forest of Okara, Pakistan
by Sadia Jabeen, Fahim Arshad, Nidaa Harun, Muhammad Waheed, Saud Alamri, Shiekh Marifatul Haq, Ivana Vitasović-Kosić, Kaneez Fatima, Abdul Shakoor Chaudhry and Rainer W. Bussmann
Plants 2024, 13(6), 832; https://doi.org/10.3390/plants13060832 - 14 Mar 2024
Cited by 1 | Viewed by 1362
Abstract
Wild fruits and vegetables (WFVs) have been vital to local communities for centuries and make an important contribution to daily life and income. However, traditional knowledge of the use of wild fruits is at risk of being lost due to inadequate documentation. This [...] Read more.
Wild fruits and vegetables (WFVs) have been vital to local communities for centuries and make an important contribution to daily life and income. However, traditional knowledge of the use of wild fruits is at risk of being lost due to inadequate documentation. This study aimed to secure this knowledge through intermittent field visits and a semi-structured questionnaire. Using various ethnobotanical data analysis tools and SPSS (IBM 25), this study identified 65 WFV species (52 genera and 29 families). These species, mostly consumed as vegetables (49%) or fruits (43%), were predominantly herbaceous (48%) in wild and semi-wild habitats (67%). 20 WFVs were known to local communities (highest RFC), Phoenix sylvestris stood out as the most utilized species (highest UV). Surprisingly, only 23% of the WFVs were sold at markets. The survey identified 21 unique WFVs that are rarely documented for human consumption in Pakistan (e.g., Ehretia obtusifolia, Euploca strigosa, Brassica juncea, Cleome brachycarpa, Gymnosporia royleana, Cucumis maderaspatanus, Croton bonplandianus, Euphorbia prostrata, Vachellia nilotica, Pongamia pinnata, Grewia asiatica, Malvastrum coromandelianum, Morus serrata, Argemone mexicana, Bambusa vulgaris, Echinochloa colonum, Solanum virginianum, Physalis angulata, Withania somnifera, Zygophyllum creticum, and Peganum harmala), as well as 14 novel uses and five novel edible parts. Despite their ecological importance, the use of WFVs has declined because local people are unaware of their cultural and economic value. Preservation of traditional knowledge through education on conservation and utilization could boost economies and livelihoods in this and similar areas worldwide. Full article
(This article belongs to the Section Plant Systematics, Taxonomy, Nomenclature and Classification)
Show Figures

Figure 1

Figure 1
<p>Chord diagram showing the species and type of utilization of wild fruits and vegetables from Pipli Pahar Reserved Forest in Okara, Pakistan. The full names of the species are listed in <a href="#plants-13-00832-t0A1" class="html-table">Table A1</a> in <a href="#app1-plants-13-00832" class="html-app">Appendix A</a>.</p>
Full article ">Figure 2
<p>Chord diagram of utilized parts of wild fruits and vegetables from Pipli Pahar Reserved Forest in Okara, Pakistan. The complete names of the species are given in <a href="#plants-13-00832-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 3
<p>Biplot analysis of Principal Component Analysis (PCA) illustrating the predominant recipes of different plant species in Pipli Pahar Reserved Forest of Okara, Pakistan.</p>
Full article ">Figure 4
<p>Biplot analysis of Principal Component Analysis (PCA) illustrating the prevalent diseases treated with different plant species in Pipli Pahar Reserved Forest of Okara, Pakistan.</p>
Full article ">Figure 5
<p>Venn diagram representing unique and common plant species among different categories.</p>
Full article ">Figure 6
<p>A pictorial representation of some typical WFVs in the studied area of Pipli Pahar where (<b>A</b>) <span class="html-italic">Argemone mexicana</span> L., (<b>B</b>) <span class="html-italic">Croton bonplandianus</span> Baill., (<b>C</b>) <span class="html-italic">Ehretia obtusifolia</span> Hochst. ex ADC., (<b>D</b>) <span class="html-italic">Euploca strigosa</span> (Willd.) Diane &amp; Hilger, (<b>E</b>) <span class="html-italic">Cleome brachycarpa</span> Vahl ex DC., (<b>F</b>) <span class="html-italic">Brassica juncea</span> (L.) Czern., (<b>G</b>) <span class="html-italic">Gymnosporia royleana</span> Wall.ex M.A. Lawson, (<b>H</b>) <span class="html-italic">Cucumis maderaspatanus</span> L., (<b>I</b>) <span class="html-italic">Euphorbia prostrata</span> Aiton, (<b>J</b>) <span class="html-italic">Grewia asiatica</span> L., (<b>K</b>) <span class="html-italic">Bambusa vulgaris</span> Schrad. ex J.C.Wendl., (<b>L</b>) <span class="html-italic">Pongamia pinnata</span> (L.), (<b>M</b>) <span class="html-italic">Vachellia nilotica</span>, (L.) P.J.H.Hurter &amp; Mabb., (<b>N</b>) <span class="html-italic">Echinochloa colonum</span> (L.) Link, (<b>O</b>) <span class="html-italic">Morus serrata</span> Roxb. (<b>P</b>) <span class="html-italic">Withania somnifera</span> (L.) Dunal, (<b>Q</b>) <span class="html-italic">Physalis angulata</span> L., (<b>R</b>) Solanum virginianum L., (<b>S</b>) <span class="html-italic">Zygophyllum creticum</span> (L.) Christenh. &amp; Byng, (<b>T</b>) <span class="html-italic">Peganum harmala</span> L., and (<b>U</b>) <span class="html-italic">Malvastrum coromandelianum</span> (L.). (These are original photographs, Jabeen 2022).</p>
Full article ">Figure 7
<p>Map of the study area Pipli Pahar reserved forest.</p>
Full article ">
19 pages, 2597 KiB  
Article
Chemical Characterization and Biological Properties Assessment of Euphorbia resinifera and Euphorbia officinarum Moroccan Propolis
by Oumaima Boutoub, Soukaina El-Guendouz, Isabel Matos, Lahsen El Ghadraoui, Maria Clara Costa, Jorge Dias Carlier, Maria Leonor Faleiro, Ana Cristina Figueiredo, Letícia M. Estevinho and Maria Graça Miguel
Antibiotics 2024, 13(3), 230; https://doi.org/10.3390/antibiotics13030230 - 29 Feb 2024
Viewed by 1425
Abstract
Although the plants of the genus Euphorbia are largely exploited by therapists in Morocco, the composition and antibacterial activities of propolis from these plants are still unknown. To address this gap, this study aimed to characterize the pollen type, the volatile compounds, and [...] Read more.
Although the plants of the genus Euphorbia are largely exploited by therapists in Morocco, the composition and antibacterial activities of propolis from these plants are still unknown. To address this gap, this study aimed to characterize the pollen type, the volatile compounds, and the phenolic and mineral profiles of three Euphorbia propolis samples collected in Morocco and evaluate their antimicrobial activities. The minimum inhibitory concentration of the propolis samples was determined by the microdilution method, and the anti-adherence activity was evaluated by the crystal violet assay. The examination of anti-quorum-sensing proprieties was performed using the biosensor Chromobacterium violaceum CV026. Pollen analysis revealed that Euphorbia resinifera pollen dominated in the P1 sample (58%), while E. officinarum pollen dominated in the P2 and P3 samples (44%). The volatile compounds were primarily composed of monoterpene hydrocarbons, constituting 35% in P1 and 31% in P2, with α-pinene being the major component in both cases, at 16% in P1 and 15% in P2. Calcium (Ca) was the predominant mineral element in both E. resinifera (P1) and E. officinarum (P2 and P3) propolis samples. Higher levels of phenols, flavonoids and dihydroflavonoids were detected in the E. officinarum P2 sample. The minimum inhibitory concentration (MIC) value ranged from 50 to 450 µL/mL against Gram-positive and Gram-negative bacteria. Euphorbia propolis displayed the ability to inhibit quorum sensing in the biosensor C. violaceum CV026 and disrupted bacterial biofilm formation, including that of resistant bacterial pathogens. In summary, the current study evidences the potential use of E. officinarum propolis (P2 and P3) to combat important features of resistant pathogenic bacteria, such as quorum sensing and biofilm formation. Full article
Show Figures

Figure 1

Figure 1
<p>Anti-QS properties of propolis. P1: <span class="html-italic">Euphorbia resinifera</span> propolis. P2 and P3: <span class="html-italic">Euphorbia officinarum</span> propolis. N-hexanoylhomoserine lactone (C6-HSL) at 0.12 µg/mL was added to the culture medium. Control: No addition of C6-HSL to the culture medium. The white arrow highlights the concentration (0.6 mg/mL) at which the zone of inhibition of the production of the pigment violacein is observed. The assay was conducted using three independent triplicates.</p>
Full article ">Figure 2
<p>Inhibition of the bacterial adherence by the propolis samples P1, P2 and P3. Data represent the mean of three biological replicates. Error bars represent the standard deviation. * <span class="html-italic">p</span> &lt; 0.05. ** <span class="html-italic">p</span> &lt; 0.01. *** <span class="html-italic">p</span> &lt; 0.001. **** <span class="html-italic">p</span> &lt; 0.0001. ns: not significant.</p>
Full article ">Figure 3
<p>The impact of the <span class="html-italic">E. resinifera</span> (P1) propolis hydro-alcoholic extract on the disruption of biofilm formed by MRSA12 ((<b>A</b>) control, no exposure to anti-biofilm agent; (<b>B</b>) after exposure to propolis P1; (<b>C</b>) after exposure to 70% ethanol) and multiresistant <span class="html-italic">E. coli</span> I73194 ((<b>D</b>) control, no exposure to anti-biofilm agent; (<b>E</b>) after exposure to propolis P1; (<b>F</b>) after exposure to 70% ethanol). The bacterial cells formed a biofilm over 24 h and were visualized after staining with LIVE/DEAD Baclight.</p>
Full article ">Figure 4
<p>From right to left, the geographical location of Morocco and of the apiaries where the samples of <span class="html-italic">Euphorbia resinifera</span> (P1) and <span class="html-italic">E. officinarum</span> propolis (P2 and P3) were obtained. On the left are the flowers and propolis of <span class="html-italic">E. resinifera</span> (P1) and <span class="html-italic">E. officinarum</span> (P2 and P3).</p>
Full article ">
40 pages, 4309 KiB  
Review
Anti-Inflammatory and Cytotoxic Compounds Isolated from Plants of Euphorbia Genus
by Sarai Rojas-Jiménez, María Guadalupe Valladares-Cisneros, David Osvaldo Salinas-Sánchez, Julia Pérez-Ramos, Leonor Sánchez-Pérez, Salud Pérez-Gutiérrez and Nimsi Campos-Xolalpa
Molecules 2024, 29(5), 1083; https://doi.org/10.3390/molecules29051083 - 29 Feb 2024
Cited by 1 | Viewed by 1213
Abstract
Euphorbia is a large genus of the Euphorbiaceae family. Around 250 species of the Euphorbia genus have been studied chemically and pharmacologically; different compounds have been isolated from these species, especially diterpenes and triterpenes. Several reports show that several species have anti-inflammatory activity, [...] Read more.
Euphorbia is a large genus of the Euphorbiaceae family. Around 250 species of the Euphorbia genus have been studied chemically and pharmacologically; different compounds have been isolated from these species, especially diterpenes and triterpenes. Several reports show that several species have anti-inflammatory activity, which can be attributed to the presence of diterpenes, such as abietanes, ingenanes, and lathyranes. In addition, it was found that some diterpenes isolated from different Euphorbia species have anti-cancer activity. In this review, we included compounds isolated from species of the Euphorbia genus with anti-inflammatory or cytotoxic effects published from 2018 to September 2023. The databases used for this review were Science Direct, Scopus, PubMed, Springer, and Google Scholar, using the keywords Euphorbia with anti-inflammatory or cytotoxic activity. In this review, 68 studies were collected and analyzed regarding the anti-inflammatory and anti-cancer activities of 264 compounds obtained from 36 species of the Euphorbia genus. The compounds included in this review are terpenes (95%), of which 68% are diterpenes, especially of the types ingenanes, abietanes, and triterpenes (approximately 15%). Full article
Show Figures

Figure 1

Figure 1
<p>Structures of compounds isolated from <span class="html-italic">E. antiquorum</span>.</p>
Full article ">Figure 2
<p>Structures of compounds isolated from <span class="html-italic">E. atoto</span> and <span class="html-italic">E. ebracteolata</span>.</p>
Full article ">Figure 3
<p>Structures of compounds isolated from <span class="html-italic">E. fischeriana</span> and <span class="html-italic">E. formasana</span>.</p>
Full article ">Figure 4
<p>Structures of compounds isolated from <span class="html-italic">E. helioscopia</span> and <span class="html-italic">E. kansuensis</span>.</p>
Full article ">Figure 5
<p>Structures of compounds isolated from <span class="html-italic">E. kansui</span>.</p>
Full article ">Figure 6
<p>Structures of compounds isolated from <span class="html-italic">E. lathyris</span>.</p>
Full article ">Figure 7
<p>Structures of compounds isolated from <span class="html-italic">E. maculata</span>.</p>
Full article ">Figure 8
<p>Structures of compounds isolated from <span class="html-italic">E. neriifolia</span>.</p>
Full article ">Figure 9
<p>Structures of compounds isolated from <span class="html-italic">E. peplus</span> and <span class="html-italic">E. pulcherrima</span>.</p>
Full article ">Figure 10
<p>Structures of compounds isolated from <span class="html-italic">E. resinifera</span>, <span class="html-italic">E. thymifolia,</span> and <span class="html-italic">E. wallichii</span>.</p>
Full article ">Figure 11
<p>Structures of compounds isolated from <span class="html-italic">E. alatavica</span>, <span class="html-italic">E. balsamifera</span>, <span class="html-italic">E. dendroides</span>, and <span class="html-italic">E. denticulata</span>.</p>
Full article ">Figure 12
<p>Structures of compounds isolated from <span class="html-italic">E. ebracteolata</span>.</p>
Full article ">Figure 13
<p>Structures of compounds isolated from <span class="html-italic">E. fisheriana</span>.</p>
Full article ">Figure 13 Cont.
<p>Structures of compounds isolated from <span class="html-italic">E. fisheriana</span>.</p>
Full article ">Figure 14
<p>Structures of compounds isolated from <span class="html-italic">E. gedrosiaca</span>, <span class="html-italic">E. glomerulans</span>, <span class="html-italic">E. grandicornis</span>, and <span class="html-italic">E. grantii</span>.</p>
Full article ">Figure 15
<p>Structures of compounds isolated from <span class="html-italic">E. helioscopia</span>.</p>
Full article ">Figure 15 Cont.
<p>Structures of compounds isolated from <span class="html-italic">E. helioscopia</span>.</p>
Full article ">Figure 16
<p>Structures of compounds from <span class="html-italic">E. hypericifolia</span>, <span class="html-italic">E. kansuensis</span>, <span class="html-italic">E. kansui</span>, and <span class="html-italic">E. kopetdaghi</span>.</p>
Full article ">Figure 17
<p>Structures of compounds isolated from <span class="html-italic">E. lactea</span>, <span class="html-italic">E. lathyris</span>, <span class="html-italic">E. microsphaera</span>, and <span class="html-italic">E. neriifolia</span>.</p>
Full article ">Figure 18
<p>Structures of compounds isolated from <span class="html-italic">E. pedroi</span> and <span class="html-italic">E. pekinensis</span>.</p>
Full article ">Figure 19
<p>Structures of compounds isolated from <span class="html-italic">E. saudiarabica</span>, <span class="html-italic">E. schimperiana</span>, <span class="html-italic">E. sororia</span>, <span class="html-italic">E. stracheyi</span>, and <span class="html-italic">E. tirucalli</span>.</p>
Full article ">Figure 20
<p>Hidrocarbon skeleton of <span class="html-italic">Euphorbia</span> diterpene classes.</p>
Full article ">
15 pages, 2600 KiB  
Article
Citrus Aphids in Algarve Region (Portugal): Species, Hosts, and Biological Control
by Paulo Eduardo Branco Paiva, Luís Mascarenhas Neto, Natália Tomás Marques, Beatriz Zarcos Duarte and Amílcar Marreiros Duarte
Ecologies 2024, 5(1), 101-115; https://doi.org/10.3390/ecologies5010007 - 19 Feb 2024
Cited by 1 | Viewed by 1663
Abstract
Aphids affect citrus by causing leaf deformations and reducing fruit production. Additionally, aphids are a great concern due to their ability to transmit Citrus tristeza virus (CTV), the cause of tristeza, one of the main citrus diseases. In the last four years, citrus [...] Read more.
Aphids affect citrus by causing leaf deformations and reducing fruit production. Additionally, aphids are a great concern due to their ability to transmit Citrus tristeza virus (CTV), the cause of tristeza, one of the main citrus diseases. In the last four years, citrus orchards in the south of Portugal (Algarve region) were sampled for aphid species identification and counting. Aphis spiraecola was the most abundant species, representing more than 80% of all identified aphids, and the damage (leaf deformation) it causes was directly proportional to its density. A. gossypii was the second most common species, followed by A. aurantii and Macrosiphum euphorbiae. The number of aphids in nymph stages was predominant over the adult stages (both wingless and winged) in all species. A. citricidus, the most efficient CTV vector, was not detected. The largest populations of A. spiraecola were observed in lemon and orange trees during spring (>100 individuals per shoot), with great damage observed in orange, lemon, and mandarin trees. A. gossypii was observed mainly in mandarin and tangor trees. There was a low activity of natural biological control agents, with the parasitism of A. spiraecola by Lysiphlebus spp. and Binodoxys spp. ranging from 0.3 to 1.5%. The numerical ratio ranged from 150 to 440 aphids per predator, and among these, syrphids were the most abundant, followed by lacewings and coccinellids (Scymnus). Full article
Show Figures

Figure 1

Figure 1
<p>Leaf deformation (curling) caused by <span class="html-italic">Aphis spiraecola</span> on young shoots (<b>left</b>) that persists as shoots mature (<b>right</b>).</p>
Full article ">Figure 2
<p>Location of sampling points in the Algarve region (Southern Portugal). <span class="html-fig-inline" id="ecologies-05-00007-i001"><img alt="Ecologies 05 00007 i001" src="/ecologies/ecologies-05-00007/article_deploy/html/images/ecologies-05-00007-i001.png"/></span> Survey 1; <span class="html-fig-inline" id="ecologies-05-00007-i002"><img alt="Ecologies 05 00007 i002" src="/ecologies/ecologies-05-00007/article_deploy/html/images/ecologies-05-00007-i002.png"/></span> Survey 2 and 3; <span class="html-fig-inline" id="ecologies-05-00007-i003"><img alt="Ecologies 05 00007 i003" src="/ecologies/ecologies-05-00007/article_deploy/html/images/ecologies-05-00007-i003.png"/></span> Survey 4.</p>
Full article ">Figure 3
<p>Relationship between the number of <span class="html-italic">Aphis spiraecola</span> per shoot and the proportion (%) of damaged leaves on the same shoot. These observations were carried out in the spring of 2019 on a citrus collection orchard in southern Portugal (Y = 0.3823X + 5.6561, R<sup>2</sup> = 0.77, ANOVA: F = 44.07, <span class="html-italic">p</span> &lt; 0.0001). The shaded area represents the standard error of the regression.</p>
Full article ">Figure 4
<p>Proportion of shoots with aphid nymphs and adults (F = 2.98; <span class="html-italic">p</span> = 0.11) on a ’Lane Late’ orange orchard managed organically (green bars) and in a conventional way (blue bars) (survey 4). Statistical analysis was performed after a data transformation of arcsin√p.</p>
Full article ">
10 pages, 5721 KiB  
Communication
Identification of One O-Methyltransferase Gene Involved in Methylated Flavonoid Biosynthesis Related to the UV-B Irradiation Response in Euphorbia lathyris
by Wanli Zhao, Long Huang, Shu Xu, Junzhi Wu, Fan Wang, Pirui Li, Linwei Li, Mei Tian, Xu Feng and Yu Chen
Int. J. Mol. Sci. 2024, 25(2), 782; https://doi.org/10.3390/ijms25020782 - 8 Jan 2024
Viewed by 906
Abstract
Flavonoids are ubiquitous polyphenolic compounds that play a vital role in plants’ defense response and medicinal efficacy. UV-B radiation is a vital environmental regulator governing flavonoid biosynthesis in plants. Many plants rapidly biosynthesize flavonoids as a response to UV-B stress conditions. Here, we [...] Read more.
Flavonoids are ubiquitous polyphenolic compounds that play a vital role in plants’ defense response and medicinal efficacy. UV-B radiation is a vital environmental regulator governing flavonoid biosynthesis in plants. Many plants rapidly biosynthesize flavonoids as a response to UV-B stress conditions. Here, we investigated the effects of flavonoid biosynthesis via UV-B irradiation in Euphorbia lathyris. We found that exposure of the E. lathyris callus to UV-B radiation sharply increased the level of one O-methyltransferase (ElOMT1) transcript and led to the biosynthesis of several methylated flavonoids. The methyltransferase ElOMT1 was expressed heterologously in E. coli, and we tested the catalytic activity of recombinant ElOMT1 with possible substrates, including caffeic acid, baicalin, and luteolin, in vitro. ElOMT1 could efficiently methylate when the hydroxyl groups were contained in the core nucleus of the flavonoid. This molecular characterization identifies a methyltransferase responsible for the chemical modification of the core flavonoid structure through methylation and helps reveal the mechanism of methylated flavonoid biosynthesis in Euphorbiaceae. This study identifies the O-methyltransferase that responds to UV-B irradiation and helps shed light on the mechanism of flavonoid biosynthesis in Euphorbia lathyris. Full article
Show Figures

Figure 1

Figure 1
<p>Volcano map of differentially expressed genes after UV-B irradiation of <span class="html-italic">E. lathyris</span> callus (ELC-UV0h vs. ELC-UV3h). ELC-UV0h: the callus of <span class="html-italic">E. lathyris</span> was not exposed to UV-B irradiation; ELC-UV3h: the callus of <span class="html-italic">E. lathyris</span> was exposed to UV-B irradiation for 3 h. FoldChange: the differential fold change in gene expression.</p>
Full article ">Figure 2
<p>KEGG metabolic pathway classification of <span class="html-italic">E. lathyris</span>.</p>
Full article ">Figure 3
<p>Relative expression levels and phylogenetic tree of methyltransferase genes. Relative expression levels of ElOMT genes measured through FPKM (<b>A</b>). The genes with low expression (FPKM less than 20) in different samples are not listed. Bars represent the mean ± SD of three biological replicates. The Student’s t-test was used to analyze gene expression data. The level of significance was set at *** <span class="html-italic">p</span> &lt; 0.001. The phylogenetic tree of candidate ElOMT1 and previously characterized plant methyltransferases was constructed using the maximum likelihood (ML) with 1000 bootstrap replicates (<b>B</b>). GenBank accession numbers and sequence of <span class="html-italic">O</span>-methyltransferase proteins in this tree are listed in <a href="#app1-ijms-25-00782" class="html-app">Table S2 and Supporting Information S2</a>.</p>
Full article ">Figure 4
<p>Results of SDS-PAGE of ElOMT1. M: marker; lane 1: pellet of pMAL-c4x (empty vector); lane 2: supernatant of pMAL-c4x; lane 3: pellet of ElOMT1; lane 4: supernatant of ElOMT1; lane 5: purified ElOMT1 (MBP Tag ≈ 42 KDa). Red box: the band of recombinant protein ElOMT1.</p>
Full article ">Figure 5
<p>Results of caffeic acid substrate catalyzed by ElOMT1 in vitro. EIC, extract ion chromatogram.</p>
Full article ">Figure 6
<p>Results of baicalein substrate catalyzed by ElOMT1 in vitro. EIC, extract ion chromatogram.</p>
Full article ">Figure 7
<p>Results of luteolin substrate catalyzed by ElOMT1 in vitro. EIC, extract ion chromatogram.</p>
Full article ">
32 pages, 5158 KiB  
Review
Myrsinane-Type Diterpenes: A Comprehensive Review on Structural Diversity, Chemistry and Biological Activities
by Eduarda Mendes, Cátia Ramalhete and Noélia Duarte
Int. J. Mol. Sci. 2024, 25(1), 147; https://doi.org/10.3390/ijms25010147 - 21 Dec 2023
Cited by 3 | Viewed by 1574
Abstract
Euphorbia species are important sources of polycyclic and macrocyclic diterpenes, which have been the focus of natural-product-based drug research due to their relevant biological properties, including anticancer, multidrug resistance reversal, antiviral, and anti-inflammatory activities. Premyrsinane, cyclomyrsinane, and myrsinane diterpenes are generally and collectively [...] Read more.
Euphorbia species are important sources of polycyclic and macrocyclic diterpenes, which have been the focus of natural-product-based drug research due to their relevant biological properties, including anticancer, multidrug resistance reversal, antiviral, and anti-inflammatory activities. Premyrsinane, cyclomyrsinane, and myrsinane diterpenes are generally and collectively designated as myrsinane-type diterpenes. These compounds are derived from the macrocyclic lathyrane structure and are characterized by having highly oxygenated rearranged polycyclic systems. This review aims to describe and summarize the distribution and diversity of 220 myrsinane-type diterpenes isolated in the last four decades from about 20 Euphorbia species. Some myrsinane diterpenes obtained from Jatropha curcas are also described. Discussion on their plausible biosynthetic pathways is presented, as well as isolation procedures and structural elucidation using nuclear magnetic resonance spectroscopy. Furthermore, the most important biological activities are highlighted, which include cytotoxic and immunomodulatory activities, the modulation of efflux pumps, the neuroprotective effects, and the inhibition of enzymes such as urease, HIV-1 reverse transcriptase, and prolyl endopeptidase, among other biological effects. Full article
(This article belongs to the Special Issue Natural Product Chemistry and Biological Research)
Show Figures

Figure 1

Figure 1
<p>From GGPP to polycyclic and macrocyclic diterpenes—a biosynthetic proposal (adapted from [<a href="#B18-ijms-25-00147" class="html-bibr">18</a>]).</p>
Full article ">Figure 2
<p>Biosynthesis of jatrophane-, lathyrane-, and myrsinane-type diterpenes.</p>
Full article ">Figure 3
<p>Reduction of the polyester mixture obtained from <span class="html-italic">E. myrsinitis</span> to obtain the 14α,β myrsinol epimers [<a href="#B39-ijms-25-00147" class="html-bibr">39</a>].</p>
Full article ">Figure 4
<p>6,17-epoxylathyrane as the myrsinol biosynthetic precursor, as proposed by Rentzea and Hecker [<a href="#B36-ijms-25-00147" class="html-bibr">36</a>].</p>
Full article ">Figure 5
<p>5,6-epoxylathyranes as premyrsinol, myrsinol, and cyclomyrsinol biosynthetic precursors, as proposed by Jeske et al. [<a href="#B37-ijms-25-00147" class="html-bibr">37</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) Chemical transformation of Euphorbia factor L1 as proposed by Wang and collaborators in 2019 [<a href="#B45-ijms-25-00147" class="html-bibr">45</a>]. (<b>b</b>) Structure of PM1 and PM2 obtained under the catalysis of iron [<a href="#B46-ijms-25-00147" class="html-bibr">46</a>].</p>
Full article ">Figure 7
<p><sup>1</sup>H- and <sup>13</sup>C-NMR spectroscopic data of the premyrsinane diterpene <b>25 [<a href="#B58-ijms-25-00147" class="html-bibr">58</a>]</b>.</p>
Full article ">Figure 8
<p><sup>1</sup>H- and <sup>13</sup>C-NMR spectroscopic data of the cyclomyrsinane diterpene <b>8</b> [<a href="#B51-ijms-25-00147" class="html-bibr">51</a>].</p>
Full article ">Figure 9
<p><sup>1</sup>H- and <sup>13</sup>C-NMR spectroscopic data of myrsinane diterpenes <b>205</b> [<a href="#B37-ijms-25-00147" class="html-bibr">37</a>] and <b>176</b> [<a href="#B59-ijms-25-00147" class="html-bibr">59</a>].</p>
Full article ">Figure 10
<p><sup>1</sup>H- and <sup>13</sup>C-NMR spectroscopic data of rearranged myrsinane diterpene <b>4</b> [<a href="#B51-ijms-25-00147" class="html-bibr">51</a>].</p>
Full article ">Figure 11
<p>Selected <sup>1</sup>H−<sup>1</sup>H COSY (bold, in red, green and blue) and HMBC (H→C) correlations for myrsinane, premyrsinane, and cyclomyrsinane skeleton.</p>
Full article ">Figure 12
<p>Chemical structures of myrsinane-type diterpenes isolated from <span class="html-italic">E. falcata</span>.</p>
Full article ">Figure 13
<p>Chemical structures of myrsinane-type diterpenes isolated from <span class="html-italic">E. prolifera.</span></p>
Full article ">Figure 13 Cont.
<p>Chemical structures of myrsinane-type diterpenes isolated from <span class="html-italic">E. prolifera.</span></p>
Full article ">Figure 14
<p>Chemical structures of myrsinane-type diterpenes isolated from <span class="html-italic">E. kopetdaghi and E. sogdiana</span>.</p>
Full article ">Figure 15
<p>Chemical structures of myrsinane-type diterpenes isolated from <span class="html-italic">E. gedrosiaca</span>.</p>
Full article ">Figure 16
<p>Chemical structures of myrsinane-type diterpenes isolated from <span class="html-italic">E. decipiens</span>.</p>
Full article ">Figure 17
<p>Chemical structures of myrsinane-type diterpenes isolated from <span class="html-italic">E. connata</span> and <span class="html-italic">E. sanctae-catharinae</span>.</p>
Full article ">Figure 18
<p>Chemical structures of myrsinane-type diterpenes isolated from <span class="html-italic">E. pithyusa and E. cupanii</span>.</p>
Full article ">Figure 19
<p>Chemical structures of myrsinanes isolated from <span class="html-italic">E. aellenii</span> and <span class="html-italic">E. aleppica</span>.</p>
Full article ">Figure 20
<p>Chemical structures of myrsinane-type diterpenes isolated from <span class="html-italic">E. lathyris</span>, <span class="html-italic">E. macrorrhiza</span> and <span class="html-italic">E. boetica</span>.</p>
Full article ">Figure 21
<p>Chemical structures of myrsinane-type diterpenes isolated from <span class="html-italic">E. myrsinites, E. cheiradenia, E. dracunculoides</span>, and <span class="html-italic">E. erythradenia</span>.</p>
Full article ">Figure 22
<p>Chemical structures of myrsinane-type diterpenes isolated from <span class="html-italic">E. macroclada</span>, <span class="html-italic">E. microsciadia</span>, <span class="html-italic">E. seguieriana</span>, <span class="html-italic">E. splendida</span>, <span class="html-italic">E. teheranica</span>, and <span class="html-italic">Jatropha curcas</span>.</p>
Full article ">
22 pages, 1482 KiB  
Article
LC-ESI QToF MS Non-Targeted Screening of Latex Extracts of Euphorbia seguieriana ssp. seguieriana Necker and Euphorbia cyparissias and Determination of Their Potential Anticancer Activity
by Milka Jadranin, Danica Savić, Ema Lupšić, Ana Podolski-Renić, Milica Pešić, Vele Tešević, Slobodan Milosavljević and Gordana Krstić
Plants 2023, 12(24), 4181; https://doi.org/10.3390/plants12244181 - 16 Dec 2023
Viewed by 1283
Abstract
Euphorbia seguieriana ssp. seguieriana Necker (ES) and Euphorbia cyparissias (EC) with a habitat in the Deliblato Sands were the subject of this examination. The latexes of these so far insufficiently investigated species of the Euphorbia genus are used in traditional medicine for the [...] Read more.
Euphorbia seguieriana ssp. seguieriana Necker (ES) and Euphorbia cyparissias (EC) with a habitat in the Deliblato Sands were the subject of this examination. The latexes of these so far insufficiently investigated species of the Euphorbia genus are used in traditional medicine for the treatment of wounds and warts on the skin. To determine their chemical composition, non-targeted screening of the latexes’ chloroform extracts was performed using liquid chromatography coupled with quadrupole time-of-flight mass spectrometry employing an electrospray ionization source (LC-ESI QTOF MS). The analysis of the obtained results showed that the latexes of ES and EC represent rich sources of diterpenes, tentatively identified as jatrophanes, ingenanes, tiglianes, myrsinanes, premyrsinanes, and others. Examination of the anticancer activity of the ES and EC latex extracts showed that both extracts significantly inhibited the growth of the non-small cell lung carcinoma NCI-H460 and glioblastoma U87 cell lines as well as of their corresponding multi-drug resistant (MDR) cell lines, NCI-H460/R and U87-TxR. The obtained results also revealed that the ES and EC extracts inhibited the function of P-glycoprotein (P-gp) in MDR cancer cells, whose overexpression is one of the main mechanisms underlying MDR. Full article
Show Figures

Figure 1

Figure 1
<p>Total ion chromatogram of the chloroform extract of the latex of <span class="html-italic">E. seguieriana</span> ssp. <span class="html-italic">seguieriana</span> Necker (ES).</p>
Full article ">Figure 2
<p>Total ion chromatogram of the chloroform extract of the latex of <span class="html-italic">E. cyparissias</span> (EC).</p>
Full article ">Figure 3
<p>Flow cytometric profiles of Rho123 accumulation in NCI-H460/R (<b>a</b>) and U87-TxR (<b>b</b>) cells untreated and treated with 20 µg mL<sup>−1</sup> of the ES and EC extracts. Sensitive NCI-H460 and U87 cells were used as a positive control for Rho 123 accumulation. Two independent experiments were performed (a minimum of 10,000 events were collected for each experimental sample).</p>
Full article ">
12 pages, 1917 KiB  
Article
Norsesquiterpenes from the Latex of Euphorbia dentata and Their Chemical Defense Mechanisms against Helicoverpa armigera
by Tong An, Dongxu Cao, Yangyang Zhang, Xiamei Han, Zhiguo Yu and Zhixiang Liu
Molecules 2023, 28(23), 7681; https://doi.org/10.3390/molecules28237681 - 21 Nov 2023
Viewed by 849
Abstract
Euphorbia dentata (Euphorbiaceae), an invasive weed, is rarely eaten by herbivorous insects and could secrete a large amount of white latex, causing a serious threat to local natural vegetation, agricultural production and human health. In order to prevent this plant from causing more [...] Read more.
Euphorbia dentata (Euphorbiaceae), an invasive weed, is rarely eaten by herbivorous insects and could secrete a large amount of white latex, causing a serious threat to local natural vegetation, agricultural production and human health. In order to prevent this plant from causing more negative effects on humans, it is necessary to understand and utilize the chemical relationships between the latex of E. dentata and herbivorous insects. In this study, three new norsesquiterpenes (13), together with seven known analogues (410), were isolated and identified from the latex of E. dentata. All norsesquiterpenes (110) showed antifeedant and growth-inhibitory effects on H. armigera with varying levels, especially compounds 1 and 2. In addition, the action mechanisms of active compounds (13) were revealed by detoxifying enzyme (AchE, CarE, GST and MFO) activities and corresponding molecular docking analyses. Our findings provide a new idea for the development and utilization of the latex of E. dentata, as well as a potential application of norsesquiterpenes in botanical insecticides. Full article
(This article belongs to the Section Natural Products Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structures of compounds <b>1</b>–<b>10</b>.</p>
Full article ">Figure 2
<p>Key HMBC correlations of compounds <b>1</b>–<b>3</b>.</p>
Full article ">Figure 3
<p>ECD curves of compound <b>1</b>–<b>3</b>.</p>
Full article ">Figure 4
<p>The effects of compounds <b>1</b>–<b>3</b> (100 μg/mL) on the detoxifying enzymes of <span class="html-italic">H. armigera</span>: (<b>A</b>) the effects of compounds <b>1</b>–<b>3</b> on AchE; (<b>B</b>) the effects of compounds <b>1</b>–<b>3</b> on CarE; (<b>C</b>) the effects of compounds <b>1</b>–<b>3</b> on MFO; and (<b>D</b>) the effects of compounds <b>1</b>–<b>3</b> on GST. CK = Control check.</p>
Full article ">Figure 5
<p>Molecular docking analyses of compounds <b>1</b>–<b>3</b> with GST: (<b>A</b>) 3D diagram of the interactions between compound <b>1</b> and the active pocket of GST; (<b>B</b>) 2D diagram of the interactions between compound <b>1</b> and amino acids of GST; (<b>C</b>) 3D diagram of the interactions between compound <b>2</b> and the active pocket of GST; (<b>D</b>) 2D diagram of the interactions between compound <b>2</b> and amino acids of GST; (<b>E</b>) 3D diagram of the interactions between compound <b>3</b> and the active pocket of GST; and (<b>F</b>) 2D diagram of the interactions between compound <b>3</b> and amino acids of GST.</p>
Full article ">
29 pages, 14724 KiB  
Article
Isolation and Identification of 12-Deoxyphorbol Esters from Euphorbia resinifera Berg Latex: Targeted and Biased Non-Targeted Identification of 12-Deoxyphorbol Esters by UHPLC-HRMSE
by Abdellah Ezzanad, Carolina De los Reyes, Antonio J. Macías-Sánchez and Rosario Hernández-Galán
Plants 2023, 12(22), 3846; https://doi.org/10.3390/plants12223846 - 14 Nov 2023
Viewed by 938
Abstract
Diterpenes from the Euphorbia genus are known for their ability to regulate the protein kinase C (PKC) family, which mediates their ability to promote the proliferation of neural precursor cells (NPCs) or neuroblast differentiation into neurons. In this work, we describe the isolation [...] Read more.
Diterpenes from the Euphorbia genus are known for their ability to regulate the protein kinase C (PKC) family, which mediates their ability to promote the proliferation of neural precursor cells (NPCs) or neuroblast differentiation into neurons. In this work, we describe the isolation from E. resinifera Berg latex of fifteen 12-deoxyphorbol esters (115). A triester of 12-deoxy-16-hydroxyphorbol (4) and a 12-deoxyphorbol 13,20-diester (13) are described here for the first time. Additionally, detailed structural elucidation is provided for compounds 3, 5, 6, 14 and 15. The absolute configuration for compounds 3, 4, 6, 13, 14 and 15 was established by the comparison of their theoretical and experimental electronic circular dichroism (ECD) spectra. Access to the above-described collection of 12-deoxyphorbol derivatives, with several substitution patterns and attached acyl moieties, allowed for the study of their fragmentation patterns in the collision-induced dissociation of multiple ions, without precursor ion isolation mass spectra experiments (HRMSE), which, in turn, revealed a correlation between specific substitution patterns and the fragmentation pathways in their HRMSE spectra. In turn, this allowed for a targeted UHPLC-HRMSE analysis and a biased non-targeted UHPLC-HRMSE analysis of 12-deoxyphorbols in E. resinifera latex which yielded the detection and identification of four additional 12-deoxyphorbols not previously isolated in the initial column fractionation work. One of them, identified as 12-deoxy-16-hydroxyphorbol 20-acetate 13-phenylacetate 16-propionate (20), has not been described before. Full article
(This article belongs to the Topic Natural Compounds in Plants, 2nd Volume)
Show Figures

Figure 1

Figure 1
<p>12-Deoxyphorbol esters (<b>1</b>–<b>15</b>) and other diterpenes (<b>16</b>–<b>18</b>) isolated from <span class="html-italic">E. resinifera.</span> 12-Deoxyphorbol esters (<b>19</b>–<b>22</b>) detected and identified from a biased untargeted UHPLC-HRMS<sup>E</sup> analysis of <span class="html-italic">E. resinifera</span> latex.</p>
Full article ">Figure 2
<p>Comparison of HRMS<sup>E</sup> spectra (DIA) [<a href="#B26-plants-12-03846" class="html-bibr">26</a>] for (<b>a</b>) DPPI (<b>1</b>), (<b>b</b>) DPPT (<b>2</b>), (<b>c</b>) DPPBz (<b>3</b>) and (<b>d</b>) DPPU<sub>1</sub> (<b>19</b>) (data acquired in positive ionization with a ramp trap collision energy of the high-energy function set at 60–120 eV). See proposed fragmentation route for selected ions in <a href="#plants-12-03846-sch001" class="html-scheme">Scheme 1</a>, together with color key.</p>
Full article ">Figure 3
<p>Selected NOESY correlations (blue arrows) for DPPBz (<b>3</b>).</p>
Full article ">Figure 4
<p>Comparison of HRMS<sup>E</sup> spectra (DIA) [<a href="#B26-plants-12-03846" class="html-bibr">26</a>] for (<b>a</b>) AcDPPI (<b>4</b>), (<b>b</b>) AcDPPT (<b>5</b>), (<b>c</b>) AcDPPBz (<b>6</b>), (<b>d</b>) AcDPPU<sub>2</sub> (<b>20</b>) and (<b>e</b>) AcDPPU<sub>3</sub> (<b>21</b>), <span class="html-italic">m</span>/<span class="html-italic">z</span> range 240–415 (<span class="html-italic">m</span>/<span class="html-italic">z</span> range 240–680 in <a href="#app1-plants-12-03846" class="html-app">Figure S21</a>) (data acquired in positive ionization with a ramp trap collision energy of the high-energy function set at 60–120 eV). See proposed fragmentation route for selected ions in <a href="#plants-12-03846-sch002" class="html-scheme">Scheme 2</a>, together with color key.</p>
Full article ">Figure 5
<p>Experimental and calculated ECD spectra of (<b>a</b>) 12-deoxy-16-hydroxyphorbol 20-acetate-16-isobutyrate-13-phenylacetate (AcDPPI (<b>4</b>)); (<b>b</b>) 12-deoxy-16-hydroxyphorbol 20-acetate-13-phenylacetate-16-tigliate (AcDPPBz (<b>6</b>)).</p>
Full article ">Figure 6
<p>Comparison of HRMS<sup>E</sup> spectra (DIA) [<a href="#B26-plants-12-03846" class="html-bibr">26</a>] for (<b>a</b>) AcDPiPn (<b>13</b>) and (<b>b</b>) AcDPMeOP (<b>14</b>), <span class="html-italic">m</span>/<span class="html-italic">z</span> range 240–410 (<span class="html-italic">m</span>/<span class="html-italic">z</span> range 240–600 in <a href="#app1-plants-12-03846" class="html-app">Figure S28</a>) (data acquired in positive ionization with a ramp trap collision energy of the high-energy function set at 10–40 eV). See proposed fragmentation route for selected ions in <a href="#plants-12-03846-sch004" class="html-scheme">Scheme 4</a>, together with color key.</p>
Full article ">Figure 7
<p>Selected NOESY correlations (blue arrows) for AcDPiPn (<b>13</b>).</p>
Full article ">Figure 8
<p>Calculated and experimental ECD spectra of (<b>a</b>) 12-deoxyphorbol 20-acetate-13-isopentanoate (<b>13</b>); (<b>b</b>) 12-deoxyphorbol 20-acetate-13-p-methoxy-phenylacetate (<b>14</b>).</p>
Full article ">Figure 9
<p>Comparison of HRMS<sup>E</sup> spectra (DIA) [<a href="#B26-plants-12-03846" class="html-bibr">26</a>] for (<b>a</b>) diDPB (<b>15</b>) and (<b>b</b>) diDPU<sub>4</sub> (<b>22</b>), <span class="html-italic">m</span>/<span class="html-italic">z</span> range 240–470 (data acquired in positive ionization with a ramp trap collision energy of the high-energy function set at 10–40 eV). See proposed fragmentation route for selected ions in <a href="#plants-12-03846-sch005" class="html-scheme">Scheme 5</a>, together with color key.</p>
Full article ">Figure 10
<p>Selected NOESY correlations (blue arrows) for diDPB (<b>15</b>).</p>
Full article ">Figure 11
<p>Calculated and experimental ECD spectra of 12,20-dideoxyphorbol 13-isobutyrate (<b>15</b>).</p>
Full article ">Scheme 1
<p>Proposed fragmentation route for selected ions on HRMS<sup>E</sup> spectra (DIA, high-energy function) for 12-deoxy-16-hydroxyphorbol 13,16-diacyl derivatives (<b>1</b>–<b>3</b>, <b>19</b>, group A compounds; see <a href="#sec2dot2-plants-12-03846" class="html-sec">Section 2.2</a>). For each ion, nominal mass and elemental composition is presented; see a more detailed interpretation of the fragmentation route in <a href="#app1-plants-12-03846" class="html-app">Figure S20</a>. In red, common daughter ions with group B compounds (see <a href="#sec2dot2-plants-12-03846" class="html-sec">Section 2.2</a> and <a href="#plants-12-03846-sch002" class="html-scheme">Scheme 2</a>). In green, highlighted ions in <a href="#plants-12-03846-f002" class="html-fig">Figure 2</a>c.</p>
Full article ">Scheme 2
<p>Proposed fragmentation route for selected ions on HRMS<sup>E</sup> spectra (DIA, high-energy function) for 16-hydroxy-12-deoxyphorbol 20-acetate-13,16-diacyl derivatives (<b>4</b>–<b>6</b>, <b>20</b>–<b>21</b>, group B compounds; see <a href="#sec2dot2-plants-12-03846" class="html-sec">Section 2.2</a>). For each ion, nominal mass and elemental composition is presented; see a more detailed interpretation of the fragmentation route in <a href="#app1-plants-12-03846" class="html-app">Figure S22</a>. In red, common daughter ions with group A compounds (see <a href="#sec2dot2-plants-12-03846" class="html-sec">Section 2.2</a> and <a href="#plants-12-03846-sch001" class="html-scheme">Scheme 1</a>). In blue, highlighted ions in <a href="#plants-12-03846-f004" class="html-fig">Figure 4</a>a.</p>
Full article ">Scheme 3
<p>Proposed fragmentation route for selected ions on HRMS<sup>E</sup> spectra (DIA, high-energy function) for 12-deoxyphorbol 13-acyl derivatives (<b>7</b>–<b>9</b>, group C compounds; see <a href="#sec2dot2-plants-12-03846" class="html-sec">Section 2.2</a>). For each ion, nominal mass and elemental composition is presented; see a more detailed interpretation of the fragmentation route in <a href="#app1-plants-12-03846" class="html-app">Figure S25</a>. In green, common daughter ions with group D compounds (see <a href="#sec2dot2-plants-12-03846" class="html-sec">Section 2.2</a> and <a href="#plants-12-03846-sch004" class="html-scheme">Scheme 4</a>). In deep red, highlighted ions in <a href="#app1-plants-12-03846" class="html-app">Figure S24a</a>.</p>
Full article ">Scheme 4
<p>Proposed fragmentation route for selected ions on HRMS<sup>E</sup> spectra (DIA, high-energy function) for 12-deoxyphorbol 20-acetate-13-acyl derivatives (<b>10</b>–<b>14</b>, group D compounds; see <a href="#sec2dot2-plants-12-03846" class="html-sec">Section 2.2</a>). For each ion, nominal mass and elemental composition is presented; see a more detailed interpretation of the fragmentation route in <a href="#app1-plants-12-03846" class="html-app">Figure S29</a>. In green, common daughter ions with group C compounds (see <a href="#sec2dot2-plants-12-03846" class="html-sec">Section 2.2</a> and <a href="#plants-12-03846-sch003" class="html-scheme">Scheme 3</a>). In purple, highlighted ions in <a href="#plants-12-03846-f006" class="html-fig">Figure 6</a>a.</p>
Full article ">Scheme 5
<p>Proposed fragmentation route for selected ions on HRMS<sup>E</sup> spectra (DIA, high-energy function) for 12,20-dideoxyphorbol 13-acyl derivatives (<b>15</b> and <b>22</b>, group E compounds; see <a href="#sec2dot2-plants-12-03846" class="html-sec">Section 2.2</a>). In brown, highlighted ions in <a href="#plants-12-03846-f009" class="html-fig">Figure 9</a>a. For each ion, nominal mass and elemental composition is presented; see a more detailed interpretation of the fragmentation route in <a href="#app1-plants-12-03846" class="html-app">Figure S30</a>.</p>
Full article ">
16 pages, 1325 KiB  
Article
Indonesian Euphorbiaceae: Ethnobotanical Survey, In Vitro Antibacterial, Antitumour Screening and Phytochemical Analysis of Euphorbia atoto
by Dyke Gita Wirasisya, Annamária Kincses, Lívia Vidács, Nikoletta Szemerédi, Gabriella Spengler, Anita Barta, I Gde Mertha and Judit Hohmann
Plants 2023, 12(22), 3836; https://doi.org/10.3390/plants12223836 - 13 Nov 2023
Cited by 1 | Viewed by 1146
Abstract
Indonesia is among the countries with the most significant biodiversity globally. Jamu, the traditional medicine of Indonesia, predominantly uses herbal materials and is an integral component of the Indonesian healthcare system. The present study reviewed the ethnobotanical data of seven Indonesian Euphorbiaceae [...] Read more.
Indonesia is among the countries with the most significant biodiversity globally. Jamu, the traditional medicine of Indonesia, predominantly uses herbal materials and is an integral component of the Indonesian healthcare system. The present study reviewed the ethnobotanical data of seven Indonesian Euphorbiaceae species, namely Euphorbia atoto, E. hypericifolia, Homalanthus giganteus, Macaranga tanarius, Mallotus mollissimus, M. rufidulus, and Shirakiopsis indica, based on the RISTOJA database and other literature sources. An antimicrobial screening of the plant extracts was performed in 15 microorganisms using the disk diffusion and broth microdilution methods, and the antiproliferative effects were examined in drug-sensitive Colo 205 and resistant Colo 320 cells by the MTT assay. The antimicrobial testing showed a high potency of M. tanarius, H. giganteus, M. rufidulus, S. indica, and E. atoto extracts (MIC = 12.5–500 µg/mL) against different bacteria. In the antitumour screening, remarkable activities (IC50 0.23–2.60 µg/mL) were demonstrated for the extracts of H. giganteus, M. rufidulus, S. indica, and E. atoto against Colo 205 cells. The n-hexane extract of E. atoto, with an IC50 value of 0.24 ± 0.06 µg/mL (Colo 205), was subjected to multistep chromatographic separation, and 24-methylene-cycloartan-3β-ol, jolkinolide E, tetra-tert-butyl-diphenyl ether, α-tocopherol, and β-sitosterol were isolated. Full article
Show Figures

Figure 1

Figure 1
<p>Frequencies of the use of medical plants parts (%).</p>
Full article ">Figure 2
<p>Structures of compounds 1–3 isolated from <span class="html-italic">E. atoto</span>.</p>
Full article ">Figure 3
<p>Map of Indonesia and the areas from which plants were collected.</p>
Full article ">
Back to TopTop