banner
News center
Our well-established business has loyal customers across the globe.

Egyptian mandarin peel oil's anti

Jul 17, 2023

Scientific Reports volume 13, Article number: 14192 (2023) Cite this article

Metrics details

The current study investigated the scabicidal potential of Egyptian mandarin peel oil (Citrus reticulata Blanco, F. Rutaceae) against sarcoptic mange-in-rabbits. Analysis of the oil's GC–MS identified a total of 20 compounds, accounting for 98.91% of all compounds found. Mandarin peel oil topical application improved all signs of infection, causing a scabicidal effect three days later, whereas in vitro application caused complete mite mortality one day later. In comparison to ivermectin, histopathological analysis showed that the epidermis' inflammatory-infiltration/hyperkeratosis-had disappeared. In addition to TIMP-1, the results of the mRNA gene expression analysis showed upregulation of I-CAM-1-and-KGF and downregulation of ILs-1, 6, 10, VEGF, MMP-9, and MCP-1. The scabies network was constructed and subjected to a comprehensive bioinformatic evaluation. TNF-, IL-1B, and IL-6, the top three hub protein-coding genes, have been identified as key therapeutic targets for scabies. From molecular docking data, compounds 15 and 16 acquired sufficient affinity towards the three screened proteins, particularly both possessing higher affinity towards the IL-6 receptor. Interestingly, it achieved a higher binding energy score than the ligand of the docked protein rather than displaying proper binding interactions like those of the ligand. Meanwhile, geraniol (15) showed the highest affinity towards the GST protein, suggesting its contribution to the acaricidal effect of the extract. The subsequent, MD simulations revealed that geraniol can achieve stable binding inside the binding site of both GST and IL-6. Our findings collectively revealed the scabicidal ability of mandarin peel extract for the first time, paving the way for an efficient, economical, and environmentally friendly herbal alternative for treating rabbits with Sarcoptes mange.

Sarcoptic mange (Sarcoptes scabiei) is a serious infectious disease that invades humans and animals all over the world1. The mites are highly adapted to contact with their host as contagious, burrowing, and obligate parasites. Sarcoptic mange Grower pig production is negatively impacted by adult female mites; because they mate on the skin's surface, burrow into the skin, lay eggs, and cause irritations that can lead to bleeding, reduced feeding and development, chronic stress, and decreased welfare2,3. The clinical picture represents chronic hyperkeratotic, which is characterized by the presence of aural crusts and many mites on the animal4. Similar to people, rabbits are susceptible to Sarcoptes infection, or mange, which reduces production and causes economic losses for rabbits, especially in the absence of effective treatment5. Therapy options include the systemic treatment of macrocyclic lactones, local administration of amitraz or pyrethroids, or both6,7. Despite their long history of effectiveness in treating mange, their extensive use has led to a decline in effectiveness because of the emergence of drug resistance. Thus, it is crucial to create novel scabicides that are both efficient and secure in order to treat and control mammalian scabies6.

In rabbits, goats, and pigs, several essential oils derived from Citrus limon, Lavandula angustifolia, Citrus aurantium amara, Pelargonium asperum, Melaleuca alternifolia, Syzygium aromaticum, Eucalyptus radiata, Leptospermum scoparium, Juniperus oxycedrus, Cryptomeria japonica, and Cymbopogon martini, were put to the test in real time against S. scabiei8,9,10,11,12. Essential oils are typically favoured over chemical acaricides since they are less harmful to animals and have a shorter environmental persistence. Also, the complex chemistry of essential oils is known to considerably impede the emergence of drug resistance against these chemicals13. Yet, because essential oils consist of a complex mixture of components, it might be challenging to attribute an essential oil's acaricidal properties to a specific ingredient or combination of compounds14. Skin irritation is yet another potential drawback that has been reported in humans15.

Some of the most coveted Citrus fruits for fresh consumption are mandarins, C. reticulata16. The more frequent name for them is "mandarin," but they are also occasionally called "tangerines." The Mandarin species includes a number of cultivars and hybrids16. Popularly grown varieties include C. unshiu Marcovitch (also known as Unshiu mikan in Japanese), C. nobilis Loureiro (also known as king mandarins), C. deliciosa Tenore (also known as Mediterranean mandarins), and C. reticulata Blanco (common mandarins)16,17. Mandarins are one of the main Citrus fruits grown in many countries such as China, Brazil, USA, India, Mexico, Spain, etc. The fruits have a great commercial worth for their essential oils and other fragrant compounds, even though they are primarily used to make pastries18. A lot of beverages, candies, cookies, and desserts use Citrus flavours19, while the peels of C. reticulata are used to flavour alcohol19. Citrus reticulata EO shown an anti-proliferative activity against rat pulmonary fibrosis produced by bleomycin (BLM) and protective properties against human embryonic lung fibroblasts (HELFs). The method is believed to involve correcting the imbalance between oxidation and antioxidation, lowering collagen deposition and fibrosis, and down-regulating lung tissue expressions of connective tissue growth factor (CTGF) and mRNA20. Due to its high d-limonene concentration21, C. reticulata EO demonstrated a moderate level of radical scavenging action22. Mandarin oil is well known for its broad spectrum antibacterial and antifungal actions. It inhibits the growth of several bacteria including Escherichia coli, Bacillus subtilis, Pseudomonas aeruginosa, and Staphylococcus aureus22,23, as well as several fungi including Penicillium italicum, P. chrysogenum, P. digitatum, Aspergillus niger,-A. flavus, Alternaria alternata, Curvularia lunata, Rhizoctonia solani, Fusarium oxysporum, and-Helminthosporium oryzae23,24,25,26.

The GC–MS profiling of mandarin peel oil has been used in the current study. Additionally, for the first time, through-in vitro,-in vivo,-histopathology,-mRNA-expression, and network/in silico analysis, the extract's scabicidal potential against-Sarcoptic-mange-in-rabbits has been investigated, allowing for the incorporation of natural candidates to proper and secure management of infectious diseases. The present investigation's framework is shown in Fig. 1.

General outflow of the study.

Plant materials and experiments were conducted in accordance with relevant institutional, national, and international guidelines. The study took place according to the ethical committee's permission number of 9/5/2022 at Deraya College. It was done in accordance with the National Institute of Health's guidelines for the care and use of laboratory animals and ARRIVE guidelines27.

In January 2021, C. reticulata cultivated fruits were harvested from a house garden on Atia Street in Beni-Suef, Egypt. A voucher specimen (2021-BuPD-88) was deposited at Pharmacognosy-Department, Faculty-of-Pharmacy, Beni-Suef-University, Egypt.

Using the Clevenger apparatus, the fresh peels (0.5 kg) were hydrodistillated for two hours at 75 °C. The oil was gathered, dried over anhydrous sodium sulphate, and kept in airtight amber glass vials at 4 °C for storage. On the basis of the plant material's fresh weight, the yield (v/w%) was computed28,29.

Gas chromatography-mass spectrometry (GC/MS) was used to perform chromatographic analysis on the oil recovered from peels28,30. The GC–MS apparatus combines a thermal mass spectrometer detector (ISQ single quadrupole mass spectrometry) with a TRACE GC ultra-high performance gas chromatograph (THERMO Scientific Corp., USA). A TR-5 MS column (30 m × 0.32 mm i.d., 0.25 mm film thickness) was installed in the GC–MS system. For the analyses, He-lium was used as the carrier gas, and the split ratio was set at 1:10 using the following temperature program: 60 C for 1 min, followed by 4.0 C/min to 240 C and a 1-min hold. At 210 °C, the injector and detector were maintained. One-liter samples of the mixes were always administered as diluted samples (1:10 hexane, v/v). By using a spectral range of m/z 40–450 and electron ionization (EI) at 70 eV, mass spectra were produced. Using AMDIS software (www.amdis.net), the chemical components of the essential oil were deconvoluted and identified by their retention indices (relative to n-alkanes C8-C22), mass spectra matching to genuine standards, and retention times (when available). (NIST Standard Reference Database, 78 Version 5.10) Wiley spectral library collection28,31,32.

The reaction with a defined amount of exogenously provided hydrogen peroxide (H2O2) was used to determine H2O2 scavenging activity that reflects the anti-oxidative capacity of the peel oil. Colorimetric analysis was used to estimate the residual H2O233. In brief, 20 µl of the sample was mixed with 500 µl of H2O2 and incubated at 37 °C for 10 min. 500 l of the enzyme/3, 5-dichloro-2-hydroxyl-benzensulfonate solution were then added, and it was incubated at 37 °C for 5 min. The colored product's intensity was quantified colorimetrically at-510-nm. A positive control was ascorbic acid. By comparing the test results to those of the control group, the percentage-of H2O2-scavenging activity-was calculated and applying the following formula:

IC50 of each sample was calculated after performing the assay at eight different concentrations : (1000, 750, 500, 375, 250, 187.5, 125 and 0 µg/mL) using Graph pad prism 7 software.

The scavenging activity of superoxide anion was measured34. In a Tris-HCL solution (16 mM, pH 8.0) containing 90 l of NBT (0.3 mM), 90 l of NADH (0.936 mM), 0.1 ml of peel oil (125, 250, 500, and 1000 g/mL), and 0.8 ml of Tris-HCl buffer, superoxide anion radicals were produced (16 mM, pH 8.0). After adding 0.1 ml of PMS solution (0.12 mM) to the mixture, the reaction was started. The mixture was then incubated at 25 °C for 5 min, during which time the absorbance was measured at 560 nm. Ascorbic acid was used as a model substance. Using the formula below, the percentage inhibition was calculated by comparing the test results to those of the control:

IC50 was estimated by doing the test at four different concentrations and using the GraphPad Prism 7 software.

Adult-mites-were collected from rabbits that were-naturally-infected, Deraya University, Minia, Egypt's Animal House. Scraped from the borders of the lesions, the infected skin samples were then shifted-to-petri plates and-incubated within a biochemical-oxygen-demand (BOD) for an incubator for 30 min at 35 °C.

A petri dish containing mites was filled with 2 ml of diluting extract (20%), along with the plates were then incubated-in-BOD. Reaction observations were made-at 1, 12, and 24 h after application. Petri plates were incubated at an ambient temperature of 25 °C and with a relative moisture of 75%, with a 5% ivermectin (1 cm3/l) group as the positive-control-and distilled water as the-negative-control. By stimulating the mites with a needle, the death of the mite was confirmed; the mite was deemed dead if it showed no response.

The study took place on male adult rabbits for 4 weeks (weighing 2.8–3.2 kg) that were infected. The animals' ears showed clinical indicators of mange infection, such as hyperkeratinization, inflammation, redness, itching, and irritability. Microscopic mite identification in skin scrapings further corroborated this. Four groups of five rabbits each were made up of twenty animals, as follows: Five rabbits made up the normal group, the paraffin oil-positive-control-group. The-ivermectin-treated-group (5%-ivermectin). The peel-oil-group (20%-peel-oil in paraffin-oil). Paraffin oil, which is a-mineral-oil, was reportedly chosen as a diluent for the peel oil because it has little impact-on-mites35. Each group were kept in a separate cage, and each group received treatment by dipping the infected ears once daily. Steel hoppers were used to feed all of the rabbits, and water was available at all times. The rabbits were observed every two days to assess their clinical recovery. The goal was to find any signs of improvement in the lesions, such as the absence of irritation and redness, cutaneous smoothing, the start of the development of hair from the infection, and the cessation of scab development10. Skin scrapings from each rabbit's sick and healed areas were taken every three days, and Throughout the course of the therapy, they were microscopically investigated to check for sarcoptic mites with a LEICA, DM1000 microscope with a digital camera (LEICA, EC3, Germany)10.

Tissue samples were collected at zero and three weeks following the start of the course of therapy via 20% peeling oil as well as ivermectin from healthy and infected ears. Following that, samples were dried in ethyl alcohols of increasing strength, sterilized via xylene, infused with paraffin that had been melted at 55–60 °C, and finally inserted into paraffin wax. The samples were then preserved in 10% buffered formalin. Deparaffinized, rehydrated, and stained with hematoxylin and eosin (H & E), “3–5 m thick” tissue sections were examined using-a-light-electron-microscope36.

Using-a-digital-homogenizer-(Branson-Digital-Homogenizer®,-Danbury,-CT,-USA), 100 mg of the tissues under investigation were homogenised in 1 ml of TRIzolTM RNA Extraction Reagent (Amresco, Solon, OH, USA). RNA extraction from the biopsy sample was done in accordance with the manufacturer's instructions. RevertAid H-minus First Strand cDNA Synthesis Kits (#K1632, Thermo Science Fermentas, St. Leon-Ro, Germany) were used to create cDNA from the extracted RNA for comparable amounts of total RNA in all samples. The-qRT-PCR-was carried out on-the-Applied-Biosystems Step One Plus system using the cDNA as a template. The primers were created using the NCBI primer blast software and were produced by Invitrogen. Using the GAPDH gene as a housekeeping gene, data were analyzed using the 2CT approach31. Table 1 lists the primer sequences that were employed.

Using Cytoscape 3.9.1 software (https://www.cytoscape.org/)37 and by lunching STRING disease query tool incorporated in it which retrieves network for the top human proteins associated with the queried disease from a weekly updated web source of diseases database (https://string-db.org/)38 Scabies was chosen as the search term, and "Human sapiens" was chosen as the species type. The confidence score was set to 0.4, and the default settings for the remaining parameters were used to create the PPI network39.

The plugin for cytoHubba the hub genes are identified using ranking techniques such as degree, edge percolated component (EPC), maximum neighbourhood component (MNC), the density of maximum neighborhood component (DMNC), and maximal clique centrality (MCC), as well as bottleneck, eccentricity, closeness, radiality, betweenness, stress, and clustering coefficient. Cytoscape is regarded as a useful exploration interface for the most significant nodes in PPI networks40,41.

We employed a freely accessible bioinformatics web tool in the current investigation (ShinyGO v0.76.3). Using the many bioinformatics databases accessible, it is possible to perform both gene ontology enrichment analysis and pathway enrichment analysis. ShinyGO was used to perform the gene ontology and enrichment analysis on the 16 genes to determine the cellular elements, molecular functions, and biological processes that were impacted by this set of genes. ShinyGO retrieves comprehensive descriptions of biological signal transduction pathways from numerous databases42.

The methodologies of molecular docking intend to predict the best binding orientation of a ligand to a receptor. It proposes several suitable poses of the ligand within the active or docking site of a receptor molecule.in this study, twenty compounds that were identified underwent an in silico study by using screening for three different important-protein-targets that are heavily involved in the scabies infection process, as well as screening for potential targets at the mite itself as an acaricidal effect, in an attempt to get deep inside the mechanistic anti-scabietic effect of orange oil. The chosen targets include IL-1, which is highly effective in stimulating T cells with regulatory functions, and IL-6, which is involved in the formation of Th17 lymphocytes and the release of IL-1743. These cy-tokines have been identified as one of the primary molecules responsible for allergic Th2-type inflammation in the immunological response to scabies, along with TNF-, which is significant in alternative macrophage activation44. GSH, which is linked to the scabies defense system, takes part in a variety of processes crucial to the preservation of cells from oxygen and free radical oxidative damage45, Its distinctive anti-oxidant action makes it a potential target for the oil's acaricidal impact46. In our docking investigation, we validated the ligand and visualized the many docked poses using the computer programme MOE 2019.010. TNF- complexed with its ligand (PDB ID code: 2AZ5) is the last one, and GST is the other protein target of the mite delta class. The first protein target is (IL-1), depicted by the protein's PDB ID code of 6Y8M in co-crystallization with IL-6, as reflected by PDB ID code 1ALU, and its inhibiting ligand SX2 (a-bromo-amido-pyridine-derivative)47 represented by proteins (PDB ID code: 3EIN), the selected targets were acquired via the web from the Database of Proteins (http://www.rcsb.org/pdb).

The MD simulations were carried out using NAMD 3.0.0. software48,49. The Charmm-36 force field is implemented in this piece of software. The protein structure was examined for missing hydrogens, the protonation states of the amino acid residues were set (pH = 7.4), and the co-crystallized water molecules were removed using the QwikMD toolkit of the VMD software. The entire assembly was then packed into a 20 solvent buffer containing 0.15 M Na + and Cl- ions in an orthorhombic box of TIP3P water. After 5 ns of equilibration, the systems were subjected to an energy minimization protocol. Force Field Toolkit (ffTK), a plugin for the VMD software, was used to determine ligand properties and topologies. After the parameters and topology files were prepared, they were imported into VMD so that the protein–ligand complexes could be read accurately, and the simulations could be run.

The data were tabulated using the statistical programme-GraphPad-Prism-version-9 (GraphPad,-La-Jolla,-CA,-USA). To evaluate statistical differences between the groups, the ANOVA test was performed, followed by the-Bonferroni-post-hoc-test-for multiple-comparisons. The threshold for statistical significance is a p-value of 0.05 or less.

Egyptian C. reticulata peels gave 2.6% v/w volatile oil fresh weight, being colourless with a characteristic odor, lighter than water, clear, transparent, and not viscous at room temperature as well as at 4 °C. GC–MS analysis was used to identify a total of 20 compounds, accounting for 98.91% of all compounds found (Table 2, Figs. 2, 3). The identified compounds 1–20 belonged to different chemical classes, including monoterpene, phenylpropene, fatty alcohol, and sesquiterpene (Table 2, Fig. 3). where monoterpenes represented 92.16% of the total identified compounds, followed by phenylpropene (3.01%), fatty alcohol (2.36%), and sesquiterpene (1.38%) (Table 2). Fourteen monoterpenes compounds (92.16%) were identified; ranging from cyclic hydrocarbon (D-limonene 4, γ-terpinene 6, 73.32%) which represented the major oil fraction, to oxygenated cyclic hydrocarbon ((-)-isomenthone 10, terpinen-4-ol 11, (-)-carvone 14, 3.32%), and oxygenated acyclic hydrocarbon (linalool 8, citronellol 13, geraniol 15, 8.78%), acyclic hydrocarbon (α-myrcene 3, α-ocimene 5, 2.84%), bicyclic hydrocarbon (α-pinene 1, sabinene 2, 3.45%), to oxygenated bicyclic hydrocarbon (camphor 9, 0.45%) (Table 2, Fig. 3). Also, phenylpropene class (3.01%) contained 2.70 and 0.31% of estragole 12, and anethole 17, respectively. The detected fatty alcohol class contained only 1-octanol 7 and 1-decanol 16, representing 2.36% (Table 2, Fig. 3). On the other hand, three sesquiterpene compounds (1.38%) were identified, varying from a bicyclic hydrocarbon (caryophyllene 19, ( +)-valencene 20, 1.00%), to a tricyclic hydrocarbon (α-copaene 18, 0.38%) (Table 2, Fig. 3).

GC/MS spectrum for Citrus reticulata peels oil.

Structures of identified compounds, using GC/MS analysis, from Citrus reticulata oil isolated from peels.

According to the literature, the chemical composition of essential oils varies depending on the age of the plant, harvesting time, geographical location, and environmental conditions50. The Indian C. reticulata peels volatile oil differently from the Egyptian, having 80 compounds, where monoterpene (63.80%), represents mainly limonene (50.42%), myrecene (3.03%), and α-terpineol (1.19%), while sesquiterpene (12.98%) represents mainly α-copaene (1.49%), β-copaene (1.30%), and α-humulene (1.23%). The Indian C. reticulata oil is characterised by its high content of fatty acids (8.73%) and aldehyde content (7.08%), mainly n-hexadecanoic acid (5.65%) α-sinensal (3.14%)51. The essential oil isolated from fully matured, ripened Indian fruit peels of C. reticulata, on the other hand, contained 37 different components (99%). The primary ingredients included limonene (46.7%), geranial (19.0%), neral (14.5%), geranyl acetate (3.9%), geraniol (3.5%), -caryophyllene (2.6%), nerol (2.3%), neryl acetate (1.1%), and others26.

The essential oil constituents reported in C. reticulata grown in Burundi contained 58 constituents52. The most prevalent chemical category was monoterpene hydrocarbons (94.7%). Limonene accounted for 84.8% of the total composition, with -terpinene (5.4%), myrcene (2.2%), and -pinene (1.1%) following. Germacrene D and valencene were the primary components of the sesquiterpene hydrocarbons, which made up only 0.2% of the total composition. Compounds containing oxygen from different chemical groups made up 2.3%52. The two main chemical groupings were terpene alcohols (0.7%) and aliphatic aldehydes (0.7%). Linalool (0.7%), octanal (0.5%), and decanal (0.2%) made up the bulk of the mixture. In concentrations of 0.1%, octyl acetate, α-sinensal, decanol, and perillaldehyde were present. Thymol, α-sinensal, methyl thymol, as well as the acetate esters bornyl, ɣ-terpinyl, geranyl, citronellyl, and decyl acetates, were all found at concentrations of less than 0.05%52.

The essential oil constituents of C. reticulata cultivated in Algeria were reported to contain 24 constituents. Monoterpene hydrocarbons accounted for the most abundant chemical group (89.56%). The main components were limonene (67.04%), -terpinene (15.50%), and -pinene (2.75%). Sesquiterpene hydrocarbons accounted for a minor quantity (3.26%), where l-caryophyllene was the main constituent53.

The literature review on essential oil components in C. reticulata cultivated in different regions corroborates some commonalities. Consequently, limonene, a hydrocarbon monoterpene, is invariably the most common ingredient in essential oils made from Citrus peels, making up typically between 60 and 70 percent of the oil. However, limonene can show lower levels, as in fully matured, ripened Indian fruit peels of C. reticulata, in which it can decrease to 46%26. Also prevalent are the following substances: monoterpenes, which typically account for less than 15%, γ-terpinene, myrcene, and α-pinene, which can reach an abundance of 6.0%, 3.6%, and 1.5%, respectively.

Non-terpenoid or terpenoid compounds (aldehydes, ketones, esters, fatty acids, and phenyl) are reported to be present (1–10%) or absent according to the cultivated region, but there are no commonalities among studies reporting these compounds to have an impact on the essential oil activity or not. Sesquiterpene hydrocarbons are the most varied group of all known chemicals, and this is true for the majority of species. The most prevalent groupings also frequently include oxygenated monoterpene alcohols and monoterpene hydrocarbons.

This study looked into the antioxidant activity of mandarin peel oil as-a-scavenger-potential-against-H2O2. The outcomes showed that mandarin peel oil had H2O2 scavenging capacity at a concentration of 1000 µg/mL increased in an exceedingly dose-dependent manner, compared with a standard-ascorbic-acid (IC50 = 139.2 µg/mL). This means that the higher the concentration of the oil, the more effectively it scavenges the H2O2 radicals (Fig. 4A).

The H2O2 scavenging activity of both the mandarin peel oil and the standard increased in a concentration-dependent manner (Fig. 4A). Interestingly, at a concentration of 1000 μg/mL, mandarin peel oil exhibited the highest superoxide removal action, with an IC50 value of 176.2 μg/mL (Fig. 4B). This indicates that the oil was more effective at scavenging the superoxide radicals than the standard, ascorbic acid.

The SOD activity of both the mandarin peel oil and the standard also increased in a concentration-dependent manner (Fig. 4B). Interestingly, at a concentration of 1000 µg/mL, mandarin peel oil exhibited the highest superoxide removal action, with an IC50 value of 176.2 µg/mL (Fig. 4B). This indicates that the oil was more effective at scavenging the superoxide radicals than the standard, ascorbic acid.

Overall, these findings indicate that mandarin peel oil is a potent antioxidant with a high ability to scavenge H2O2 and superoxide.

According to in vitro data, the mandarin peel oil (20%) achieved a remarkable acaricidal impact. The mites displayed a slow movement that began at one-hour post-application-(PA) and terminated at 24 PA via 99 percent death rates, as determined by microscopic analysis.

Sarcoptic mange, some chronic lesions, and scabs were visible on and inside the ears of rabbits infected with Sarcoptes scabiei. These animals suffered from itching, congestion, scratching, and anorexia, while those treated with mandarin peel oil (20% peel oil in paraffin oil) showed a gradual improvement in clinical symptoms from the fourth day PA through the experiment's conclusion (three weeks-PA). The lack of irritation, bleeding, scale formation, restlessness, and the appearance of smooth skin and new hair growth were signs of the recovery54. The ivermectin-treated animals, on the other hand, gradually improved but did not completely eradicate the condition from the seventh day PA till the investigation's conclusion (Fig. 5).

Inspection of mange-infected rabbits under a microscope, (A) control group (paraffin oil), (B) mandarin peels oil group (20% peels oil in paraffin oil), (C) ivermectin group (5% ivermectin).

On the fifth day PA, each the peels oil along with ivermectin groups of infected animals' skin scrapings contained dead mites. By the time the animals were checked once more, on day 10, the dead mites had totally disappeared.

The normal skin's epidermis and dermis were clearly visible in the histological analyses of the normal group. The stratum corneum and stratum granulosum made up the epidermis, and the reticular layer, hair follicles, sebaceous glands, and sweat glands are visible in the dermis (Fig. 6A). Skin samples from the control group, on the other hand, displayed a changed histology, which is usual for this parasite infection55. Skin erosion could be seen as a result of the stratified squamous epithelial sloughing, hyperkeratosis, akanthosis, and folded, seemingly injured skin. Moreover, the epidermis, inflammatory cellular infiltration, and hypergranulating dermis all displayed necrotic debris mixed with various stages of mites (Fig. 6B).

Microscopical-examination-of-skin from-different-groups of-animals, (A) normal-architecture-of the-skin: e; epidermis, d; dermis, h.f.; hair-follicles, (B) control-group-showing-skin-damage-with-hyperkeratosis (red arrows), mites-remnants-embedded-in the-skin (blue arrows), hypergranulation-of-dermis (green-arrows), severe-akanthosis-with-cellular-infiltration (black-arrows), (C) mandarin-peels-oil group showing-restoration of-normal-architecture, with-mild-infiltration (red-arrow), healthy-sebaceous-glands (yellow-arrow) and hair-follicles (black-arrows), (D) ivermectin-group showing-moderate damage-with-hyperkeratosis (red-arrow), mature-mites with-eggs-remnants-embedded in-the-dermis (black-arrow) surrounded-by-cellular-infiltration (green-arrow), and some-sebaceous-adenitis (yellow-arrows).

Also, biopsy samples from the animals given mandarin peel oil (20% peel oil in paraffin oil) demonstrated a slight cellular infiltration, a lack of mites, an increase in the number of hair growth follicles, and the appearance of normal sebaceous glands are signs of progress in the skin's surface layers (Fig. 6C). The skin condition was improved in the group that had ivermectin treatment, where only a few inflammatory cells and hyperkeratosis were seen. In the outermost layer of skin, the mites' remains could be seen embedded. Cellular filtration and sebaceous adenitis were present in certain regions (Fig. 6D).

The-pro-inflammatory-cytokines-(IL-1β,-IL-6),-the-pleiotropic-cytokine-(IL-10)-and-the-monocyte chemoattractant-protein-1-(MCP-1) were all downregulated in the animals treated with mandarin peel oil (20% peel oil in paraffin oil), according to the results of q-PCR. On the other hand, 2–sevenfold increases in ICAM-1, MMP-9, VEGF, KGF, and TIMP-1 were seen (Fig. 7).

Relative gene expression in skin tissue of different animal groups using qRT-PCR After normalisation to GAPDH, (A) I-CAM, (B) IL-1, (C) IL-10, (D) MCP-1, (E) TIMP-1, (F) MMP-9, (G) KGF, (H) IL-6, and (I) VEGF. In comparison to the healthy control group, the data show an increase in expression by a factor of two. The mean ± SD are shown as bars. A one-way ANOVA test is used to determine whether there's a significant difference between categories, via (a) p < 0.05 in contrast to the normal control grouping and (b) p < 0.05 in contrast to the marketplace drug-induced category.

The methodologies of molecular docking intend to predict the best binding orientation of a ligand to a receptor. It proposes several suitable poses of the ligand within the active or docking site of a receptor molecule.

The created PPI network comprised of 296 nodes and 1725 edges, which is illustrated in Figs. S1–S3.

The cytoHubba plugin Cytoscape is considered a useful exploring interface for the most important nodes in the PPI networks, it used to determine the hub genes using ranking methods ,The results shown in (Table 3) demonstrated that 16 nodes repeated in more than analysis method, regarding the occurrence, IL1B possessed the highest score as it appeared in 10 methods from the 12 methods followed by Il6 and TNF-α with score of 9 for each, while CD4 appeared 8 times, IL10 and IL2 seven times (Figs. S2 and S3). In protein–protein interaction networks, it is believed that the most connected nodes (hubs) are the key players, being responsible for the most extensive pathological effects56 a circular layout for the filtered nodes revealed that TNF-α and IL6 possessed the highest node degree in the 16 nodes (Fig. 8A), the highest occurred protein in cytoHubba analysis IL1B with TNF-α and IL6 were chosen for in silico molecular modelling57,58.

(A) Through a circular network design, the margins represent interactions between proteins, and the nodes serve as the hub protein criteria. Each protein's connectivity is represented by the dimension of the nodes; the larger the node, the greater its connection with other nodes in the network, (B) Functional enrichment analysis of filtered 16 protein coding genes by ShinyGO (https://www.genome.jp/kegg/, accessed on 12 September 2022); (http://bioinformatics.sdstate.edu/go/, accessed on 13 September 2022, a graphical gene set enrichment tool).

The Gene Ontology (GO) is considered a computational bioinformatic model of biological systems, beginning with the molecular level reaching to the organism level, GO aims to provide knowledge about the functions of gene products, namely, proteins and non-coding RNA molecules. GO is organized in three aspects. GO Molecular Functions (MF) describe activities that occur at the molecular level, Biological Processes (BP) represent the larger processes or ‘biological programs’ accomplished by multiple molecular activities and Cellular Components (CC) which are the cellular structures in which a gene product performs a function, the specific genes expressed in each cell define the identity and functionalities of that cell. Regulation of transcription is highly complex and leads to differential gene expression in specific cells or under specific conditions59. The analysis of the selected genes revealed that Positive regulation, and phosphorylation of STAT on JAK/STAT pathway were the top biological process in the same order while Interleukin 6 receptor complex was the top molecular component followed by keratohyalin granule and T cell receptor complex. For the Molecular function category screened genes were correlated with interleukin 4 and 8 receptor binding followed by toxic substance binding, Finally, the KEGG pathway for the selected protein coding genes were found to be involved in inflammatory bowel disease, Malaria and Legionellosis. (Fig. 8B).

The X-ray crystallographic structure of (IL-1β) complexed with its ligand was obtained from the Protein Data Bank (http://www.rcsb.org/pdb/,code 6Y8M). The ligand was re-docked in an active pocket at an acceptable RMSD of 1.311 and an energy score of -5.870 kcal/mol in five interactions of hydrogen bonds along with one ionic bond interaction to verify the results of our research. The involved amino acid residues in the-H-bond-interactions were Thr 147, Met 148, Gln 149, and-Arg 11 as H-acceptor, and another one with-Met-148 as-H-donor, while the ionic interaction was encountered with Arg 11. The-dock-score-of-the-twenty compounds against 6Y8M is summarized in Tables S1–S4. According to docking outcomes, compound 15 (geraniol) had a docking score of  − 5.881 kcal/mol, which was less favorable than the kinetic energy obtained by the co-crystallized ligand (Table S5), with-two-hydrogen-bond-interaction one as H-donor with Asn 108 and other as H-acceptor with Lys 109, meanwhile compound 16 (1-decanol) achieved approximately similar energy score of  − 5.625 kcal/mol when compared with the co-crystalized ligand showing three hydrogen bond interactions as H- acceptor through the hydroxyl group of 1-decanol and the amino acid residues Gln 149,Thr 147 and Arg 11 which resemble the interactions of the-co-crystalized-ligand (Fig. 9A,B).

(A) 2D actions and 3D docking represent compound 15 (geraniol) in the successful pocket location of IL-1 (PDB: 6Y8M), (B) 2D actions and 3D docking represent compound 16 (1-decanol) in the successful pocket location of IL-1 (PDB: 6Y8M), (C) 2D actions and 3D docking present in compound 15 in the successful pocket location of IL-6 (PDB: 1ALU), (D) 2D relationships and 3D docking present in compound 16 in the successful pocket location of IL-6 (PDB: 1ALU), (E) 2D relationships and 3D docking present in compound 17 in the successful pocket location of IL-6 (PDB: 1ALU), (E) 2D relationships and 3D docking represent compound 16 in TNF- successful pocket location (PDB: 2AZ5), and (F) 2D relationships and 3D docking present compound 15 in GST successful pocket location (PDB: 2AZ5) (PDB: 3EIN).

The X-ray crystallographic structure of IL-6 complexed with its ligand was made available by the Protein Data Bank (http://www.rcsb.org/pdb/, code 1ALU) (tartaric acid). The co-crystallized ligand (l-( +)-tartaric acid) posture was predicted by the docking method with an RMSD-of-1.758-and-an-energy-score-of  − 4.191 kcal/mol. In the style of contacts shown in Table S6, H-acceptor-interactions with Arg-182 and-Arg-179 were present, as were ionic interactions and one-hydrogen-bond-with-Gln-acting as-the-H-donor. It's interesting to note that many the 20 phytochemicals' docking results showed strong affinity for the receptor, with scores similar to the co-crystallized ligand (Table S2). It is worth mentioning that both compounds 15 and 16 exhibited better affinity towards the binding site of IL-6 than the ligand, as they showed ΔG of  − 4.372 and  − 4.401 kcal/mol respectively. The hydrogen bond interactions appeared as three hydrogen bond acceptors with amino acid residues Arg 179 and Arg 182 in both, which match the interaction pattern of the co-crystallized ligand, moreover, compound 8 also achieved a good energy score of  − 4.151 kcal/mol when compared with the ligand score of  − 4.191 kcal/mol. (Fig. 9C,D).

The Protein Data Bank (http://www.rcsb.org/pdb/, code 2AZ5) provided the X-ray crystallographic structure of (TNF-) complexed with its ligand. It was demonstrated that the co-crystallized ligand was associated with 16 residues of amino acids and attached inside a small pocket, with seven of those residues coming from-chain-A and the remaining nine-from-chain-B, including-six-tyrosine-residues, from each subunit of the TNF- dimer. This inhibitor works by attaching to the cytokine's active trimer form, stimulating its dissociation into the inactive dimer form, and stabilizing it60. The ligand was re-docked in the active pocket to verify our research. During interactions with receptors, the ligand established hydrogen-bonds-with-Gln-61 as an H-donor-and-with-Tyr-119 as a pi-H interaction. The co-crystallized ligand pose was predicted by the docking method with the least RMSD and an energy score of  − 6.923 kcal/mol. Figure 9E, Table S7. Table S3 summarizes the dock scores of the 20 compounds against 2AZ5. Due to its hydroxyl moiety and the amino acid sequence Gln 61, compound 16 was the only one to obtain a dock score of  − 5.129 kcal/mol with a single hydrogen bond interaction as an H-donor (Fig. 9).

The-Protein-Data-Bank-(http://www.rcsb.org/pdb/,code 3EIN) has-the-X-ray-crystallographic-structure-of-Drosophila melanogaster's delta-class-GST. When glutathione was redocked, it revealed four hydrogen bond interactions, two of which included Arg-67-and-Ser-66 as H-acceptors and the-other-two-involving Glu 65-and Ile 53-as H-donors. The ligand's energy score was  − 5.945 kcal/mol in addition to the two ionic interactions with Arg 67 and Glu 65 (Table S8). based on the investigated compounds' docking results reported in Table S4, compound 15 (geraniol), which displayed a G of  − 5.861 kcal/mol, exhibits strong similarities to glutathione in terms of energy score. In a similar manner to the co-crystallized ligand, geraniol interacted with the binding site of the GST receptor by forming two hydrogen bonds, one with the amino acid residue Ser-66 as an H-acceptor-and the-other with-Glu 65 as an-H-donor (Fig. 9F).

To validate the docking outcomes, the best-scoring docking pose of geraniol with both GST and IL-6 were subjected to 50 ns-long MD simulation. As shown in Fig. 10, geraniol modelled structure achieved stable binding inside the binding site of each protein with RMSD profiles (~ 2.5 Å) comparable with that of the co-crystallized ligands (~ 1.7 Å). These results suggest geraniol as a probable inhibitor of both GST and IL-6.

RMSDs of geraniol inside both GST and IL-6 in comparison with the co-crystalized ligand of each protein [(A) and (B) respectively] over the course of 50 ns-long MD simulation.

Various physicochemical properties of a given drug may have a significant impact on its bioactivity, as they are closely related to interactions between the drug and its potentially suspected target. Recently, in silico approaches introduced a powerful tool for drug discovery to assess the proposed pharmacokinetics (ADME) of compounds, which play a vital role in their pharmacological activities, especially at the early stages of screening for lead compounds. Consequently, the measurement of these parameters is of great value in the selection of an efficient drug candidate. Lipinski and Veber’s rules are successful tools to perform such screening, as Lipinski’s rule of five states that a compound has drug-like activity if at least three of the following criteria are achieved: a molecular mass less than 500 Da, a maximum of five hydrogen donors, a maximum of 10 hydrogen bond acceptors, and a partition coefficient between octanol and water (LogP (o/w)) smaller than 5. According to Veber’s rule, a compound is orally active if it has 10 or fewer rotatable bonds and a polar surface area (PSA) greater than 140 Å. For predicting drug-like properties, we used Reaxys. The screening of compounds 15 and 16 revealed that all of them complied with Lipinski and Veber’s rules (Fig. 11).

In silico druglikeness (Lipinski and Veber rules) of compounds 15 and 16.

Essential oils (EOs) are generally well tolerated, as evidenced by their widespread use in food, hair, and skin preparations61. In comparison to conventional drugs, EOs are less likely to cause resistance due to their multiple active components61. EOs may have antibacterial, anti-inflammatory, and antipruritic properties in addition to their scabicide properties8,14. All these adjuvant properties are especially appealing for the treatment of scabies.

As a result of mites burrowing deeply into the skin, scabies pathogenesis is complicated and involves a number of mechanisms, including parasite persistence, which has an impact on both the structure and function of skin62. All these elements work together to make treatment ineffective, especially given that most synthetic medications kill mites rather than altering the immune system or promoting tissue repair. In light of this, plant-derived phytochemicals can operate as safe substitutes for synthetic options for the eradication of infectious diseases due to their broad therapeutic potential and negligible adverse effects63. Citrus fruits have reportedly been found to have immunostimulatory, anti-inflammatory, antimicrobial, and antioxidant properties64,65,66. It had considerable antibacterial efficacy against S. aureus and Candida-skin-infections, including-oral-and-vaginal-candidiasis65,67,68,69,70. Fascinating studies have shown that Citrus oil may change the way that inflammatory responses are expressed, suppressing pro-inflammatory cytokines and enhancing skin's defensive barriers71. No investigation has yet been conducted, to the best of the information we have, on the acaricidal-potential-of mandarin peel oil against Sarcoptes scabiei.

Therefore, the current study examined the GC/MS makeup of mandarin peel oil and assessed the oil's ability to kill Scabies-mites-in-both in vitro-and in vivo testing. There were no symptoms of skin irritability, inflammation, or unease during or after the application of mandarin peel oil. Our findings showed that orange peel oil could have a substantial acaricidal effect on Sarcoptes scabiei mites 24 h after application. The animals' skin began to exhibit healthy symptoms after the mites died, including the cessation of inflammation-and-hyperkeratosis, the emergence of new-skin-layers, and-the beginning of new hair growth. This occurred at the same time as reports of the effective treatment of rabbit mange72. This full recovery was seen after 3 weeks, whereas the ivermectin group's healing continued until the completion of the experiment (4 weeks) without leading to full recovery. The histopathological findings revealed that the dermis and epidermis of the treated animals improved, inflammation cells decreased, and mite remains were not present in the skin layers. The death of mites, as well as the absence of inflammation, pruritis, skin damage, and scale formation, are the primary causes of the improvement55. In contrast, the skin layers of the ivermectin group saw gradual alterations throughout therapy, and at the end of the investigation, some inflammatory cells as well as traces of deceased mites were still visible. This could be explained by the potent anti-mite effects of ivermectin as well as the common itching and allergic reactions brought on by topical deltamethrin use, which extends inflammation and causes additional delays the emergence of good indicators73.

Epidermal-keratinocytes-as the-first line-of defense against hazardous external invaders must be used to obtain understanding of the-modulative-effects-of mandarin peel oil on the-pathophysiology-of scabies. To recognize various infections and launch immune responses, keratinocytes generate recognition receptors on their surfaces. These receptors allow them to secrete cytokines, chemokines, and anti-microbial peptides that help attract inflammatory cells74. Any imbalance in the activity of keratinocytes, which is crucial for the control of skin immunological homeostasis, might lead to illness. Our results show that when exposed to live digging scabies mites or their waste products (including saliva or eggs as well), a significant amount of genes in the skin fibroblasts and keratinocytes change their expression, which further activates other cell types75. Hence, in response to scabies, many other skins cell categories, such as lymphoid cells, endothelial cells, or LCs, and dendritic cells, have complex interactions (cross talk), which results in the development of inflammatory and oxidative stress states. This could increase reactive oxygen species like H2O2, which leads to lipid peroxidation and negatively affects the skin's structure and permeability. In our investigation, faster clinical and parasitological recovery confirmed the potential antioxidant action of mandarin peel oil by restoring the altered oxidant/antioxidant balance in treated animals to normal. Antioxidants are believed to hasten wound healing by reducing oxidative stress on the wound. They are essential in preventing harm from being done to biological elements like DNA, proteins, lipids, and bodily tissue when reactive species are present31. Because the elevated levels of ROS at the site of injury are the main promoters of collagen disintegration, the breakdown of the extracellular matrix (ECM), a decrease in vascular development and re-epithelialization, and a rise in cytokines that are pro-inflammatory, all of which extend inflammation, an extract with ROS scavenging potential could be a key component of the healing protocol31.

The anticipated mechanism of action of mandarin peel oil on scabies-infected rabbits was shown schematically in Fig. 12. Infiltrating mites, based on reports, activate the skin's keratinocytes, and exhibit a capacity to suppress the immune system's response by lowering the expression of genes of i-CAM-1, an intracellular adhesion molecular structure noticed on endothelial surface cells. This decreases the blood-supply and-immune cells to the penetration site and lessens the protective abilities of both lymphocytes and neutrophils. On the other hand, an infection increases MCP-1, a chemokine that stimulates immune cells and causes inflammation76. Clinical signs are not noticed for 4–6 weeks after an individual has been diagnosed with scabies mites. This is since the regulatory T cells (type 1) are induced to produce IL-10, a cytokine with anti-inflammatory properties that is required by humans to prevent inflammatory and autoimmune illnesses77. Furthermore, mite products that sensitize keratinocytes are likely to raise the production of VEGF, which is additionally induced by the mites to raise angiogenesis. The mites raise the flow of blood in the region in order to get the nourishment they require from the food being consumed, which worsens inflammation77. A delayed re-epithelialization of the wound is the result of decreased KGF receptor signaling, which also lowers the rate of proliferation of epidermal keratinocytes along the wound edge. Matrix metalloproteinase (MMP-9) is one of a set of hydrolase enzymes that are expressed in many severe conditions, including wounds, osteoarthritis, ischemia, and viral disorders. The inflammation also significantly increases MMP-9 levels78. Nearly all parasite infections use MMP-9 to remodel tissue, which often slows down the production of ECM molecules including collagen II and aggrecan79. TIMP-1 (tissue inhibitor of met-alloproteinase) tightly regulates the biological activities of MMPs, and proteolysis results from an imbalance in the MMPs/TIMPs ratio80. Reversing the activation of this network of interrelated genes may therefore be a useful treatment approach to slow the spread of scabies. When CSE was applied topically, the expression of IL-1, 6, 10, VEGF, MMP-9, and MCP-1 significantly decreased, whereas the expression of i-CAM-1, KGF, and TIMP-1 significantly increased.

The suggested mechanism for the effect of mandarin peel oil on scabies-infected rabbits.

The outcome was an improvement in host immunity against invading mites, a decrease in pro-inflammatory cytokines and an increase in anti-inflammatory ones, which could reverse the unfavorable symptoms and lead to improved re-epithelialization, rapid recovery, and a decrease in inflammation.

With instances of treatment failure and the emergence of resistance, controlling scabies effectively using the available acaricidal medicines has proven to be extremely difficult. With a biocidal performance comparable to that of traditional synthetic treatments, this study demonstrated the mandarin peel oil's acaricidal efficacy against mange mites in rabbits. The work examined the composition of the oil and revealed the presence of different hydrocarbons and their oxygenated forms, with proved biocidal activities. Additionally, the oil has been tested against naturally infected rabbits with mange using different techniques and proved higher efficacy and safety compared to market agents. The mandarin peel oil presents an ideal alternative to commercial medications used for the control of arachnids that can harm humans and animals while being economical, safe, and environmentally friendly. These candidates can be successfully employed to create novel biocides for applications in agricultural improvement and livestock protection.

All data generated or analyzed during this study are included in this article (and its supplementary information files).

Smith, K. F., Acevedo-Whitehouse, K. & Pedersen, A. B. The role of infectious diseases in biological conservation. Anim. Conserv. 12, 1–12 (2009).

Article Google Scholar

Fain, A. Epidemiological problems of scabies. Int J Dermatol 17, 20–30 (1978).

Article CAS PubMed Google Scholar

Currier, R. W., Walton, S. F. & Currie, B. J. Scabies in animals and humans: History, evolutionary perspectives, and modern clinical management. Ann. N. Y. Acad. Sci. 1230, E50–E60 (2011).

Article ADS PubMed Google Scholar

Arlian, L. G. & Morgan, M. S. A review of Sarcoptes scabiei: Past, present and future. Parasit. Vectors 10, 1–22 (2017).

Article Google Scholar

Abu Hafsa, S. H., Senbill, H., Basyony, M. M. & Hassan, A. A. Amelioration of sarcoptic mange-induced oxidative stress and growth performance in ivermectin-treated growing rabbits using turmeric extract supplementation. Animals 11, 2984 (2021).

Article PubMed PubMed Central Google Scholar

Rodriguez-Vivas, R. I., Jonsson, N. N. & Bhushan, C. Strategies for the control of Rhipicephalus microplus ticks in a world of conventional acaricide and macrocyclic lactone resistance. Parasitol. Res. 117, 3–29 (2018).

Article PubMed Google Scholar

Rehbein, S. et al. Productivity effects of bovine mange and control with ivermectin. Vet. Parasitol. 114, 267–284 (2003).

Article CAS PubMed Google Scholar

Andriantsoanirina, V. et al. In vitro efficacy of essential oils against Sarcoptes scabiei. Sci. Rep. 12, 1–8 (2022).

Article Google Scholar

Fang, F. et al. In vitro activity of ten essential oils against Sarcoptes scabiei. Parasit. Vectors 9, 1–7 (2016).

Article Google Scholar

Aboelhadid, S., Mahrous, L. N., Hashem, S. A., Abdel-Kafy, E. & Miller, R. J. In vitro and in vivo effect of Citrus limon essential oil against sarcoptic mange in rabbits. Parasitol. Res. 115, 3013–3020 (2016).

Article CAS PubMed Google Scholar

Zhou, Y. et al. Composition and acaricidal activity of essential oil from Elsholtzia densa Benth against Sarcoptes scabiei mites in vitro. Veterinární medicína 64, 178–183 (2019).

Article CAS Google Scholar

Pasay, C. et al. Acaricidal activity of eugenol based compounds against scabies mites. PLoS ONE 5, e12079 (2010).

Article ADS PubMed PubMed Central Google Scholar

Chen, Z., van Mol, W., Vanhecke, M., Duchateau, L. & Claerebout, E. Acaricidal activity of plant-derived essential oil components against Psoroptes ovis in vitro and in vivo. Parasit. Vectors 12, 1–11 (2019).

Article Google Scholar

Dhifi, W., Bellili, S., Jazi, S., Bahloul, N. & Mnif, W. Essential oils’ chemical characterization and investigation of some biological activities: A critical review. Medicines 3, 25 (2016).

Article PubMed PubMed Central Google Scholar

Türkmenoğlu, A. & Özmen, D. Allergenic components, biocides, and analysis techniques of some essential oils used in food products. J. Food Sci. 86, 2225–2241 (2021).

Article PubMed Google Scholar

Hodgson, R. W. Horticultural varieties of Citrus. History, world distribution, botany and varieties, 431–591 (1967).

Njoroge, S. M., Koaze, H., Mwaniki, M., Minh Tu, N. & Sawamura, M. Essential oils of Kenyan Citrus fruits: Volatile components of two varieties of mandarins (Citrus reticulate) and a tangelo (C. paradisi × C. tangerine). Flavour Fragr. J. 20, 74–79 (2005).

Article CAS Google Scholar

Minh Tu, N., Thanh, L., Une, A., Ukeda, H. & Sawamura, M. Volatile constituents of Vietnamese pummelo, orange, tangerine and lime peel oils. Flavour Fragr. J. 17, 169–174 (2002).

Article CAS Google Scholar

Dharmawan, J., Kasapis, S., Curran, P. & Johnson, J. R. Characterization of volatile compounds in selected Citrus fruits from Asia. Part I: Freshly-squeezed juice. Flavour Fragr. J. 22, 228–232 (2007).

Article CAS Google Scholar

Zhou, X., Zhao, Y., He, C. C. & Li, J. Preventive effects of Citrus reticulata essential oil on bleomycin-induced pulmonary fibrosis in rats and the mechanism. Zhong xi yi jie he xue bao= J. Chin. Integr. Med. 10, 200–209 (2012).

Google Scholar

Yu, L., Yan, J. & Sun, Z. D-limonene exhibits anti-inflammatory and antioxidant properties in an ulcerative colitis rat model via regulation of iNOS, COX-2, PGE2 and ERK signaling pathways. Mol. Med. Rep. 15, 2339–2346 (2017).

Article CAS PubMed Google Scholar

Yi, F., Jin, R., Sun, J., Ma, B. & Bao, X. Evaluation of mechanical-pressed essential oil from Nanfeng mandarin (Citrus reticulata Blanco cv. Kinokuni) as a food preservative based on antimicrobial and antioxidant activities. Lwt 95, 346–353 (2018).

Article CAS Google Scholar

Tao, N., Jia, L. & Zhou, H. Anti-fungal activity of Citrus reticulata Blanco essential oil against Penicillium italicum and Penicillium digitatum. Food Chem. 153, 265–271 (2014).

Article CAS PubMed Google Scholar

Matan, N. & Matan, N. Antifungal activities of anise oil, lime oil, and tangerine oil against molds on rubberwood (Hevea brasiliensis). Int. Biodeterior. Biodegrad. 62, 75–78 (2008).

Article CAS Google Scholar

Wu, T. et al. Antifungal action and inhibitory mechanism of polymethoxylated flavones from Citrus reticulata Blanco peel against Aspergillus niger. Food Control 35, 354–359 (2014).

Article CAS Google Scholar

Chutia, M., Bhuyan, P. D., Pathak, M., Sarma, T. & Boruah, P. Antifungal activity and chemical composition of Citrus reticulata Blanco essential oil against phytopathogens from North East India. LWT-Food Sci. Technol. 42, 777–780 (2009).

Article CAS Google Scholar

Ren, Y., Shen, L. & Dai, S. Studies on flavonoids and amides from herbs of Solanum lyratum. Zhongguo Zhong yao za zhi= Zhongguo Zhongyao Zazhi= China J. Chin. Mater. Medica 34, 721–723 (2009).

CAS Google Scholar

Elmaidomy, A. H., Hassan, H. M., Amin, E., Mohamed, W. & Hetta, M. H. Premna odorata volatile oil as a new mycobacterium tuberculosis growth inhibitor for the control of tuberculosis disease. Eur. J. Med. Plants 21, 1–11 (2017).

Article Google Scholar

Mohammad, W. A. et al. Immunomodulatory effect of Premna odorata volatile oils in Mycobacterium tuberculosis by inhibiting TLR4/NF-κB pathway. J. Herbmed Pharmacol. 8, 1–7 (2019).

Article CAS Google Scholar

Alsenani, F. et al. Wound healing metabolites from Peters’ elephant-nose fish oil: An in vivo investigation supported by in vitro and in silico studies. Mar. Drugs 19, 605 (2021).

Article CAS PubMed PubMed Central Google Scholar

Al-Warhi, T. et al. Antioxidant and wound healing potential of vitis vinifera seeds supported by phytochemical characterization and docking studies. Antioxidants 11, 881 (2022).

Article CAS PubMed PubMed Central Google Scholar

Zahran, E. M. et al. Scabicidal potential of coconut seed extract in rabbits via downregulating inflammatory/immune cross talk: A comprehensive phytochemical/GC-MS and in silico proof. Antibiotics 12, 43 (2022).

Article PubMed PubMed Central Google Scholar

Hassan, H. & Abdel-Aziz, A. Evaluation of free radical-scavenging and anti-oxidant properties of black berry against fluoride toxicity in rats. Food Chem. Toxicol. 48, 1999–2004 (2010).

Article CAS PubMed Google Scholar

Sreenivasan, S., Ibrahim, D. & Mohd Kassim, M. J. N. Free radical Scavenging Activity and Total Phenolic Compounds of Gracilaria changii. Int. J. Nat. Eng. Sci. 1, 115–117 (2007).

CAS Google Scholar

Andriantsoanirina, V. et al. In vitro efficacy of essential oils against Sarcoptes scabiei. Scientific Rep. 12, 7176. https://doi.org/10.1038/s41598-022-11176-x (2022).

Article ADS CAS Google Scholar

Zahran, E. M. et al. Metabolic profiling, histopathological anti-ulcer study, molecular docking and molecular dynamics of ursolic acid isolated from Ocimum forskolei Benth.(family Lamiaceae). South African J. Botany 131, 311–319 (2020).

Article CAS Google Scholar

Shannon, P. et al. Cytoscape: A software environment for integrated models of biomolecular interaction networks. Genome Res. 13, 2498–2504 (2003).

Article CAS PubMed PubMed Central Google Scholar

Szklarczyk, D. et al. The STRING database in 2017: Quality-controlled protein–protein association networks, made broadly accessible. Nucleic Acids Re. 45, gkw937 (2016).

Google Scholar

Shady, N. H. et al. Mechanistic wound healing and antioxidant potential of moringa oleifera seeds extract supported by metabolic profiling, in silico network design, molecular docking, and in vivo studies. Antioxidants 11, 1743 (2022).

Article CAS PubMed PubMed Central Google Scholar

Chin, C.-H. et al. cytoHubba: Identifying hub objects and sub-networks from complex interactome. BMC Syst. Biol. 8, 1–7 (2014).

Article Google Scholar

Vaghasia, H., Sakaria, S., Prajapati, J., Saraf, M. & Rawal, R. M. Interactive bioinformatics analysis for the screening of hub genes and molecular docking of phytochemicals present in kitchen spices to inhibit CDK1 in cervical cancer. Comput. Biol. Med. 149, 105994 (2022).

Article CAS PubMed Google Scholar

Ge, S. X., Jung, D. & Yao, R. ShinyGO: A graphical gene-set enrichment tool for animals and plants. Bioinformatics 36, 2628–2629 (2020).

Article CAS PubMed Google Scholar

Fischer, K., Holt, D., Currie, B. & Kemp, D. Scabies: Important clinical consequences explained by new molecular studies. Adv. Parasitol. 79, 339–373 (2012).

Article PubMed Google Scholar

Bhat, S. A., Mounsey, K. E., Liu, X. & Walton, S. F. Host immune responses to the itch mite, Sarcoptes scabiei, in humans. Parasit. Vectors 10, 1–12 (2017).

Article Google Scholar

Sharp, P., Smith, D., Bach, W., Wagland, B. & Cobon, G. Purified glutathione S-transferases from parasites as candidate protective antigens. Int. J. Parasitol. 21, 839–846 (1991).

Article CAS PubMed Google Scholar

Hu, Z. et al. In vitro acaricidal activity of 1, 8-cineole against Sarcoptes scabiei var. cuniculi and regulating effects on enzyme activity. Parasitol. Res. 114, 2959–2967 (2015).

Article PubMed Google Scholar

Pettersson, E. U., Ljunggren, E. L., Morrison, D. A. & Mattsson, J. G. Functional analysis and localisation of a delta-class glutathione S-transferase from Sarcoptes scabiei. Int. J. Parasitol. 35, 39–48 (2005).

Article CAS PubMed Google Scholar

Phillips, J. C. et al. Scalable molecular dynamics with NAMD. J. Comput. Chem. 26, 1781–1802 (2005).

Article CAS PubMed PubMed Central Google Scholar

Ribeiro, J. V., Bernardi, R. C., Rudack, T., Schulten, K. & Tajkhorshid, E. QwikMD-gateway for easy simulation with VMD and NAMD. Biophys. J . 114, 673a–674a (2018).

Article Google Scholar

Huang, B. et al. Comparative analysis of essential oil components and antioxidant activity of extracts of Nelumbo nucifera from various areas of China. J. Agric. Food Chem. 58, 441–448 (2010).

Article CAS PubMed Google Scholar

Goyal, L. & Kaushal, S. Evaluation of chemical composition and antioxidant potential of essential oil from Citrus reticulata fruit peels. Adv. Res. 15, 1–9 (2018).

Article Google Scholar

Njoroge, S. M., Mungai, H. N., Koaze, H., Phi, N. T. L. & Sawamura, M. Volatile constituents of mandarin (Citrus reticulata Blanco) peel oil from Burundi. J. Essent. Oil Res. 18, 659–662 (2006).

Article CAS Google Scholar

Boughendjioua, H. & Boughendjioua, Z. Chemical composition and biological activity of essential oil of mandarin (Citrus reticulata) cultivated in Algeria. Int. J. Pharm. Sci. Revi. 40, 179–184 (2017).

Google Scholar

Hu, Z. et al. In vitro acaricidal activity of 1,8-cineole against Sarcoptes scabiei var. cuniculi and regulating effects on enzyme activity. Parasitol. Res. 114, 2959–2967. https://doi.org/10.1007/s00436-015-4498-8 (2015).

Article PubMed Google Scholar

Mohamed, R., Ibrahim, M. D., Khalil, M., Masry, E. L. & Rasheed, N. Histopathological, clinico-biochemical and therapeutic studies on different types of mange in domestic rabbits. Assiut Vet. Med. J. 63, 90–101 (2017).

Article Google Scholar

Zhang, X., Zhu, J., Wang, Q. & Zhao, H. Identifying influential nodes in complex networks with community structure. Knowl.-Based Syst. 42, 74–84 (2013).

Article CAS Google Scholar

Walton, S. F. The immunology of susceptibility and resistance to scabies. Parasite Immunol. 32, 532–540 (2010).

CAS PubMed Google Scholar

Arlian, L. G., Morgan, M. S. & Neal, J. S. Modulation of cytokine expression in human keratinocytes and fibroblasts by extracts of scabies mites. Am. J. Trop. Med. Hyg. 69, 652–656 (2003).

Article PubMed Google Scholar

Bogenhagen, D. F. Mitochondrial DNA nucleoid structure. Biochimica et Biophysica Acta (BBA)-Gene Regul. Mech. 1819, 914–920 (2012).

Article CAS Google Scholar

He, M. M. et al. Small-molecule inhibition of TNF-α. Science 310, 1022–1025 (2005).

Article ADS CAS PubMed Google Scholar

Tisserand, R. & Young, R. Essential oil safety: A guide for health care professionals (Elsevier Health Sciences, Netherlands, 2013).

Google Scholar

Walton, S. F., Beroukas, D., Roberts-Thomson, P. & Currie, B. New insights into disease pathogenesis in crusted (Norwegian) scabies: The skin immune response in crusted scabies. Br. J. Dermatol. 158, 1247–1255 (2008).

Article CAS PubMed Google Scholar

Mohamad, S. A., Zahran, E. M., Fadeel, M. R. A., Albohy, A. & Safwat, M. A. New acaciin-loaded self-assembled nanofibers as mpro inhibitors against bcv as a surrogate model for sars-cov-2. Int. J. Nanomed. 16, 1789 (2021).

Article Google Scholar

Tomasello, B. et al. Anti-adipogenic and anti-oxidant effects of a standardised extract of Moro blood oranges (Citrus sinensis (L.) Osbeck) during adipocyte differentiation of 3T3-L1 preadipocytes. Nat. Product Res. 35, 2660–2667 (2021).

Article CAS Google Scholar

Fisher, K. & Phillips, C. Potential antimicrobial uses of essential oils in food: Is Citrus the answer?. Trends Food Sci. Technol. 19, 156–164 (2008).

Article CAS Google Scholar

Shehwana, H. et al. Transcriptome analysis of host inflammatory responses to the Ectoparasitic Mite Sarcoptes scabiei var. hominis. Front. Immunol. 12, 778840 (2021).

Article CAS PubMed PubMed Central Google Scholar

Lalouckova, K. et al. In vitro antagonistic inhibitory effects of palm seed crude oils and their main constituent, lauric acid, with oxacillin in Staphylococcus aureus. Sci. Rep. 11, 1–12 (2021).

Article Google Scholar

Benavente-Garcia, O. & Castillo, J. Update on uses and properties of Citrus flavonoids: New findings in anticancer, cardiovascular, and anti-inflammatory activity. J. Agric. Food Chem. 56, 6185–6205 (2008).

Article CAS PubMed Google Scholar

Song, X. et al. Antibacterial effects and mechanism of mandarin (Citrus reticulata L.) essential oil against Staphylococcus aureus. Molecules 25, 4956 (2020).

Article CAS PubMed PubMed Central Google Scholar

Boukhennoufa, A., Louiza, M.-B.H., Aicha, T. & Boumediene, M. In vivo Anti-Candida activity of Artemisia herba alba asso and Citrus aurantium Extracts formulated in different ointments. Anti-Infect. Agents 19, 57–65 (2021).

Article Google Scholar

Mitoshi, M. et al. Suppression of allergic and inflammatory responses by essential oils derived from herbal plants and Citrus fruits. Int. J. Mol. Med. 33, 1643–1651 (2014).

Article CAS PubMed Google Scholar

Abd El-Ghany, W. A. Mange in rabbits: An Ectoparasitic disease with a zoonotic potential. Veter. Med. Int. 2022, 5506272. https://doi.org/10.1155/2022/5506272 (2022).

Article Google Scholar

Fujimoto, K., Kawasaki, Y., Morimoto, K., Kikuchi, I. & Kawana, S. Treatment for crusted scabies: Limitations and side effects of treatment with ivermectin. J. Nippon Med. Sch. 81, 157–163. https://doi.org/10.1272/jnms.81.157 (2014).

Article PubMed Google Scholar

Piipponen, M., Li, D. & Landén, N. X. The Immune Functions of Keratinocytes in Skin Wound Healing. Int. J. Mol. Sci. https://doi.org/10.3390/ijms21228790 (2020).

Article PubMed PubMed Central Google Scholar

Dai, X. et al. Mite allergen is a danger signal for the skin via activation of inflammasome in keratinocytes. J. Allergy Clin. Immunol. 127, 806–814 (2011).

Article CAS PubMed Google Scholar

Reale, M. et al. Induction of monocyte chemotactic protein-1 (MCP-1) and TNF alpha by Trichinella spiralis in serum of mice in vivo. Mol. Cell Biochem. 179, 1–5. https://doi.org/10.1023/a:1006875429323 (1998).

Article CAS PubMed Google Scholar

Morgan, M. S., Arlian, L. G. & Markey, M. P. Sarcoptes scabiei mites modulate gene expression in human skin equivalents. PLoS ONE 8, e71143. https://doi.org/10.1371/journal.pone.0071143 (2013).

Article ADS CAS PubMed PubMed Central Google Scholar

Zahran, E. et al. Bioactives and functional food ingredients with promising potential for management of cerebral and myocardial ischemia: A comprehensive mechanistic review. Food Funct. https://doi.org/10.1039/D2FO00834C (2022).

Article PubMed Google Scholar

Livshits, G. & Kalinkovich, A. Hierarchical, imbalanced pro-inflammatory cytokine networks govern the pathogenesis of chronic arthropathies. Osteoarthr. Cartil. 26, 7–17 (2018).

Article CAS Google Scholar

Takeshita, S. et al. Elevated serum levels of matrix metalloproteinase-9 (MMP-9) in Kawasaki disease. Clin. Exp. Immunol. 125, 340–344. https://doi.org/10.1046/j.1365-2249.2001.01608.x (2001).

Article CAS PubMed PubMed Central Google Scholar

Download references

The Deanship of Scientific Research at Umm Al-Qura University supported this work via Grant Code: 22UQU4310013DSR08, and the authors would like to thank them for that. Also, the identifications of the plant’s fruit name was kindly provided by Abdel Haleem A. Mohammed of the Department-of-Flora-and-Phytotaxonomic-Research-at-the-Horticulture-Research-Institute, Dokki, Cairo, Egypt.

Department of Pharmacognosy, Faculty of Pharmacy, Beni-Suef University, Beni-Suef, Egypt

Abeer H. Elmaidomy

Department of Pathology, Faculty of Pharmacy, Deraya University, Minya, Egypt

Nehad M. Reda Abdel-Maqsoud

Department of Biochemistry, Faculty of Pharmacy, New Valley University, Kharga, New Valley, Egypt

Omar. Y. Tammam

Department of Pharmaceutical Chemistry, Faculty of Pharmacy, Deraya University, Minya, Egypt

Islam M. Abdel-Rahman

Department of Biochemistry, Faculty of Pharmacy, Deraya University, New Minya, Egypt

Mahmoud A. Elrehany

Department of Pharmacy Practice, Faculty of Pharmacy, King Abdulaziz University, Jeddah, Saudi Arabia

Hussain T. Bakhsh

Department of Medical Laboratory Technology, Faculty of Applied Medical Sciences, University of Tabuk, Tabuk, Saudi Arabia

Faisal H. Altemani & Naseh A. Algehainy

Department of Biological Sciences, Faculty of Science, King Abdulaziz University, Jeddah, Saudi Arabia

Mubarak A. Alzubaidi

Department of Pharmacognosy, College of Pharmacy, Umm Al-Qura University, Mecca, Saudi Arabia

Faisal Alsenani

Department of Pharmacognosy, Faculty of Pharmacy, Nahda University, Beni-Suef, 62513, Egypt

Ahmed M. Sayed

Department of Pharmacognosy, Faculty of Pharmacy, Deraya University, Minya, Egypt

Usama Ramadan Abdelmohsen & Eman Maher Zahran

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

Conceptualization: U.R.A., A.H.E.; methodology: A.H.E., N.M.R.A., O.Y.T., I.M.A.R.; software: M.A.E., A.M.S., H.T.B., F.H.A., N.A.A., M.A.A.; formal analysis: A.H.E., N.M.R.A., O.Y.T., I.M.A.R.; investigation: U.R.A., A.H.E., E.M.Z.; resources: M.A.E., H.T.B., F.H.A., N.A.A., M.A.A.; data curation: U.R.A., A.H.E., E.M.Z., M.A.E., H.T.B., F.H.A., N.A.A., M.A.A.; writing—original draft: A.H.E., N.M.R.A., O.Y.T., I.M.A.R.; writing—review and editing: A.H.E., N.M.R.A., O.Y.T., I.M.A.R., U.R.A., E.M.Z. The final manuscript has been approved for publication by all authors after reading it.

Correspondence to Abeer H. Elmaidomy or Usama Ramadan Abdelmohsen.

The authors declare no competing interests.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and Permissions

Elmaidomy, A.H., Abdel-Maqsoud, N.M.R., Tammam, O.Y. et al. Egyptian mandarin peel oil's anti-scabies potential via downregulation-of-inflammatory/immune-cross-talk: GC–MS and PPI network studies. Sci Rep 13, 14192 (2023). https://doi.org/10.1038/s41598-023-38390-5

Download citation

Received: 29 March 2023

Accepted: 07 July 2023

Published: 30 August 2023

DOI: https://doi.org/10.1038/s41598-023-38390-5

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.