Preparation of silymarin-loaded zein polysaccharide core–shell nanostructures and evaluation of their biological potentials
-
Muhammad Saqib Khalil
, Farhat Ali Khan
, Sadiq Azam
Abstract
Silymarin-loaded zein polysaccharide core–shell nanoparticles (SZPCS-NPs) were synthesized where sodium alginate and pectin offer stability and controlled release qualities to zein, a maize protein, having excellent biocompatibility. The present study is an attempt to develop zein–silymarin polysaccharide core–shell nanostructures to enhance water solubility, thereby improving bioavailability and producing enhanced biological responses in living systems. SZPCS-NPs were prepared using pH-induced antisolvent precipitation method. Five different types of SZPCS-NPs were synthesized using different combinations of sodium alginate and pectin, namely P100–A00 (non-uniform size ranging from 20 to 100 nm), P70–A30 (spherical and uniform size measuring approximately 80 nm in diameter), P50–A50, P30–A70, and P00–A100 exhibited irregular shapes with the presence of some triangular and oval structures and non-uniform size ranging from 20 to 100 nm. The SZPCS-NPs P70–A30 possessed the best results in terms of shape, size, and other characterization studies. Furthermore, the SZPCS-NPs possessed a percent drug loading of 72.5% and entrapment efficiency of 51.7%, respectively. The resulting SZPCS-NPs exhibited an excellent relative bioavailability percentage of 97.4% in comparison to commercial silymarin, having 58.1%, and crude silymarin, having 46.97% bioavailability percentage, correspondingly. In addition, SZPCS-NPs possessed an almost two folds’ increase in antioxidant activity in comparison to crude and commercially available silymarin. Similarly, SZPCS-NPs also showed better stabilization in hepatic biomarker enzymes and possessed better hepatoprotective activity for a period of 6 weeks, in contrast to commercial and crude silymarin formulations.
Graphical Abstract

1 Introduction
Plant-based natural products have been widely used as treatments against numerous diseases since ancient times [1]. Around 40% of the marketed medicines have hydro-solubility issues, making it difficult to achieve enough drug absorption from the gastrointestinal tract (GIT) [1,2]. Drugs having low hydro-solubility and cytopermeability have lower absorption and bioavailability, which has an impact on drug safety and efficacy. To produce a therapeutic product with better pharmacokinetic features from these active pharmaceutical ingredients, innovative formulation efforts are required [3]. The poor water solubility of the majority of therapeutic medicines might decrease their dissolution rate and oral bioavailability, posing a problem for pharmaceutical development [4]. In order to get the appropriate drug concentration in the systemic circulation, solubility is an important attribute [5]. Over 90% of the drugs that have been manufactured possess poor oral bioavailability due to limited solubility in water and/or a slow disintegration rate [6]. Using poorly soluble drugs has several limitations, including the need to increase the dose and administration frequency of the drug product as well as the likelihood of side effects. The solubility of poorly soluble drugs can be improved using various techniques [7].
One method is the application of nanotechnology, which reduces the size of the drug particle to the nanoscale. This increases the surface area and enhances the solubility and dissolution rate, leading to an increase in bioavailability [8]. Pharmaceutical nanoparticles (NPs) are solid except spongosomes, submicron-sized (<100 nm in diameter), and submicron-sized biodegradable drug carriers [9,10]. The latest nanotechnology technique plays a significant role in improving the water dispersibility, bioavailability, and therapeutic properties of compounds, particularly plant compounds [11,12]. The protection of drugs or biological compounds by incorporating them in the form of nanostructures to enhance the solubility and dissolution rate, leading to an increase in bioavailability, has been extensively investigated [13]. Furthermore, few colloidal carrier solutions have been reported to understand the problem of poor hydro-solubility and low bioavailability of active pharmaceutical ingredients. However, these colloidal carrier systems possess numerous drawbacks, such as inadequate physical stability, employment of residual organic solvents, drug expulsion during storage, low drug loading, and polymer cytotoxicity [14]. According to previous studies, the benefits of encapsulation of the drug compounds into zein polysaccharide core–shell nanostructures are specifically shown to possess excellent loading capacity and encapsulation efficiency, good dispersion in water, better stability in hard conditions, control release and improved bioavailability, and maintenance or improvement of biological activities of the encapsulated drug or biological compound [15,16,17]. According to Shuai et al., under in vitro GIT conditions, zein polysaccharide NPs significantly delayed the release and the photo- and thermal-degradation of the encapsulated curcumin and piperine [18]. Quercitin-loaded zein polysaccharide NPs demonstrate good physicochemical stability and biological activity. The resulting NPs showed good stability and slow release throughout simulated GIT digestion in vitro, and it was found that they have increased antioxidant and antibacterial activity [19]. Huang et al. incorporated curcumin into core–shell zein polysaccharide NPs by pH-induced antisolvent precipitation to improve the water dispersibility and antioxidant activity of curcumin [20].
Zein is a prominent biological carrier that has effective self-assembly and monodispersive characteristics when it comes to encapsulating bioactive components. The Food and Drug Administration (FDA), USA, has recognized zein as a safe human food ingredient that has the capacity to act as an edible carrier, making it a drug delivery system [21]. Many functional and nutritional ingredients have been successfully encapsulated in zein, and zein NPs have been synthesized to increase their physicochemical properties [22]. By embedding active compounds in the zein carrier, their microstructure, chemical stability, and bioavailability are considerably increased. All these efficacies make it a preferable option for developing NP delivery systems that encapsulate lipophilic medicinal compounds [21,22]. However, zein as a solitary material is often unsuitable for technological applications due to its weak water solubility and inability to effectively regulate the release of encapsulated substances. Widespread usage of polysaccharides as natural materials in the food business, which are very stable, biocompatible, and degradable. Polysaccharides may contribute to constructing nanocarriers with proteins through the following mechanisms: chemical cross-linking, physical cross-linking, poly-ion complex formation, and self-assembly [23]. Combining zein with polysaccharides to generate nanodelivery systems offers greater opportunities for the use of hydrophobic bioactivities than polymers alone [24]. Various polysaccharides, including chitosan, alginate, and pectin, have been combined with zein to develop nanodelivery systems. Silybum marianum is a plant that belongs to the family “Asteraceae” and is trivially known as milk thistle. It is native to Europe, Asia, and Northern Africa. Parts of plants can be consumed as raw vegetables and cultivated as ornamental plants. It is thought to be an ancient medicinal weed possessing ethno-pharmacological significance for the treatment of various liver ailments [25]. Many therapeutic compounds such as silymarin and their functional analogues like silybins, isosilybins, silychristin, and silydianin are documented to possess effective hepatoprotective properties by neutralizing the effects of harmful free radicals [26]. Along with this, silymarin has been reported to have excellent anti-arthritic, anti-cancer, anti-diabetic, anti-viral, antimicrobial, and immune-modulatory properties [27]. It has been reported that about 23% of the total silymarin is accessible to be absorbed from GIT due to very low aqueous solubility, hence resulting in very little bioavailability. Tea preparations of silymarin are less preferred due to their reduced hydro-solubility; therefore, oral encapsulated forms are readily administered by herbal practitioners to treat the ailment [28].
Keeping in view the importance of zein and silymarin, the present study aimed to develop zein–silymarin polysaccharide core–shell nanostructures to increase its water solubility that will enhance the bioavailability and biological responses in the living system.
2 Materials and methods
2.1 Plants collection
The S. marianum plants of white flowering and pink flowering were collected from two different regions of Peshawar (Hayatabad and University town Peshawar) and dried in shade separately. The plants and seeds were properly authenticated by Prof. Dr. Farhat Ali Khan and Prof. Dr. Muhammad Zahoor. The seeds of white and pink flowering plants were collected and ground into pulverized form (pulverization) separately. The ground samples were kept in a cool, dry place for the extraction procedure.
2.2 Extraction of oil
In blue-capped bottles, 250 g of powdered seeds of white and pink flowering S. marianum were soaked separately in 150 ml of n-hexane and kept at room temperature for 48 h. Each extract was then filtered by using filter paper and transferred into a rotary evaporator for oil extraction. Both the oils were then shifted into separate amber glass containers and stored at 25°C for further analysis. This oil extraction step is intended to obtain a comprehensive understanding of the overall composition of the seed extract, including lipids and all other non-polar components, and the percent yield by white and pink flowering S. marianum varieties.
2.3 Characterization of oil
2.3.1 Analysis by gas chromatography–mass spectrometry (GC–MS)
For active constituent’s detection, 1 µl of extracted oil was injected into the Thermo Scientific GC Focus Series DSQ model within the GC–MS instrument (Shimadzu-Lab) and undergone analysis utilizing a column with dimensions of 0.1 mm diameter, 0.5 μm film thickness, and 50 m length, using helium as the carrier gas flowing at a rate of 1 ml·min−1. The column temperature ranged from 80°C to 310°C at a rate of 10°C·min−1, facilitating the separation of silymarin constituents based on their volatility and interaction with the stationary phase, composed of a polar-embedded material polyethylene glycol (PEG).
2.3.2 Fourier transform infrared (FT-IR) analysis
FT-IR analysis was performed via IRTracer-100, Shimadzu, Japan, instrument to assess functional groups present in extracted oils at variable IR spectra, i.e., 400–45,000 cm−1. The analysis was performed utilizing a specific attenuated total reflectance (ATR) acquisition mode due to its suitability for analyzing oils directly without extensive sample preparation.
2.3.3 Extraction of silymarin
For silymarin extraction from both pink and white flowering S. marianum seeds, 70 g of dried powder of S. marianum seeds were dissolved in 120 ml of 70% ethanol in a blue-capped bottle. The solution was kept at room temperature for 48 h and then filtered into a flask using filter paper. Moreover, 10% (w/v) of NaCl solution was prepared and added to the ethanolic extract drop wise with continuous stirring till precipitation of silymarin occurred. Following that, the resulting precipitate was washed with distilled water to remove residual NaCl and impurities. The precipitates were collected by passing through filter paper, dried, and stored in a suitable container at 25°C for further use.
2.4 Preparation of silymarin–zein stock solution
The stock solution of the silymarin–zein complex was prepared by dissolving 2.0 g zein in 100 ml of ethanol (85% v/v) using a magnetic stirrer at 500 rpm (IKA R05, USA) for 1 h. Then, dried powdered silymarin (0.24 g) was added to the solution and stirred for 30 min in the dark. The insoluble impurities were removed from the solution through centrifugation at 1,006 × g for 15 min.
2.4.1 Preparation of pectin/sodium alginate solution
Aqueous solutions (0.2% w/v) of pectin alginate were prepared by mixing pectin and sodium alginate in different ratios, i.e., (100–0), (70–30), (50–50), (30–70), and (0–100). For proper dissolution, the solutions were stirred at 700 rpm using a magnetic stirrer for 30 min at 70°C. The solutions were then cooled in tap water and stirred for another hour. Non-soluble particles were removed from the solutions through filtration and then pH was adjusted at 4.0 with 1 mol·l−1 of HCL.
2.4.2 Preparation of silymarin-loaded zein-pectin/alginate core–shell NPs
Silymarin-loaded NPs were prepared following the literature method, specifically pH-induced antisolvent precipitation, with slight modification [29]. Initially, a silymarin–zein stock solution was prepared. About 20.0 ml of this stock solution was mixed with 80.0 ml of acidified water (pH 4.0) under constant stirring at 900 rpm for 3 min in the dark. This step aimed to disperse the silymarin–zein solution uniformly in the aqueous medium. The resultant dispersion solution underwent ethanol evaporation using a rotary evaporator (Model RE-2000A, Yarong Bio-instrument, Shanghai, China). This process removed ethanol from the solution, leaving behind a mixture of silymarin and zein in the aqueous phase. Following ethanol evaporation, the remaining solution was adjusted to a final volume of 100 ml by adding more acidified water (pH 4.0) with constant stirring at 900 rpm for 3 min in the dark. This step ensured the desired concentration and pH for subsequent reactions. The synthesis of the core–shell NPs involved combining portions of this solution with pectin–alginate solutions to form the desired silymarin-loaded zein–polysaccharide core–shell NPs.
2.5 Characterization of silymarin-loaded zein polysaccharide core–shell nanoparticles (SZPCS-NPs)
The prepared SZPCS-NPs were then passed through different sophisticated instruments for characterization.
2.5.1 Characterization by transmission electron microscope (TEM) and scanning electron microscope (SEM)
The size and morphology of SZPCS-NPs were analyzed via TEM (FEI Techni, G2, 300 kV) and SEM (Hitachi-s 3400 N, Japan). A thin film of test solution was prepared on the carbon-coated copper grid, which was then vaporized using a mercuric vapor lamp for 5 min. Finally, a 2-D micrograph of the test NPs manifesting their size was observed through TEM. Furthermore, a sample of SZPCS-NPs was placed on a copper-coated grid and allowed to dry. After drying, the SEM was operated at a high voltage of 100 kV. SEM images of prepared SZPCS-NPs, which were prepared in different concentrations of pectin and sodium alginate solution, were depicted.
2.5.2 Characterization by X-ray diffractometer (XRD)
The nature of SZPCS-NPs was analyzed through an XRD (Jeol JDX-3532, Japan), which contains an Ni filter and CuKα as a source of monochromatic radiation at 1.5418 A° wavelength. The machine was operated at 40 kV voltage and 30 mA current. For scanning, 2θ/θ range was selected, and for precise determination, the scanning speed of 10 min−1 was used.
2.5.3 Characterization by FT-IR
For the identification of functional groups, the SZPCS-NPs were subjected to infrared spectroscopic analysis. Infrared spectrometer Prestige FT-IR spectrophotometer (IRTracer-100, Shimadzu, Japan) was used for this purpose. The wavelength used for analysis was set in the 4,000–600 cm−1 range.
2.5.4 Particle size and charge measurements
The zeta potential of the particle size distribution of the colloidal dispersions was measured by dynamic light scattering using a commercial instrument (Malvern Zetasizer Nano-ZS, Malvern Instruments, Worcestershire, UK). The Z-average diameter and polydispersity index (PDI) were calculated from the light scattering measurements. Each sample was analyzed in duplicate. The electrical properties (ζ-potential) of the particles in the colloidal dispersions were characterized using a micro-electrophoresis device (Malvern Zetasizer Nano-ZS, Malvern Instruments, Worcestershire, UK). Colloidal dispersions were diluted with 5 mM phosphate-citric acid buffers at the same pH as the samples prior to analysis to avoid multiple scattering effects. Each sample was analyzed in duplicate.
2.6 Antioxidant potential of SZPCS-NPs
2.6.1 2,2-Diphenyl-1-picrylhydrazyl (DPPH) scavenging activity
The antioxidant activity of SZPCS-NPs was evaluated using literature with slight modifications [30]. For this purpose, differently concentrated ethanolic solutions of the encapsulated silymarin NPs as well as of pure silymarin were prepared. Then, 100 μM DPPH solutions in ethanol were prepared. From the DPPH–ethanol solution, 2 ml were mixed with equivalent volumes of the differently concentrated ethanolic solutions of encapsulated silymarin and pure silymarin. The resulting solutions were subsequently incubated for 30 min in the dark at room temperature. After 30 min, the solutions were filtered and subjected to spectrophotometric absorbance at 517 nm. The free radical scavenging activity from the sample was calculated using Eq. 1:
where A s and A c are the absorbance of sample and control (ethanol) solutions, respectively. The scavenging activity (SC50) was used to compare the levels of free radical scavenging activity of different samples, which was taken to be the concentration of sample required to reduce the scavenging by 50%. The experiment was performed in triplicate.
2.6.2 ABTS·+ scavenging capacity
The scavenging activity towards the ABTS˙+ radical was measured by modifying a previously used methodology. First, the ABTS˙+ radical was produced by the reaction of ABTS stock solution (7 mM) with 2.45 mM of potassium persulfate in darkness at room temperature for 16 h before use. Subsequently, radical ABTS˙+ solution was diluted with buffer (5 mM, pH 7.4) to adjust the absorbance to 0.700 ± 0.002 AU measured at 734 nm wavelength and 30°C. The mixtures in the 96-well microplate contained the following reagents: 250 µl of ABTS˙+ radical solution and 100 µl of S. marianum extract serially diluted from 0.014 to 0.0014 µg·ml−1 in 50% aqueous EtOH was incubated at room temperature in the absence of the light for 45 min. The absorbance readings were measured at 734 nm and 30°C. Trolox was used as the reference standard. ABTS˙+ inhibition percentage from S. marianum extracts was calculated according to the formula:
where ABTS control is the absorbance of ABTS˙+ radical in buffer at t = 0 min; A sample is the absorbance of an ABTS˙+ solution mixed with extracts, and A sample blank is the absorbance of samples without ABTS˙+. Comparisons between extracts were done by IC50 value representing the concentration.
2.7 Hepatoprotective activity
Silymarin and SZPCS-NPs were formulated in carboxymethyl cellulose (0.5%) to be administered through the oral route (oral gavages).
2.7.1 Design of study
Adult rats of both sexes weighing 200–220 g were indiscriminately allocated into 5 equal groups (A–E), having 15 rats in each class. Each group was allocated to separate cages and labeled accordingly. Seven days of acclimatization were allowed. The therapeutic groups, denoted as Groups C, D, and E, received a dose of 200 mg·kg−1 body weight of silymarin, a dosage selected based on a comprehensive review of literature and established practices in hepatoprotection research [31,32]. The rats were euthanized using the physical method of cervical dislocation. No anesthetic agents were employed in the euthanasia procedure [33].
Group A. Normal untreated control
Group B. Carbon tetrachloride (CCl4)-treated control
Group C. Commercial silymarin
Group D. Crude silymarin
Group E. SZPCS-NPs
2.7.2 Biochemical assay of blood
To compare the hepatoprotective effects of crude silymarin, commercially available and SZPCS-NPs, blood biochemical tests were carried out every 2, 4, and 6 weeks. Rats were anesthetized, and 2 ml of blood was drawn by cardiac puncture from every experimental rat and transferred into labelled tubes. Samples were made to spin to separate as much serum as possible. The collected serum was kept at 4°C until it was time to use it for the standard test procedures for measuring the parameters of the liver function tests (LFTs).
2.7.3 Biochemistry of blood
Serum biochemical parameters, such as amino alanine transferase (ALT), serum alkaline phosphatase (ALP), serum albumin, total proteins, serum globulin, urea, and creatinine, were determined through commercially accessible kits of RANDOX.
-
Ethical approval: The research related to animals’ use has been complied with all the relevant national regulations and institutional policies for the care and use of animals.
3 Results and discussion
3.1 Determination of oil content from the seed (capitulum) of S. marianum
From 1 kg of pink flower plant seed, it was observed that 24.3 ml of oil was extracted, and from 1 kg of white flower plant seed, 22.7 ml of seed was extracted. It is evident that the difference is 1.6 ml, so there is no significant difference between these two different flowering seeds of S. marianum.
Lahlah et al. studied the estimation of active principle pharmacological entities of white and blue capitulum and the whole plant of S. marianum. They also found no significant difference and concluded that S. marianum (blue and white flowering varieties) showed nearly the same values of phytochemical constituents and possessed similar antimicrobial activities [34]. Shaker et al. also concluded that there was less phenotypic difference in purple and white varieties of S. marianum [35].
3.2 GC–MS analysis of oil
The composition analysis of oil extracted from S. marianum seeds – both white and pink flowering varieties – was conducted using GC–MS. The GC–MS spectra for both extracts are illustrated in Figure 1a (white flowering seeds) and Figure 1b (pink flowering seeds). In Figure 1a, the GC–MS analysis revealed prominent peaks at specific retention times (RTs) of 18.85, 20.28, 21.97, and 25.96. These peaks corresponded to significant constituents present in the S. marianum oil extracted from white flowering seeds. The respective peak areas associated with these RTs were 56,066,569.30, 120,574,395.29, 56,546.66, and 8,915,044.66, indicating the relative abundance of these compounds within the oil extract. In contrast, Figure 1b depicts the GC–MS analysis of S. marianum oil obtained from pink flowering seeds, exhibiting major peaks at RTs of 18.89, 20.27, 20.32, 21.99, and 25.98. These RTs signify the presence of significant chemical components within the pink flowering seed extract. The corresponding peak areas for these RTs were measured at 13,987,946.94, 176,668,263.01, 63,484.01, 952,602.82, and 2,663,794.40, indicating the relative quantities of these compounds in the oil extract from the pink flowering seeds.

GC–MS of oil extracted from (a) white and (b) pink flowering seeds of S. marianum.
Additionally, the exact composition of the oil obtained from S. marianum has previously been reported in the literature. In the fatty acid fraction, for example, linoleic (50.5%) and oleic (30.2%) acids were found to be predominant. In contrast, essential oil contains 14 volatile components, with the two main ones being γ-cadinene (49.8%) and α-pinene (24.5%). The main ingredients in the seed oil extracts were linoleic acid (43.5–64.4%), oleic acid (20.8–29.8%), palmitic acid (7.2–9.7%), and stearic acid (2–6.6%). A dwarf S. marianum genotype from India was used to produce around 20% of the yellow oil, with the seeds having 9.7% palmitic, 4.7% stearic, 30.9% oleic, 45.8% linoleic, and 5.2% linolenic acids [36].
The comparison of the major peaks observed through GC–MS analysis between the white and pink flowering seed extracts provides insights into the differences in chemical composition and relative abundance of key compounds present in the two varieties of S. marianum seeds. Our GC–MS results closely correspond with the results of the GC–MS analysis of Iraqi S. marianum flowers, leaves, and seed extracts by Eldalawy and Al-ani [37].
3.3 FT-IR analysis of oil
Figure 2a and b represents the FT-IR analysis of oil derived from S. marianum of two distinct flowering seeds, namely pink and white flowers. FT-IR of oil extracted from S. marianum (pink flower capitulum seeds) displays strong absorption bands at 3,009.1, 2,923.4, 2,886.46, 2,853.8, 1,737.9, 1,710.7, 1,557.6, 1,463.6, 1,243.8, 1,182.7, and 466.0 cm−1. FT-IR of oil extracted from S. marianum (white flower capitulum seeds) displays strong absorption bands at 3,009.0, 2,923.4, 2,853.8, 1,737.9, 1,710.0, 1,556.8, 1,463.5, 842.7, and 466.4 cm−1. The existence of saturated hydrocarbon chains is indicated by strong absorption bands at around 3,009.1, 2,923.4, 2,886.46, and 2,853.8 cm−1. These bands most likely correspond to the C–H stretching vibrations of aliphatic hydrocarbons. The stretching vibration of carbonyl (C═O) groups, perhaps originating from esters or other compounds containing carbonyl, might be represented by the absorption band at 1,737.9 cm−1. The absorption bands at 1,557.6 cm−1, 1,463.6, 1,243.8, 1,182.7, and 466.0 cm−1 may be indicative of aromatic chemicals or unsaturated hydrocarbons that have C═C stretching. The bands located at 1,243.8 and 1,182.7 cm−1 may be linked to C–O stretching, which might suggest the existence of alcohols or esters. The band at 710.7 cm−1 may be associated with aromatic compounds or substituent out-of-plane bending vibrations.

FT-IR spectra of oil extracted from seeds of (a) white and (b) pink flowering S. marianum.
Both the FT-IR spectra of white and pink flowering S. marianum seeds oil show approximately the same values as already reported by Fadhil et al. [38]. Our results are in close accordance with the study conducted by Radusheva et al., in which they studied the physicochemical characteristics of seed oil of different medicinal plants, including S. marianum [39].
3.4 Quantification of silymarin content in the seeds of S. marianum
From 1 kg of pink flower seed, it was observed that 38.6 g of silymarin was extracted, and from 1 kg of white flower seed, 35.7 g of silymarin was extracted. It can be observed that the difference is only 2.9 g, so there is no significant difference between these two different flowers of S. marianum. No significant difference was found among S. marianum blue and white varieties and they possess almost similar amounts of phytochemical constituents. It has been reported that the contents of silymarin in blue and white capitulum seeds of S. marianum show a very negligible variation, and the difference was noted as insignificant [40]. Studies on blue and white flower varieties of S. marianum with special reference to the silymarin content were carried out by Maryam et al., which confirmed nearly the same values of silymarin content in the seeds of both S. marianum varieties [41].
3.5 FT-IR analysis of silymarin
FT-IR analysis of silymarin extracted from seeds of S. marianum of two different flowerings, i.e., pink and white flowers, is shown in Figure 3a and b, respectively. FT-IR of silymarin extracted from S. marianum (pink flower capitulum seeds) displays strong absorption bands at 3,449.72, 2,938.63, 1,686.91, 1,634.12, 1,542.14, 1,412.57, 1,160.63, and 717.50 cm−1 (Figure 3b). FT-IR of silymarin extracted from S. marianum (white flower capitulum seeds) displays strong absorption bands at 3,465.0, 2,997.6, 1,670.2, 1,639.9, 1,543.5, 1,412.57, 1,314.15, 1,269.8, 913.68, and 809.48 cm−1 (Figure 3a). According to Mukherjee et al., FT-IR analysis of silymarin extracted from S. marianum displayed strong absorption bands at 2,849, 1,733, 1,636, and 1,509 cm−1 [42]. In addition, the spectrum of silymarin demonstrated by Maryana et al. showed bands at 3,550, 3,001, and 1,776 cm−1 [43]. According to Maryana et al., the FT-IR spectrum of silymarin showed strong bands at 3,550, 3,001, and 1,776 cm−1 whose intensity was decreased or even disappeared in the complex form [44]. Additionally, vibrational bands of silymarin were observed at 1,633, 1,510, and 1,271 cm−1, which correspond to –C–O (in-ring), C–C (in-ring), and C–O of polyol group, respectively [45].

FT-IR spectra of silymarin extracted from (a) white and (b) pink flowering S. marianum seeds.
3.6 Preparation of SZPCS-NPs
SZPCS-NPs were prepared by using different ratios of pectin and alginate solutions and then characterized by TEM, XRD, and FT-IR to select the best ratio.
3.7 Characterization of SZPCS-NPs
SZPCS-NPs were analyzed by TEM, XRD, and FT-IR. Further SZPCS-NPs were evaluated for antimicrobial, antioxidant, and hepatoprotective activities. In the literature, zein–NPs have been successfully synthesized for entrapment of medicinal entities like Curcuma Longa (curcumin) and Quercitin [46,47,48].
3.7.1 TEM analysis
In this study, different ratios of alginate and pectin were used to form the shell around zein NPs to get uniform size and polydispersity. Figure 4a–e shows TEM images of SZPCS-NPs produced in various concentrations of pectin and sodium alginate solution. The dark areas in the photos represent the size of the NPs. According to the TEM study in Figure 4a, the size of silymarin NPs produced in a polysaccharide solution with a pectin ratio of 100% is uneven and not uniform in size ranges of 20–80 nm. Figure 4b demonstrates the best results, the shape of the created NPs was spherical and uniform in size, around 80 nm in diameter (average), and they were made in a 70% pectin and 30% alginate (P70–A30) solution. Furthermore, in Figure 4c, SZPCS-NPs were prepared in a ratio of 50% pectin and 50% sodium alginate solution, possessed an average diameter of 50 nm; in Figure 4d, SZPCS-NPs were prepared in P70–A30 solution possessed an average diameter of 70 nm; and in Figure 4e, the NPs had an irregular shape with an average diameter of 80 nm. Our results are similar to previous studies. Muzamil et al. determined the morphology for both NP1 and NP3. In the case of both NP1 and NP3, small rounded or spherical-shaped morphology with regular shapes was noticed. The particles were highly monodispersed [46]. Huang et al. also used a combination of alginate and pectin to form the shell around zein NPs. They concluded that the replacement of 30% of pectin with alginate greatly improved aggregation stability at pH 5–7. It was observed that the shape of the prepared NPs was spherical and uniform in size, about 80 nm in diameter [20].

TEM images of ZSPCS-NPs made with different ratios of pectin and alginate. (a) 50% pectin and 50% alginate solution, (b) 70% pectin and 30% alginate solution, (c) 30% pectin and 70% alginate solution, (d) 100% alginate solution, and (e) 100% pectin solution.
3.7.2 SEM analysis
Figure 5a–e shows the SEM images of prepared SZPCS-NPs, which were prepared in different concentrations of pectin and sodium alginate solution. The white spots indicate the size of NPs in the images. According to SEM analysis in Figure 5a, the size of SZPCS-NPs prepared in a polysaccharide solution with a pectin ratio of 100% is irregular and not uniform in size, ranging from 20 to 100 nm. While Figure 5b shows the best results, it was observed that the shape of the prepared SZPCS-NPs was spherical and uniform in size, about 80 nm in diameter, and that they were prepared in P70–A30 solution. Moreover, in Figure 5c, in which the SZPCS-NPs were prepared in a ratio of 50% pectin and 50% sodium alginate solution, in Figure 5d, where the SZPCS-NPs were prepared in P70–A30 solution, and in Figure 5e, where the SZPCS-NPs were synthesized in 100% sodium alginate solution, they also showed an irregular shape; some triangle and oval shapes were also observed; and in the above-mentioned figures, the NPs were not all of the same size, ranging from 20 to 100 nm. Paul and Yadav prepared silver NPs having 50 nm size. On SEM analysis, it was observed that the generated silver NPs were spherical in shape with a size ranging from 3 to 36 nm [49]. Silybum et al. synthesized the silymarin-loaded TiO2 NPs having an average size of 20 nm. The size of our prepared SZPCS-NPs was a little larger than silymarin-loaded TiO2 NPs [50]. While silibinin-loaded nanoparticles (SILNPs) prepared by Guhagarkar et al. were ranging from 60 to 70 nm [51], Sajadi et al. synthesized silymarin-loaded Cu/Fe3O4 NPs having spherical shape and size ranging from 8.5 to 60 nm [52]. Li et al. demonstrated that plain zein colloidal NPs showed a spherical shape and smooth surface with diameters ranging from 80 to 100 nm [49]. Abbasi et al. determined the size and surface morphology of the prepared ZnO-NPs, P-ZNPs, C-ZNPs, and S-ZNPs of S. marianum extract through SEM analysis. The average size of ZnO NPs evaluated from SEM analysis was 61.3, 51.7, 64, and 61.9 nm synthesized for W-ZNPs, P-ZNPs, C-ZNPs, and S-ZNPs, respectively [53]. The average size of silymarin-zein NPs (SMN-Zein) evaluated from SEM analysis was around 100 nm in diameter, having spherical and uniform size [54].

SEM images of ZSPCS-NPs made with different ratios of pectin and alginate. (a) 50% pectin and 50% alginate solution, (b) 70% pectin and 30% alginate solution, (c) 30% pectin and 70% alginate solution, (d) 100% alginate solution, and (e) 100% pectin solution.
3.7.3 XRD analysis
Powder XRD was used to verify the crystallinity and phase distribution of the synthesized SZPCS-NPs. Figure 6a–e represents the photograph of the XRD pattern. Sharp peaks for SZPCS-NPs (P100–A00) were recorded at 20.0, 25.0, and 40.0 (Figure 6a). For these NPs, the presence of distinct peaks at 20.0, 25.0, and 40.0 implies a particular crystalline structure. The peaks for SZPCS-NPs (P70–A30) were recorded at 30.0, 39.0, and 45.0, showing sharper peaks, although some additional peaks were also noted at 20.5, 21.0, and 39.0. Increased peak sharpness at 30.0, 39.0, and 45.0 indicates a distinct crystalline structure. Extra peaks at 20.5, 21.0, and 39.0 indicate possible differences in phase distribution as well as composition (Figure 6b). The peaks for SZPCS-NPs (P50–A50) were recorded at 25.0, 28.0, and 31.5 (Figure 6c). The peaks located at 25.0, 28.0, and 31.5 indicate a characteristic crystalline pattern; however, it may not be as prominent as in other compositions as mentioned above. The peaks for SZPCS-NPs (P30–A70) were recorded at 13.5, 20.0, 30.0, and 45.0. In contrast to other compositions, peaks at 13.5, 20.0, 30.0, and 45.0 imply a distinct crystalline structure, perhaps signifying changes in the composition or crystalline phases present (Figure 6d). Major peaks for SZPCS-NPs (P00–A100) were recorded at 29.5, 31.5, and 45.0. Prominent peaks at 29.5, 31.5, and 45.0 have a particular crystalline structure, which may suggest an alternate phase distribution in this structure (Figure 6e).

XRD spectra of ZSPCS-NPs made with different ratios of pectin and alginate. (a) 50% pectin and 50% alginate solution, (b) 70% pectin and 30% alginate solution, (c) 30% pectin and 70% alginate solution, (d) 100% alginate solution, and (e) 100% pectin solution.
P70–A30 exhibits sharper and extra peaks among the five different types of SZPCS-NPs, which may indicate a more complicated structure and composition. It might serve effectively in situations when a combination of multiple phases or enhanced properties is required. Given that, the goal of the present study was to improve water solubility and dispersibility of silymarin, the XRD pattern clearly indicated that P70–A30 provided the most favorable outcomes. Abbasi et al. studied the XRD pattern of S. marianum extract-loaded NPs, namely W-ZNPs, S-ZNPs, C-ZNPs, and P-ZNPs, peaks for W-ZNPs were recorded at 31.6, 34.2, 36.2, 47.3, 56.4, 62.7, 67.8, and 68.9, while peaks for S-ZNPs were recorded at 31.7, 34.3, 36.1, 47.4, 56.5, 62.8, 66.2, 67.9, 69.1, and 76.9. C-ZNPs showed peaks at 31.5, 34.2, 36.0, 47.3, 56.3, 62.5, 67.8, and 68.9. P-ZNPs showed peaks at 31.6, 34.3, 36.1, 47.3, 56.4, 62.7, 66.2, 67.8, 69.0, 72.4, and 76.8 [53]. The XRD pattern of Fe3O4 S. marianum NPs showed characteristic peaks at 2θ = 30.5, 35.8, 43.2, 53.7, 57.2, and 62.9 [55]. Similarly, the XRD pattern of silymarin-loaded Ag-NPs distinguished diffraction peaks at 2θ of 38.5° (111), 44.1° (200), 64.5° (220), and 77.3° (311) that strongly indicate the face-centered cubic crystalline structure of Ag-NPs [56].
3.7.4 FT-IR analysis
Figure 7a shows that the FT-IR spectrum of SZPCS-NPs displays strong absorption bands at 3,264.3, 2,923.0, 1,729.3, 1,610.4, 1,526.7, 1,195.1, 1,010.1, and 861.5 cm−1. Figure 7b demonstrates that the FT-IR spectrum of SZPCS-NPs displays strong absorption bands at 2,997.6, 1,603.4, 1,492.2, 1,217.7, 1,017.1, 853.6, and 601.8 cm−1. Figure 7c reveals that the FT-IR spectrum of SZPCS-NPs displays strong absorption bands at 2,923.0, 2,849.1, 1,595.6, 1,538.5, 1,195.7, 1,017.1, and 846.8 cm−1. Figure 7d shows that the FT-IR spectrum of SZPCS-NPs displays strong absorption bands at 2,916.01, 2,856.9, 1,610.4, 1,484.44, 1,017.1, and 853.61 cm−1. Figure 7e shows the FT-IR spectrum of SZPCS-NPs displays strong absorption bands at 3,220.0, 2,916.0, 2,856.9, 1,751.1, 1,603.4, 1,543.5, 1,239, 846.81, and 609.6 cm−1.

Overlaped images of FTIR spectra of ZSPCS-NPs made with different ratios of pectin and alginate. (A) 50% pectin and 50% alginate solution, (B) 70% pectin and 30% alginate solution, (C) 30% pectin and 70% alginate solution, (D) 100% alginate solution, and (E) 100% pectin solution.
The FT-IR spectrum obtained for silymarin Zein (SMN-Zein) NPs showed characteristic peaks at 1,693 and 1,702 cm−1 [57]. Our results were pretty much closure with that of Gopalakrishnan and Raghu, and the FT-IR spectrum of gold NPs of S. marianum extract prepared during their study displayed strong absorption bands at 3,416, 2,924, 1,630, 1,383, and 1,031 cm−1 and the spectrum of silver NPs of the S. marianum extract displays strong bands at 3,421, 2,925, 1,654, 1,399, 1,102, and 706 cm−1 [58]. Iqbal et al. determined that the FT-IR spectra of Cu NPs of S. marianum produced during polymerization exhibit peaks at 2,163 and 1,990 cm−1. The peak at 1,004 cm−1 may be associated with ether, pyranose ring, and glycosidic linkage, as demonstrated by the C–O(C–O–C) stretch. Cu–O stretching is observed by the peaks at the fingerprint areas at 518, 523, 528, and 542 cm−1 [59]. In addition, Gopalakrishnan et al. determined the FT-IR analysis of S. marianum seed extract, which shows prominent peaks located at 3,415, 2,925, 1,654, 1,383, and 1,115 cm−1. While S. marianum-based Au–Ag bimetallic nanocomposites exhibit robust absorbance bands at 3,423, 2,918, 1,603, 1,383, 1,105, and 824 cm−1. Strong and wide bands at 3,415 and 3,423 cm−1 are the stretching vibrations of the O–H group in alcohols and phenols that are intermolecular hydrogen bound. They can also be attributed to the stretching of the hydrogen-bonded N–H group of proteins [55]. For chemically synthesized S. marianum-based ZnO-NPs, sharp absorption peaks were seen at 3,220.63, 1,551.90, 1,415.86, 1,088.28, 1,033.43, 848.48, 776.23, 716.20, 620.72, and 610.06 cm−1. For WPE-mediated ZnO-NPs, intense absorption peaks were seen at 3,244.37, 2,930.31, 1,416.63, 1,088.89, 1,036.02, 854.96, 778.42, 715.84, and 611.70 cm−1 [60].
3.7.5 Analysis by thermal gravimetric analysis (TGA)
The thermal stability and mass loss of SZPCS-NPs were monitored using thermogravimetry differential thermal analysis in the temperature range of 0–800°C (Figure 8). The prepared NPs from selected plants were thermally stable ≤200°C; the loss in particle aptitude and mass was witnessed by a gradual increase in temperature (200–800°C), thus affirming it as a thermally sensitive compound.

Overlapped image of TG/DTA of ZSPCS-NPs made with different ratios of pectin and alginate. (A) 50% pectin and 50% alginate solution, (B) 70% pectin and 30% alginate solution, (C) 30% pectin and 70% alginate solution, (D) 100% alginate solution, and (E) 100% pectin solution.
The thermal stability of SZPCS-NPs (P100–A00) sample weight 6.5 mg was evaluated via TGA by subjecting the NPs to a temperature range of 0–800°C. Major weight loss was observed at 280–380°C, and loss of weight in this event was ∼75%.
In Figure 8, A–D represents the thermal stability of SZPCS-NPs. (A) represents the thermal stability of SZPCS-NPs (P70–A30) sample weight 6.5 mg was evaluated via TGA by subjecting the NPs to a temperature range 0–800°C. Major weight loss was observed at 220–320°C; the loss of weight in this event was ∼75%. (B) The thermal stability of SZPCS-NPs (P50–A50) sample weight 6.5 mg was evaluated via TGA by subjecting the NPs to a temperature range of 0–800°C. Major weight loss was observed at 200–310°C, which is attributed to the decomposition of the compound, while the loss of weight in this event was ∼70–78%. The minor decomposition occurred between 500°C and 700°C. (C) The thermal stability of SZPCS-NPs (P30–A70) sample weight 6.5 mg was evaluated via TGA by subjecting the NPs to a temperature range of 0–800°C. Initial weight loss was observed at 200–310°C, which is attributed to the loss of moisture, while the loss of weight in this event was ∼50%. The minor decomposition occurred between 500°C and 700°C. (D) The thermal stability of SZPCS-NPs (P0–A100) sample weight 6.5 mg was evaluated via TGA by subjecting the NPs to a temperature range of 0–800°C. Initial weight loss was observed at 200–400°C, which is attributed to the loss of moisture, while the loss of weight in this event was ∼70%. The second minor decomposition occurred between 500°C and 700°C. (E) A minor weight loss was observed for all SZPCS-NPs below 100°C, which is attributed to the loss of moisture.
According to these results, SZPCS-NPs maintained their thermal stability up to 160°C. Nonetheless, the loss of particle mass and aptitude increased gradually over 200°C. This means that the compound loses bulk and structural integrity when the temperature rises over 200–800°C since it becomes more sensitive to heat. P70–A30 maintains an appropriate balance between stability and decomposition temperatures. Many applications might benefit from this balance, particularly those that need stability in temperatures that are moderately high. According to the results, P70–A30 appears to be a promising option that maintains a balance between thermal stability and the beginning of decomposition. As such, it is a candidate that is worth researching further for certain applications where this balance is important. In the meanwhile, P30–A70 exhibits little degradation at higher temperatures and moderate stability with initial moisture loss (200–310°C). Even though P100–A00 loses weight at temperatures between 280°C and 380°C, it may still be taken into consideration if greater thermal stability is a top concern. P50–A50, on the contrary, decomposes at lower temperatures (200–310°C) and loses weight at a faster rate (70–78%), making it less suitable for applications that demand strong stability. Abbasi et al. also studied the thermal stability of the green synthesized silymarin NPs via TGA by subjecting the NPs to a temperature of 0–1,100°C. Weight loss was observed at 130–140°C [61].
3.7.6 Analysis of zeta potential
Figure 9a shows zeta potential of SZPCS-NPs (P100–A0). The zeta potential analysis of SZPCS-NPs (P100–A0) revealed a % intensity of 15> and a particle size ranging below 100 nm. Figure 9b shows zeta potential of SZPCS-NPs (P70–A30). The zeta potential analysis of SZPCS-NPs (P70–A30) revealed a % intensity of 15> and a particle size ranging from 90 to 100 nm. The observed zeta potential suggests a suitable surface charge for stability. Figure 9c shows the zeta potential of SZPCS-NPs (P50–A50). The zeta potential analysis of SZPCS-NPs (P50–A50) revealed an intensity of 16% and a particle size ranged from 100 nm. Figure 9d shows the zeta potential of SZPCS-NPs (P70–A30). The zeta potential analysis of SZPCS-NPs (P70–A30) revealed an intensity of the NPs of 17%, indicating the relative strength of their scattering signals. Furthermore, the size distribution of the NPs ranged from 100 nm. Figure 9e shows the zeta potential of SZPCS-NPs (P0–A100). The zeta potential analysis of SZPCS-NPs (P0–A100) revealed an intensity of the NPs of 17%, indicating the relative strength of their scattering signals. Furthermore, the size distribution of the NPs ranged from 100 nm. According to Mohammadi et al., the particle size of silymarin-loaded silver NPs was 1–25 nm, having a zeta potential of −24.2 mV [62]. According to Ma et al., the zeta potential analysis of silymarin-loaded self-assembled NPs of Bletilla striata polysaccharide revealed a % intensity of 12 > and a particle size ranging from 50 to 200 nm [63]. The average particle size of gold and silver NPs was found to be 120 and 64 nm, having a zeta potential distribution of −15.4 and −15.8, respectively [55].

Zeta potential of SZPCS-NPs made with different ratios of pectin and alginate. (a) 50% pectin and 50% alginate solution, (b) 70% pectin and 30% alginate solution, (c) 30% pectin and 70% alginate solution, (d) 100% alginate solution, and (e) 100% pectin solution.
Overall, based on the previous characterization results, SZPCS-NPs that had a 70% pectin/30% alginate (P70–A30) were selected for further assessments, including antioxidant activity and hepatoprotective activity. The SZPCS-NPs (P70–A30) demonstrated the best results; the shape of the created NPs was spherical and uniform in size, around 80 nm in diameter. The XRD analysis suggests that the P70–A30 formulation may have the best combination of crystallinity and phase distribution among the tested formulations, as it exhibited sharper peaks compared to P100–A00 and P50–A50. In addition, the FT-IR analysis demonstrated a prominent absorption band at 2,997.6 cm−1, indicating C–H stretching vibrations. The band at 1,603.4 cm−1 suggests the presence of C═C stretching vibrations. Peaks at 1,492.2, 1,217.7, 1,017.1, 853.6, and 601.8 cm−1 may correspond to other functional groups.
3.7.7 % Drug loading and % entrapment efficiency
The data provided represent the percent drug loading and percent entrapment efficiency of silymarin into SZPCS-NPs at different time points, as derived from the percent absorbance measurements (Table 1).
Percent drug loading and percent entrapment efficiency of silymarin into zein silymarin NPs at different time points, as derived from the percent absorbance measurements
% Absorbance | ||
---|---|---|
Time (min) |
|
|
10 | 54.67 | 38.54 |
20 | 65.43 | 44.67 |
30 | 72.43 | 51.78 |
40 | 72.42 | 51.54 |
50 | 72.55 | 51. 46 |
60 | 72.46 | 51.76 |
After 10 min, the percent drug loading was found to be 54.67%, indicating the amount of silymarin loaded into SZPCS-NPs. Simultaneously, the percent entrapment efficiency was calculated to be 38.54%, representing the efficiency of trapping or encapsulating silymarin within the NPs at this time point. At 20 min, there was an increase in both percent drug loading and percent entrapment efficiency. The percent drug loading reached 65.43%, indicating a higher amount of silymarin incorporated into SZPCS-NPs. The percent entrapment efficiency also increased to 44.67%, suggesting an improved efficiency in trapping silymarin within the NPs.
The trend of increasing percent drug loading and percent entrapment efficiency continued at subsequent time points. After 30 min, the percent drug loading was 72.43%, while the percent entrapment efficiency was 51.78%. After 40 min, the values were 72.42% and 51.54% for percent drug loading and percent entrapment efficiency, respectively. At 50 min, the percent drug loading and percent entrapment efficiency were measured to be 72.55% and 51.46%, respectively. Similarly, at 60 min, the values were 72.46% for percent drug loading and 51.76% for percent entrapment efficiency. These findings indicate that the zein silymarin NPs effectively loaded silymarin, with the percent drug loading consistently increasing over time. The percent entrapment efficiency remained relatively stable throughout the studied time points, suggesting consistent trapping of silymarin within the NPs.
Overall, the results highlight the capability of SZPCS-NPs to efficiently load and entrap silymarin, offering a promising approach for the delivery and controlled release of silymarin in various applications. Our observations regarding the entrapment efficiency align with previous studies in the field. A study conducted by Gupta et al. investigated the loading and encapsulation efficiency of silymarin chitosan NPs. Entrapment efficiency of silymarin chitosan NPs was observed in the range of 52.68–85.28% [64]. In addition, solid lipid NPs of silymarin possessed % drug loading of 79.5–92.4% and entrapment efficiency of 83.9–92.3%, respectively [65].
Eqs. 3 and 4 were used to determine drug loading and encapsulation, efficiency, respectively.
4 Potential activities of SZPCS-NPs (P70–A30)
4.1 Antioxidant activity
Finally, the impact of the encapsulation of silymarin within SZPCS-NPs on its antioxidant activity was determined. This kind of analysis is important because any strategy used to improve the water dispersibility and physical stability of silymarin should not adversely affect its functional attributes. In these experiments, SZPCS-NPs that had a P70–A30 coating were prepared since this system gave the best physical stability and were compared with the antioxidant activity of commercially available silymarin and crude silymarin.
4.1.1 DPPH scavenging capacity
The percent absorbance of commercially available silymarin, crude silymarin, and SZPCS-NPs at a sample concentration of 10 µl·ml−1 was recorded at 39.2%, 37.8%, and 81.8%, respectively (Figure 10). Similarly, at sample concentration 20 µl·ml−1, the % absorbance (517 nm) of commercially available silymarin, crude silymarin, and SZPCS-NPs were 43.4%, 41.7%, and 91.3%, correspondingly. In addition, at sample concentrations 30 and 60 µl·ml−1, the % absorbance (517 nm) of commercially available silymarin, crude silymarin, and SZPCS-NPs were 42.2%, 39.9%, 92.1%, and 62.8%, 58.2%, and 96.3%, individually. Finally, at sample concentrations 70 and 80 µL·ml−1, the % absorbance (517 nm) of commercially available silymarin, crude silymarin, and SZPCS-NPs were 66.3%, 63.4%, and 97.8% and 67.0%, 65.8%, and 98.8%, respectively. An aliquot of the samples was mixed with DPPH solution (5 ml, 23.6 μg·ml−1 in ethanol), followed by incubation of 30 min. It can be concluded from the results that zein silymarin NPs consistently exhibit the highest percent absorbance at all sample concentrations. This indicates that the SZPCS-NPs possess higher solubility or better optical properties at the tested wavelength compared to the other forms of silymarin. In addition, SZPCS-NPs showed better antioxidant activity compared to both commercially available silymarin and crude silymarin. The absorbance of each sample was read at 517 nm [66]. Seeds of S. marianum with purple flower showed 62.67%, 74.02%, and 78.34% activity after 10, 20, and 30 min, respectively, while white flowering seeds exhibited 37.79%, 42.97%, and 49.18% absorbance activity, respectively [34]. DPPH free radical scavenging activity of methanol extracts of S. marianum at sample concentrations 0.25, 0.05, 0.1, 0.25, and 0.5 µg·ml−1 was 25%, 30%, 55%, 75%, and 85%, respectively [66]. Seeds of S. marianum with purple flower showed 62.67%, 74.02%, and 78.34% activity after 10, 20, and 30 min, respectively, while white flowering seeds exhibited 37.79%, 42.97% and 49.18% activity, respectively [67].

DPPH scavenging capacity of commercially available silymarin, crude silymarin, and SZPCS-NPs.
4.1.2 ABTS radical scavenging assay
The ABTS radical cation decolorization test was employed to assess the free radical scavenging activity of commercial, crude, and SZPCS-NPs (Figure 11). The ABTS cation radical was developed by combining 10 mg of ABTS and 2 mg of potassium persulfate in water. The solution was stored in the dark at room temperature for 12–16 h before use. After that, the ABTS solution (1 ml) was diluted with 60 ml of methanol. The measurements were then gradually performed in triplicate. The percent inhibition was determined using the equation as follow;

DPPH scavenging capacity of commercially available silymarin, crude silymarin, and SZPCS-NPs.
Sample SM (S. marianum extracts; commercial, crude, and zein silymarin NPs)
% Absorbance (734 nm) of commercially available silymarin, crude silymarin, and SZPCS-NPs at sample concentration 10 µl·ml−1 was recorded at 31.45%, 30.67%, and 78.98%, respectively (Figure 11). Similarly, at sample concentration 20 µl·ml−1, the % absorbance (734 nm) of commercially available silymarin, crude silymarin, and SZPCS-NPs were recorded as 38.17%, 33.76%, and 91.3%, respectively. In addition, at sample concentrations 30 and 60 µl·ml−1, the % absorbance (734 nm) of commercially available silymarin, crude silymarin, and SZPCS-NPs increased to 42.27%, 39.98%, 92.12% and 62.89, 58.23, and 96.34, respectively. Finally, at sample concentrations 70 and 80 µl·ml−1, the % absorbance (734 nm) of commercially available silymarin, crude silymarin, and SZPCS-NPs further increased to 64.78%, 63.47%, 96.98% and 68.56%, 66.87%, and 98.67%, respectively. The maximum values for total antioxidants in terms of ascorbic acid equivalents per mg were found to be 67.6 ± 1.44 and 72.6 ± 1.32 for the tested concentration of 1,000 μg·ml−1 for ZnO and Ag–ZnO, respectively [60].
4.2 Hepatoprotective activity
Up to date, many studies have revealed that S. marianum extracts possess potential hepatoprotective activity [67,69,70,71,72]. However, we were interested in checking the relative hepatoprotective effect of commercial, crude, and SZPCS-NPs. For induction of hepatotoxicity and liver damage, CCl4 was used to induce liver fibrosis in the experimental animals (rats). In order to evaluate the liver cell damage, the level of the hepatocytolytic marker enzymes, such as ALT, AST, ALP, and TP levels of bilirubin and albumin, which leak into the blood stream from the liver tissue, was assessed. In addition, the effects of commercial, crude, and SZPCS-NPs were assessed for 6 weeks from hepatocytolytic marker enzymes and concentration of bilirubin and albumin proteins. Hermenean et al. [73] and Kim et al. [74] injected the experimental mice with CCl4 to induce hepatotoxicity, while Mandegary et al. used hydrogen sulfide (H2S) [68]. Clichici et al. also used the same method (ALT and AST levels) for evaluating the hepatoprotective effect of silymarin-loaded gold NPs [75]. Kalantari et al. used AST, ALT, and ALP levels for the assessment of the hepatoprotective action of S. marianum [76]. In addition, some studies also focused on evaluating the hepatoprotective effects of S. marianum from levels of gamma-glutamyl transferase (GGT), ALT, AST [73,77], lactate dehydrogenase and creatine kinase [78], malondialdehyde (MDA), and CYP450 levels [79]. On the contrary, Abu-zaiton investigated the hepatoprotective activity from serum glutamate pyruvate transaminase and serum glutamic-oxaloacetic transaminase levels of experimental animals [80].
Group A. Normal untreated control
Group B. CCl4-treated control
Group C. Commercial silymarin
Group D. Crude silymarin
Group E. SZPCS-NPs
Tables 2–7 show a 6-week detailed analysis of the levels of hepatocytolytic marker enzymes such as ALT, AST, ALP, and TP, and levels of bilirubin and albumin, which leak into the blood stream from the liver tissue. Group A animals correspond to normal untreated control and physically healthy rats, with no administration of CCl4 or either silymarin (crude, commercial, or SZPCS-NPs). In addition, Group B animals relate to CCl4-treated control rats, with no administration of silymarin (crude, commercial, or SZPCS-NPs were done for this group). Similarly, Group C animals relate to CCl4-treated rats along with co-administration of commercial silymarin. Group D animals relate to CCl4-treated rats along with co-administration of crude silymarin. Finally, Group E animals relate to CCl4-treated rats along with co-administration of SZPCS-NPs.
Blood biochemical parameters after Week 1 (Day 7)
S. No | ALT (U·l−1) | AST (U·l−1) | ALP (U·l−1) | TP (mg·dl−1) | Bilirubin (mg·dl−1) | Albumin (g·dl−1) |
---|---|---|---|---|---|---|
Group A | 38.68 ± 5.93 | 110.67 ± 12.53 | 48.68 ± 7.46 | 4.35 ± 0.62 | 15.67 ± 2.24 | 2.67 ± 0.38 |
Group B | 72.67 ± 10.41 | 254.68 ± 36.47 | 87.63 ± 12.55 | 6.27 ± 0.90 | 31.67 ± 4.54 | 2.98 ± 0.43 |
Group C | 60.78 ± 6.88 | 207.67 ± 23.51 | 73.87 ± 8.36 | 5.97 ± 0.68 | 27.89 ± 3.16 | 2.81 ± 0.32 |
Group D | 65.89 ± 10.09 | 215.67 ± 33.04 | 81.67 ± 12.51 | 6.27 ± 0.96 | 29.68 ± 4.55 | 2.92 ± 0.45 |
Group E | 47.68 ± 7.78 | 130.89 ± 18.74 | 54.31 ± 8.86 | 4.41 ± 0.72 | 17.39 ± 2.84 | 2.62 ± 0.30 |
ALT: alanine aminotransferase, AST: aspartate aminotransferase, ALP: alkaline phosphatase, TP: total protein.
Blood biochemical parameters after Week 2 (Day 14)
S. No | ALT (U·l−1) | AST (U·l−1) | ALP (U·l−1) | TP (mg·dl−1) | Bilirubin (mg·dl−1) | Albumin (g·dl−1) |
---|---|---|---|---|---|---|
Group A | 38.68 ± 5.44 | 114.89 ± 12.43 | 48.65 ± 6.94 | 4.37 ± 0.61 | 15.71 ± 2.35 | 2.68 ± 0.43 |
Group B | 145.67 ± 18.98 | 350.13 ± 28.49 | 128.67 ± 18.38 | 8.53 ± 1.32 | 47.35 ± 7.43 | 3.27 ± 0.34 |
Group C | 97.67 ± 13.76 | 289.31 ± 24.34 | 103.68 ± 15.32 | 7.47 ± 1.03 | 35.63 ± 5.11 | 3.07 ± 0.27 |
Group D | 110.35 ± 14.39 | 310.45 ± 19.87 | 115.64 ± 14.65 | 7.54 ± 1.12 | 37.13 ± 6.34 | 3.14 ± 0.40 |
Group E | 51.37 ± 8.45 | 133.67 ± 13.24 | 54.38 ± 8.62 | 4.46 ± 0.47 | 17.38 ± 2.98 | 2.69 ± 0.32 |
Blood biochemical parameters after Week 3 (Day 21)
S. No | ALT (U·l−1) | AST (U·l−1) | ALP (U·l−1) | TP (mg·dl−1) | Bilirubin (mg·dl−1) | Albumin (g·dl−1) |
---|---|---|---|---|---|---|
Group A | 35.67 ± 4.56 | 115.83 ± 17.09 | 43.38 ± 6.62 | 4.21 ± 0.65 | 14.37 ± 2.32 | 2.28 ± 0.23 |
Group B | 207.83 ± 31.23 | 423.63 ± 60.34 | 163.67 ± 23.43 | 8.91 ± 1.32 | 59.63 ± 9.89 | 3.94 ± 0.45 |
Group C | 113.67 ± 19.23 | 327.87 ± 51.21 | 119.87 ± 16.45 | 8.08 ± 1.02 | 45.21 ± 5.65 | 3.34 ± 0.19 |
Group D | 130.83 ± 21.42 | 398.76 ± 55.34 | 163.83 ± 20.02 | 8.23 ± 0.98 | 47.38 ± 7.01 | 3.53 ± 0.47 |
Group E | 43.67 ± 5.65 | 110.83 ± 17.34 | 42.37 ± 6.56 | 4.28 ± 0.45 | 15.65 ± 2.10 | 2.45 ± 0.52 |
Blood biochemical parameters after Week 4 (Day 28)
S. No | ALT (U·l−1) | AST (U·l−1) | ALP (U·l−1) | TP (mg·dl−1) | Bilirubin (mg·dl−1) | Albumin (g·dl−1) |
---|---|---|---|---|---|---|
Group A | 47.68 ± 7.15 | 160.38 ± 24.06 | 53.83 ± 8.07 | 4.67 ± 0.75 | 21.63 ± 3.24 | 2.71 ± 0.35 |
Group B | 316.38 ± 37.97 | 502.38 ± 60.29 | 183.56 ± 22.03 | 9.63 ± 1.16 | 68.38 ± 8.21 | 4.31 ± 0.52 |
Group C | 128.38 ± 20.54 | 397.63 ± 63.62 | 136.81 ± 19.15 | 8.64 ± 1.38 | 57.91 ± 6.95 | 4.43 ± 0.71 |
Group D | 164.75 ± 18.12 | 467.81 ± 51.46 | 196.38 ± 21.60 | 9.03 ± 0.99 | 67.63 ± 7.44 | 4.54 ± 0.50 |
Group E | 46.18 ± 6.47 | 123.76 ± 17.33 | 47.83 ± 6.70 | 4.81 ± 0.67 | 18.97 ± 2.66 | 2.74 ± 0.38 |
Blood biochemical parameters after Week 5 (Day 35)
S. No | ALT (U·l−1) | AST (U·l−1) | ALP (U·l−1) | TP (mg·dl−1) | Bilirubin (mg·dl−1) | Albumin (g·dl−1) |
---|---|---|---|---|---|---|
Group A | 41.67 ± 6.82 | 116.93 ± 16.81 | 47.94 ± 7.85 | 4.17 ± 0.64 | 18.98 ± 2.16 | 2.73 ± 0.45 |
Group B | 453.65 ± 60.67 | 623.78 ± 83.42 | 243.87 ± 32.62 | 9.71 ± 1.30 | 73.67 ± 9.85 | 4.68 ± 0.63 |
Group C | 167.92 ± 20.78 | 420.65 ± 52.05 | 153.87 ± 19.04 | 8.96 ± 1.29 | 61.37 ± 8.82 | 4.49 ± 0.69 |
Group D | 217.65 ± 35.64 | 476.98 ± 78.10 | 192.76 ± 31.56 | 9.06 ± 1.21 | 67.97 ± 11.13 | 4.61 ± 0.75 |
Group E | 43.91 ± 4.99 | 125.87 ± 14.32 | 48.31 ± 5.49 | 4.19 ± 0.48 | 19.35 ± 2.20 | 2.75 ± 0.40 |
Blood biochemical parameters after Week 6 (Day 42)
S. No | ALT (U·l−1) | AST (U·l−1) | ALP (U·l−1) | TP (mg·dl−1) | Bilirubin (mg·dl−1) | Albumin (g·dl−1) |
---|---|---|---|---|---|---|
Group A | 40.93 ± 5.47 | 118.97 ± 17.10 | 47.91 ± 7.84 | 4.18 ± 0.64 | 18.96 ± 2.16 | 2.63 ± 0.43 |
Group B | 621.87 ± 83.17 | 708.93 ± 94.81 | 376.54 ± 50.36 | 9.87 ± 1.32 | 91.68 ± 12.26 | 4.93 ± 0.66 |
Group C | 235.67 ± 29.16 | 676.98 ± 7.7 | 193.67 ± 23.96 | 9.12 ± 1.31 | 71.87 ± 10.33 | 4.54 ± 0.70 |
Group D | 312.21 ± 51.12 | 721.78 ± 118.18 | 224.21 ± 36.71 | 9.23 ± 1.23 | 76.54 ± 12.3 | 4.61 ± 0.75 |
Group E | 42.54 ± 4.84 | 124.67 ± 14.18 | 48.12 ± 5.47 | 4.21 ± 0.48 | 19.24 ± 2.19 | 2.67 ± 0.38 |
It is evident from the results (Tables 2–7) that SZPCS-NPs showed a maximal hepatoprotective effect in comparison to crude silymarin and commercial silymarin. Clichici et al. also concluded that the effect of silymarin gold NPs was significantly better than silymarin alone in terms of its hepatoprotective effects [75]. The concentrations of the liver enzymes such as AST and ALT were significantly reduced by the S. marianum extract [80]. Pre-treatment with a self-micro-emulsifying drug delivery system containing S. marianum native seed oil (SMEDDS-SM) decreases the CCl4-induced elevation both in biochemical and in morphological parameters associated with liver damage or toxicity [81]. Guhagarkar et al. also concluded that polyethylene sebacate-silymarin nanoparticles (PES-SIM NPs) possess better hepato-protection in comparison to crude silymarin [51]. Gupta et al. also concluded that Silymarin chitosan NPs successfully enhance its hepatoprotective effect by passive targeting and sustained release [66]. Yang et al. determined that silymarin-loaded NPs significantly reduced carbon tetrachloride-induced hepatotoxicity, indicating improved bioactivity compared with silymarin powder and the commercial product [82].
4.3 Bioavailability
The results obtained from the percent absorbance measurements at different contact times provide valuable information about the rate and extent of absorption of commercial silymarin, crude silymarin, and SZPCS-NPs (Table 8).
Percent absorbance measurements at different contact times about the rate and extent of absorption of commercial silymarin, crude silymarin, and zein silymarin NPs
% Absorbance | |||
---|---|---|---|
Time (min) | Commercial silymarin
|
Crude silymarin
|
SZPCS-NPs
|
10 | 22.34 | 15.87 | 45.85 |
20 | 34.87 | 25.76 | 66.64 |
30 | 45.32 | 38.67 | 82.14 |
40 | 52.65 | 42.54 | 93.24 |
50 | 54.43 | 46.44 | 96.55 |
60 | 58.15 | 46.97 | 97.48 |
After 10 min of contact time, the percent absorbance values were 22.34%, 15.87%, and 45.85% for commercial silymarin, crude silymarin, and SZPCS-NPs, respectively. These values indicate that SZPCS-NPs exhibited the highest absorbance, suggesting a more efficient and rapid absorption compared to the other two formulations. At 20 min, the percent absorbance values increased to 34.87%, 25.76%, and 66.64% for commercial silymarin, crude silymarin, and SZPCS-NPs, respectively. Again, SZPCS-NPs showed the highest absorbance, indicating a continued increase in their absorption rate. After 30 min, the percent absorbance values further increased to 45.32%, 38.67%, and 82.14% for commercial silymarin, crude silymarin, and zein silymarin NPs, respectively. SZPCS-NPs continued to exhibit the highest absorbance, demonstrating a greater extent of absorption compared to the other formulations. The absorbance values at 50 min were 54.43%, 46.44%, and 96.55% for commercial silymarin, crude silymarin, and zein silymarin NPs, respectively. Once again, SZPCS-NPs displayed the highest absorbance, indicating a significant increase in their absorption compared to the other two formulations. After 60 min of contact time, the recorded percent absorbance values were 58.15%, 46.97%, and 97.48% for commercial silymarin, crude silymarin, and SZPCS-NPs, respectively. SZPCS-NPs maintained their superior absorbance, demonstrating the highest rate and extent of absorption among the three substances.
Based on the percent absorbance results at different contact times, it can be inferred that zein silymarin NPs exhibit a faster and more extensive absorption compared to commercial silymarin and crude silymarin. These findings suggest that SZPCS-NPs have the potential for enhanced bioavailability and improved therapeutic efficacy. These findings are consistent with other studies that have investigated the absorption of silymarin NPs. A study by Zhao et al. reported similar results, showing that silymarin-loaded NPs had higher absorption rates compared to raw formulation [83]. Additionally, a study by Onodera et al. demonstrated that silymarin NPs exhibited improved bioavailability in in vitro release studies [84].
4.4 Histopathological evaluation
All animals from Group A showed normal liver histology. The animals assigned in the other three experimental groups such as Groups B, C, D, and E presented marked morphological changes, characteristic of a liver injury associated with the administration of CCl4.
4.4.1 Week 1 study
Light microscopic studies of the liver tissues, stained with hematoxylin–eosin (H&E) in each group. After 1 week, Group A untreated control group shows normal lobular architecture and cell structure with no evidence of portal inflammation and hepatocellular necrosis or liver tissue damage. Group B CCl4-treated control group shows extensive hepatocellular damage with the presence of portal inflammation, hepatocellular necrosis, and liver tissue damage. Group C treated with commercial silymarin, and Group D treated with crude silymarin show less hepatocellular damage in comparison to Group B. Group E, treated with SZPCS-NPs, possesses much lesser hepatocellular damage in comparison to Groups B, C, and D. All the hepatocytes are visible within the microscopic pictures shown above, closely resemble Group A subjects showing normal lobular architecture and cell structure with no evidence of portal inflammation and hepatocellular necrosis. SZPCS-NPs hold a better hepatoprotective effect than commercial and crude silymarin (Figure 12).

Histopathological assessment of hepatotoxicity after 1 week, through light microscopic studies of the liver tissues, stained with H&E in each group. Specifically, (a) represents (H&E) stained hepatocytes of normal untreated control, (b) CCl4 treated control, (c) commercial silymarin, (d) crude silymarin, and (e) SZPCS-NPs.
4.4.2 Week 2 study
Light microscopic studies of the liver tissues, stained with H&E in each group. After 2 weeks, Group A untreated control group shows normal hepatocellular physiology with no evidence of necrosis or liver tissue damage. Group B CCl4-treated control group shows much more hepatocellular damage (red encircled) in comparison to week 1, with the presence of portal inflammation, cellular necrosis, and liver tissue damage. Group C treated with commercial silymarin and Group D treated with crude silymarin show a little bit of hepatocellular damage in comparison to Group B. Group E treated with SZPCS-NPs possesses much lesser hepatocellular damage in comparison to Groups B, C, and D. All the hepatocytes are visible within the microscopic pictures shown above, closely resemble Group A subjects showing normal lobular architecture and cell structure with no evidence of portal inflammation and hepatocellular necrosis. SZPCS-NPs holds a better hepatoprotective effect than commercial and crude silymarin (Figure 13).

Histopathological assessment of hepatotoxicity after 2 weeks, through light microscopic studies of the liver tissues, stained with H&E in each group. Specifically, (a) represents (H&E) stained hepatocytes of normal untreated control, (b) CCl4-treated control, (c) commercial silymarin, (d) crude silymarin, and (e) SZPCS-NPs.
4.4.3 Week 3 study
Light microscopic studies of the liver tissues, stained with H&E in each group. After 3 weeks, Group A untreated control group shows normal hepatocellular physiology with no evidence of necrosis or liver tissue damage (TN; tissue necrosis). Group B CCl4-treated control group shows much more hepatocellular damage (red encircled) in comparison to weeks 1 and 2, with the presence of portal inflammation, cellular necrosis, and liver tissue damage. Group C treated with commercial silymarin and Group D treated with crude silymarin show a little bit hepatocellular damage (red encircled) in comparison to Group B. Group E treated with SZPCS-NPs possesses much lesser hepatocellular damage in comparison to Groups B, C, and D. All the hepatocytes are visible within the microscopic pictures shown above, closely resemble Group A subjects showing normal lobular architecture and cell structure with no evidence of portal inflammation and hepatocellular necrosis. SZPCS-NPs hold a better hepatoprotective effect than commercial and crude silymarin (Figure 14).

Histopathological assessment of hepatotoxicity after 3 weeks, through light microscopic studies of the liver tissues, stained with H&E in each group. Specifically, (a) represents H&E stained hepatocytes of normal untreated control, (b) CCl4-treated control, (c) commercial silymarin, (d) crude silymarin, and (e) SZPCS-NPs.
4.4.4 Week 4 study
Light microscopic studies of the liver tissues, stained with H&E in each group. After 4 weeks, Group A untreated control group shows normal hepatocellular physiology with no evidence of necrosis or liver tissue damage (TN; tissue necrosis). Group B CCl4-treated control group shows much more hepatocellular damage in comparison to weeks 1 and 2, with the presence of portal inflammation, cellular necrosis, and liver tissue damage (red encircled). Group C treated with commercial silymarin and Group D treated with crude silymarin show a little bit hepatocellular damage in comparison to Group B. Group E treated with SZPCS-NPs possesses much lesser hepatocellular damage in comparison to Groups B, C, and D. All the hepatocytes are visible within the microscopic pictures shown above, closely resemble Group A subjects showing normal lobular architecture and cell structure with no evidence of portal inflammation and hepatocellular necrosis. SZPCS-NPs hold better hepatoprotective effects than commercial and crude silymarin (Figure 15).

Histopathological assessment of hepatotoxicity after 4 weeks, through light microscopic studies of the liver tissues, stained with H&E in each group. Specifically, (a) represents (H&E) stained hepatocytes of normal untreated control, (b) CCl4-treated control, (c) commercial silymarin, (d) crude silymarin, and (e) SZPCS-NPs.
4.4.5 Week 5 study
Light microscopic studies of the liver tissues, stained with H&E in each group. After 5 weeks, Group A untreated control group shows normal lobular architecture and cell structure with no evidence of portal inflammation and hepatocellular necrosis or liver tissue damage. Group B CCl4-treated control group, showing extensive hepatocellular damage (encircled red) with the presence of portal inflammation, hepatocellular necrosis, and liver tissue damage. Group C treated with commercial silymarin and Group D treated with crude silymarin showed less hepatocellular damage in comparison to Group B. While Group E treated with SZPCS-NPs possesses much lesser hepatocellular damage in comparison to Groups B, C, and D. All the hepatocytes are visible within the microscopic pictures. SZPCS-NPs hold better hepatoprotective effects than commercial and crude silymarin (Figure 16).

Histopathological assessment of hepatotoxicity after 5 weeks, through light microscopic studies of the liver tissues, stained with H&E in each group. Specifically, (a) represents (H&E) stained hepatocytes of normal untreated control, (b) CCl4-treated control, (c) commercial silymarin, (d) crude silymarin, and (e) SZPCS-NPs.
4.4.6 Week 6 study
Light microscopic studies of the liver tissues, stained with H&E in each group. After 6 weeks, Group A untreated control group shows normal lobular architecture and cell structure with no evidence of portal inflammation and hepatocellular necrosis or liver tissue damage. Group B CCl4-treated control group shows extensive hepatocellular damage with the presence of portal inflammation, hepatocellular necrosis, and liver tissue damage. The damaged cells are not properly stained with H&E stained (red encircled). Group C treated with commercial silymarin and Group D treated with crude silymarin show less hepatocellular damage in comparison to Group B. Group E treated with SZPCS-NPs possesses much lesser hepatocellular damage in comparison to Groups B, C, and D. All the hepatocytes are visible within the microscopic pictures. SZPCS-NPs hold better hepatoprotective effects than commercial and crude silymarin (Figure 17).

Histopathological assessment of hepatotoxicity after 6 weeks, through light microscopic studies of the liver tissues, stained with H&E in each group. Specifically, (a) represents (H&E) stained hepatocytes of normal untreated control, (b) CCl4-treated control, (c) commercial silymarin, (d) crude silymarin, and (e) SZPCS-NPs.
Overall, these findings suggest that SZPCS-NPs hold promise as a potential therapeutic approach for liver diseases and as a drug delivery system for silymarin. It is evident from the results that SZPCS-NPs demonstrated a maximal hepatoprotective effect in comparison to crude silymarin and commercial silymarin.
Clichici et al. also concluded that the effect of silymarin gold NPs possessed significantly better hepatoprotective effects in contrast to crude silymarin [75]. Yang et al. determined that silymarin-loaded NPs significantly reduced CCl4-induced hepatotoxicity, indicating improved bioactivity compared with silymarin crude powder and the commercial silymarin product [82]. In addition, pre-treatment with a self-micro-emulsifying drug delivery system containing SMEDDS-SM decreases the CCl4-induced elevation both in biochemical and in morphological parameters associated with liver damage or toxicity [81]. In addition, Guhagarkar et al. also concluded that PES-SIM NPs possess better hepato-protection in comparison to crude silymarin [51]. Gupta et al. also concluded that silymarin chitosan NPs successfully enhance its hepatoprotective effect [66]. Shriram et al. revealed that the nano-delivery system of silymarin offers enhanced hepatoprotective activity [85]. Yang et al. found that silymarin-loaded NPs dramatically reduced CCL4-induced hepatotoxicity in contrast to commercial and crude silymarin formulations [82]. Nasr et al. also concluded that silymarin-loaded mesoporous silica NPs possess better hepatoprotective activity in comparison to crude silymarin [86].
5 Conclusion
In conclusion, there were no significant differences noted in the contents of oil and silymarin in the pink and white flowering S. marianum seeds can be used in any variety for oil and silymarin extraction. SZPCS-NPs were found best in pectin/sodium alginate solution with a ratio of 30% pectin and 70% alginate, whereas the antioxidant activity and hepatoprotective activity were enhanced with many folds.
These findings suggest that SZPCS-NPs hold promise as a potential therapeutic approach for liver diseases and as a drug delivery system for silymarin. Further research and development in this area are warranted to explore their full potential in clinical applications.
Acknowledgments
Authors wish to thank Researchers Supporting Project Number (RSP2024R346) at King Saud University Riyadh, Saudi Arabia, for financial support.
-
Funding information: This research was supported by Researchers Supporting Project Number (RSP2024R346) at King Saud University, Riyadh, Saudi Arabia.
-
Author contributions: Conceptualization, MSK, IK, MZ, RU, AB, FAK; writing – original draft preparation, MSK, IB, ME; writing – review and editing, FS, MZ, SA, SK, RU, ME; supervision, IK, FAK, MZ, SA, SK, RU; funding acquisition, RU, and AB. All authors have read and agreed to the submitted version of the manuscript.
-
Conflict of interest: Authors state no conflict of interest.
-
Data availability statement: The datasets generated during and/or analysed during the current study are available from the corresponding author on reasonable request.
References
[1] Mohanty SK, Swamy MK, Sinniah UR, Anuradha M. Leptadenia reticulata (Retz.) Wight & Arn.(Jivanti): botanical, agronomical, phytochemical, pharmacological, and biotechnological aspects. Molecules. 2017;22(6):1019.10.3390/molecules22061019Search in Google Scholar PubMed PubMed Central
[2] Patel AA, Patel RJ, Patel SR. Nanomedicine for intranasal delivery to improve brain uptake. Curr Drug Delivery. 2018;15(4):461–9.10.2174/1567201814666171013150534Search in Google Scholar PubMed
[3] Siddiqui AA, Iram F, Siddiqui S, Sahu K. Role of natural products in drug discovery process. Int J Drug Dev Res. 2014;6(2):172–204.Search in Google Scholar
[4] Da Silva FLO, Marques MBDF, Kato KC, Carneiro G. Nanonization techniques to overcome poor water-solubility with drugs. Exp Opin Drug Disc. 2020;15:853–64.10.1080/17460441.2020.1750591Search in Google Scholar PubMed
[5] Das B, Baidya AT, Mathew AT, Yadav AK, Kumar R. Structural modification aimed for improving solubility of lead compounds in early phase drug discovery. Bioorg Med Chem. 2022;56:116614.10.1016/j.bmc.2022.116614Search in Google Scholar PubMed
[6] Khan KU, Minhas MU, Badshah SF, Suhail M, Ahmad A, Ijaz S. Overview of nanoparticulate strategies for solubility enhancement of poorly soluble drugs. Life Sci. 2022;291:120301.10.1016/j.lfs.2022.120301Search in Google Scholar PubMed
[7] Smita S, O’Brien F, Cheng Thiam DT, Harris D, Math MC, Ariën T, et al. Oral drug delivery strategies for development of poorly water soluble drugs in paediatric patient population. Adv Drug Delivery Rev. 2022;190:114507.10.1016/j.addr.2022.114507Search in Google Scholar PubMed
[8] Mengran G, Qin S, Wang S, Sun M, Yang H, Wang X, et al. Herbal medicine nanocrystals: A potential novel therapeutic strategy. Molecules. 2023;28(17):6370.10.3390/molecules28176370Search in Google Scholar PubMed PubMed Central
[9] Shiyi Z, Huang C, Yue X, Li X, Zhou P, Wu A, et al. Application advance of electrosprayed micro/nanoparticles based on natural or synthetic polymers for drug delivery system. Mat Des. 2022;220:110850.10.1016/j.matdes.2022.110850Search in Google Scholar
[10] Suyal J, Kumar B, Jakhmola V. Novel approach self-nanoemulsifying drug delivery system: A review. Adv Pharmacol Pharm. 2023;11(2):131–9.10.13189/app.2023.110205Search in Google Scholar
[11] Ayala-Fuentes JC, Chavez-Santoscoy RA. Nanotechnology as a key to enhance the benefits and improve the bioavailability of flavonoids in the food industry. Foods. 2021;10(11):2701.10.3390/foods10112701Search in Google Scholar PubMed PubMed Central
[12] Mirian P, Gómez B, Munekata PES, Barba FJ, Putnik P, Kovačević DB, et al. Nanoencapsulation of promising bioactive compounds to improve their absorption, stability, functionality and the appearance of the final food products. Molecules. 2021;26(6):1547.10.3390/molecules26061547Search in Google Scholar PubMed PubMed Central
[13] Rocha B, de Morais LA, Viana MC, Carneiro G. Promising strategies for improving oral bioavailability of poor water-soluble drugs. Exp Opin Drug Discovery. 2023;18(6):615–27.10.1080/17460441.2023.2211801Search in Google Scholar PubMed
[14] Elzoghby AO, Samy WM, Elgindy NA. Protein-based nanocarriers as promising drug and gene delivery systems. J Cont Rel. 2012;161(1):38–49.10.1016/j.jconrel.2012.04.036Search in Google Scholar PubMed
[15] Ding R, Zhang M, Zhu Q, Qu Y, Jia X, Yin L. Curcumin loaded Zein-alginate nanogels with “core-shell” structure: Formation, characterization and simulated digestion. Int J Biol Macromol. 2023;251:126201.10.1016/j.ijbiomac.2023.126201Search in Google Scholar PubMed
[16] Takma DK, Bozkurt S, Koç M, Korel F, Nadeem HŞ. Characterization and encapsulation efficiency of zein nanoparticles loaded with chestnut fruit shell, cedar and sweetgum bark extracts. Food Hydrocoll Health. 2023;4:100151.10.1016/j.fhfh.2023.100151Search in Google Scholar
[17] Zhiheng Z, Meng Y, Wang J, Qiu C, Miao W, Lin Q, et al. Preparation and characterization of zein-based core-shell nanoparticles for encapsulation and delivery of hydrophobic nutrient molecules. Food Hydrocoll. 2024;148:109524.10.1016/j.foodhyd.2023.109524Search in Google Scholar
[18] Shuai C, Li Q, McClements DJ, Han Y, Dai L, Mao L, et al. Co-delivery of curcumin and piperine in zein-carrageenan core-shell nanoparticles: Formation, structure, stability and in vitro gastrointestinal digestion. Food Hydrocoll. 2020;99:105334.10.1016/j.foodhyd.2019.105334Search in Google Scholar
[19] Zhiheng Z, Hu Y, Ji H, Lin Q, Li X, Sang S, et al. Physicochemical stability, antioxidant activity, and antimicrobial activity of quercetin-loaded zein nanoparticles coated with dextrin-modified anionic polysaccharides. Food chem. 2023;415:135736.10.1016/j.foodchem.2023.135736Search in Google Scholar PubMed
[20] Huang X, Huang X, Gong Y, Xiao H, McClements DJ, Hu K. Enhancement of curcumin water dispersibility and antioxidant activity using core–shell protein–polysaccharide nanoparticles. Food Res Int. 2016;87:1–9.10.1016/j.foodres.2016.06.009Search in Google Scholar PubMed
[21] Kasaai MR. Zein and zein-based nano-materials for food and nutrition applications: A review. Trends Food Sci Tech. 2018;79:184–97.10.1016/j.tifs.2018.07.015Search in Google Scholar
[22] Xue F, Li C, Liu Y, Zhu X, Pan S, Wang L. Encapsulation of tomato oleoresin with zein prepared from corn gluten meal. J Food Eng. 2013;119(3):439–45.10.1016/j.jfoodeng.2013.06.012Search in Google Scholar
[23] Nitta SK, Numata K. Biopolymer-based nanoparticles for drug/gene delivery and tissue engineering. Int J Mol Sci. 2013;14(1):1629–54.10.3390/ijms14011629Search in Google Scholar PubMed PubMed Central
[24] Wang H, Pengfei S, Shan C, Jian W, Lijun G, Yun W, et al. Nitrogen-doped carbon dots for “green” quantum dot solar cells. Nanoscale Res Lett. 2016;11:1–6.10.1186/s11671-016-1231-1Search in Google Scholar PubMed PubMed Central
[25] Le Q-U, Lay H-L, Wu M-C, Joshi RK. Phytoconstituents and pharmacological activities of Silybum marianum (Milk Thistle): A critical review. Am J Ess Oils Nat Prod. 2018;6(4):41–7.Search in Google Scholar
[26] Porwal O, Ameen MSM, Anwer ET, Uthirapathy S, Ahamad J, Tahsin A. Silybum marianum (Milk Thistle): Review on Its chemistry, morphology, ethno medical uses, phytochemistry and pharmacological activities. J Drug Deliv Ther. 2019;9(5):199–206.10.22270/jddt.v9i5.3666Search in Google Scholar
[27] Wang X, Zhang Z, Wu S-C. Health benefits of Silybum marianum: Phytochemistry, pharmacology, and applications. J Agric Food Chem. 2020;68(42):11644–64.10.1021/acs.jafc.0c04791Search in Google Scholar PubMed
[28] Kesharwani SS, Jain V, Dey S, Sharma S, Mallya P, Kumar VA. An overview of advanced formulation and nanotechnology-based approaches for solubility and bioavailability enhancement of silymarin. J Drug Delivery Sci Tech. 2020;60:102021.10.1016/j.jddst.2020.102021Search in Google Scholar
[29] Shangguan M, Qi J, Lu Y, Wu W. Comparison of the oral bioavailability of silymarin-loaded lipid nanoparticles with their artificial lipolysate counterparts: Implications on the contribution of integral structure. Int J Pharm. 2015;489(1–2):195–202.10.1016/j.ijpharm.2015.05.005Search in Google Scholar PubMed
[30] Ayad ZM, Ibrahim OMS, Omar LW. Biosynthesis and characterization of silver nanoparticles by Silybum marianum (silymarin) fruit extract. Adv Anim Vet Sci. 2019;7(2):122–130.10.17582/journal.aavs/2019/7.2.122.130Search in Google Scholar
[31] Sahin E, Bagci R, Bektur Aykanat NE, Kacar S, Sahinturk V. Silymarin attenuated nonalcoholic fatty liver disease through the regulation of endoplasmic reticulum stress proteins GRP78 and XBP‐1 in mice. J Food Biochem. 2020;44(6):13194.10.1111/jfbc.13194Search in Google Scholar PubMed
[32] Veisi S, Johari SA, Tyler CR, Mansouri B, Esmaeilbeigi M. Antioxidant properties of dietary supplements of free and nanoencapsulated silymarin and their ameliorative effects on silver nanoparticles induced oxidative stress in Nile tilapia (Oreochromis niloticus). Env Sci Pollut Res. 2021;28:26055–63.10.1007/s11356-021-12568-8Search in Google Scholar PubMed
[33] Shomer NH, Nirah H, Krystal H, Worthington A, Debra L, Hickman H, et al. Review of rodent euthanasia methods. J Am Assoc Lab Anim Sci. 2020;59(3):242–53.10.30802/AALAS-JAALAS-19-000084Search in Google Scholar PubMed PubMed Central
[34] Lahlah ZF, Meziani M, Maza A. Silymarin natural antimicrobial agent extracted from Silybum marianum. J Acad. 2012;2:164–9.Search in Google Scholar
[35] Shaker E, Mahmoud H, Mnaa S. Silymarin, the antioxidant component and Silybum marianum extracts prevent liver damage. Food Chem Toxic. 2010;48(3):803–6.10.1016/j.fct.2009.12.011Search in Google Scholar PubMed
[36] Chambers CS, Holečková V, Petrásková L, Biedermann D, Valentová K, Buchta M, et al. The silymarin composition… and why does it matter??? Food Res Int. 2017;339–53.10.1016/j.foodres.2017.07.017Search in Google Scholar PubMed
[37] Eldalawy R, Al-Ani WM. GC–MS analysis of Iraqi Silybum marianum Flowers, Leaves and Seeds Extracts. Al Must J Pharm Sci. 2020;20:93–112.10.32947/ajps.v20i4.780Search in Google Scholar
[38] Fadhil AB, Ahmed KM, Dheyab MM. Silybum marianum L. seed oil: A novel feedstock for biodiesel production. Arab J Chem. 2017;10:S683–90.10.1016/j.arabjc.2012.11.009Search in Google Scholar
[39] Radusheva P, Pashev A, Uzunova G, Nikolova K, Gentscheva G, Perifanova M, et al. Physicochemical characteristics of seed oil of Sambucus ebulus, Coriandrum sativum L. and Silybum marianum L. Bulg. Chem Commun. 2019;51(S.I A):144–9.Search in Google Scholar
[40] Hussain I, Khan FA, Ullah R, Khan N, Muhammad Z. Analysis of silymarin in the seeds of white and blue capitulum of Silybum marianum plants. Afr J Pharm Pharmacol. 2011;5:742–5.10.5897/AJPP11.160Search in Google Scholar
[41] Maryam A, Saeed F, Ahmad N, Ahmad A, Afzaal M, Hussain S, et al. Biochemical profile of milk thistle (Silybum Marianum L.) with special reference to silymarin content. Food Sci Nutr. 2021;9:244–50.10.1002/fsn3.1990Search in Google Scholar PubMed PubMed Central
[42] Das S, Roy P, Auddy RG, Mukherjee A. Silymarin nanoparticle prevents paracetamol-induced hepatotoxicity. Int J Nanomed. 2011;6:1291–301.10.2147/IJN.S15160Search in Google Scholar PubMed PubMed Central
[43] Maryana W, Rahma A, Mudhakir D, Rachmawati H. Phytosome containing silymarin for oral administration: Formulation and physical evaluation. J Biomimetics Biomater Biomed Eng. 2015;25:54–65.10.4028/www.scientific.net/JBBBE.25.54Search in Google Scholar
[44] Maryana W, Rachmawati H, Mudhakir D. Formation of phytosome containing silymarin using thin layer-hydration technique aimed for oral delivery. Mat Today: Proc. 2016;3:855–66.10.1016/j.matpr.2016.02.019Search in Google Scholar
[45] Raghuram K, Dash S, Bhowal AC, Kalita S, Talukdar NC, Kundu S, et al. Glycogen–gold nanohybrid escalates the potency of silymarin. Int J Nanomed. 2017;7025–38.10.2147/IJN.S142497Search in Google Scholar PubMed PubMed Central
[46] Muzamil S, Nawaz S, Jan H, Uddin N, Ali A, Anjum S, et al. Synthesis of bio-mediated silver nanoparticles from Silybum marianum and their biological and clinical activities. Mat Sci Eng. 2020;112:110889.10.1016/j.msec.2020.110889Search in Google Scholar PubMed
[47] Feng S, Sun Y, Wang D, Sun P, Shao P. Effect of adjusting pH and chondroitin sulfate on the formation of curcumin-zein nanoparticles: Synthesis, characterization and morphology. Carb Polym. 2020;250:116970.10.1016/j.carbpol.2020.116970Search in Google Scholar PubMed
[48] Songlin X, Xiao L, Dong X, Li X, Wang Y, Hu X, et al. Preparation of chitosan/curcumin nanoparticles based zein and potato starch composite films for Schizothorax prenati fillet preservation. Int J Bio Macromol. 2020;164:211–21.10.1016/j.ijbiomac.2020.07.082Search in Google Scholar PubMed
[49] Paul NS, Yadav RP. Biosynthesis of silver nanoparticles using plant seeds and their antimicrobial activity. Asian J Biomed Pharm Sci. 2015;5(45):26.10.15272/ajbps.v5i45.711Search in Google Scholar
[50] Samira J, Fard SM, Nejad AR, Mumivand H, Sorkheh K, Nikoloudakis N, et al. Chitosan and titanium dioxide are more effective in improving seed yield and quality in nanoparticle compared to non-structured form: A case study in five milk thistle ecotypes (Silybum marianum (L.) Gaertn.). Agronomy . 2022;12(8):1827.10.3390/agronomy12081827Search in Google Scholar
[51] Guhagarkar SA, Shah D, Patel MD, Sathaye SS, Devarajan PV. Polyethylene sebacate-silymarin nanoparticles with enhanced hepatoprotective activity. J Nanosci Nanotech. 2015;15:4090–3.10.1166/jnn.2015.9518Search in Google Scholar PubMed
[52] Sajadi SM, Nasrollahzadeh M, Maham M. Aqueous extract from seeds of Silybum marianum L. as a green material for preparation of the Cu/Fe3O4 nanoparticles: a magnetically recoverable and reusable catalyst for the reduction of nitroarenes. J Colloid Interface Sci. 2016;469:93–8.10.1016/j.jcis.2016.02.009Search in Google Scholar PubMed
[53] Abbasi BH, Muzamil S, Syed SH, Munazza N, Sania N, Waqar A, et al. Green bio-assisted synthesis, characterization and biological evaluation of biocompatible ZnO NPs synthesized from different tissues of milk thistle (Silybum marianum). Nanomaterials. 2019;9(8):1171.10.3390/nano9081171Search in Google Scholar PubMed PubMed Central
[54] Tsai Y-H, Yang Y-N, Ho Y-C, Tsai M-L, Mi F-L. Drug release and antioxidant/antibacterial activities of silymarin-zein nanoparticle/bacterial cellulose nanofiber composite films. Carbohydr Polym. 2018;180:286–96.10.1016/j.carbpol.2017.09.100Search in Google Scholar PubMed
[55] Gopalakrishnan R, Raghu K. Biosynthesis and characterization of gold and silver nanoparticles using milk thistle (Silybum marianum) seed extract. J Nanosci. 2014;15(6):4090–3.10.1155/2014/905404Search in Google Scholar
[56] Safarpoor M, Ghaedi M, Asfaram A, Yousefi-Nejad M, Javadian H, Khafri HZ, et al. Ultrasound-assisted extraction of antimicrobial compounds from Thymus daenensis and Silybum marianum: Antimicrobial activity with and without the presence of natural silver nanoparticles. Ultrason Sonochem. 2018;42:76–83.10.1016/j.ultsonch.2017.11.001Search in Google Scholar PubMed
[57] Zare M, Sarkati MN. Chitosan‐functionalized Fe3O4 nanoparticles as an excellent biocompatible nanocarrier for silymarin delivery. Polym Adv Tech. 2021;32(10):4094–100.10.1002/pat.5416Search in Google Scholar
[58] Gopalakrishnan R, Loganathan B, Raghu K. Green synthesis of Au–Ag bimetallic nanocomposites using Silybum marianum seed extract and their application as a catalyst. RSC Adv. 2015;5(40):31691–99.10.1039/C5RA03571FSearch in Google Scholar
[59] Iqbal J, Andleeb A, Ashraf H, Meer B, Mehmood A, Jan H, et al. Potential antimicrobial, antidiabetic, catalytic, antioxidant and ROS/RNS inhibitory activities of Silybum marianum mediated biosynthesized copper oxide nanoparticles. RSC Adv. 2022;12(22):14069–83.10.1039/D2RA01929ASearch in Google Scholar PubMed PubMed Central
[60] Hameed S, Khalil AT, Ali M, Numan M, Khamlich S, Shinwari ZK, et al. Greener synthesis of ZnO and Ag–ZnO nanoparticles using Silybum marianum for diverse biomedical applications. Nanomedicine. 2019;14(6):655–73.10.2217/nnm-2018-0279Search in Google Scholar PubMed
[61] Abbasi R, Shineh G, Mobaraki M, Doughty S, Tayebi L. Structural parameters of nanoparticles affecting their toxicity for biomedical applications: a review. J Nanopart Res. 2023;25(3):43.10.1007/s11051-023-05690-wSearch in Google Scholar PubMed PubMed Central
[62] Mohammadinejad R, Pourseyedi S, Baghizadeh A, Ranjbar S, Mansoori GA. Synthesis of silver nanoparticles using Silybum marianum seed extract. Int J Nanosci Nanotech. 2013;9(4):221–6.Search in Google Scholar
[63] Ma Y, He S, Ma X, Hong T, Li Z, Park K, et al. Silymarin-loaded nanoparticles based on stearic acid-modified Bletilla striata polysaccharide for hepatic targeting. Molecules . 2016;21(3):265.10.3390/molecules21030265Search in Google Scholar PubMed PubMed Central
[64] Zhao X, Deng Y, Zhang Y, Zu Y, Lian B, Wu M, et al. Silymarin nanoparticles through emulsion solvent evaporation method for oral delivery with high antioxidant activities, bioavailability, and absorption in the liver. RSC Adv. 2016;6(95):93137–46.Search in Google Scholar
[65] Onodera R, Hayashi T, Motoyama K, Tahara K, Takeuchi H. Hydroxypropyl-β-cyclodextrin enhances oral absorption of silymarin nanoparticles prepared using PureNano™ continuous crystallizer. Pharmaceutics. 2022;14(2):394.Search in Google Scholar
[66] Gupta S, Singh SK, Girotra P. Targeting silymarin for improved hepatoprotective activity through chitosan nanoparticles. Int J Pharm Invest. 2014;4(4):156.10.4103/2230-973X.143113Search in Google Scholar PubMed PubMed Central
[67] Netto MPharm G, Jose J. Development, characterization, and evaluation of sunscreen cream containing solid lipid nanoparticles of silymarin. J Cosmetic Derm. 2018;7(6):1073–83.10.1111/jocd.12470Search in Google Scholar PubMed
[68] Kosalai D, Chandran M. Phytochemical analysis and anti-oxidant activity of gold nanoparticles synthesizing plant-Silybum marianum. Int J Curr Microb Appl Sci. 2016;5(4):469–75.10.20546/ijcmas.2016.504.055Search in Google Scholar
[69] Aghazadeh S, Amini R, Yazdanparast R, Ghaffari SH. Anti-apoptotic and anti-inflammatory effects of Silybum marianum in treatment of experimental steatohepatitis. Exp Toxic Pathol. 2011;63(6):569–74.10.1016/j.etp.2010.04.009Search in Google Scholar PubMed
[70] Mandegary A, Saeedi A, Eftekhari A, Montazeri V, Sharif E. Hepatoprotective effect of silyamarin in individuals chronically exposed to hydrogen sulfide; modulating influence of TNF-α cytokine genetic polymorphism. D J Pharm Sci. 2013;21:1–5.10.1186/2008-2231-21-28Search in Google Scholar PubMed PubMed Central
[71] Singh S, Mehta A, Ahirwal L, Dubey MK, Mishra A, Kumar R, et al. Combinational effect of Cajanus cajan, Silybum marianum and Andrographis paniculata on in vivo antioxidant and hepatoprotective activities of carbon tetrachloride intoxicated albino rats. Annu Res Rev Bio. 2014;665–74.10.9734/ARRB/2014/6022Search in Google Scholar
[72] Behrouj H, Ziamajidi N, Abbasalipourkabir R, Nasiri A, Asl SS. Therapeutic effect of Silybum marianum plant extract on tamoxifen-induced fatty liver in rats. Avicenna J Med Biochem. 2015;3:5–27160.10.17795/ajmb-27160Search in Google Scholar
[73] Hermenean A, Stan M, Ardelean A, Pilat L, Mihali CV, Popescu C, et al. Antioxidant and hepatoprotective activity of milk thistle (Silybum marianum L. Gaertn.) seed oil. Open Life Sci. 2015;10(1):225–36.10.1515/biol-2015-0017Search in Google Scholar
[74] Kim SH, Cheon HJ, Yun N, Oh ST, Shin E, Shim KS, et al. Protective effect of a mixture of Aloe vera and Silybum marianum against carbon tetrachloride–induced acute hepatotoxicity and liver fibrosis. J Pharmacol Sci. 2009;109(1):119–27.10.1254/jphs.08189FPSearch in Google Scholar PubMed
[75] Clichici S, David L, Moldovan B, Baldea I, Olteanu D, Filip M, et al. Hepatoprotective effects of silymarin coated gold nanoparticles in experimental cholestasis. Mat Sci Eng. 2020;115:111117.10.1016/j.msec.2020.111117Search in Google Scholar PubMed
[76] Kalantari A, Salimi A, Kalantari H, Broojeni JE, Rashidi I, Vanani AR, et al. The hepatoprotective effect of livergol microemulsion preparation (nanoparticle) against bromobenzene induced toxicity in mice. Toxicol Rep. 2019;6:444–8.10.1016/j.toxrep.2019.05.005Search in Google Scholar PubMed PubMed Central
[77] Gordon A, Hobbs DA, Bowden DS, Bailey MJ, Mitchell J, Francis AJ, et al. Effects of Silybum marianum on serum hepatitis C virus RNA, alanine aminotransferase levels and well‐being in patients with chronic hepatitis C. J Gastroenterol Hepatol. 2006;21(1):275–80.10.1111/j.1440-1746.2006.04138.xSearch in Google Scholar PubMed
[78] Zaker‐Esteghamati H, Seidavi A, Bouyeh M. The effects of Cynara scolymus and Silybum marianum on growth, carcass and organ characteristics, immunity, blood constitutes, liver enzymes, jejunum morphology, and fatty acid profile of breast meat in broilers. Food Sci Nutr. 2021;9(12):6692–706.10.1002/fsn3.2620Search in Google Scholar PubMed PubMed Central
[79] Eldemerdash R, El-Gayar HA, Salim SA, Salah A, Abdel-Aziz A. Hepato-protective effect of aqueous extract of silybum-marianum against carbon tetrachloride induced liver injury in rats. Int J Res Stud Biosci. 2016;4(2):13–24.10.20431/2349-0365.0402003Search in Google Scholar
[80] Abu-zaiton AS. Evaluating the effect of Silybum marianum extract on blood glucose, liver and kidney functions in diabetic rats. Adv studies in biology. 2013;5(10):447–54.10.12988/asb.2013.3936Search in Google Scholar
[81] Fehér P, Ujhelyi Z, Vecsernyés M, Fenyvesi F, Damache G, Ardelean A, et al. Hepatoprotective effects of a self-micro emulsifying drug delivery system containing Silybum marianum native seed oil against experimentally induced liver injury. Die Pharmazie-Int J Pharm Sci. 2015;70(4):231–8.Search in Google Scholar
[82] Yang KY, Hwang DH, Yousaf AM, Kim DW, Shin YJ, Bae ON, et al. Silymarin-loaded solid nanoparticles provide excellent hepatic protection: physicochemical ch aracterization and in vivo evaluation. Int J Nanomed. 2013;3333–43.10.2147/IJN.S50683Search in Google Scholar PubMed PubMed Central
[83] Zhao X, Yiping D, Ying Z, Yuangang Z, Bolin L, Mingfang W, et al. Silymarin nanoparticles through emulsion solvent evaporation method for oral delivery with high antioxidant activities, bioavailability, and absorption in the liver. RSC Adv. 2016;6(95):93137–46.10.1039/C6RA12896CSearch in Google Scholar
[84] Onodera R, Hayashi T, Motoyama K, Tahara K, Takeuchi H. Hydroxypropyl-β-cyclodextrin enhances oral absorption of silymarin nanoparticles prepared using PureNano™ continuous crystallizer. Pharmaceutics. 2022;14(2):394.10.3390/pharmaceutics14020394Search in Google Scholar PubMed PubMed Central
[85] Shriram RG, Moin A, Alotaibi HF, Khafagy ES, Al Saqr A, Abu Lila AS, et al. Phytosomes as a plausible nano-delivery system for enhanced oral bioavailability and improved hepatoprotective activity of silymarin. Pharmaceuticals. 2022;15(7):790.10.3390/ph15070790Search in Google Scholar PubMed PubMed Central
[86] Nasr SS, Nasra MM, Hazzah HA, Abdallah OY, Mesoporous silica nanoparticles, a safe option for silymarin delivery: Preparation, characterization, and in vivo evaluation. Drug Delivery and Translational Research. 2019 Oct;9:968.10.1007/s13346-019-00640-3Search in Google Scholar PubMed
© 2024 the author(s), published by De Gruyter
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Research Articles
- Green polymer electrolyte and activated charcoal-based supercapacitor for energy harvesting application: Electrochemical characteristics
- Research on the adsorption of Co2+ ions using halloysite clay and the ability to recover them by electrodeposition method
- Simultaneous estimation of ibuprofen, caffeine, and paracetamol in commercial products using a green reverse-phase HPTLC method
- Isolation, screening and optimization of alkaliphilic cellulolytic fungi for production of cellulase
- Functionalized gold nanoparticles coated with bacterial alginate and their antibacterial and anticancer activities
- Comparative analysis of bio-based amino acid surfactants obtained via Diels–Alder reaction of cyclic anhydrides
- Biosynthesis of silver nanoparticles on yellow phosphorus slag and its application in organic coatings
- Exploring antioxidant potential and phenolic compound extraction from Vitis vinifera L. using ultrasound-assisted extraction
- Manganese and copper-coated nickel oxide nanoparticles synthesized from Carica papaya leaf extract induce antimicrobial activity and breast cancer cell death by triggering mitochondrial caspases and p53
- Insight into heating method and Mozafari method as green processing techniques for the synthesis of micro- and nano-drug carriers
- Silicotungstic acid supported on Bi-based MOF-derived metal oxide for photodegradation of organic dyes
- Synthesis and characterization of capsaicin nanoparticles: An attempt to enhance its bioavailability and pharmacological actions
- Synthesis of Lawsonia inermis-encased silver–copper bimetallic nanoparticles with antioxidant, antibacterial, and cytotoxic activity
- Facile, polyherbal drug-mediated green synthesis of CuO nanoparticles and their potent biological applications
- Zinc oxide-manganese oxide/carboxymethyl cellulose-folic acid-sesamol hybrid nanomaterials: A molecularly targeted strategy for advanced triple-negative breast cancer therapy
- Exploring the antimicrobial potential of biogenically synthesized graphene oxide nanoparticles against targeted bacterial and fungal pathogens
- Biofabrication of silver nanoparticles using Uncaria tomentosa L.: Insight into characterization, antibacterial activities combined with antibiotics, and effect on Triticum aestivum germination
- Membrane distillation of synthetic urine for use in space structural habitat systems
- Investigation on mechanical properties of the green synthesis bamboo fiber/eggshell/coconut shell powder-based hybrid biocomposites under NaOH conditions
- Green synthesis of magnesium oxide nanoparticles using endophytic fungal strain to improve the growth, metabolic activities, yield traits, and phenolic compounds content of Nigella sativa L.
- Estimation of greenhouse gas emissions from rice and annual upland crops in Red River Delta of Vietnam using the denitrification–decomposition model
- Synthesis of humic acid with the obtaining of potassium humate based on coal waste from the Lenger deposit, Kazakhstan
- Ascorbic acid-mediated selenium nanoparticles as potential antihyperuricemic, antioxidant, anticoagulant, and thrombolytic agents
- Green synthesis of silver nanoparticles using Illicium verum extract: Optimization and characterization for biomedical applications
- Antibacterial and dynamical behaviour of silicon nanoparticles influenced sustainable waste flax fibre-reinforced epoxy composite for biomedical application
- Optimising coagulation/flocculation using response surface methodology and application of floc in biofertilisation
- Green synthesis and multifaceted characterization of iron oxide nanoparticles derived from Senna bicapsularis for enhanced in vitro and in vivo biological investigation
- Potent antibacterial nanocomposites from okra mucilage/chitosan/silver nanoparticles for multidrug-resistant Salmonella Typhimurium eradication
- Trachyspermum copticum aqueous seed extract-derived silver nanoparticles: Exploration of their structural characterization and comparative antibacterial performance against gram-positive and gram-negative bacteria
- Microwave-assisted ultrafine silver nanoparticle synthesis using Mitragyna speciosa for antimalarial applications
- Green synthesis and characterisation of spherical structure Ag/Fe2O3/TiO2 nanocomposite using acacia in the presence of neem and tulsi oils
- Green quantitative methods for linagliptin and empagliflozin in dosage forms
- Enhancement efficacy of omeprazole by conjugation with silver nanoparticles as a urease inhibitor
- Residual, sequential extraction, and ecological risk assessment of some metals in ash from municipal solid waste incineration, Vietnam
- Green synthesis of ZnO nanoparticles using the mangosteen (Garcinia mangostana L.) leaf extract: Comparative preliminary in vitro antibacterial study
- Simultaneous determination of lesinurad and febuxostat in commercial fixed-dose combinations using a greener normal-phase HPTLC method
- A greener RP-HPLC method for quaternary estimation of caffeine, paracetamol, levocetirizine, and phenylephrine acquiring AQbD with stability studies
- Optimization of biomass durian peel as a heterogeneous catalyst in biodiesel production using microwave irradiation
- Thermal treatment impact on the evolution of active phases in layered double hydroxide-based ZnCr photocatalysts: Photodegradation and antibacterial performance
- Preparation of silymarin-loaded zein polysaccharide core–shell nanostructures and evaluation of their biological potentials
- Preparation and characterization of composite-modified PA6 fiber for spectral heating and heat storage applications
- Preparation and electrocatalytic oxygen evolution of bimetallic phosphates (NiFe)2P/NF
- Rod-shaped Mo(vi) trichalcogenide–Mo(vi) oxide decorated on poly(1-H pyrrole) as a promising nanocomposite photoelectrode for green hydrogen generation from sewage water with high efficiency
- Green synthesis and studies on citrus medica leaf extract-mediated Au–ZnO nanocomposites: A sustainable approach for efficient photocatalytic degradation of rhodamine B dye in aqueous media
- Cellulosic materials for the removal of ciprofloxacin from aqueous environments
- The analytical assessment of metal contamination in industrial soils of Saudi Arabia using the inductively coupled plasma technology
- The effect of modified oily sludge on the slurry ability and combustion performance of coal water slurry
- Eggshell waste transformation to calcium chloride anhydride as food-grade additive and eggshell membranes as enzyme immobilization carrier
- Synthesis of EPAN and applications in the encapsulation of potassium humate
- Biosynthesis and characterization of silver nanoparticles from Cedrela toona leaf extracts: An exploration into their antibacterial, anticancer, and antioxidant potential
- Enhancing mechanical and rheological properties of HDPE films through annealing for eco-friendly agricultural applications
- Immobilisation of catalase purified from mushroom (Hydnum repandum) onto glutaraldehyde-activated chitosan and characterisation: Its application for the removal of hydrogen peroxide from artificial wastewater
- Sodium titanium oxide/zinc oxide (STO/ZnO) photocomposites for efficient dye degradation applications
- Effect of ex situ, eco-friendly ZnONPs incorporating green synthesised Moringa oleifera leaf extract in enhancing biochemical and molecular aspects of Vicia faba L. under salt stress
- Biosynthesis and characterization of selenium and silver nanoparticles using Trichoderma viride filtrate and their impact on Culex pipiens
- Photocatalytic degradation of organic dyes and biological potentials of biogenic zinc oxide nanoparticles synthesized using the polar extract of Cyperus scariosus R.Br. (Cyperaceae)
- Assessment of antiproliferative activity of green-synthesized nickel oxide nanoparticles against glioblastoma cells using Terminalia chebula
- Chlorine-free synthesis of phosphinic derivatives by change in the P-function
- Anticancer, antioxidant, and antimicrobial activities of nanoemulsions based on water-in-olive oil and loaded on biogenic silver nanoparticles
- Study and mechanism of formation of phosphorus production waste in Kazakhstan
- Synthesis and stabilization of anatase form of biomimetic TiO2 nanoparticles for enhancing anti-tumor potential
- Microwave-supported one-pot reaction for the synthesis of 5-alkyl/arylidene-2-(morpholin/thiomorpholin-4-yl)-1,3-thiazol-4(5H)-one derivatives over MgO solid base
- Screening the phytochemicals in Perilla leaves and phytosynthesis of bioactive silver nanoparticles for potential antioxidant and wound-healing application
- Graphene oxide/chitosan/manganese/folic acid-brucine functionalized nanocomposites show anticancer activity against liver cancer cells
- Nature of serpentinite interactions with low-concentration sulfuric acid solutions
- Multi-objective statistical optimisation utilising response surface methodology to predict engine performance using biofuels from waste plastic oil in CRDi engines
- Microwave-assisted extraction of acetosolv lignin from sugarcane bagasse and electrospinning of lignin/PEO nanofibres for carbon fibre production
- Biosynthesis, characterization, and investigation of cytotoxic activities of selenium nanoparticles utilizing Limosilactobacillus fermentum
- Highly photocatalytic materials based on the decoration of poly(O-chloroaniline) with molybdenum trichalcogenide oxide for green hydrogen generation from Red Sea water
- Highly efficient oil–water separation using superhydrophobic cellulose aerogels derived from corn straw
- Beta-cyclodextrin–Phyllanthus emblica emulsion for zinc oxide nanoparticles: Characteristics and photocatalysis
- Assessment of antimicrobial activity and methyl orange dye removal by Klebsiella pneumoniae-mediated silver nanoparticles
- Influential eradication of resistant Salmonella Typhimurium using bioactive nanocomposites from chitosan and radish seed-synthesized nanoselenium
- Antimicrobial activities and neuroprotective potential for Alzheimer’s disease of pure, Mn, Co, and Al-doped ZnO ultra-small nanoparticles
- Green synthesis of silver nanoparticles from Bauhinia variegata and their biological applications
- Synthesis and optimization of long-chain fatty acids via the oxidation of long-chain fatty alcohols
- Eminent Red Sea water hydrogen generation via a Pb(ii)-iodide/poly(1H-pyrrole) nanocomposite photocathode
- Green synthesis and effective genistein production by fungal β-glucosidase immobilized on Al2O3 nanocrystals synthesized in Cajanus cajan L. (Millsp.) leaf extracts
- Green stability-indicating RP-HPTLC technique for determining croconazole hydrochloride
- Green synthesis of La2O3–LaPO4 nanocomposites using Charybdis natator for DNA binding, cytotoxic, catalytic, and luminescence applications
- Eco-friendly drugs induce cellular changes in colistin-resistant bacteria
- Tangerine fruit peel extract mediated biogenic synthesized silver nanoparticles and their potential antimicrobial, antioxidant, and cytotoxic assessments
- Green synthesis on performance characteristics of a direct injection diesel engine using sandbox seed oil
- A highly sensitive β-AKBA-Ag-based fluorescent “turn off” chemosensor for rapid detection of abamectin in tomatoes
- Green synthesis and physical characterization of zinc oxide nanoparticles (ZnO NPs) derived from the methanol extract of Euphorbia dracunculoides Lam. (Euphorbiaceae) with enhanced biosafe applications
- Detection of morphine and data processing using surface plasmon resonance imaging sensor
- Effects of nanoparticles on the anaerobic digestion properties of sulfamethoxazole-containing chicken manure and analysis of bio-enzymes
- Bromic acid-thiourea synergistic leaching of sulfide gold ore
- Green chemistry approach to synthesize titanium dioxide nanoparticles using Fagonia Cretica extract, novel strategy for developing antimicrobial and antidiabetic therapies
- Green synthesis and effective utilization of biogenic Al2O3-nanocoupled fungal lipase in the resolution of active homochiral 2-octanol and its immobilization via aluminium oxide nanoparticles
- Eco-friendly RP-HPLC approach for simultaneously estimating the promising combination of pentoxifylline and simvastatin in therapeutic potential for breast cancer: Appraisal of greenness, whiteness, and Box–Behnken design
- Use of a humidity adsorbent derived from cockleshell waste in Thai fried fish crackers (Keropok)
- One-pot green synthesis, biological evaluation, and in silico study of pyrazole derivatives obtained from chalcones
- Bio-sorption of methylene blue and production of biofuel by brown alga Cystoseira sp. collected from Neom region, Kingdom of Saudi Arabia
- Synthesis of motexafin gadolinium: A promising radiosensitizer and imaging agent for cancer therapy
- The impact of varying sizes of silver nanoparticles on the induction of cellular damage in Klebsiella pneumoniae involving diverse mechanisms
- Microwave-assisted green synthesis, characterization, and in vitro antibacterial activity of NiO nanoparticles obtained from lemon peel extract
- Rhus microphylla-mediated biosynthesis of copper oxide nanoparticles for enhanced antibacterial and antibiofilm efficacy
- Harnessing trichalcogenide–molybdenum(vi) sulfide and molybdenum(vi) oxide within poly(1-amino-2-mercaptobenzene) frameworks as a photocathode for sustainable green hydrogen production from seawater without sacrificial agents
- Magnetically recyclable Fe3O4@SiO2 supported phosphonium ionic liquids for efficient and sustainable transformation of CO2 into oxazolidinones
- A comparative study of Fagonia arabica fabricated silver sulfide nanoparticles (Ag2S) and silver nanoparticles (AgNPs) with distinct antimicrobial, anticancer, and antioxidant properties
- Visible light photocatalytic degradation and biological activities of Aegle marmelos-mediated cerium oxide nanoparticles
- Physical intrinsic characteristics of spheroidal particles in coal gasification fine slag
- Exploring the effect of tea dust magnetic biochar on agricultural crops grown in polycyclic aromatic hydrocarbon contaminated soil
- Crosslinked chitosan-modified ultrafiltration membranes for efficient surface water treatment and enhanced anti-fouling performances
- Study on adsorption characteristics of biochars and their modified biochars for removal of organic dyes from aqueous solution
- Zein polymer nanocarrier for Ocimum basilicum var. purpurascens extract: Potential biomedical use
- Green synthesis, characterization, and in vitro and in vivo biological screening of iron oxide nanoparticles (Fe3O4) generated with hydroalcoholic extract of aerial parts of Euphorbia milii
- Novel microwave-based green approach for the synthesis of dual-loaded cyclodextrin nanosponges: Characterization, pharmacodynamics, and pharmacokinetics evaluation
- Bi2O3–BiOCl/poly-m-methyl aniline nanocomposite thin film for broad-spectrum light-sensing
- Green synthesis and characterization of CuO/ZnO nanocomposite using Musa acuminata leaf extract for cytotoxic studies on colorectal cancer cells (HCC2998)
- Review Articles
- Materials-based drug delivery approaches: Recent advances and future perspectives
- A review of thermal treatment for bamboo and its composites
- An overview of the role of nanoherbicides in tackling challenges of weed management in wheat: A novel approach
- An updated review on carbon nanomaterials: Types, synthesis, functionalization and applications, degradation and toxicity
- Special Issue: Emerging green nanomaterials for sustainable waste management and biomedical applications
- Green synthesis of silver nanoparticles using mature-pseudostem extracts of Alpinia nigra and their bioactivities
- Special Issue: New insights into nanopythotechnology: current trends and future prospects
- Green synthesis of FeO nanoparticles from coffee and its application for antibacterial, antifungal, and anti-oxidation activity
- Dye degradation activity of biogenically synthesized Cu/Fe/Ag trimetallic nanoparticles
- Special Issue: Composites and green composites
- Recent trends and advancements in the utilization of green composites and polymeric nanocarriers for enhancing food quality and sustainable processing
- Retraction
- Retraction of “Biosynthesis and characterization of silver nanoparticles from Cedrela toona leaf extracts: An exploration into their antibacterial, anticancer, and antioxidant potential”
- Retraction of “Photocatalytic degradation of organic dyes and biological potentials of biogenic zinc oxide nanoparticles synthesized using the polar extract of Cyperus scariosus R.Br. (Cyperaceae)”
- Retraction to “Green synthesis on performance characteristics of a direct injection diesel engine using sandbox seed oil”
Articles in the same Issue
- Research Articles
- Green polymer electrolyte and activated charcoal-based supercapacitor for energy harvesting application: Electrochemical characteristics
- Research on the adsorption of Co2+ ions using halloysite clay and the ability to recover them by electrodeposition method
- Simultaneous estimation of ibuprofen, caffeine, and paracetamol in commercial products using a green reverse-phase HPTLC method
- Isolation, screening and optimization of alkaliphilic cellulolytic fungi for production of cellulase
- Functionalized gold nanoparticles coated with bacterial alginate and their antibacterial and anticancer activities
- Comparative analysis of bio-based amino acid surfactants obtained via Diels–Alder reaction of cyclic anhydrides
- Biosynthesis of silver nanoparticles on yellow phosphorus slag and its application in organic coatings
- Exploring antioxidant potential and phenolic compound extraction from Vitis vinifera L. using ultrasound-assisted extraction
- Manganese and copper-coated nickel oxide nanoparticles synthesized from Carica papaya leaf extract induce antimicrobial activity and breast cancer cell death by triggering mitochondrial caspases and p53
- Insight into heating method and Mozafari method as green processing techniques for the synthesis of micro- and nano-drug carriers
- Silicotungstic acid supported on Bi-based MOF-derived metal oxide for photodegradation of organic dyes
- Synthesis and characterization of capsaicin nanoparticles: An attempt to enhance its bioavailability and pharmacological actions
- Synthesis of Lawsonia inermis-encased silver–copper bimetallic nanoparticles with antioxidant, antibacterial, and cytotoxic activity
- Facile, polyherbal drug-mediated green synthesis of CuO nanoparticles and their potent biological applications
- Zinc oxide-manganese oxide/carboxymethyl cellulose-folic acid-sesamol hybrid nanomaterials: A molecularly targeted strategy for advanced triple-negative breast cancer therapy
- Exploring the antimicrobial potential of biogenically synthesized graphene oxide nanoparticles against targeted bacterial and fungal pathogens
- Biofabrication of silver nanoparticles using Uncaria tomentosa L.: Insight into characterization, antibacterial activities combined with antibiotics, and effect on Triticum aestivum germination
- Membrane distillation of synthetic urine for use in space structural habitat systems
- Investigation on mechanical properties of the green synthesis bamboo fiber/eggshell/coconut shell powder-based hybrid biocomposites under NaOH conditions
- Green synthesis of magnesium oxide nanoparticles using endophytic fungal strain to improve the growth, metabolic activities, yield traits, and phenolic compounds content of Nigella sativa L.
- Estimation of greenhouse gas emissions from rice and annual upland crops in Red River Delta of Vietnam using the denitrification–decomposition model
- Synthesis of humic acid with the obtaining of potassium humate based on coal waste from the Lenger deposit, Kazakhstan
- Ascorbic acid-mediated selenium nanoparticles as potential antihyperuricemic, antioxidant, anticoagulant, and thrombolytic agents
- Green synthesis of silver nanoparticles using Illicium verum extract: Optimization and characterization for biomedical applications
- Antibacterial and dynamical behaviour of silicon nanoparticles influenced sustainable waste flax fibre-reinforced epoxy composite for biomedical application
- Optimising coagulation/flocculation using response surface methodology and application of floc in biofertilisation
- Green synthesis and multifaceted characterization of iron oxide nanoparticles derived from Senna bicapsularis for enhanced in vitro and in vivo biological investigation
- Potent antibacterial nanocomposites from okra mucilage/chitosan/silver nanoparticles for multidrug-resistant Salmonella Typhimurium eradication
- Trachyspermum copticum aqueous seed extract-derived silver nanoparticles: Exploration of their structural characterization and comparative antibacterial performance against gram-positive and gram-negative bacteria
- Microwave-assisted ultrafine silver nanoparticle synthesis using Mitragyna speciosa for antimalarial applications
- Green synthesis and characterisation of spherical structure Ag/Fe2O3/TiO2 nanocomposite using acacia in the presence of neem and tulsi oils
- Green quantitative methods for linagliptin and empagliflozin in dosage forms
- Enhancement efficacy of omeprazole by conjugation with silver nanoparticles as a urease inhibitor
- Residual, sequential extraction, and ecological risk assessment of some metals in ash from municipal solid waste incineration, Vietnam
- Green synthesis of ZnO nanoparticles using the mangosteen (Garcinia mangostana L.) leaf extract: Comparative preliminary in vitro antibacterial study
- Simultaneous determination of lesinurad and febuxostat in commercial fixed-dose combinations using a greener normal-phase HPTLC method
- A greener RP-HPLC method for quaternary estimation of caffeine, paracetamol, levocetirizine, and phenylephrine acquiring AQbD with stability studies
- Optimization of biomass durian peel as a heterogeneous catalyst in biodiesel production using microwave irradiation
- Thermal treatment impact on the evolution of active phases in layered double hydroxide-based ZnCr photocatalysts: Photodegradation and antibacterial performance
- Preparation of silymarin-loaded zein polysaccharide core–shell nanostructures and evaluation of their biological potentials
- Preparation and characterization of composite-modified PA6 fiber for spectral heating and heat storage applications
- Preparation and electrocatalytic oxygen evolution of bimetallic phosphates (NiFe)2P/NF
- Rod-shaped Mo(vi) trichalcogenide–Mo(vi) oxide decorated on poly(1-H pyrrole) as a promising nanocomposite photoelectrode for green hydrogen generation from sewage water with high efficiency
- Green synthesis and studies on citrus medica leaf extract-mediated Au–ZnO nanocomposites: A sustainable approach for efficient photocatalytic degradation of rhodamine B dye in aqueous media
- Cellulosic materials for the removal of ciprofloxacin from aqueous environments
- The analytical assessment of metal contamination in industrial soils of Saudi Arabia using the inductively coupled plasma technology
- The effect of modified oily sludge on the slurry ability and combustion performance of coal water slurry
- Eggshell waste transformation to calcium chloride anhydride as food-grade additive and eggshell membranes as enzyme immobilization carrier
- Synthesis of EPAN and applications in the encapsulation of potassium humate
- Biosynthesis and characterization of silver nanoparticles from Cedrela toona leaf extracts: An exploration into their antibacterial, anticancer, and antioxidant potential
- Enhancing mechanical and rheological properties of HDPE films through annealing for eco-friendly agricultural applications
- Immobilisation of catalase purified from mushroom (Hydnum repandum) onto glutaraldehyde-activated chitosan and characterisation: Its application for the removal of hydrogen peroxide from artificial wastewater
- Sodium titanium oxide/zinc oxide (STO/ZnO) photocomposites for efficient dye degradation applications
- Effect of ex situ, eco-friendly ZnONPs incorporating green synthesised Moringa oleifera leaf extract in enhancing biochemical and molecular aspects of Vicia faba L. under salt stress
- Biosynthesis and characterization of selenium and silver nanoparticles using Trichoderma viride filtrate and their impact on Culex pipiens
- Photocatalytic degradation of organic dyes and biological potentials of biogenic zinc oxide nanoparticles synthesized using the polar extract of Cyperus scariosus R.Br. (Cyperaceae)
- Assessment of antiproliferative activity of green-synthesized nickel oxide nanoparticles against glioblastoma cells using Terminalia chebula
- Chlorine-free synthesis of phosphinic derivatives by change in the P-function
- Anticancer, antioxidant, and antimicrobial activities of nanoemulsions based on water-in-olive oil and loaded on biogenic silver nanoparticles
- Study and mechanism of formation of phosphorus production waste in Kazakhstan
- Synthesis and stabilization of anatase form of biomimetic TiO2 nanoparticles for enhancing anti-tumor potential
- Microwave-supported one-pot reaction for the synthesis of 5-alkyl/arylidene-2-(morpholin/thiomorpholin-4-yl)-1,3-thiazol-4(5H)-one derivatives over MgO solid base
- Screening the phytochemicals in Perilla leaves and phytosynthesis of bioactive silver nanoparticles for potential antioxidant and wound-healing application
- Graphene oxide/chitosan/manganese/folic acid-brucine functionalized nanocomposites show anticancer activity against liver cancer cells
- Nature of serpentinite interactions with low-concentration sulfuric acid solutions
- Multi-objective statistical optimisation utilising response surface methodology to predict engine performance using biofuels from waste plastic oil in CRDi engines
- Microwave-assisted extraction of acetosolv lignin from sugarcane bagasse and electrospinning of lignin/PEO nanofibres for carbon fibre production
- Biosynthesis, characterization, and investigation of cytotoxic activities of selenium nanoparticles utilizing Limosilactobacillus fermentum
- Highly photocatalytic materials based on the decoration of poly(O-chloroaniline) with molybdenum trichalcogenide oxide for green hydrogen generation from Red Sea water
- Highly efficient oil–water separation using superhydrophobic cellulose aerogels derived from corn straw
- Beta-cyclodextrin–Phyllanthus emblica emulsion for zinc oxide nanoparticles: Characteristics and photocatalysis
- Assessment of antimicrobial activity and methyl orange dye removal by Klebsiella pneumoniae-mediated silver nanoparticles
- Influential eradication of resistant Salmonella Typhimurium using bioactive nanocomposites from chitosan and radish seed-synthesized nanoselenium
- Antimicrobial activities and neuroprotective potential for Alzheimer’s disease of pure, Mn, Co, and Al-doped ZnO ultra-small nanoparticles
- Green synthesis of silver nanoparticles from Bauhinia variegata and their biological applications
- Synthesis and optimization of long-chain fatty acids via the oxidation of long-chain fatty alcohols
- Eminent Red Sea water hydrogen generation via a Pb(ii)-iodide/poly(1H-pyrrole) nanocomposite photocathode
- Green synthesis and effective genistein production by fungal β-glucosidase immobilized on Al2O3 nanocrystals synthesized in Cajanus cajan L. (Millsp.) leaf extracts
- Green stability-indicating RP-HPTLC technique for determining croconazole hydrochloride
- Green synthesis of La2O3–LaPO4 nanocomposites using Charybdis natator for DNA binding, cytotoxic, catalytic, and luminescence applications
- Eco-friendly drugs induce cellular changes in colistin-resistant bacteria
- Tangerine fruit peel extract mediated biogenic synthesized silver nanoparticles and their potential antimicrobial, antioxidant, and cytotoxic assessments
- Green synthesis on performance characteristics of a direct injection diesel engine using sandbox seed oil
- A highly sensitive β-AKBA-Ag-based fluorescent “turn off” chemosensor for rapid detection of abamectin in tomatoes
- Green synthesis and physical characterization of zinc oxide nanoparticles (ZnO NPs) derived from the methanol extract of Euphorbia dracunculoides Lam. (Euphorbiaceae) with enhanced biosafe applications
- Detection of morphine and data processing using surface plasmon resonance imaging sensor
- Effects of nanoparticles on the anaerobic digestion properties of sulfamethoxazole-containing chicken manure and analysis of bio-enzymes
- Bromic acid-thiourea synergistic leaching of sulfide gold ore
- Green chemistry approach to synthesize titanium dioxide nanoparticles using Fagonia Cretica extract, novel strategy for developing antimicrobial and antidiabetic therapies
- Green synthesis and effective utilization of biogenic Al2O3-nanocoupled fungal lipase in the resolution of active homochiral 2-octanol and its immobilization via aluminium oxide nanoparticles
- Eco-friendly RP-HPLC approach for simultaneously estimating the promising combination of pentoxifylline and simvastatin in therapeutic potential for breast cancer: Appraisal of greenness, whiteness, and Box–Behnken design
- Use of a humidity adsorbent derived from cockleshell waste in Thai fried fish crackers (Keropok)
- One-pot green synthesis, biological evaluation, and in silico study of pyrazole derivatives obtained from chalcones
- Bio-sorption of methylene blue and production of biofuel by brown alga Cystoseira sp. collected from Neom region, Kingdom of Saudi Arabia
- Synthesis of motexafin gadolinium: A promising radiosensitizer and imaging agent for cancer therapy
- The impact of varying sizes of silver nanoparticles on the induction of cellular damage in Klebsiella pneumoniae involving diverse mechanisms
- Microwave-assisted green synthesis, characterization, and in vitro antibacterial activity of NiO nanoparticles obtained from lemon peel extract
- Rhus microphylla-mediated biosynthesis of copper oxide nanoparticles for enhanced antibacterial and antibiofilm efficacy
- Harnessing trichalcogenide–molybdenum(vi) sulfide and molybdenum(vi) oxide within poly(1-amino-2-mercaptobenzene) frameworks as a photocathode for sustainable green hydrogen production from seawater without sacrificial agents
- Magnetically recyclable Fe3O4@SiO2 supported phosphonium ionic liquids for efficient and sustainable transformation of CO2 into oxazolidinones
- A comparative study of Fagonia arabica fabricated silver sulfide nanoparticles (Ag2S) and silver nanoparticles (AgNPs) with distinct antimicrobial, anticancer, and antioxidant properties
- Visible light photocatalytic degradation and biological activities of Aegle marmelos-mediated cerium oxide nanoparticles
- Physical intrinsic characteristics of spheroidal particles in coal gasification fine slag
- Exploring the effect of tea dust magnetic biochar on agricultural crops grown in polycyclic aromatic hydrocarbon contaminated soil
- Crosslinked chitosan-modified ultrafiltration membranes for efficient surface water treatment and enhanced anti-fouling performances
- Study on adsorption characteristics of biochars and their modified biochars for removal of organic dyes from aqueous solution
- Zein polymer nanocarrier for Ocimum basilicum var. purpurascens extract: Potential biomedical use
- Green synthesis, characterization, and in vitro and in vivo biological screening of iron oxide nanoparticles (Fe3O4) generated with hydroalcoholic extract of aerial parts of Euphorbia milii
- Novel microwave-based green approach for the synthesis of dual-loaded cyclodextrin nanosponges: Characterization, pharmacodynamics, and pharmacokinetics evaluation
- Bi2O3–BiOCl/poly-m-methyl aniline nanocomposite thin film for broad-spectrum light-sensing
- Green synthesis and characterization of CuO/ZnO nanocomposite using Musa acuminata leaf extract for cytotoxic studies on colorectal cancer cells (HCC2998)
- Review Articles
- Materials-based drug delivery approaches: Recent advances and future perspectives
- A review of thermal treatment for bamboo and its composites
- An overview of the role of nanoherbicides in tackling challenges of weed management in wheat: A novel approach
- An updated review on carbon nanomaterials: Types, synthesis, functionalization and applications, degradation and toxicity
- Special Issue: Emerging green nanomaterials for sustainable waste management and biomedical applications
- Green synthesis of silver nanoparticles using mature-pseudostem extracts of Alpinia nigra and their bioactivities
- Special Issue: New insights into nanopythotechnology: current trends and future prospects
- Green synthesis of FeO nanoparticles from coffee and its application for antibacterial, antifungal, and anti-oxidation activity
- Dye degradation activity of biogenically synthesized Cu/Fe/Ag trimetallic nanoparticles
- Special Issue: Composites and green composites
- Recent trends and advancements in the utilization of green composites and polymeric nanocarriers for enhancing food quality and sustainable processing
- Retraction
- Retraction of “Biosynthesis and characterization of silver nanoparticles from Cedrela toona leaf extracts: An exploration into their antibacterial, anticancer, and antioxidant potential”
- Retraction of “Photocatalytic degradation of organic dyes and biological potentials of biogenic zinc oxide nanoparticles synthesized using the polar extract of Cyperus scariosus R.Br. (Cyperaceae)”
- Retraction to “Green synthesis on performance characteristics of a direct injection diesel engine using sandbox seed oil”