Next Article in Journal
Biological Application of the Allopathic Characteristics of the Genus Maclura: A Review
Next Article in Special Issue
First Chemical Profile Analysis of Acacia Pods
Previous Article in Journal
Anatomical and Transcriptome Analyses of Moso Bamboo Culm Neck Growth: Unveiling Key Insights
Previous Article in Special Issue
Quantification of Dry Matter Content in Hass Avocado by Near-Infrared Spectroscopy (NIRS) Scanning Different Fruit Zones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Wild-Grown Romanian Helleborus purpurascens Approach to Novel Chitosan Phyto-Nanocarriers—Metabolite Profile and Antioxidant Properties

by
Adina-Elena Segneanu
1,*,†,
Gabriela Vlase
1,2,
Titus Vlase
1,2,
Crina Andreea Sicoe
3,
Maria Viorica Ciocalteu
4,
Dumitru Daniel Herea
5,
Ovidiu-Florin Ghirlea
6,
Ioan Grozescu
7 and
Valentin Nanescu
4,†
1
Institute for Advanced Environmental Research-West, University of Timisoara (ICAM-WUT), Oituz Nr. 4, 300086 Timisoara, Romania
2
Research Centre for Thermal Analysis Environmental Problems, West University of Timisoara, Pestalozzi St. 16, 300115 Timisoara, Romania
3
Faculty of Chemistry, Biology, Geography, West University of Timisoara, Pestalozzi St. 16, 300115 Timisoara, Romania
4
Faculty of Pharmacy, University of Medicine and Pharmacy Craiova, St. Petru Rareș 2, 200349 Craiova, Romania
5
National Institute of Research and Development for Technical Physics, 47 Mangeron Blvd., 700050 Iasi, Romania
6
Faculty of Medicine, “Victor Babes” University of Medicine and Pharmacy, square Eftimie Murgu No. 2, 300041 Timisoara, Romania
7
CAICON Department, University Politehnica Timisoara, 300006 Timisoara, Romania
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Plants 2023, 12(19), 3479; https://doi.org/10.3390/plants12193479
Submission received: 11 September 2023 / Revised: 27 September 2023 / Accepted: 29 September 2023 / Published: 4 October 2023
(This article belongs to the Special Issue Spectra Analysis and Plants Research 2.0)

Abstract

:
The current nanomedicinal approach combines medicinal plants and nanotechnology to create new scaffolds with enhanced bioavailability, biodistribution and controlled release. In an innovative approach to herb encapsulation in nanosized chitosan matrices, wild-grown Romanian Helleborus purpurascens was used to prepare two new chitosan nanocarriers. The first carrier preparation involved the nanoencapsulation of hellebore in chitosan. The second carrier emerged from two distinct stages: hellebore-AgNPs phyto-carrier system succeeded by nanoencapsulation in chitosan. The morphostructural characteristics and thermal behavior of these newly prepared nanocarriers were examined using FT-IR, XRD, DLS, SEM, EDS and thermogravimetric analyses. In addition, the encapsulation yield, encapsulation efficiency and encapsulation contents were investigated. The antioxidant activity was estimated using four in vitro, noncompetitive methods: total phenolic assay; 2,2-diphenyl-1-picrylhydrazyl (DPPH) radical scavenging assay; phosphomolybdate (i.e., total antioxidant capacity); and iron(III)-phenanthroline antioxidant assay. Moreover, this study reports the first low-molecular-weight metabolite profile of wild-grown Romanian Helleborus purpurascens Waldst. & Kit. A total of one hundred and five secondary metabolites were identified in the mass spectra (MS)-positive mode from fourteen secondary metabolite categories (alkaloids, butenolides, bufadienolides, phytoecdysteroids, amino acids and peptides, terpenoids, fatty acids, flavonoids, phenolic acids, sterols, glycosides, carbohydrates, nucleosides and miscellaneous). The collective results suggest the potential application is a promising new antioxidant vehicle candidate in tumor therapeutic strategy.

1. Introduction

Helleborus purpurascens Waldst. & Kit. (Helleborus purpurascens L.) from the Ranunculaceae family has been used as a medicinal plant in Romanian and European ethnomedicine for centuries [1,2]. One of the first mentions of the therapeutic effect of this plant occurs in Diseases of Women by Hippocrates. Also, the use of the plant in Dacian medicine is reported in De Materia Medica [2]. In traditional European medicine, hellebore is used for its antimicrobial, anti-inflammatory, antitumor, cardiotonic, analgesic, antiseizure, laxative and other effects [2,3,4]. Modern Romanian studies from the end of the last century have demonstrated the therapeutic potential of the plant in the treatment of severe forms of rheumatism [1,2,5,6,7]. Recent research has confirmed their unique antitumor and immunomodulating activity [8,9,10,11]. Hellebore’s outstanding biological activity is the outcome of the combined and synergistic action of its multiple bioactive compounds (bufadienolides, butenolides, phytoecdysteroids, terpenoids, alkaloids, sterols, flavonoids, fatty acids, peptides, lactones, phenolic acids, carbohydrates and others) [1,2].
Synthetic drugs are the primary option in the therapeutic strategy for severe ailments. In addition, factors such as pollution, self-medication or others can induce drug resistance and, thus, treatment failure [12]. Conversely, ethnomedicine had an essential role in the development of humanity and is currently the first choice as a treatment strategy for various diseases in over 80% of all countries, according to the WHO’s estimation [13]. In this context, urgent alternative, efficient strategies are required. To this end, intensive efforts have been made to develop new and performant plant-based drug delivery systems. Consequently, substantially modern pharmaceuticals are based on natural products despite the numerous technical barriers related to bioactive compounds, namely, total synthesis and optimization [14,15]. Furthermore, many phytoconstituents with exceptional biological activity display limited in vivo bioavailability and reduced membrane transport [14,16,17]. As a result, numerous studies have focused on new drug delivery systems with designs that encapsulate secondary metabolites in different biomolecules [14,16,17,18].
Alternatively, metal nanoparticles, particularly magnetic or precious metals, have been exploited in numerous biomedical applications because of their remarkable physicochemical and biological characteristics [19,20,21,22]. Among these, silver nanoparticles show huge biopotential (high stability and flexibility, biocompatibility, ability to cross the blood–brain barrier (BBB)) for detection, diagnosis and treatment (drug delivery, antibacterial, antiviral, antifungal, antitumor and antioxidant agents) [23,24,25]. Nonetheless, the main drawback of AgNPs is the genotoxicity and cytotoxicity mechanisms of AgNPs that occur in different tissues or organs under certain conditions (concentration, contact time, dimensions, etc.). The different latest approaches based on AgNP coating/encapsulation or functionalization have overcome these shortcomings. Among the biomaterials used for a targeted drug delivery design, chitosan stands out for its performance and versatility. Chitosan is an incredibly versatile, low-cost substance with numerous applications in diverse fields, such as agriculture, biotechnology, cosmetics, food preservation, medicine, nanotechnology, textiles and wastewater treatment. It is a natural biopolymer composed of various units (2-acetamido-2-deoxy-D-glucopyranose and 2-amino-2-deoxy-D-glucopyranose), and is the next plentifullest natural polymer after cellulose [26,27]. This biopolymer is produced using deacetylating chitin in an alkaline environment. Chitosan is the most suitable material for biomedical applications because of its outstanding features: biocompatibility, nontoxicity, mucoadhesive properties and antibacterial activity [26,27,28]. Moreover, chitosan nanoparticles demonstrate significant potential for cancer and osteoarthritis by providing target specificity, drug stability and efficiency [26,27,28].
Great effort has been devoted to the study of antioxidants’ role in the prevention of cancer and its cure [29,30,31,32]. However, despite their critical potential in tumor therapy, their effectiveness is still a controversial issue in the medical field due to several limitations, namely, low bioavailability and transmembrane permeability, inadequate doses, uneven distribution, among others, which can decrease their beneficial outcomes [31]. Therefore, a novel approach is necessary to ensure the successful delivery of antioxidants and maximize their therapeutic effects. Furthermore, the activity of plant biology is a complex process that relies on the synergistic action of various phytoconstituents in different proportions [32,33,34]. Factors such as climate, soil composition and the plant’s harvesting stage can significantly impact the antioxidant activity of these phytoconstituents. The antioxidant activities of phytoconstituents varies with certain factors: ecological parameters (sunlight, temperature, aridity, humidity, soil composition and pH), harvest growth stage and others [35].
Antioxidant agents are usually classified based on their physical–chemical properties, source and mechanism of action. Additionally, it is worth noting that various criteria, such as metabolism pathway, bioactive phytocompound chemical structure, amount, bioavailability and specific rate constant could significantly affect the effectiveness of an antioxidant agent [35,36,37,38,39]. In conclusion, to gain a comprehensive understanding of phytoconstituents’ biologic activity, it is crucial to simultaneously examine the factors encompassing the origin and composition of the phytoconstituents with the techniques utilized to assess their antioxidant properties.
An herb or plant extract loaded into biopolymer nanoparticles ensures the controlled intake of highly active phytoconstituents, cancels the effects of overdose-induced tissue damage and increases the therapeutic efficiency compared to individual components because of the complementary action of the plant and the biopolymer [26,27]. Moreover, nanocarrier systems present multiple advantages compared to other drug delivery systems: improved efficiency through the complementary action between encapsulated biomolecules and chitosan nanoparticles; higher stability; fast, selective, targeted and sustained release; and excellent biocompatibility. In this sense, the research on developing successful phyto-carriers based on nanoparticles reflects a substantial improvement in the therapeutic performance, target specificity, control released and overcoming of drug resistance.
In this study, the approach to developing new carrier systems moved to a different level using two different chitosan phyto-nanocarriers from hellebore. The first nanocarrier involved the encapsulation of hellebore in chitosan nanoparticles. The second one emerged in a new phyto-carrier system based on AgNPs encapsulated in chitosan nanoparticles to ensure the complementary and synergistic action of AgNPs and the secondary metabolites from hellebore. To the best of our knowledge, this is the first study reporting on the concurrent use of hellebore, chitosan and AgNPs. The physical, chemical and antioxidant properties of the new nanocarriers were evaluated.

2. Results and Discussion

Plants encompass a multitude of secondary metabolites with highly complex chemical structures. There is a close relationship between environmental factors and the configuration and proportion of bioactive compounds in a specific plant [40,41]. Furthermore, the amount of active phytoconstituents depends on the selected experimental conditions (extraction procedure, temperature, time, etc.) [4,11,42].
The collective action of all secondary metabolites dictates the biological activity of a particular plant [43,44]. Consequently, determining a direct connection between the chemical composition of a plant and its therapeutic activity is arduous. Additionally, the physical–chemical properties (size, chemical or thermic stability, etc.) of some phytochemicals isolated from plant material impose restrictions on the biological environment (e.g., biodisponibility) and, thus, significantly diminish their bioefficiency [45].
Studies on Romanian Helleborus purpurascens are few and target only the content of a few classes of phytochemicals (solely isolated from the plant root) and their pharmacological activities [5,6,7,8,9,10,11]. In addition, the integrated chemical screening of several plant parts (stem, leaves, and flowers) of this medicinal plant from Romanian natural vegetation has not been carried out yet. The secondary metabolite chemical screening of hellebore samples was performed using gas-chromatography coupled with mass spectroscopy (GC-MS) and electrospray ionization-quadrupole time-of-flight mass spectrometry (ESI-QTOF-MS) analysis.
GC-MS is one of the most accurate and fast techniques used for analyzing various phytochemicals with relatively low molecular mass separation and to confirm identification [46]. To this end, the separation of the secondary metabolites from the hellebore samples was investigated (Figure S1 and Table 1)
The results of the tentative compound identification from the hellebore sample using the GC-MS method are shown in Table 1.
The GC-MS analysis (Figure S1) indicated that eight bioactive phytoconstituents were separated (96.75% of the total peak area), taking into consideration that additional procedural methods conducted before the GC-MS analysis (for instance, derivatization) are requested thermally unstable phytochemicals [46].
Conversely, the electrospray ionization-quadrupole time-of-flight mass spectrometry (EIS-QTOF-MS) analyzer allows for the fast, affordable and highly accurate identification of secondary metabolites from complex matrices. To this end, the hellebore’s metabolite profile screening was performed using the mass spectrometry method. The mass spectra depicted a sophisticated mix of low-molecular-weight biomolecules, of which some were detected and compared for identification with those of the NIST/EPA/NIH Mass Spectral Library 3.0 database and in the literature review [2,3,4,6,7,8,9,10,11].
The mass spectrum and the hellebore’s secondary metabolites identified in the ESI–QTOF–MS analysis are presented in Figure S2 and Table 2, respectively.

2.1. Screening and Classification of the Differential Metabolites

A total of 105 secondary metabolites identified through mass spectroscopy were assigned to different chemical classes: amino acids and peptides (60.95%), terpenoids (6.67%), fatty acids (6.67%), flavonoids (5.71%), phenolic acids (4.76%), bufadienolides (3.80%), sterols (1.90%), glycosides (0.95%), alkaloids (0.95%), phytoecdysteroids (0.95%), nucleosides (0.95%), butenolides (0.95%), carbohydrates (0.95%) and miscellaneous. The assignment of the identified phytoconstituents into different chemical categories is presented in Table 2.
According to Figure 1, amino acids and peptides are the largest categories of phytochemicals, constituting about 60.95%. An equal proportion of essential amino acids (phenylalanine, isoleucine, tryptophan, threonine and histidine) and nonessential amino acids (alanine, ornithine, glutamic acid, cysteine and glycine) were found in the hellebore sample. A large percentage (over 70%) of this category of secondary metabolites (isoleucine, tryptophan, phenylalanine, arginine, histidine, glycine, alanine and glutamic acid) exhibit cytotoxic, antiproliferative and immunomodulating activity [65,66,67,68,69,70,71].
Over 50 small peptides (di-, tri- and tetrapeptides) were identified in the hellebore sample. Over two-thirds of these are sulfur-containing peptides. Various studies have reported on the antitumoral, antimicrobial, neuroregulating and antiviral (anti-HIV) properties of phytopeptides and thionins [68,72,73,74,75]. This difference in the compositions of the small peptides in the hellebore samples compared to our previously reported results [11] confirms that the profile of the secondary metabolites is determined by various endogenous or ecological factors, plant components, extraction methods and solvent polarities [40,41,42,76,77,78].
Bufadienolides are a category of metabolites with remarkable antitumoral, antiviral and anti-inflammatory properties [79,80,81]
Butanolides (anemonin) act as anti-inflammatory and antimicrobial agents [82,83].
Phytoecdysteroids are a category of secondary metabolites with notable antitumoral, antimicrobial, antiviral, anti-inflammatory, antioxidant, hepatoprotective and antidiabetic activities [84].
Terpenoids and sesquiterpenoids are another category of phytochemicals that were found in the hellebore samples. Numerous studies have demonstrated their notable therapeutic properties: antimicrobial, antioxidant, anticonvulsant, analgesic, neuroprotective, anti-inflammatory, anti-allergic and antitumor [66,85,86,87,88,89]. Furthermore, protoanemonin exerts antimicrobial and anti-HIV effects [2,90,91].
Fatty acids represent approximately 7% of the phytochemicals identified in the hellebore samples. Numerous research has reported on their antioxidant, anti-inflammatory, neuroprotective and cardiovascular protective properties [66,89,92].
Flavonoids are secondary metabolites with outstanding pharmacological effects: antioxidant, anti-inflammatory, antimicrobial, antitumoral and neuroprotective [93,94,95,96].
Phenolic acids represent 4.7% of the biomolecules identified in the hellebore samples. Extensive studies have demonstrated their significant therapeutic properties (antioxidant, antimicrobial, antimicrobial, antitumoral, neuroprotective and antidiabetic) [96,97,98].
Phytosterols are steroids have antioxidant, antitumoral, anti-inflammatory, immunomodulatory, neuroprotective, antidiabetic, cardiovascular protective and osteoporosis protection properties [99].
Carbohydrates act as antioxidant, anti-inflammatory, antibacterial, antiviral, antitumoral, immunomodulatory, antidiabetic and cardioprotective agents [65,100,101].
Glycosides are biomolecules with notable antitumoral (leukemia and gastric tumors) properties [102].
Magnoflorine, the alkaloid identified in the hellebore samples, has shown antitumoral, neuropsychopharmacological, immunomodulatory, anti-inflammatory, antioxidant and antifungal activities [103].
The nucleoside uridine act as a neuroregenerative and neuroprotective agent [104].
Among the miscellaneous compounds identified in the hellebore samples, resveratrol has antioxidant, antitumoral anti-inflammatory, immunomodulatory, antiobesity, antidiabetic, neuroprotective and cardiovascular protective effects [105].

2.2. Phyto-Nanocarriers

Secondary metabolites from plants have excellent therapeutic properties. However, most of these biomolecules exhibit reduced bioavailability in vivo and an inability to cross biological membranes because of their physicochemical properties (size, chemical stability, polarity, etc.) [12,68,106,107,108].
The foremost hurdles of the current antitumoral therapy are target-specificity, retention and permeability, and need to minimize the side effects and overcome chemoresistance [12,68,108]. Therefore, the development of cutting-edge phyto-nanocarrier systems through hellebore’s encapsulation in chitosan nanocapsules (HCs) and a hellebore-AgNP system’s entrapment in chitosan nanocapsules (HAgCs) combines the biological activity of phytoconstituents and a natural biopolymer in the case of the first phyto-nanocarrier. In addition, the second photo-nanocarrier (HAgC) is added to the silver nanoparticles’ biological activity to achieve a new therapeutic strategy that could overcome the resistance of bacteria to antibiotics or the drug resistance of cancer cells. Furthermore, this new approach increases the target specificity, precise control release, stability and permeability.

2.3. FT-IR Spectroscopy

FT-IR is an essential analytical tool that is widely used for the rapid, efficient and sensitive analysis of multicomponent systems [109].
The first phyto-nanocarrier’s (HC) preparation and the following preparation of the second phyto-nanocarrier occurred in two stages: (a) the metallic nanoparticles were incorporated into the pores of the hellebore; (b) the encapsulation of the hellebore-AgNPs into the chitosan nanoparticles was investigated using FT-IR spectroscopy to identify the functional groups specific to each component. The individual FT-IR spectrum of the hellebore, chitosan, citrate-coated AgNPs, hellebore-AgNPs system, and HC and HAgC phyto-nanocarrier(s) are shown in Figure 2 and Figure 3. The FT-IR absorption bands identified in the hellebore samples are presented in Table 3.
The FT-IR peak of the AgNPs coated with trisodium citrate (surfactant) (Figure 3A) presented vibrational bands characteristic of a surfactant at 3435 cm−1 (H-OH stretching vibration) and 2927 cm−1, as well as at 2852 cm−1 (CH- asymmetric and symmetric stretching vibrations), 1631 cm−1 (COO- stretching vibration) and 1387 cm−1 (C-H bending) [130].
The FT-IR peak of the chitosan (Figure 3B) showed characteristic vibrational bands at 3363 cm−1 (associated to N-H and O-H stretching); 2923 and 2877 cm−1 (attributed to C-H stretching); 1647 cm−1 (corresponding to the C=O stretching of amide); 1427 and 1375 cm−1 (associated to C-H bending); 1325 cm−1 (associated to a tertiary amide); 1265 cm−1 (corresponding to C-OH vibration); 890 cm−1 (attributed to C-H bending); and 663 cm−1 (associated to N-H vibration) [131,132,133,134].
Regarding the spectra of the HC phyto-nanocarrier (Figure 3B), all of the characteristic peaks of chitosan (2924, 2877, 1650, 1430, 1377, 1266 and 890 cm−1) and the plant phytoconstituents were observed. Nevertheless, the intensity of the band changes in the C-O and C-H regions (890 and 1380 cm−1) indicated the encapsulation and, probably, the hydrogen bond formation between the hellebore and chitosan [135].
The FT-IR spectra of the hellebore-AgNPs system (Figure 3A) included the characteristic peaks of hellebore at 3383 cm−1, attributed to the -OH group; at 2928 cm−1, assigned to C-H aromatic; at 1742, assigned to δ-lactone C=O and the lactone ring of bufadienolides; at 1709, assigned to carboxyl acid C=O stretching; at 1653, assigned to the β-unsaturated keto group and ecdysteroids; at 1642, corresponding to amino acids group N-H; at 1244 and 1017, attributed to the C-N of amine; at 881.34 and 815 cm−1, assigned to C-O and CH vibrations of aromatic rings, as well as AgNPs coated with surfactant (trisodium citrate) [130]. In addition, the absorption bands at 1635, 1388, 1115 and 676 cm−1 that appeared in the synthesized AgNPs solution (Figure 3B) shifted to higher wavenumbers (1639.7, 1428.3, 1151 and 683 cm−1), suggesting the binding of AgNPs to the O–H, C=O, N-H and C–O functional groups of the hellebore phytoconstituents (Figure 3 and Table 3) [130,136]. Moreover, several detectable changes occurred in the hellebore spectra, in particular, in the case of the hydroxyl vibrations (O-H stretching (the absorption bands at 1575 (N-H), 1415 and 1382 (O-H), and 1294, 1267 and 1081 (C-O)) which shifted to lower wavenumbers, indicating that this functional group is involved in the binding of AgNPs (Figure 3 and Table 3). Based on the analysis of the FTIR spectra of the HAgC phyto-nanocarrier (Figure 3B), it appeared to exhibit the characteristic vibrational bands of the chitosan sample and hellebore-AgNPs system with small shifts. A pronounced sharpening of the O–H and N-H stretching regions in the FT-IR supports the preparation of the HAgC phyto-nanocarrier. Additionally, these changes in the band intensity indicate that this functional group is involved in the binding of the hellebore-AgNPs system, probably via hydrogen bonding [135]. The FT-IR data for both of the phyto-nanocarriers’ development (HC and HAgC) is important for understanding their behavior and can aid in their further analysis.

2.4. X-ray Diffraction Spectroscopy

The XRD analysis provided valuable insight into the crystalline nature of the samples, the degree of order and the arrangement of the atoms in the crystal lattice of the initial and final products.
Figure 4 shows the XRD patterns of the chitosan, hellebore-AgNPs system, HC and HAgC phyto-nanocarriers. The chitosan exhibited peaks at 2θ: 11.55°, 16.37°, 18.33°, 22.4° and 34.4°, indicating the crystalline form II structure of chitosan [137,138,139].
The XRD pattern for the hellebore-AgNPs system presented a mix of crystalline and amorphous patterns with peaks at 2θ: 15.82° (weak), 22.23°, 27.87° and 38.15°.
The peaks at 15.82° (weak) and 22.23° (strong) are associated with hellebore constituents. In addition, the two distinct diffraction peaks of the 2θ: values of 27.87°and 38.15°, as well as 64.4° and 78.5°, indicate the crystalline structure of the AgNPs [39,139], while the specific XRD spectrum of the HC phyto-nanocarrier displays characteristic peaks of chitosan (at 2θ: 11.55°, 16.37°, 18.33°, 22.4° and 34.4°) and hellebore (at 2θ: 15.82° and 22.23°), which are considerably attenuated. Accompanying the XRD pattern for the HAgC phyto-nanocarrier, a few differences were found, despite the similarities to that of hellebore and chitosan. For example, the characteristic bands of the plant at 2θ = 15.8° and 22.3° were slightly weaker, and the AgNP peaks can easily be observed at 2θ = 38.2°, 44.1°, 64.4° and 78.5°.
The XRD results demonstrate that the new phyto-nanocarriers were successfully prepared through encapsulation into chitosan nanoparticles [140,141]. The mean crystallite size obtained via XRD analysis for CN (18.87 nm), HC (21.23 nm) and HAgC (23.66 nm) were confirmed again using SEM.

2.5. Scanning Electron Microscopy (SEM)

The surface morphology, shape and particle size of both phyto-nanocarriers (i.e., HC and HAgC), as well as their components (i.e., hellebore and chitosan) were studied with SEM (Figure 5).
The hellebore micrograph (Figure 5A) exhibits a porous, irregular structure. The morphology of the chitosan particles (Figure 5B) indicates the presence of clusters of rod-shaped crystals (~20 nm). The HC phyto-nanocarrier micrographs (Figure 5C,D) each show a spherical cluster of nanoparticles. A notable difference is apparent, particularly in terms of the particle size (~21.2 nm). Another visible aspect is the presence of numerous uniform spherical-shaped nanoparticles. The morphology of the hellebore-AgNPs system (Figure 5E–G) indicates the changes in the hellebore that appeared after the preparation of the hellebore-AgNPs system. A noticeable element is the presence of numerous uniform, nanosized silver particles (~28 nm) loaded in the porous structure, which has an irregular shape and is the same size as the hellebore particles. After encapsulation, SEM images of the HAgC phyto-nanocarrier (Figure 5I–J) show the appearance of clusters of nanospherical particles (~23.5 nm). Additionally, the SEM (Figure 5C,D,H–J) analysis indicated a modification of the shape, a decrease in the size of the particles and a tendency toward agglomeration for both of the encapsulated phyto-materials (hellebore and phyto-carrier system). These changes in the morphology suggest that under the employed experimental conditions, the preparation of the HC and HAgC phyto-nanocarriers was successfully achieved [142].
Accompanying the SEM spectra were EDX analyses on the elemental composition of the HC, hellebore-AgNPs system and HAgC samples investigated (Figure 6).
The EDX spectra of the HC phyto-nanocarrier indicated the presence of both hellebore and chitosan components (Figure 6A). Analogous work with the hellebore-AgNPs system indicated the introduction of AgNPs in the hellebore’s pores (Figure 6B). An elemental analysis of the HAgC phyto-nanocarrier was conducted using EDX spectroscopy (Figure 6C). It appears that there were differences compared to the EDX spectra before encapsulation, thus, confirming the HAgC phyto-nanocarrier’s preparation.

2.6. Dynamic Light Scattering

Dynamic light scattering (DLS) is a fast, simple and accurate method that is widely used to determine the size and distribution profile of nanoparticles suspensions.
In this study, the DLS method was employed to determine the average mean of each phyto-nanocarrier prepared and their raw components (Figure 7).
The results indicate a mean hydrodynamic diameter for the chitosan sample of 100.2 nm (Figure 7A). The DLS pattern the HC phyto-nanocarrier (Figure 7B) exhibited two peaks that can be attributed to chitosan and hellebore nanoparticles. Moreover, the HAgC phyto-nanocarrier (Figure 7C) results showed an additional peak that can be attributed to the AgNPs. Nonetheless, a notable difference appeared, specifically in the intensity peaks of the HC and HAgC phyto-nanocarriers, suggesting the formation of an agglomerate. Furthermore, the increase in the sample’s diameter size to between the values determined via the XRD and SEM analyses could be attributed to the manner in which the particles interacted with each other in the suspension (i.e., ultrapure water). In addition, various studies have reported that a shift toward larger average sizes is associated with encapsulation [131,143]. Another visible aspect was the distribution of particles in a narrow range, demonstrating the good stability of the newly prepared phyto-nanocarriers.

2.7. Thermal Behavior Study

The thermogravimetric analysis (TG)/differential thermogravimetric analysis (DTG) method was used to investigate the thermal properties of the chitosan matrix and HC and HAgC phyto-nanocarriers. The results are presented in Figure 8 and Figure 9.
According to Figure 8A and Figure 9, the chitosan followed three thermal decomposition steps in the temperature range of 25–400 °C. The first weight loss at a temperature up to 100 °C corresponds to the loss of adsorbed water. The second loss at a temperature range of 100–230 °C was due to the onset of chitosan decay (i.e., pyrolysis of the polysaccharide structure), which begins with a cleavage of glycosidic bonds. The final stage of weight loss occurred at temperatures above 230 °C and can be attributed to the advanced decomposition of polysaccharides with the formation of acetic and butyric acids and a series of lower fatty acids [144]. Chitosan’s degradation was not completed in the studied temperature range.
The results for the HC phyto-nanocarrier (HC) (Figure 8B and Figure 9) exhibited a minimum of weight loss at temperatures up to 100 °C, which was related to moisture evaporation. Then, the decomposition of the HC compound occurred in two stages of decomposition represented by separate thermoanalytical curves. The first occurred in the range of 170–200 °C, reaching a maximum of 178 °C, and showed a weight loss of 3.58%. The second weight loss (8.19%) step involved a complex process and took place in the temperature range between 200 and 278 °C. The higher weight loss (19.68%) in the temperature range of 170–400 °C can be attributed to chitosan’s degradation and, to a smaller extent, its deacetylation [135].
Regarding the HAgC nanocarrier, the TG/DTG curve (Figure 8C and Figure 9) followed the thermal trends of its components (i.e., chitosan and hellebore), with several exothermic decomposition steps in the temperature range of 70–380 °C.
The first weight loss, at temperatures up to 70 °C, corresponded to the evaporation of adsorbed water. The higher weight loss (10.76%) occurred in the temperature range of 73–126 °C, followed by a continuous loss of mass with stages that are problematic to separate. Nonetheless, four maxima on the DTG curve were present. This behavior could be attributed to supplementary groups and a higher hydroxyl group number from the encapsulated hellebore-AgNPs in the chitosan matrix [135,145]. The DSC analysis of the HAgC phyto-nanocarrier is presented in Figure 10.
The data reveal an endothermic process with a maximum of 119 °C and two exothermic processes at 328 °C and 380 °C. The thermic profile of the HAgC phyto-nanocarrier showed a sharp endothermic peak attributed to the various water-holding capacities and interaction strengths between the water and polymer. In addition, the exothermic peaks reflected the chitosan’s degradation (glycosidic ring dehydration, depolymerization and chemical decomposition of the monosaccharide units) [146,147,148,149]. The changes in the DTG thermograms observed for the HC and HAgC phyto-nanocarriers (Figure 9) (compared to the chitosan sample (shift of the endothermic effects) correspond to the interaction between chitosan and the hellebore and hellebore-AgNPs system, respectively. These data point to the (i) successful preparation of the HC and HAgC phyto-nanocarriers into chitosan nanoparticles and (ii) the overall improved thermal stability of these materials. These results corroborate those reported in the literature [146,149,150,151].

2.8. Screening of the Antioxidant Activity

The highly complex composition of bioactive compounds from a specific herb imposes the selection of several assays for the evaluation of the antioxidant capacity, since their separation and quantification are arduous because of their chemical diversity [37]. Therefore, methods for the direct estimation of the total antioxidant activity potential from plant matrices are preferred [37,38,152,153]. There are several chemical and biochemical antioxidant capacity assays. The most commonly used in vitro tests include the hydrogen atom transfer (HAT) and electron transfer (ET) methods. HAT methods include ORAC, TRAP and TOSCA, among others, while the primary ET methods include the Folin–Ciocalteu reagent assay, DPPH, TEAC, FRAP, and others [37,38,152,153]. The selection of the most appropriate method depends on a diverse set of factors, like simplicity, sensitivity, cost and reproducibility. To gain a comprehensive understanding of a plant’s antioxidant potential [37,38,152,153,154,155], and, therefore, for a more accurate assessment of the antioxidant potential of hellebore microencapsulated in chitosan and the new phyto-carrier system in vitro, noncompetitive methods—DPPH, o-phenanthroline (OPM), Folin–Ciocalteu and phosphomolybdate (total antioxidant capacity)—were chosen.

2.9. DPPH (1,1-Diphenyl-2-Picrylhydrazyl) Free Radical Scavenging Assay

The DPPH assay is one of the simplest, most accurate and cheapest methods based on a single electron transfer (ET) for evaluating the antioxidant capacity of complex plant matrices. It is based on the stable, free radical DPPH, which acts as a hydrogen donor [35,37,39,153,156]. Hence, the antioxidant activity of the new carrier(s) and its components were evaluated using this technique. The data obtained are presented in Table 4 and Figure 11A.
The IC50 values obtained indicate a significant increase in the antioxidant activity of the hellebore-AgNPs system correlated with the plant sample. Thus, the IC50 value was lower by over 60% than that of the hellebore samples, attributed to the synergistic action of the AgNPs and the phenolic compounds in the hellebore. Concomitantly, there is a slight decrease in the antioxidant activity of both prepared phyto-nanocarriers compared to the hellebore samples and the hellebore-AgNPs system. The increase in the IC50 values for the HC by 7% and HAgC by almost 8% compared to the nonencapsulated components might be due to the protective effect of the chitosan nanocapsules, which partially limits direct contact with the DPPH reagent, which is in good accordance with the results reported in the literature [143,157].
The chitosan sample presented relatively low scavenging activity against DPPH, and the IC50 value obtained was close to that reported in the specialized literature [158,159].

2.10. Total Polyphenolic Contents (TPCs) Assay—Folin–Ciocalteu

The Folin–Ciocalteu assay is a classic electron transfer (ET) technique used to establish the total phenol/polyphenol content from complex vegetable matrices or from food [39,153,160,161].
The total polyphenolic contents (TPCs) of the new phyto-nanocarrier(s) and its raw materials were determined, and the obtained results are presented in Table 4 and Figure 11B.
According to the data obtained, with the first phyto-nanocarrier, HC, more than 94% of the polyphenols in the hellebore sample reacted with the Folin–Ciocalteu reagent due to the density of the chitosan layer [151]. The antioxidant activity of the HAgC represented approximately 95% that of the hellebore-AgNPs system’s because of the synergistic action of the AgNPs, with a catalytic role played by the hellebore’s polyphenolic compounds. As reported in the literature [157], this slight decrease in the antioxidant activity detected for both phyto-nanocarriers indicates the good encapsulation efficiency of the bioactive compounds. The total phenolic content value determined for the chitosan sample is in good agreement with the literature [162].

2.11. Phosphomolybdate Assay (Total Antioxidant Capacity)

Phosphomolybdate (total antioxidant capacity) is a simple, sensitive and common method to determine the total antioxidant potential of plant extracts or other complex mixtures of biomolecules. The assay is based on the molybdenum (VI) to molybdenum (V) reduction through an ET or HAT mechanism [163].
The phosphomolybdate assay (total antioxidant capacity) was used to determine the total antioxidant potential of the prepared phyto-carrier system and hellebore encapsulated in chitosan compared to those of the hellebore, hellebore-AgNPs, chitosan and ascorbic acid. The obtained experimental results are displayed in Table 4 and Figure 11C.
Table 4 indicates that both phyto-nanocarriers displayed a slight decrease in antioxidant activity (less than 2–5%) than the nonencapsulated components suggesting that, under the employed experimental conditions, the chitosan matrix protected the core bioactive compound and also acted as an antioxidant activity booster agent [154]. The superior antioxidant activity of the hellebore-AgNPs complex compared to the plant sample can be attributed to the synergistic and complementary action of the secondary metabolites from the hellebore and AgNPs [164,165].

2.12. Iron(III)-Phenanthroline Antioxidant Assay

The o-phenanthroline method (OPM) is a simple, fast, sensitive and accurate method for hydroxyl radical scavenging activity evaluation in complex samples. The assay is based on the inhibition of the formation of Fe (II)-1,10-phenanthroline using hydroxyl radical as a quantitative measure of the scavenging activity [163,166,167,168,169,170].
The iron(III)-phenanthroline antioxidant assay was used to measure the antioxidant activity of the new phyto-nanocarriers and their components. The results are presented in Table 4 and Figure 11D.
The hellebore-AgNPs system exhibited a higher antioxidant activity than the hellebore sample (more than 70%) due to the complementary action of the bioactive compounds and the antioxidant mechanism of the AgNPs (quenched free radicals by donating or accepting electrons) in the presence of the hellebore’s various phytoconstituents. Additionally, studies have reported that AgNPs’ structural features (i.e., shape and size) affect antioxidant activity [130,171]. In the case of phyto-nanocarriers, there was a decrease in the antioxidant activity (about 20% for HC and almost 14% for HAgC), which indicates that the encapsulation was successful, and the chitosan matrix prevented the complete release of all antioxidant agents from the core phyto-materials [157].
Altogether, the evaluation of both phyto-nanocarriers’ antioxidant potential through the four assays shows a slight decrease compared to the nonencapsulated components. However, the antioxidant activity of the two nanocarriers is still very high, considering that their antioxidant performance depends on the physicochemical characteristics of chitosan and the experimental conditions (reagent ratio, pH, temperature, solvent polarity, etc.) [172]. The collective results indicate the success of the encapsulation and, thus, an enhancement of the in vivo biocompatibility and stability, as well as prolonged controlled-release [143,157,173].

2.13. Encapsulation Efficiency, Loading Capacity and Encapsulation Yield (EY%)

The encapsulation efficiency, loading capacity and encapsulation yield are the determining parameters for establishing the quality of nanocapsules and their application potential and economic feasibility [145,174].
The data obtained are shown in the Table 5.
According to Table 5, the encapsulation yield decreased in the following order: HAgC > HC > CN.
These results indicate that the hellebore and hellebore-AgNPs were efficiently entrapped in the chitosan nanoparticles, which can be attributed to electrostatic interactions between the positively charged (amino) groups in the chitosan nanoparticles and the negatively charged groups of the different phytoconstituents from the hellebore sample [139,175,176].
Moreover, the higher values for the encapsulation parameters in the case of the HAgC compared to the HC indicate the occurrence of various interactions (electrostatic, hydrogen bonds, Van der Waals) between the amino groups of the chitosan matrix and the AgNPs [177].

3. Materials and Methods

All reagents used in this study were of analytical grade, purchased from commercial sources and used without further purification. Ethanol, methanol, chloroform, dichloromethane, FeCl3·6H2O, FeSO4·7H2O and anhydrous Na2CO3 were from Sigma Aldrich (München, Germany). 1,10-Phenantroline; Folin–Ciocalteu phenol reagent (2 N); 2,2-diphenyl-1-picrylhydrazyl reagent; gallic acid; ascorbic acid; AgNO3; sodium citrate; sodium carbonate; potassium persulfate; sodium phosphate; ammonium molybdate; and potassium chloride of 99% were also purchased from Sigma-Aldrich (München, Germany). Propyl gallate (purum) was purchased from Fluka (Buchs, Switzerland). Chitosan (MW 100,000–300,000) was purchased from Acros Organics (Verona, Italy). Ultrapure water (DDW) was used throughout the experiments.
The Helleborus purpurascens Waldst. & Kit. samples (whole plant: stem (28 cm in height), leaves, roots and rhizome (8 cm)) were harvested in March 2023 from an area in Prahova County, Romania (geographic coordinates: 45°20’46.4689750856423” N, 25°33’22.9191970825194” E), and taxonomically authenticated at the University of Medicine and Pharmacy Craiova, Romania.

3.1. Phyto-Carrier-System Components’ Preparation

The AgNPs’ synthesis was carried out according to a procedure described in our earlier publication [130].

Plant Samples’ Preparation for Chemical Screening

The plant samples were chopped and frozen in liquid nitrogen (−185 °C). Subsequently, they were milled using a planetary mill (Fritsch Pulverisette mill, Idar-Oberstein, Germany), sieved to obtain particle diameters ranging from 0.30 to 0.35 mm and then stored at −30 °C to avoid the phytoconstituents’ enzymatic degradation. The ultrasonic-assisted extraction’s experimental procedure consisted of the following: round-bottom flasks (50 mL) containing 1.8 g dried plant sample and 18 mL of solvent (methanol/chloroform = 1:1) were kept at 40 °C and 50 Hz for 35 min (Elmasonic, Singen, Germany). The solvent was evaporated at 40 °C using a rotary evaporator (Rotavapor; BÜCHI, Switzerland), followed by the dissolution of the residue obtained with MeOH (10 mL), filtration through a 0.2 µm syringe filter and storage at −30 °C until further analysis. All experiments were prepared in triplicate.

3.2. GC-MS Analysis

Gas chromatography was carried out on a GCMS-QP2020NX Shimadzu apparatus with a ZB-5MS capillary column (30 m × 0.25 mm id × 0.25 µm) (Agilent Technologies, Santa Clara, CA, USA), helium, and flow of 1 mL/min.

GC–MS Separation Conditions

The oven temperature was increased from 50 °C (held for 2 min) to 300 °C at a rate of 6 °C/min (held for 5 min). The temperature of the injector was 280 °C, and the temperature at the interface was 220 °C. The mass of the compounds was registered at a 70 eV ionization energy. The mass spectrometer was source-heated at 230 °C, and the MS Quad was heated at 150 °C. Compounds were identified based on their mass spectra, which were compared to the NIST 3.0 database mass spectra library database (USA National Institute of Science and Technology software) (NIST, Gaithersburg, MD, USA). Retention indices (RIs) were also calculated for each compound based on a C8–C20 alkane standard mixture calibration curve, which was compared to Adams indices in the literature [178].
A percentage area of each compound was calculated by summing all of the experimental integrated peak areas (Table 1).

3.3. Mass Spectrometry

The MS experiments were performed using an EIS-QTOF-MS (Bruker Daltonics, Bremen, Germany). The mass spectra were acquired in the positive ion mode in a mass range of 40–3000 m/z. The scan speed was 2.0 scans/s, the collision energy was 10–85 eV, and the temperature of the source block was 82 °C. The standard library NIST/NBS-3 (National Institute of Standards and Technology/National Bureau of Standards) (NIST, Gaithersburg, MD, USA) was used for their assignment to specific bioactive compounds. The results are presented in Table 2.

3.3.1. Phyto-Carrier System Preparation (Hellebore-AgNPs System)

To prepare the hellebore-AgNPs system, the dried plant sample and AgNPs solution were mixed at a 1:2.5 mass ratio. Then, they were shaken (500 rpm) for 24 h at room temperature (22 °C). Subsequently, the obtained solution was filtered (0.45 μm) and dried in an oven at 30 °C for 4 h. Each experiment was repeated three times.

3.3.2. Chitosan Phyto-Nanocarrier Preparation (Encapsulation Procedure)

The chitosan solution was prepared by dissolving 1.4 g of chitosan into 100 mL acetic acid solution (5% (v/v)). The resulting mixture was incubated at 80 °C for 48 h, under continuous stirring.
Chitosan nanoparticles (CN) (used to evaluated the encapsulation process)
The chitosan solution, cooled to room temperature, was added to 25 mL (2 M) NaOH solution dropwise with a Pasteur pipette, filtered and washed with deionized water.
To prepare the first nanocarrier (HC), a 0.25 g dried hellebore sample was added to 18 mL chitosan solution under continuous stirring at room temperature (22 °C). The mixture was added under stirring to 25 mL (2 M) NaOH solution dropwise with a Pasteur pipette, filtered and washed with deionized water.
The hellebore-AgNPs nanocarrier (HAgC) was prepared from 0.25 g hellebore-AgNPs and 18 mL chitosan solution, following an identical experimental procedure. The experiment was prepared in triplicate.

3.4. Characterization of Nanocarriers

3.4.1. Fourier Transform Infrared (FT-IR) Spectroscopy

The data collection was performed after 20 recordings at a resolution of 4 cm−1 in the range of 4000–400 cm−1 on a Shimadzu AIM-9000 with ATR devices.

3.4.2. XRD Spectroscopy

The X-ray powder diffraction (XRD) was performed using a Bruker AXS D8-Advance X-ray diffractometer (Bruker AXS GmbH, Karlsruhe, Germany) equipped with a rotating sample stage, Anton Paar TTK low-temperature cell (−180 °C ÷ 450 °C), high-vacuum, inert atmosphere, relative humidity control, and an Anton Paar TTK high-temperature cell (up to 1600 °C). The XRD patterns were compared with those from the ICDD Powder Diffraction Database (ICDD file 04-015-9120). The average crystallite size and the phase content were calculated using the whole-pattern profile-fitting method (WPPF).

3.4.3. Scanning Electron Microscopy (SEM)

The SEM micrographs were obtained with an SEM-EDS system (JEOL JSM-IT200 Field Emission, Nieuw-Vennep, The Netherlands) equipped with a high-resolution electron gun and energy-dispersive X-ray spectrometer (EDS) with a MnK resolution of 133 eV.

3.4.4. Dynamic Light Scattering (DLS) Particle Size Distribution Analysis

The DLS analysis was carried on a Microtrac/Nanotrac 252 (Montgomeryville, PA, USA), at room temperature (22 °C) and a scattering angle of 172°. Each experiment was repeated three times.

3.4.5. Thermal Analysis

The thermal stability of the samples was investigated using a Thermal Analyzer produced by METTLER TOLEDO, model TGA/DSC3+ STARe System. The analyses were performed between 25 and 400 °C in a dynamic air atmosphere (20 mL/min, synthetic air) with a 10 °/min heating rate. Small amounts of every sample were placed in 40 μL aluminum melting crucibles. The DSC analysis was performed using a DSC 3+ Mettler Toledo in an air atmosphere (50 mL/min) in same range (25–400 °C).

3.4.6. Antioxidant Activity

The antioxidant activity of the new nanocarriers (HC and HagC) and their components were evaluated using four different assays: 2,2-diphenyl-1-picrylhydrazyl (DPPH) radical scavenging assay; Folin–Ciocalteu assay; phosphomolybdate assay (total antioxidant capacity); and iron(III)-phenanthroline antioxidant assay. The experiments were carried out in 100 mL conical flasks containing 0.4 (±0.001) g dried sample to which 10 mL ethanol (70%) was added. The flasks were stirred (200 rpm), subjected to an ultrasonic extraction (45 min) and centrifuged (3500× g rpm, 10 min). Subsequently, the supernatant was collected and kept at −18 °C until further use in the selected antioxidant assays for this study.

3.5. Total Polyphenol Content Using the Folin–Ciocalteu Method

The total phenolic content in the new nanocarrier(s) and its components were determined spectrophotometrically according to the Folin–Ciocalteu procedure adapted from the literature [179]. A volume of 1.5 mL Folin–Ciocalteu reagent (2 M) and 0.5 mL Na2CO3 (2%) were added, dropwise with a Pasteur pipette, under stirring to 0.5 mL of each alcohol sample. The resulting solutions were kept at room temperature in a dark place for 60 min. All solutions were prepared in triplicate. The absorbance was measured at 725 nm using a UV-Vis spectrometer (DLAB SP-UV1000, Penjuru, Singapore). The phenol content was expressed in gallic acid equivalents (mg GAE/g sample) using a propyl gallate standard calibration curve between 9.375 µg/mL and 300 µg/mL in ethanol. The sample extract concentrations were calculated based on the linear equation obtained from the standard curve (y = 0.0015x + 0.2145 and the correlation coefficient (R2 = 0.9904).

3.6. DPPH Radical Scavenging Assay

The antioxidant activities of the new nanocarrier(s) and its components were comparatively evaluated using a DPPH (2,2-diphenyl-1-picrylhydrazyl) free radical elimination assay. A stock solution of 0.25 mM DPPH in ethanol was prepared. The sample extracts had dilutions between 20 µg/mL and 0.625 mg/mL. The ratio (v/v) of the DPPH to the samples was 3:1. The employed experimental procedure was the following: a volume of 1.5 mL 2,2-diphenyl-1-picrylhydrazyl reagent (2 M) was added, dropwise with a Pasteur pipette, under stirring to 0.5 mL of each alcohol sample. The resulting solutions were kept at room temperature (22 °C) for 30 min. The absorbance was measured at 515 nm using a UV-Vis spectrometer (DLAB SP-UV1000). All solutions were prepared in triplicate in ethanol, and 1:1 DPPH 0.1 mg/mL and ethanol was used as a control sample. The obtained results were used to calculate the average and the inhibition percentage (Inh%) (Equation (1)) [180].
Inh% = (A0 − Al)/A0 × 10
where:
  • A0 = vehicle control absorbance;
  • Al = sample absorbance.
Further, the IC50 value was obtained from the inhibition percentage using the equation from the calibration curve generated for each sample and standard. The results are presented as the Inh% versus the concentration (mg/mL) [181].

3.7. Phosphomolybdate Assay (Total Antioxidant Capacity)

The total antioxidant capacity assay of the new nanocarrier(s) and its components was carried out with the phosphomolybdenum procedure using ascorbic acid as the standard [182]. A volume of 6 mL reagent solution (0.6 M sulfuric acid, 28 mM sodium phosphate and 4 mM ammonium molybdate) and 0.6 mL of each sample were placed into a water bath at 90 °C for 120 min. Next, the mixed solutions were cooled at room temperature (22 °C). The absorbance was measured at 765 nm using a UV-Vis Perkin-Elmer Lambda 35 (Perkin Elmer, Waltham, MA, USA).
A blank solution was used in which 6 mL of reagent was added to 0.6 mL of methanol, and then the mixture was incubated under the same experimental conditions: 90 °C for 120 min and then cooled at room temperature (22 °C). The total antioxidant capacity was determined according to the following equation (Equation (2)):
Total antioxidant capacity (%) = (Abs. of control − Abs. of sample)/(Abs. of control) × 100
The results are presented as μg/mL of ascorbic acid equivalents (AAE).

3.8. Iron(III)-Phenanthroline Antioxidant Assay (OPM)

The ferric ion reduction capacity of the new nanocarrier(s) and its components were comparatively evaluated according to the following procedure: in 25 mL mL Erlenmeyer flasks containing 1 mL 1,10-phenanthroline (0.1%), 1 mL FeCl3•6H2O (0.1%) and 7 mL distilled water, 1 mL of each sample was added. The resulting solution was stirred at room temperature (22 °C) for 25 min. The absorbance was measured at 510 nm using a UV-Vis spectrometer (DLAB SP-UV1000, DLAB Scientific Inc., Reverside, CA, USA). The ferric ion reduction capacity of the extracts is expressed in mmol Fe/g dry product using a FeSO4•7H2O calibration curve between 0 and 500 µmol/L in distilled water [183].
The results are reported according to the concentrations on the standard curve (y = 0.0015x + 0.1795, (correlation coefficient) R2 = 0.9727.

3.9. Encapsulation Yield, Encapsulation Efficiency and Loading Capacity

The encapsulation yield (EY%) was determined as the ratio of the weight of the material nanoencapsulated into chitosan and the raw materials’ weight, according to the following equation (Equation (3)) [174]:
E Y   % = w e i g h t   o f   n a n o c a r r i e r   ( g ) w e i g h   r a w   m a t e r i a l s   ( g )   ×   100

3.10. Encapsulation Efficiency (EE%)

The encapsulation efficiency (EE%) was calculated as the percent of the total amount (g) of the hellebore and hellebore-AgNPs system based on the total amount (g) used as raw materials in the encapsulation process (Equation (4)) [184,185].
E E   % = a m o u n t   o f   c o m p o u n d   e n c a p s u l a t e d   ( g ) a m o u n t   c o m p o u n d   u s e d   a s   r a w   m a t e r i a l s   ( g ) × 100  
The compound contents and encapsulation efficiency (EE%) of the chitosan nanoparticles were determined using a UV-Vis Perkin-Elmer Lambda 35 (Perkin Elmer, Waltham, MA, USA). All absorbance measurements were conducted with a 10 mm UV/Vis spectroscopy cell at room temperature, using a solvent (ethanol/chloroform = 1:1) as a blank. The employed experimental procedure was the following: The samples (15 mg) were subject to sonication extraction (frequency of 50 kHz) in 20 mL of solvent (HCl/ethanol/chloroform = 3:2:2) for 40 min at room temperature (22 °C) and then centrifuged. Subsequently, the supernatant’s concentration was determined using atomic UV-Vis spectroscopy. Each experiment was repeated three times.

3.11. Preparation of the Curves of the Concentrations of the Loaded Materials (Hellebore and Hellebore-AgNPs System)

The ultrasonic-assisted extraction experiments were carried out with five different amounts of each sample (0.01–0.20 mg) using a constant volume of solvent (ethanol/chloroform = 1:1) for 40 min at 22 °C, followed by centrifugation. Subsequently, the concentrations of the loaded materials in the filtrate were determined with UV-Vis spectroscopy. Calibration curves were plotted for hellebore and the hellebore-AgNPs system. The correlation coefficients, which were R2 = 0.9896 (hellebore) and R2 = 0.9932 (hellebore-AgNPs system), demonstrate the good linear relationship of the data. Each experiment was repeated three times. The amount of each compound encapsulated in the chitosan nanoparticles was obtained from the standard curve of the concentration versus its absorbency.
The encapsulated content (EC%) was calculated according to the following equation (Equation (5)) [184,185]:
E C   % = c o n c e n t r a t i o n   c o m p o u n d   e n c a p s u l a t e d   x   v o l u m e   s o l u t i o n   a m o u n t   c o m p o u n d   u s e d   a s   r a w   m a t e r i a l s   g × 100  
The concentration of each compound was determined using UV-Vis spectroscopy according to the above experimental procedure.

3.12. Statistical Analysis

Each experimental set was performed in triplicate, using one-way analysis of variance (ANOVA) without replication with Scheffe’s post hoc test comparison; p < 0.05 was taken as statistically significant.

4. Conclusions

The undertaken study describes the low-molecular-mass metabolite profiling of Helleborus purpurascens growing wild in Romania. The biological activities of each category of phytoconstituents were discussed. Furthermore, a new phyto-carrier system (hellebore-AgNPs system) with unique physical–chemical characteristics and high antioxidant activity was prepared. The encapsulation potential of the plant samples (i.e., hellebore) and new phyto-carrier system (hellebore-AgNPs system) in chitosan nanoparticles (CNs) was evaluated. The development of two new phyto-nanocarriers, HC and HAgC, was confirmed through FTIR, EDX, XRD, DLS and SEM studies. The thermal stability was investigated. Several different antioxidant assays (DPPH, Folin–Ciocalteu, phosphomolybdate (total antioxidant capacity) and iron(III)-phenanthroline antioxidant) were carried out to ensure that HC and HAgC exhibited potent antioxidant activity. Collectively, this study shows the potential advantages of these phyto-nanocarriers encapsulated as selective therapeutic delivery systems. Nonetheless, further studies are necessary to investigate the in vitro release, biological properties and bioavailability of these new phyto-nanocarriers.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/plants12193479/s1. Figure 1S: TIC chromatogram of hellebore extract. Figure 2S: The mass spectrum of Helleborus purpurascens sample.

Author Contributions

Conception and study design, A.-E.S.; methodology, A.-E.S.; data acquisition, G.V., T.V., M.V.C., D.D.H., C.A.S., O.-F.G., I.G. and V.N.; analysis and data interpretation, G.V., T.V., M.V.C. and D.D.H.; writing—original draft preparation, A.-E.S.; writing—review and editing, A.-E.S.; investigation, G.V., T.V., I.G., M.V.C., V.N. and D.D.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant of the Ministry of Research, Innovation and Digitization CNCS–UEFISCDI, project number PN-III-P4-PCE-2021-1081 within PNCDI III (Contract no. 75/2022).

Data Availability Statement

All data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Balázs, V.L.; Filep, R.; Ambrus, T.; Kocsis, M.; Farkas, Á.; Stranczinger, S.; Papp, N. Ethnobotanical, historical and histological evaluation of Helleborus L. genetic resources used in veterinary and human ethnomedicine. Genet. Resour. Crop. Evol. 2020, 67, 781–797. [Google Scholar] [CrossRef]
  2. Maior, M.C.; Dobrotă, C. Natural compounds with important medical potential found in Helleborus sp. Cent. Eur. J. Biol. 2013, 8, 272–285. [Google Scholar]
  3. Borozan, A.B.; Popescu, S.; Moldovan, C.; Sărac, I.; Bordean, D. Helleborus—Phytochemistry and antimicrobial properties. A review. J. Hortic. For. Biotechnol. 2022, 26, 69–75. [Google Scholar]
  4. Segneanu, A.-E.; Damian, D.; Hulka, I.; Grozescu, I.; Salifoglou, A. A simple and rapid method for calixarene-based selective extraction of bioactive molecules from natural products. Amino Acids 2016, 48, 849–858. [Google Scholar] [CrossRef]
  5. Stroescu, V. Pharmacological Basis of Medical Practice; Medical Printing House: Bucharest, Romania, 1998; pp. 236–241. [Google Scholar]
  6. Neacsu, C.; Ciobanu, C.; Barbu, I.; Toader, O.; Szegli, G.; Kerek, F.; Babes, A. Substance MCS-18 isolated from Helleborus purpurascens is a potent antagonist of the capsaicin receptor, TRPV1, in rat cultured sensory neurons. Physiol. Res. 2010, 59, 289–298. [Google Scholar] [CrossRef] [PubMed]
  7. Kerek, F. Boicil, a new and very efficient antialgic spasmolytic and blood vessel regulating drug obtained from the plant Helleborus. Int. Conf. Chem. Biotechnol. Biol. Nat. Prod. 1981, 2, 22–37. [Google Scholar]
  8. Paun, G.; Neagu, E.; Radu, G.L.; Rotinberg, P.; Mihai, C. Evaluation of cytostatic potential of Helleborus purpurascens extracts concentrated by membrane techniques. Planta Med. 2011, 77, PF78. [Google Scholar] [CrossRef]
  9. Grigore, A.; Bubueanu, C.; Pirvu, L.; Neagu, G.; Bejanaru, I.; Vulturescu, V.; Panteli, M.; Rasit, I. Immunomodulatory effect of Helleborus purpurascens Waldst & Kit. Plants 2021, 10, 1990. [Google Scholar]
  10. Pilut, C.N.; Manea, A.; Macasoi, I.; Dobrescu, A.; Georgescu, D.; Buzatu, R.; Faur, A.; Dinu, S.; Chioran, D.; Pinzaru, I.; et al. Comparative evaluation of the potential antitumor of Helleborus purpurascens in skin and breast cancer. Plants 2022, 1, 194. [Google Scholar] [CrossRef]
  11. Segneanu, A.E.; Grozescu, I.; Cziple, F.; Berki, D.; Damian, D.; Niculite, C.M.; Florea, A.; Leabu, M. Helleborus purpurascens—Amino acid and peptide analysis linked to the chemical and antiproliferative properties of the extracted compounds. Molecules 2015, 20, 22170–22187. [Google Scholar] [CrossRef]
  12. Xie, P.; Wang, Y.; Wei, D.; Zhang, L.; Zhang, B.; Xiao, H.; Song, H.; Mao, X. Nanoparticle-based drug delivery systems with platinum drugs for overcoming cancer drug resistance. J. Mater. Chem. B 2021, 9, 5173–5194. [Google Scholar] [CrossRef] [PubMed]
  13. Available online: https://www.who.int/initiatives/who-global-centre-for-traditional-medicine (accessed on 18 June 2023).
  14. Atanasov, A.G.; Zotchev, S.B.; Dirsch, V.M.; Supuran, C.T. Natural products in drug discovery: Advances and opportunities. Nat. Rev. Drug Discov. 2021, 20, 200–216. [Google Scholar] [PubMed]
  15. Singh, I.P.; Ahmad, F.; Chatterjee, D.; Bajpai, R.; Sengar, N. Natural products: Drug discovery and development. In Drug Discovery and Development from Targets and Molecules to Medicines; Poduri, R., Ed.; Springer: Singapore, 2021; ISBN 978-981-15-5533-6. [Google Scholar]
  16. Nath, R.; Roy, R.; Barai, G.; Bairagi, S.; Manna, S.; Chakraborty, R. Modern developments of nano based grug delivery system by combined with phytochemicals-presenting new aspects. Int. J. Sci. Res. Sci. Technol. 2021, 8, 107–129. [Google Scholar]
  17. Patra, J.K.; Das, G.; Fraceto, L.F.; Campos, E.V.R.; del Pilar Rodriguez-Torres, M.; Acosta-Torres, L.S.; Diaz-Torres, L.A.; Grillo, R.; Swamy, M.K.; Sharma, S.; et al. Nano based drug delivery systems: Recent developments and future prospects. J. Nanobiotechnol. 2018, 16, 71. [Google Scholar]
  18. Rahman, H.S.; Othman, H.H.; Hammadi, N.I.; Yeap, S.K.; Amin, K.M.; Abdul Samad, N.; Alitheen, N.B. Novel drug delivery systems for loading of natural plant extracts and their biomedical applications. Int. J. Nanomed. 2020, 15, 2439–2483. [Google Scholar]
  19. Cele, T. Preparation of nanoparticles. In Engineered Nanomaterials-Health and Safety; IntechOpen: London, UK, 2020. [Google Scholar]
  20. Baig, N.; Kammakakam, I.; Falath, W. Nanomaterials: A review of synthesis methods, properties, recent progress, and challenges. Mater. Adv. 2021, 2, 1821–1871. [Google Scholar]
  21. Rane, A.V.; Kanny, K.; Abitha, V.K.; Thomas, S. Chapter 5—Methods for Synthesis of Nanoparticles and Fabrication of Nanocomposites. In Synthesis of Inorganic Nanomaterials; Woodhead Publishing: Cambridge, UK, 2018; pp. 121–139. [Google Scholar] [CrossRef]
  22. Chopra, H.; Bibi, S.; Singh, I.; Hasan, M.M.; Khan, M.S.; Yousafi, Q.; Baig, A.A.; Rahman, M.M.; Islam, F.; Emran, T.B.; et al. Green metallic nanoparticles: Biosynthesis to applications. Front. Bioeng. Biotechnol. 2022, 10, 874742. [Google Scholar]
  23. Hussein, H.A.; Abdullah, M.A. Novel drug delivery systems based on silver nanoparticles, hyaluronic acid, lipid nanoparticles and liposomes for cancer treatment. Appl. Nanosci. 2022, 12, 3071–3096. [Google Scholar] [CrossRef]
  24. Gomes, H.I.O.; Martins, C.S.M.; Prior, J.A.V. Silver nanoparticles as carriers of anticancer drugs for efficient target treatment of cancer cells. Nanomaterials 2021, 11, 964. [Google Scholar] [CrossRef]
  25. Galatage, S.T.; Hebalkar, S.A.; Dhobale, S.V.; Mali, O.R.; Kumbhar, P.S.; Nikade, S.V.; Killedar, S.G. Silver nanoparticles: Properties, synthesis, characterization, applications and future trends. In Silver Micro-Nanoparticles—Properties, Synthesis, Characterization, and Applications; Kumar, S., Kumar, P., Pathak, C.S., Eds.; IntechOpen: London, UK, 2021; ISBN 978-1-83968-660-3. [Google Scholar]
  26. Riva, R.; Ragelle, H.; des Rieux, A.; Duhem, N.; Jérôme, C.; Préat, V. Chitosan and chitosan derivatives in drug delivery and tissue engineering. Chitosan Biomater. 2011, II, 19–44. [Google Scholar]
  27. Domingues, J.M.; Miranda, C.S.; Silva, A.F.G.; Homem, N.C.; Amorim, M.T.P.; Felgueiras, H.P. Bioactivity of chitosan-based particles loaded with plant-derived extracts for biomedical applications: Emphasis on antimicrobial fiber- based systems. Mar. Drugs 2021, 19, 359. [Google Scholar]
  28. Ibrahim, H.M.; El-Zairy, E.M.R. Chitosan as a biomaterial—Structure, Properties and Electrospun Nanofibers. Concepts Compd. Altern. Antibact. 2015, 1, 81–101. [Google Scholar]
  29. Fuchs-Tarlovsky, V. Role of antioxidants in cancer therapy. Nutrition 2013, 29, 15–21. [Google Scholar] [CrossRef] [PubMed]
  30. Sardas, S. The Role of Antioxidants in Cancer Prevention and Treatment. Indoor Built Environ. 2003, 12, 401–404. [Google Scholar] [CrossRef]
  31. Luo, M.; Zhou, L.; Huang, Z.; Li, B.; Nice, E.C.; Xu, J.; Huang, C. Antioxidant Therapy in Cancer: Rationale and Progress. Antioxidants 2022, 11, 1128. [Google Scholar] [CrossRef]
  32. Akanji, M.A.; Fatinukun, H.D.; Rotimi, E.D.; Afolabi, B.L.; Adeyemi, O.S. The two sides of dietary antioxidants in cancer therapy. In Antioxidants-Benefits, Sources, Mechanisms of Action; Waisundara, V., Ed.; IntechOpen: London, UK, 2021; ISBN 978-1-83968-865-2. [Google Scholar]
  33. Yuan, H.; Ma, Q.; Ye, L.; Piao, G. The traditional medicine and modern medicine from natural products. Molecules 2016, 21, 559. [Google Scholar] [CrossRef]
  34. Yarnell, E. Synergy in herbal medicines: Part 1. J. Restor. Med. 2015, 4, 60–73. [Google Scholar] [CrossRef]
  35. Flieger, J.; Flieger, W.; Baj, J.; Maciejewski, R. Antioxidants: Classification, natural sources, activity/capacity measurements, and usefulness for the synthesis of nanoparticles. Materials 2021, 14, 4135. [Google Scholar] [CrossRef]
  36. Amorati, R.; Valgimigli, L. Methods to measure the antioxidant activity of phytochemicals and plant extracts. J. Agric. Food Chem. 2018, 66, 3324–3329. [Google Scholar] [CrossRef]
  37. Dinkova-Kostova, A.T.; Talalay, P. Direct and indirect antioxidant properties of inducers of cytoprotective proteins. Mol. Nutr. Food Res. 2008, 52, S128–S138. [Google Scholar] [CrossRef]
  38. Losada-Barreiro, S.; Sezgin-Bayindir, Z.; Paiva-Martins, F.; Bravo-Díaz, C. Biochemistry of antioxidants: Mechanisms and pharmaceutical applications. Biomedicines 2022, 10, 3051. [Google Scholar] [CrossRef] [PubMed]
  39. Shahidi, F.; Zhong, Y. Measurement of antioxidant activity. J. Funct. Foods 2015, 18, 757–781. [Google Scholar] [CrossRef]
  40. Wawrosch, C.; Zotchev, S.B. Production of bioactive plant secondary metabolites through in vitro technologies—Status and outlook. Appl. Microbiol. Biotechnol. 2021, 105, 6649–6668. [Google Scholar] [CrossRef]
  41. Ashraf, M.A.; Iqbal, M.; Rasheed, R.; Hussain, I.; Riaz, M.; Arif, M.S. Environmental stress and secondary metabolites in plants. In Plant Metabolites and Regulation under Environmental Stress; Academic Press: Cambridge, MA, USA, 2018; pp. 153–167. [Google Scholar]
  42. Segneanu, A.E.; Grozescu, I.; Sfirloaga, P. The influence of extraction process parameters of some biomaterials precursors from Helianthus annuus. Dig. J. Nanomater. Biostruct. 2013, 8, 1423–1433. [Google Scholar]
  43. Isah, T. Stress and defense responses in plant secondary metabolites production. Biol. Res. 2019, 52, 39. [Google Scholar] [CrossRef] [PubMed]
  44. Yang, L.; Wen, K.-S.; Ruan, X.; Zhao, Y.-X.; Wei, F.; Wang, Q. Response of plant secondary metabolites to environmental factors. Molecules 2018, 23, 762. [Google Scholar] [CrossRef]
  45. Ahmad, R.; Srivastava, S.; Ghosh, S.; Khare, S.K. Phytochemical delivery through nanocarriers: A review. Colloids Surf. B Biointerfaces 2021, 197, 111389. [Google Scholar] [CrossRef]
  46. Hill, C.B.; Roessner, U. Metabolic profiling of plants by GC–MS. In The Handbook of Plant Metabolomics, 1st ed.; Weckwerth, W., Kahl, G., Eds.; Wiley-VCH Verlag GmbH: Weinheim, Germany, 2013. [Google Scholar]
  47. Rossellia, S.; Maggio, A.; Formisano, C.; Napolitano, F.; Senatore, F.; Spadaro, V.; Bruno, M. Chemical composition and antibacterial activity of extracts of Helleborus bocconei Ten. subsp. intermedius. Nat. Prod. Commun. 2007, 2, 675–679. [Google Scholar] [CrossRef]
  48. Kusmenoglu, S.; Bingol, F.; Sener, B. Gas-chromatographic determination of the seed oil of Helleborus orientalis Lam. J. Fac. Pharm. Gazi 1995, 12, 23–27. [Google Scholar]
  49. Aziz, M.; Ahmad, S.; Khurshid, U.; Pervaiz, I.; Lodhi, A.H.; Jan, N.; Khurshid, S.; Arshad, M.A.; Ibrahim, M.M.; Mersal, G.A.M.; et al. Comprehensive biological potential, phytochemical profiling using GC-MS and LC-ESI-MS, and in-silico assessment of Strobilanthes glutinosus Nees: An important medicinal plant. Molecules 2022, 27, 6885. [Google Scholar] [CrossRef]
  50. Wu, L.; Zhao, J.; Wu, L.; Zhang, Y.; Li, J. Simultaneous determination of squalene, tocopherols and phytosterols in edible vegetable oil by SPE combined with saponification and GC-MS. LWT 2022, 169, 114026. [Google Scholar] [CrossRef]
  51. Hao, D.C. Mining Chemodiversity from biodiversity: Pharmacophylogeny of Ranunculaceae medicinal plants. In Ranunculales Medicinal Plants; Academic Press: Cambridge, MA, USA, 2019; pp. 35–71. [Google Scholar] [CrossRef]
  52. Vitalini, S.; Braca, A.; Fico, G. Study on secondary metabolite content of Helleborus niger L. leaves. Fitoterapia 2011, 82, 152–154. [Google Scholar] [CrossRef] [PubMed]
  53. Duckstein, S.M.; Stintzing, F.C. Comprehensive study of the phenolics and saponins from Helleborus niger L. leaves and stems by liquid chromatography/tandem mass spectrometry. Chem. Biodivers. 2014, 11, 276–298. [Google Scholar] [CrossRef] [PubMed]
  54. Deore, S.; Khadabadi, S. Isolation and characterization of phytoconstituents from Chlorophytum Borivilianum. Pharmacogn. Res. 2010, 2, 343. [Google Scholar] [CrossRef] [PubMed]
  55. Tsiftsoglou, O.S.; Stefanakis, M.K.; Lazari, D.M. Chemical constituents isolated from the rhizomes of Helleborus odorus subsp. cyclophyllus (Ranunculaceae). Biochem. Syst. Ecol. 2018, 79, 8–11. [Google Scholar] [CrossRef]
  56. Beridze, M.; Kalandia, A.; Japaridze, I.; Vanidze, M.; Varshanidze, N.; Turmanidze, N.; Dolidze, K.; Diasamidze, I.; Jakeli, E. Phytochemical study of endemic species Helleborus caucasicus, Helleborus abchasicus and Ficaria popovii spread in Southern Colchis. Proc. CBU Med. Pharm. 2020, 1, 1–7. [Google Scholar] [CrossRef]
  57. Jensen, U. Secondary compounds of the Ranunculiflorae. In Systematics and Evolution of the Ranunculiflorae; Plant Systematics and Evolution Supplement; Jensen, U., Kadereit, J.W., Eds.; Springer: Vienna, Austria, 1995; Volume 9. [Google Scholar]
  58. Cuny, E.; Klingler, D. Efficient isolation and structure analysis of (+)-ranuncoside, a unique tricyclic spiroacetal glycoside, from Christmas rose (Helleborus niger L.). Nat. Prod. Commun. 2022, 17, 1934578X2110694. [Google Scholar] [CrossRef]
  59. Schmitzer, V.; Mikulic-Petkovsek, M.; Stampar, F. Sepal phenolic profile during Helleborus niger flower development. J. Plant Physiol. 2013, 170, 1407–1415. [Google Scholar] [CrossRef]
  60. Li, Y.; Zhang, H.; Liang, X.; Song, B.; Zheng, X.; Wang, R.; Liu, L.; Song, X.; Liu, J. New cytotoxic bufadienolides from the roots and rhizomes of Helleborus thibetanus Franch. Nat. Prod. Res. 2018, 34, 950–957. [Google Scholar] [CrossRef]
  61. Brillatz, T.; Jacmin, M.; Vougogiannopoulou, K.; Petrakis, E.A.; Kalpoutzakis, E.; Houriet, J.; Pellissier, L.; Rutz, A.; Marcourt, L.; Queiroz, E.F.; et al. Antiseizure potential of the ancient Greek medicinal plant Helleborus odorus subsp. cyclophyllus and identification of its main active principles. J. Ethnopharmacol. 2020, 259, 112954. [Google Scholar] [CrossRef] [PubMed]
  62. Cheng, W.; Tan, Y.F.; Tian, H.Y.; Gong, X.W.; Chen, K.L.; Jiang, R.W. Two new bufadienolides from the rhizomes of Helleborus thibetanus with inhibitory activities against prostate cancer cells. Nat. Prod. Res. 2014, 28, 901–908. [Google Scholar] [CrossRef]
  63. Yokosuka, A.; Inomata, M.; Yoshizawa, Y.; Iguchi, T.; Mimaki, Y. Bufadienolides and ecdysteroids from the whole plants of Helleborus niger and their cytotoxicity. J. Nat. Med. 2021, 75, 393–402. [Google Scholar] [CrossRef] [PubMed]
  64. Colombo, M.L.; Tome, F.; Servettaz, O.; Bugatti, C. Phytochemical evaluation of Helleborus species growing in northern Italy. Int. J. Crude Drug Res. 1990, 28, 219–223. [Google Scholar] [CrossRef]
  65. Gu, C.; Mao, X.; Chen, D.; Yu, B.; Yang, Q. Isoleucine plays an iImportant role for maintaining immune function. Curr. Protein Pept. Sci. 2019, 20, 644–651. [Google Scholar] [CrossRef] [PubMed]
  66. Dhama, K.; Karthik, K.; Khandia, R.; Munjal, A.; Tiwari, R.; Rana, R.; Khurana, S.K.; Ullah, S.; Khan, R.U.; Alagawany, M.; et al. Medicinal and therapeutic potential of herbs and plant metabolites/Extracts countering viral pathogens-current knowledge and future prospects. Curr. Drug Metab. 2018, 19, 236–263. [Google Scholar] [CrossRef] [PubMed]
  67. Kim, S.H.; Roszik, J.; Grimm, E.A.; Ekmekcioglu, S. Impact of L-arginine metabolism on immune response and anticancer immunotherapy. Front. Oncol. 2018, 8, 67. [Google Scholar] [CrossRef]
  68. Chiangjong, W.; Chutipongtanate, S.; Hongeng, S. Anticancer peptide: Physicochemical property, functional aspect and trend in clinical application (Review). Int. J. Oncol. 2020, 57, 678–696. [Google Scholar] [CrossRef]
  69. Lieu, E.L.; Nguyen, T.; Rhyne, S.; Kim, J. Amino acids in cancer. Exp. Mol. Med. 2020, 52, 15–30. [Google Scholar] [CrossRef]
  70. Albaugh, V.L.; Pinzon-Guzman, C.; Barbul, A. Arginine metabolism and cancer. Surg. Oncol. March 2017, 115, 273–280. [Google Scholar] [CrossRef]
  71. Wei, Z.; Liu, X.; Cheng, C.; Yu, W.; Yi, P. Metabolism of amino acids in cancer. Front. Cell Dev. Biol. 2021, 8, 603837. [Google Scholar] [CrossRef]
  72. Azmi, S.; Hussain, M.K. Analysis of structures, functions, and transgenicity of phytopeptides defensin and thionin: A review. Beni-Suef Univ. J. Basic Appl. Sci. 2021, 10, 1–11. [Google Scholar] [CrossRef]
  73. Mani, S.; Bhatt, S.B.; Vasudevan, V.; Prabhu, D.; Rajamanikandan, S.; Velusamy, P.; Ramasamy, P.; Raman, P. The updated review on plant peptides and their applications in human health. Int. J. Pept. Res. Ther. 2022, 28, 135. [Google Scholar] [CrossRef]
  74. Slezina, M.P.; Odintsova, T.I. Plant antimicrobial peptides: Insights into structure-function relationships for practical applications. Curr. Issues Mol. Biol. 2023, 45, 3674–3704. [Google Scholar] [CrossRef]
  75. Liao, Y.; Wang, M.; Jiang, X. Sulfur-containing peptides: Synthesis and application in the discovery of potential drug candidates. Curr. Opin. Chem. Biol. 2023, 75, 102336. [Google Scholar] [CrossRef]
  76. Pirintsos, S.; Panagiotopoulos, A.; Bariotakis, M.; Daskalakis, V.; Lionis, C.; Sourvinos, G.; Karakasiliotis, I.; Kampa, M.; Castanas, E. From traditional ethnopharmacology to modern natural drug discovery: A methodology discussion and specific examples. Molecules 2022, 27, 4060. [Google Scholar] [CrossRef] [PubMed]
  77. Pang, Z.; Chen, J.; Wang, T.; Gao, C.; Li, Z.; Guo, L.; Xu, J.; Cheng, Y. Linking plant secondary metabolites and plant microbiomes: A Review. Front. Plant. Sci. 2021, 12, 621276. [Google Scholar] [CrossRef] [PubMed]
  78. Chen, G.; Wang, S.; Huang, X.; Hong, J.; Du, L.; Zhang, L.; Ye, L. Environmental factors affecting growth and development of banlangen (Radix Isatidis) in China. Afr. J. Plant. Sci. 2015, 9, 421–426. [Google Scholar]
  79. Kamboj, A.; Rathour, A.; Kaur, M. Bufadienolides and their medicinal utility: A review. Int. J. Pharm. Pharm. Sci. 2013, 5, 20–27. [Google Scholar]
  80. Li, F.; Weng, Y.; Wang, L.; He, H.; Yang, J.; Tang, X. The efficacy and safety of bufadienolides-loaded nanostructured lipid carriers. Int. J. Pharm. 2010, 393, 204–212. [Google Scholar] [CrossRef]
  81. Novotny, L.; Abbas Ghuloom, H.A.; Al-Hasawi, N.A. Structural features and biological activities of bufadienolides. J. Pharm. Biol. Chem. Sci. Res. 2019, 10, 1147–1157. [Google Scholar]
  82. Jiang, L.; Chi, C.; Yuan, F.; Lu, M.; Hu, D.; Wang, L.; Liu, X. Anti-inflammatory effects of anemonin on acute ulcerative colitis via targeted regulation of protein kinase C-θ. Chin. Med. 2022, 17, 1–12. [Google Scholar] [CrossRef] [PubMed]
  83. Husain, A.; Alam, M.M.; Shaharyar, M.; Lal, S. Antimicrobial activities of some synthetic butenolides and their pyrrolone derivatives. J. Enzym. Inhib. Med. Chem. 2009, 25, 54–61. [Google Scholar] [CrossRef] [PubMed]
  84. Das, N.; Mishra, S.K.; Bishayee, A.; Ali, E.S.; Bishayee, A. The phytochemical, biological, and medicinal attributes of phytoecdysteroids: An updated review. Acta Pharm. Sin. B 2021, 11, 1740–1766. [Google Scholar] [CrossRef]
  85. Cox-Georgian, D.; Ramadoss, N.; Dona, C.; Basu, C. Therapeutic and medicinal uses of terpenes. In Medicinal Plants; Joshee, N., Dhekney, S., Parajuli, P., Eds.; Springer: Cham, Switzerland, 2019; pp. 333–359. [Google Scholar]
  86. Fongang Fotsing, Y.S.; Bankeu Kezetas, J.J. Terpenoids as important bioactive constituents of essential oils. In Essential oils—Bioactive Compounds, New Perspectives and Application; Santana de Oliveira, M., Almeida da Costa, W., Gomes Silva, S., Eds.; Intechopen: London, UK, 2020; ISBN 978–1-83962–698–2. [Google Scholar]
  87. Islam, M.T.; Ali, E.S.; Uddin, S.J.; Shaw, S.; Islam, A.; Ahmed, I.; Shill, M.C.; Karmakar, U.K.; Yarla, N.S.; Khan, I.N.; et al. Phytol: A review of biomedical activities. Food Chem Toxicol. 2018, 121, 82–94. [Google Scholar] [CrossRef]
  88. Jahangeer, M.; Fatima, R.; Ashiq, M.; Basharat, A.; Qamar, S.A.; Bilal, M.; Iqbal, H.M. Therapeutic and biomedical potentialities of terpenoids–A review. J. Pure Appl. Microbiol. 2021, 15, 471–483. [Google Scholar] [CrossRef]
  89. Das, K.; Gezici, S. Secondary plant metabolites, their separation and identification, and role in human disease prevention. Ann. Phytomed. Int. J. 2018, 7, 13–24. [Google Scholar] [CrossRef]
  90. Martin, M.L.; Ortíz de Urbina, A.V.; Montero, M.J.; Carrón, R.; San Román, L. Pharmacologic effects of lactones isolated from Pulsatilla alpina subsp. apiifolia. J. Ethnopharmacol. 1988, 24, 185–191. [Google Scholar] [CrossRef]
  91. Mares, D. Antimicrobial activity of protoanemonin, a lactone from Ranunculaceous plants. Mycopathologia 1987, 98, 133–140. [Google Scholar] [CrossRef]
  92. Nagy, K.; Tiuca, I.D. Importance of fatty acids in physiopathology of human body. In Fatty Acids; Catala, A., Ed.; IntechOpen: London, UK, 2017. [Google Scholar]
  93. Kozłowska, A.; Szostak-Węgierek, D. Flavonoids–food sources, health benefits, and mechanisms involved. In Bioactive Molecules in Food; Reference Series in Phytochemistry; Mérillon, J.M., Ramawat, K., Eds.; Springer: Cham, Switzerland, 2018. [Google Scholar]
  94. Kozłowska, A.; Szostak-Wegierek, D. Flavonoids-food sources and health benefits. Rocz. Panstw. Zakl. Hig. 2014, 65, 79–85. [Google Scholar] [PubMed]
  95. Panche, A.N.; Diwan, A.D.; Chandra, S.R. Flavonoids: An overview. J. Nutr. Sci. 2016, 5, e47. [Google Scholar] [CrossRef]
  96. Rasouli, H.; Farzaei, M.H.; Khodarahmi, R. Polyphenols and their benefits: A review. Int. J. Food Prop. 2017, 20, 1700–1741. [Google Scholar] [CrossRef]
  97. Kumar, N.; Goel, N. Phenolic acids: Natural versatile molecules with promising therapeutic applications. Biotechnol. Rep. Amst. 2019, 24, e00370. [Google Scholar] [CrossRef] [PubMed]
  98. Minatel, I.O.; Vanz Borges, C.; Ferreira, M.I.; Gomez, H.A.G.; Chen, C.Y.O.; Pace Pereira Lima, G. Phenolic compounds: Functional properties, impact of processing and bioavailability. In Phenolic Compounds-Biological Activity; Soto-Hernandez, M., Palma-Tenango, M., del Rosario Garcia-Mateos, M., Eds.; IntechOpen: Rijeka, Croatia, 2017. [Google Scholar]
  99. Salehi, B.; Quispe, C.; Sharifi-Rad, J.; Cruz-Martins, N.; Nigam, M.; Mishra, A.P.; Konovalov, D.A.; Orobinskaya, V.; Abu-Reidah, I.M.; Zam, W.; et al. Phytosterols: From preclinical evidence to potential clinical applications. Front. Pharmacol. 2021, 11, 599959. [Google Scholar] [CrossRef] [PubMed]
  100. Kilcoyne, M.; Joshi, L. Carbohydrates in therapeutics. Cardiovasc. Hematol. Agents Med. Chem. 2007, 5, 186–197. [Google Scholar] [CrossRef] [PubMed]
  101. Singh, D.; Rajput, A.; Bhatia, A.; Kumar, A.; Kaur, H.; Sharma, P.; Kaur, P.; Singh, S.; Attri, S.; Buttar, H.; et al. Plant-based polysaccharides and their health functions. Funct. Foods Health Dis. 2021, 11, 179–200. [Google Scholar]
  102. Khan, H.; Saeedi, M.; Nabavi, S.; Mubarak, M.; Bishayee, A. Glycosides from medicinal plants as potential anticancer agents: Emerging trends towards future drugs. Curr. Med. Chem. 2018, 26, 2389–2406. [Google Scholar] [CrossRef]
  103. Okon, E.; Kukula-Koch, W.; Halasa, M.; Jarzab, A.; Baran, M.; Dmoszynska-Graniczka, M.; Angelis, A.; Kalpoutzakis, E.; Guz, M.; Stepulak, A.; et al. Magnoflorine—Isolation and the anticancer potential against NCI-H1299 lung, MDA-MB-468 breast, T98G glioma, and TE671 rhabdomyosarcoma cancer cells. Biomolecules 2020, 10, 1532. [Google Scholar] [CrossRef]
  104. Baumel, B.S.; Doraiswamy, P.M.; Sabbagh, M.; Wurtman, R. Potential neuroregenerative and neuroprotective effects of uridine/choline-enriched multinutrient dietary intervention for mild cognitive impairment: A narrative review. Neurol. Ther. 2021, 10, 43–60. [Google Scholar] [CrossRef]
  105. Meng, X.; Zhou, J.; Zhao, N.; Gan, Y.; Li, B. Health benefits and molecular mechanisms of resveratrol: A narrative review. Foods 2020, 9, 340. [Google Scholar] [CrossRef]
  106. Afzal, O.; Altamimi, A.S.A.; Nadeem, M.S.; Alzarea, S.I.; Almalki, W.H.; Tariq, A.; Mubeen, B.; Murtaza, B.N.; Iftikhar, S.; Riaz, N.; et al. Nanoparticles in drug delivery: From history to therapeutic applications. Nanomaterials 2022, 12, 4494. [Google Scholar] [CrossRef] [PubMed]
  107. Alhadrami, H.A.; Orfali, R.; Hamed, A.A.; Ghoneim, M.M.; Hassan, H.M.; Hassane, A.S.I.; Rateb, M.E.; Sayed, A.M.; Gamaleldin, N.M. Flavonoid-coated gold nanoparticles as efficient antibiotics against gram-negative bacteria-evidence from in silico-supported in vitro studies. Antibiotics 2021, 10, 968. [Google Scholar] [CrossRef] [PubMed]
  108. Emran, T.B.; Shahriar, A.; Mahmud, A.R.; Rahman, T.; Abir, M.H.; Siddiquee, M.F.; Ahmed, H.; Rahman, N.; Nainu, F.; Wahyudin, E.; et al. Multidrug resistance in cancer: Understanding molecular mechanisms, immunoprevention and therapeutic approaches. Front. Oncol. 2022, 12, 2581. [Google Scholar] [CrossRef] [PubMed]
  109. Stuart, B.H. Infrared Spectroscopy: Fundamentals and Applications; John Wiley & Sons, Ltd: Chichester, UK, 2004; ISBN 9780470854273. [Google Scholar]
  110. Pecio, Ł.; Kowalczyk, M.; Stochmal, A. New bufadienolides isolated from the roots of Kalanchoe daigremontiana (Crassulaceae). Molecules 2016, 21, 243. [Google Scholar]
  111. Knickelbein, R.G. Pharmacological Effects of a Bufadienolide Isolated from the Toad Bufo Boreas Halophilus. Master’s Thesis, University of the Pacific, Stockton, CA, USA, 1972. Available online: https://scholarlycommons.pacific.edu/uop_etds/1766 (accessed on 18 June 2023).
  112. Guo, F.; Li, Z.; Xu, X.; Wang, K.; Shao, M.; Zhao, F.; Wang, H.; Hua, H.; Pei, Y.; Bai, J. Butenolide derivatives from the plant endophytic fungus Aspergillus Terreus. Fitoterapia 2016, 113, 44–50. [Google Scholar] [CrossRef]
  113. Zaldo Castro, A.; Tacoronte, J.E.; Coll Mancha, F.; Aguilera de la Paz, L.; Cabrera, M.T. Ecdysteroid analogs based on steroidal sapogenins I. Synthesis of bromo-derivatives from diosgenin. Preliminary study of their biological activity. Rev. CENIC. Cienc. Químicas 2002, 33, 19–24. [Google Scholar]
  114. Sreejit, C.M.; Bose, C.; Banerji, A.; Thomas, M.P. Isolation, quantification and chemical characterisation of ecdysterone from medicinal plants of Kerala, Western Ghats. J. Pharmacogn. Phytochem. 2019, 8, 254–257. [Google Scholar]
  115. Bakshi, K.; Liyanage, M.R.; Volkin, D.B.; Middaugh, C.R. Fourier Transform Infrared spectroscopy of peptides. Ther. Pept. 2013, 1088, 255–269. [Google Scholar]
  116. Fabian, H.; Schultz, C.P. Fourier Transform Infrared Spectroscopy in Peptide and Protein Analysis. Encycl. Anal. Chem. 2006. [Google Scholar] [CrossRef]
  117. Segneanu, A.E.; Marin, C.N.; Herea, D.D.; Stanusoiu, I.; Muntean, C.; Grozescu, I. Romanian Viscum album L.—Untargeted low-molecular metabolomic approach to engineered Viscum–AuNPs carrier assembly. Plants 2022, 11, 820. [Google Scholar] [CrossRef]
  118. Scarsini, M.; Thurotte, A.; Veidl, B.; Amiard, F.; Niepceron, F.; Badawi, M.; Lagarde, F.; Schoefs, B.; Marchand, J. Metabolite quantification by Fourier transform infrared spectroscopy in diatoms: Proof of concept on Phaeodactylum tricornutum. Front. Plant Sci. 2021, 12, 756421. [Google Scholar] [CrossRef] [PubMed]
  119. Topala, C.M.; Tatarua, L.D.; Ducu, C. ATR-FTIR spectra fingerprinting of medicinal herbs extracts prepared using microwave extraction. Arab. J. Med. Aromat. Plants 2017, 3, 1–9. [Google Scholar]
  120. Heneczkowski, M.; Kopacz, M.; Nowak, D.; Kuźniar, A. Infrared spectrum analysis of some flavonoids. Acta Pol. Pharm. Drug Res. 2001, 58, 415–420. [Google Scholar]
  121. Noh, C.H.C.; Azmin, N.F.M.; Amid, A. Principal component analysis application on flavonoids characterization. Adv. Sci. Technol. Eng. Syst. J. 2017, 2, 435–440. [Google Scholar] [CrossRef]
  122. Pang, M.; Jiang, S.; Cao, L.; Pan, L. Novel synthesis of steryl esters from phytosterols and amino acid. J. Agric. Food Chem. 2011, 59, 10732–10736. [Google Scholar] [CrossRef] [PubMed]
  123. Wiercigroch, E.; Szafraniec, E.; Czamara, K.; Pacia, M.Z.; Majzner, K.; Kochan, K.; Kaczor, A.; Baranska, M.; Malek, K. Raman and infrared spectroscopy of carbohydrates: A review. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2017, 185, 317–335. [Google Scholar] [CrossRef]
  124. Segneanu, A.; Velciov, S.M.; Olariu, S.; Cziple, F.; Damian, D.; Grozescu, I. Bioactive molecules profile from natural compounds. In Amino Acid—New Insights and Roles in Plant and Animal; Asao, T., Asaduzzaman, M., Eds.; IntechOpen: Rijeka, Croatia, 2017. [Google Scholar]
  125. Li, B.; Han, L.; Cao, B.; Yang, X.; Zhu, X.; Yang, B.; Zhao, H.; Qiao, W. Use of magnoflorine-phospholipid complex to permeate blood-brain barrier and treat depression in the CUMS animal model. Drug Deliv. 2019, 26, 566–574. [Google Scholar] [CrossRef]
  126. Olejniczak, A.B.; Sut, A.; Wróblewski, A.E.; Leśnikowski, Z.J. Infrared spectroscopy of nucleoside and DNA-oligonucleotide conjugates labeled with carborane or metallacarborane cage. Vib. Spectrosc. 2005, 39, 177–185. [Google Scholar] [CrossRef]
  127. Trung, B.V.; Thi Thao, D.; Anh, D.H.; Van Kiem, P.; Viet, P.H. Antioxidant and hepatoprotective activity of phenyl glycosides isolated from Heliciopsis lobata. Nat. Prod. Commun. 2020, 15, 1934578X20946255. [Google Scholar] [CrossRef]
  128. Billes, F.; Mohammed-Ziegler, I.; Mikosch, H.; Tyihák, E. Vibrational spectroscopy of resveratrol. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2007, 68, 669–679. [Google Scholar] [CrossRef]
  129. Kumpugdee-Vollrath, M.; Ibold, Y.; Sriamornsak, P. Solid state characterization of trans resveratrol complexes with different cyclodextrins. JAASP 2012, 1, 125–136. [Google Scholar]
  130. Segneanu, A.E.; Vlase, G.; Lukinich-Gruia, A.T.; Herea, D.D.; Grozescu, I. Untargeted metabolomic approach of Curcuma longa to neurodegenerative phytocarrier system based on silver nanoparticles. Antioxidants 2022, 11, 2261. [Google Scholar] [CrossRef]
  131. Hong, T.; Yin, J.Y.; Nie, S.P.; Xie, M.Y. Applications of infrared spectroscopy in polysaccharide structral analysis: Progress, challenge and perspective. Food Chem. X 2021, 12, 100168. [Google Scholar] [CrossRef] [PubMed]
  132. Fernandes Queiroz, M.; Melo, K.R.; Sabry, D.A.; Sassaki, G.L.; Rocha, H.A. Does the use of chitosan contribute to oxalate kidney stone formation? Mar. Drugs 2014, 3, 141–158. [Google Scholar] [CrossRef] [PubMed]
  133. Suédina, M.L.S.; Braga, C.R.C.; Fook, M.V.L.; Raposo, C.M.O.; Carvalho, L.H.; Canedo, E.L. Application of Infrared Spectroscopy to Analysis of Chitosan/Clay Nanocomposites, Infrared Spectroscopy—Materials Science, Engineering and Technology; Theophile, T., Ed.; InTech: London, UK, 2012; ISBN 978-953-51-0537-4. [Google Scholar]
  134. Kavaz, D.; Kirac, F.; Kirac, M.; Vaseashta, A. Low releasing mitomycin C molecule encapsulated with chitosan nanoparticles for intravesical installation. J. Biomater. Nanobiotechnol. 2017, 8, 203–219. [Google Scholar] [CrossRef]
  135. Ainali, N.M.; Xanthopoulou, E.; Michailidou, G.; Zamboulis, A.; Bikiaris, D.N. Microencapsulation of fluticasone propionate and salmeterol xinafoate in modified chitosan microparticles for release optimization. Molecules 2020, 25, 3888. [Google Scholar] [CrossRef]
  136. Danish, M.S.S.; Estrella-Pajulas, L.L.; Alemaida, I.M.; Grilli, M.L.; Mikhaylov, A.; Senjyu, T. Green synthesis of silver oxide nanoparticles for photocatalytic environmental remediation and biomedical Applications. Metals 2022, 12, 769. [Google Scholar] [CrossRef]
  137. Morsy, M.; Khaled, M.; Amyn, H.; El Ebissy, A.; Salah, A.; Youssef, M. Synthesis and characterization of freeze dryer chitosan nano particles as multi functional eco-friendly finish for fabricating easy care and antibacterial cotton textiles. Egypt. J. Chem. 2019, 62, 1277–1293. [Google Scholar] [CrossRef]
  138. Soleymanfallah, S.; Khoshkhoo, Z.; Hosseini, S.E.; Azizi, M.H. Preparation, physical properties, and evaluation of antioxidant capacity of aqueous grape extract loaded in chitosan-TPP nanoparticles. Food Sci. Nutr. 2022, 10, 3272–3281. [Google Scholar] [CrossRef]
  139. Cui, S.F.; Wang, J.W.; Li, H.F.; Fang, R.; Yu, X.; Lu, Y.J. Microencapsulation of capsaicin in chitosan microcapsules: Characterization, release behavior, and pesticidal properties against Tribolium castaneum (Herbst). Insects 2023, 14, 27. [Google Scholar] [CrossRef]
  140. Vanaja, M.; Annadurai, G. Coleus aromaticus leaf extract mediated synthesis of silver nanoparticles and its bactericidal activity. Appl. Nanosci. 2012, 3, 217–223. [Google Scholar] [CrossRef]
  141. Jaya, S.; Durance, T.D.; Wang, R. Physical characterization of drug loaded microcapsules and controlled in vitro release study. Open Biomater. J. 2010, 2, 9–17. [Google Scholar]
  142. Chen, C.; Li, Z.; Wang, C.; Liu, S.; Wang, Y.; Zhang, M.; Tian, Y.; Lv, J.; Xu, H.; Xia, G. Stability and antioxidant activity of chitosan/β-lactoglobulin on anthocyanins from Aronia melanocarpa. LWT Food Sci. Technol. 2023, 173, 114335. [Google Scholar] [CrossRef]
  143. Da Silva, N.C.; Garrido Assis, O.B.; de Oliveira Sartori, A.G.; de Alencar, S.M.; Martelli-Tosi, M. Chitosan suspension as extractor and encapsulating agent of phenolics from acerola by-product. Food Res. Int. 2022, 161, 111855. [Google Scholar] [CrossRef] [PubMed]
  144. Georgieva, V.; Zvezdova, D.; Vlaev, L. Non-isothermal kinetics of thermal degradation of chitosan. Chem. Cent. J. 2012, 6, 81–91. [Google Scholar] [CrossRef]
  145. Ochiuz, L.; Popa, G.; Stoleriu, I.; Tomoiagă, A.M.; Popa, M. Microencapsulation of metoprolol tartrate into chitosan for improved oral administration and patient compliance. Ind. Eng. Chem. Res. 2013, 52, 17432–17441. [Google Scholar] [CrossRef]
  146. Negi, A.; Kesari, K.K. Chitosan nanoparticle encapsulation of antibacterial essential oils. Micromachines 2022, 13, 1265. [Google Scholar] [CrossRef]
  147. Alehosseini, E.; Shahiri Tabarestani, H.; Kharazmi, M.S.; Jafari, S.M. Physicochemical, thermal, and morphological properties of chitosan nanoparticles produced by ionic gelation. Foods 2022, 11, 3841. [Google Scholar] [CrossRef]
  148. Eulalio, H.Y.C.; Rodrigues, J.F.B.; Santos, K.O.; Peniche, C.; LiaFook, M.V. Characterization and thermal properties of chitosan films prepared with different acid solvents. Rev. Cuba Química 2019, 31, 309–323. [Google Scholar]
  149. Othayoth, R.; Khatri, K.; Gadicherla, R.; Kodandapani, S.; Botlagunta, M. Multivitamin–cisplatin encapsulated chitosan nanoparticles modulate DDX3X expression in cancer cell lines. Nano Biomed. Eng. 2023, 15, 74–85. [Google Scholar] [CrossRef]
  150. Chanaj-Kaczmarek, J.; Rosiak, N.; Szymanowska, D.; Rajewski, M.; Wender-Ozegowska, E.; Cielecka-Piontek, J. The chitosan-based system with Scutellariae baicalensis radix extract for the local treatment of vaginal infections. Pharmaceutics 2022, 14, 740. [Google Scholar] [CrossRef] [PubMed]
  151. Moreno-Vásquez, M.J.; Valenzuela-Buitimea, E.L.; Plascencia-Jatomea, M.; Encinas-Encinas, J.C.; Rodríguez-Félix, F.; Sánchez-Valdes, S.; Rosas-Burgos, E.C.; Ocaño-Higuera, V.M.; Graciano-Verdugo, A.Z. Functionalization of chitosan by a free radical reaction: Characterization, antioxidant and antibacterial potential. Carbohydr. Polym. 2017, 155, 117–127. [Google Scholar] [CrossRef] [PubMed]
  152. Gulcin, İ. Antioxidants and antioxidant methods: An updated overview. Arch. Toxicol. 2020, 94, 651–715. [Google Scholar] [CrossRef] [PubMed]
  153. Dontha, S. A review on antioxidant methods. Asian J. Pharm. Clin. Res. 2016, 9, 14–32. [Google Scholar]
  154. Kedare, S.B.; Singh, R.P. Genesis and development of DPPH method of antioxidant assay. J. Food Sci. Technol. 2011, 48, 412–422. [Google Scholar] [CrossRef] [PubMed]
  155. Haida, Z.; Hakiman, M. A comprehensive review on the determination of enzymatic assay and nonenzymatic antioxidant activities. Food Sci. Nutr. 2019, 7, 1555–1563. [Google Scholar] [CrossRef] [PubMed]
  156. Rahman, M.M.; Islam, M.B.; Biswas, M.; Alam, A.H.M.K. In vitro antioxidant and free radical scavenging activity of different parts of Tabebuia pallida growing in Bangladesh. BMC Res. Notes 2015, 8, 621. [Google Scholar] [CrossRef] [PubMed]
  157. Soltanzadeh, M.; Peighambardoust, S.H.; Ghanbarzadeh, B.; Mohammadi, M.; Lorenzo, J.M. Chitosan nanoparticles as a promising nanomaterial for encapsulation of pomegranate (Punica granatum L.) peel extract as a natural source of antioxidants. Nanomaterials 2021, 11, 1439. [Google Scholar] [CrossRef]
  158. Purgiyanti, K.; Kumoro, A.C.; Legowo, A.M. The antioxidant and antibacterial activities of chitosan extract from white shrimp shell (Penaeus indicus) in the waters north of Brebes, Indonesia. Biodiversitas 2022, 23, 1267–1272. [Google Scholar]
  159. Li, Q.; Wei, L.; Zhang, J.; Gu, G.; Guo, Z. Significantly enhanced antioxidant activity of chitosan through chemical modification with coumarins. Polym. Chem. 2019, 10, 1480–1488. [Google Scholar] [CrossRef]
  160. Lamuela-Raventós, R.M. Folin-Ciocalteu method for the measurement of total phenolic content and antioxidant capacity. In Measurement of Antioxidant Activity & Capacity; Wiley Online Library: New York, NY, USA, 2017; pp. 107–115. [Google Scholar]
  161. Carmona-Hernandez, J.C.; Taborda-Ocampo, G.; González-Correa, C.H. Folin-Ciocalteu reaction alternatives for higher polyphenol quantitation in Colombian passion fruits. Int. J. Food Sci. 2021, 2021, 8871301. [Google Scholar] [CrossRef]
  162. Casettari, L.; Gennari, L.; Angelino, D.; Ninfali, P.; Castagnino, E. ORAC of chitosan and its derivatives. Food Hydrocoll. 2012, 28, 243–247. [Google Scholar] [CrossRef]
  163. Sadeer, N.B.; Montesano, D.; Albrizio, S.; Zengin, G.; Mahomoodally, M.F. The versatility of antioxidant assays in food science and safety—Chemistry, applications, strengths, and limitations. Antioxidants 2020, 9, 709. [Google Scholar] [CrossRef] [PubMed]
  164. Zhao, X.; Zhou, L.; Riaz Rajoka, M.S.; Yan, L.; Jiang, C.; Shao, D.; Zhu, J.; Shi, J.; Huang, Q.; Yang, H.; et al. Fungal silver nanoparticles: Synthesis, application and challenges. Crit. Rev. Biotechnol. 2017, 38, 1–19. [Google Scholar] [CrossRef] [PubMed]
  165. Bedlovičová, Z.; Strapáč, I.; Baláž, M.; Salayová, A. A brief overview on antioxidant activity determination of silver nanoparticles. Molecules 2020, 25, 3191. [Google Scholar] [CrossRef]
  166. George, S.; Abrahamse, H. Redox potential of antioxidants in cancer progression and prevention. Antioxidants 2020, 9, 1156. [Google Scholar] [CrossRef]
  167. Shivakumar, A.; Yogendra Kumar, M.S. Critical review on the analytical mechanistic steps in the evaluation of antioxidant activity. Crit. Rev. Anal. Chem. 2018, 48, 214–236. [Google Scholar] [CrossRef]
  168. Özyürek, M.; Çelik, S.E.; Berker, K.I.; Güçlü, K.; Tor, İ.; Apak, R. Sensitivity enhancement of CUPRAC and iron(III)-phenanthroline antioxidant assays by preconcentration of colored reaction products on a weakly acidic cation exchanger. React. Funct. Polym. 2007, 67, 1478–1486. [Google Scholar] [CrossRef]
  169. Mashentseva, A.; Dehaen, W.; Seitembetov, T.; Seitembetova, A. Comparison of the antioxidant activity of the different Betula pendula Roth. extracts from Northern Kazakhstan. J. Phytol. 2011, 3, 18–25. [Google Scholar]
  170. Jing, Y.; Zhang, S.; Li, M.; Zhang, R.; Zhang, H.; Zheng, Y.; Zhang, D.; Wu, L. Structural characterization and biological activities of polysaccharide iron complex synthesized by plant polysaccharides: A review. Front. Nutr. 2022, 9, 1013067. [Google Scholar] [CrossRef]
  171. Miranda, R.R.; Sampaio, I.; Zucolotto, V. Exploring silver nanoparticles for cancer therapy and diagnosis. Colloids Surf. B Biointerfaces 2021, 210, 112254. [Google Scholar] [CrossRef] [PubMed]
  172. Gibis, M.; Rahn, N.; Weiss, J. Physical and oxidative stability of uncoated and chitosan-coated liposomes containing grape seed extract. Pharmaceutics 2013, 5, 421–433. [Google Scholar] [CrossRef]
  173. Morales-Olán, G.; Luna-Suárez, S.; De Dios Figueroa-Cárdenas, J.; Corea, M.; Rojas-López, M. Synthesis and characterization of chitosan particles loaded with antioxidants extracted from chia (Salvia hispanica L.) seeds. Int. J. Anal. Chem. 2021, 2012, 5540543. [Google Scholar] [CrossRef] [PubMed]
  174. Khorshidian, N.; Mahboubi, A.; Kalantari, N.; Hosseini, H.; Yousefi, M.; Arab, M.; Gomez da Cruzf, A.; Mortazavian, A.M.; Sadat Mahdavi, F. Chitosan-coated alginate microcapsules loaded with herbal galactagogue extract: Formulation optimization and characterization. Iran. J. Pharm. Res. 2019, 18, 1180–1195. [Google Scholar] [PubMed]
  175. Ribeiro, E.F.; Polachini, T.C.; Alvim, I.D.; Quiles, A.; Hernando, I.; Nicoletti, V.R. Microencapsulation of roasted coffee oil pickering emulsions using spray- and freeze-drying: Physical, structural and in vitro bioaccessibility studies. Int. J. Food Sci. Technol. 2022, 57, 145–153. [Google Scholar] [CrossRef]
  176. Maghraby, Y.R.; Farag, M.A.; Kontominas, M.G.; Shakour, Z.T.; Ramadan, A.R. Nanoencapsulated extract of a red seaweed (Rhodophyta) species as a promising source of natural antioxidants. ACS Omega 2022, 7, 6539–6548. [Google Scholar] [CrossRef] [PubMed]
  177. Kulikouskaya, V.; Hileuskaya, K.; Kraskouski, A.; Kozerozhets, I.; Stepanova, E.; Kuzminski, I.; You, L.; Agabekov, V. Chitosan-capped silver nanoparticles: A comprehensive study of polymer molecular weight effect on the reaction kinetic, physicochemical properties, and synergetic antibacterial potential. SPE Polym. 2022, 3, 77–90. [Google Scholar] [CrossRef]
  178. Adams, R.P. Identification of Essential Oil Components by Gas Chromatography/Mass Spectrometry; Allured Publishing Corporation: Carol Stream, IL, USA, 2007; Volume 456. [Google Scholar]
  179. Öztürk, M.; Bulduk, İ.; Korcan, S.; Liman, R.; Karabag, F.; Kargıoğlu, M.; Konuk, M. Total phenolics, flavonoids contents, antioxidant activity and DNA protective effect of lenten rose (Helleborus orientalis). Asian J. Biochem. Genet. Mol. Biol. 2019, 1, 1–12. [Google Scholar] [CrossRef]
  180. Baliyan, S.; Mukherjee, R.; Priyadarshini, A.; Vibhuti, A.; Gupta, A.; Pandey, R.P.; Chang, C.M. determination of antioxidants by DPPH radical scavenging activity and quantitative phytochemical analysis of Ficus religiosa. Molecules 2022, 27, 1326. [Google Scholar] [CrossRef]
  181. Christodoulou, M.C.; Orellana Palacios, J.C.; Hesami, G.; Jafarzadeh, S.; Lorenzo, J.M.; Domínguez, R.; Moreno, A.; Hadidi, M. Spectrophotometric methods for measurement of antioxidant activity in food and pharmaceuticals. Antioxidants 2022, 11, 2213. [Google Scholar] [CrossRef]
  182. Khatoon, M.; Islam, E.; Islam, R.; Rahman, A.A.; Alam, A.K.; Khondkar, P.; Rashid, M.; Parvin, S. Estimation of total phenol and in vitro antioxidant activity of Albizia procera leaves. BMC Res. Notes 2013, 6, 121. [Google Scholar] [CrossRef] [PubMed]
  183. Yefrida, H.; Suyani, A.; Efdi, M.; Hermansyah, A. Modification of phenanthroline method to determine antioxidant content in tropical fruits methanolic extract. Res. J. Chem. Environ. 2018, 22, 28–35. [Google Scholar]
  184. Garcia, L.G.S.; da Rocha, M.G.; Lima, L.R.; Cunha, A.P.; de Oliveira, J.S.; de Andrade, A.R.C.; Ricardo, N.M.P.S.; Pereira-Neto, W.A.; Sidrim, J.J.C.; Rocha, M.F.G.; et al. Essential oils encapsulated in chitosan microparticles against Candida albicans biofilms. Int. J. Biol. Macromol. 2020, 166, 621–632. [Google Scholar] [CrossRef] [PubMed]
  185. Villate, A.; San Nicolas, M.; Olivares, M.; Aizpurua-Olaizola, O.; Usobiaga, A. Chitosan-coated alginate microcapsules of a full-spectrum Cannabis extract: Characterization, long-term stability and in vitro bioaccessibility. Pharmaceutics 2023, 15, 859. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Phytoconstituent classification bar chart for Helleborus purpurascens.
Figure 1. Phytoconstituent classification bar chart for Helleborus purpurascens.
Plants 12 03479 g001
Figure 2. FT-IR spectra: (A) chitosan; (B) hellebore; (C) HC phyto-nanocarrier; (D) hellebore-AgNPs system.
Figure 2. FT-IR spectra: (A) chitosan; (B) hellebore; (C) HC phyto-nanocarrier; (D) hellebore-AgNPs system.
Plants 12 03479 g002
Figure 3. FT-IR spectra: (A) citrate-coated AgNPs, hellebore and hellebore-AgNPs; (B) chitosan, HC, hellebore-AgNPs system and HAgC.
Figure 3. FT-IR spectra: (A) citrate-coated AgNPs, hellebore and hellebore-AgNPs; (B) chitosan, HC, hellebore-AgNPs system and HAgC.
Plants 12 03479 g003
Figure 4. Powder XRD patterns of the CN, hellebore-AgNPs system, and HC and HAgC phyto-nanocarriers.
Figure 4. Powder XRD patterns of the CN, hellebore-AgNPs system, and HC and HAgC phyto-nanocarriers.
Plants 12 03479 g004
Figure 5. SEM images: (A) hellebore; (B) chitosan; (C,D) HC phyto-nanocarrier; (EG) hellebore-AgNPs system; (HJ) HAgC phyto-nanocarrier.
Figure 5. SEM images: (A) hellebore; (B) chitosan; (C,D) HC phyto-nanocarrier; (EG) hellebore-AgNPs system; (HJ) HAgC phyto-nanocarrier.
Plants 12 03479 g005
Figure 6. EDX composition: (A) HC phyto-nanocarrier; hellebore-AgNPs system (B) before and (C) after encapsulation.
Figure 6. EDX composition: (A) HC phyto-nanocarrier; hellebore-AgNPs system (B) before and (C) after encapsulation.
Plants 12 03479 g006
Figure 7. DLS patterns: (A) chitosan particles; (B) HC and (C) HAgC phyto-nanocarriers.
Figure 7. DLS patterns: (A) chitosan particles; (B) HC and (C) HAgC phyto-nanocarriers.
Plants 12 03479 g007
Figure 8. TG/DTG thermograms: (A) chitosan sample; (B) HC; (C) HAgC.
Figure 8. TG/DTG thermograms: (A) chitosan sample; (B) HC; (C) HAgC.
Plants 12 03479 g008aPlants 12 03479 g008b
Figure 9. Comparison of the thermal stability of both phyto-nanocarriers (HC and HAgC).
Figure 9. Comparison of the thermal stability of both phyto-nanocarriers (HC and HAgC).
Plants 12 03479 g009
Figure 10. DSC curve of the HAgC phyto-nanocarrier.
Figure 10. DSC curve of the HAgC phyto-nanocarrier.
Plants 12 03479 g010
Figure 11. Graphic representation: (A) DPPH results expressed as IC50 (mg/mL); (B) total polyphenolic contents assay results; (C) phosphomolybdate (total antioxidant capacity) results; (D) iron(III)-phenanthroline assay results.
Figure 11. Graphic representation: (A) DPPH results expressed as IC50 (mg/mL); (B) total polyphenolic contents assay results; (C) phosphomolybdate (total antioxidant capacity) results; (D) iron(III)-phenanthroline assay results.
Plants 12 03479 g011aPlants 12 03479 g011b
Table 1. Phytochemicals identified via GC–MS analysis of the hellebore sample.
Table 1. Phytochemicals identified via GC–MS analysis of the hellebore sample.
No. Retention Time (RT) (min)Retention Index (RI) DeterminedAdams Index (AI)Area (%)Compound NameRef.
111.43197219832.18palmitic acid[47,48,49]
211.98159816025.32hexadecane[47]
314.61316031686.76stigmasterol[47,50]
415.87319532139.85sitosterol[47,50]
517.22254326348.53stearic acid[47,48,49]
625.962078212321.84phytol[47,50]
726.882848286135.69squalene[47,50]
828.45268826946.58heptacosane[47,48]
RI—retention index calculated based on a calibration curve of a C8–C20 alkane standard mixture.
Table 2. The secondary metabolites identified in the mass spectrometry analysis in hellebore sample.
Table 2. The secondary metabolites identified in the mass spectrometry analysis in hellebore sample.
No.m/z DetectedTheoretic m/zFormulaTentative of IdentificationCategoryRef.
176.0775.07C2H5NO2glycineamino acids[3,4,11]
289.8889.09C3H7NO2alanineamino acids[4,11,13]
396.0996.08C5H4O2protoanemoninlactones[2,6,11]
4116.15116.16C6H12O2caproic acidfatty acids[47]
5119.11119.12C4H9NO3threonineamino acids[3,4,11]
6131.17131.17C6H13NO2isoleucineamino acids[3,4,11]
7132.15132.16C5H12N2O2ornithineamino acids[3,4,11]
8147.13146.14C5H10N2O3alanylglycinepeptides[4,11]
9148.12147.13C5H9NO4glutamic acidamino acids[3,4,11]
10151.14150.13C5H10O5arabinosecarbohydrates[51,52]
11156.16155.15C6H9N3O2histidineamino acids[3,4,11]
12155.15164.16C9H8O3p-coumaric acidphenolic acids[3,10,53]
13166.18165.19 C9H11NO2phenylalanineamino acids[3,4,11]
14167.16166.17C9H10O33-phenyllactic acidorganic acid[52,53]
15171.11170.12C7H6O5gallic acidphenolic acids[3,10,53]
16175.19174.20C6H14N4O2arginineamino acids[4,11]
17177.18176.17C6H12N2O4glycylthreoninepeptides[4,11]
18181.17180.16C9H8O4caffeic acidphenolic acids[3,10,53]
19193.16192.17C10H8O4anemoninbutenolides[2,3,51]
20195.18194.18C10H10O4ferulic acidphenolic acids[3,10,52]
21201.33200.32C12H24O2lauric acidfatty acids[47]
22203.26202.25C9H18N2O3alanyl-isoleucinepeptides[4,11]
23205.21204.22C11H12N2O2tryptophanamino acids[3,4,11]
24213.22212.21C8H12N4O3glycyl-histidinepeptides[4,11]
25218.26217.27C9H19N3O3alanyl-lysinepeptides[4,11]
26220.23219.24C8H17N3O4ornithinoalaninepeptides[4,11]
27221.29220.29C8H16N2O3Salanylmethioninepeptides[4,11]
28225.29224.30C6H12N2O3S2cysteylcysteinepeptides[4,11]
29227.45226.44C16H34hexadecanehydrocarbons[54]
30229.23228.24C14H12O3resveratrol stilbenoids[3,10]
31241.29240.30C6H12N2O4S2cystineamino acids[3,4,11]
32244.31243.30C11H21N3O3lysylprolinepeptides[4,11]
33245.19244.20C9H12N2O6uridinenucleosides[55]
34249.35248.34C10H20N2O3Svalyl-methioninepeptides[4,11]
35253.39252.40C8H16N2O3S2cysteylmethioninepeptides[4,11]
36255.28254.29C11H18N4O3histidylvalinepeptides[4,11]
37257.24256.25C12H16O6phenylglucosideglycosides[55]
38257.43256.42C16H32O2palmitic acidfatty acids[47]
39260.35259.35C12H25N3O3leucyllysine peptides[4,11]
40262.27261.28C13H15N3O3tryptophylglycinepeptides[4,11]
41269.33268.33C12H16N2O3Sphenylalanylcysteinepeptides[4,11]
42270.25269.26C10H15N5O4asparaginyl-Histidinepeptides[4,11]
43271.25270.24C15H10O5apigeninflavonoids[55]
44271.49270.50C17H34O2margaric acidfatty acids[47]
45276.31275.30C14H17N3O3tryptophylalaninepeptides[4,11]
46277.25276.24C11H16O8ranunculin glycosides[51,56,57]
47278.35277.35C9H19N5O3Scysteinyl-argininepeptides[4,11]
48279.34278.35C15H22N2O3phenylalanylleucinepeptides[4,11]
49284.28283.28C11H17N5O4alanyl-glycyl-histidinepeptides[4,11]
50284.49284.50C18H36O2stearic acid fatty acids[46]
51287.23286.24C15H10O6kaempferolflavonoids[3,10,58,59]
52287.34286.35C11H18N4O3Shistidylmethioninepeptides[4,11]
53288.37287.36C12H25N5O3leucylargininepeptides[4,11]
54291.26290.27C15H14O6epicatechinflavonoids[3,10]
55294.31293.32C14H19N3O4phenylalanylglutaminepeptides[4,11]
56295.29294.30C14H18N2O5glutamyl-phenylalaninepeptides[4,11]
57296.28295.29C13H17N3O5tyrosylglycylglycinepeptides[4,11]
58297.49296.50C20H40Ophytol terpenoids[57]
59303.24302.23C15H10O7quercetin flavonoids[3,59]
60306.39305.40C12H23N3O4Sarginylmethioninepeptides[4,11]
61308.36307.37C11H21N3O5Sglycyl-threonyl-methioninepeptides[4,11]
62309.51308.50C20H36O2eicosadienoic acidfatty acids[54]
63312.33311.34C12H21N7O3histidinyl-argininepeptides[4,11]
64320.41319.42C13H25N3O4Sleucyl-methionyl-glycinepeptides[4,11]
65323.36321.37C15H23N5O3arginylphenylalaninepeptides[4,11]
66333.34332.35C16H20N4O4glycyl-tryptophanyl-alaninepeptides[4,11]
67335.41334.40C11H22N6O4Sglycyl-arginyl-cysteinepeptides[4,11]
68338.37337.37C16H23N3O5threonyl-phenylalanyl-alaninepeptides[4,11]
69341.59340.60C22H44O2behenic acidfatty acids[47]
70343.39342.40C20H24NO4+magnoflorinealkaloids[51]
71349.41348.42C12H24N6O4Sarginyl-alanyl-cysteinepeptides[4,11]
72355.49354.50C24H34O2bufadienolideterpenoids[60]
73357.34356.33C13H20N6O6asparaginyl-seryl-histidinepeptides[4,11]
74361.29360.30C18H16O8rosmarinic acidphenolic acids[10]
75365.41364.40C18H28N4O4lysyl-phenylalanyl-alaninepeptides[4,11]
76368.51367.50C17H25N3O4Salanyl-methionyl-phenylalaninepeptides[4,11]
77370.41369.40C16H23N3O5Smethionyl-tyrosyl-glycinepeptides[4,11]
78374.41373.40C18H23N5O4phenylalanyl-histidyl-alaninepeptides[4,11]
79381.69380.70C27H56heptacosanehydrocarbons[47]
80383.43382.42C16H26N6O5leucyl-asparaginyl-histidinepeptides[4,11]
81384.37383.36C14H21N7O6histidyl-asparaginyl-asparaginepeptides[4,11]
82367.71386.70C27H46Ofurostanterpenoids[2,3]
83393.51392.50C18H24N4O4Stryptophanyl-methionyl-glycinepeptides[4,11]
84398.49397.50C18H27N3O5Sthreonyl-phenylalanyl-methioninepeptides[4,11]
85401.59400.60C27H44O2spirostansaponins[2,3]
86406.51405.50C15H31N7O4Slysyl-arginyl-cysteinepeptides[4,11]
87408.50407.50C20H29N3O6glutamyl-isoleucyl-phenylalaninepeptides[4,11]
88411.69410.70C30H50squaleneterpenoids[47]
89413.70412.70C29H48Ostigmasterolsterols[47]
90415.69414.70C29H50Ositosterolsterols[47]
91417.51416.50C24H32O6hellebrigeninterpenoids[2,3,61]
92434.49433.50C15H31N9O4Sarginyl-arginyl-cysteinepeptides[4,11]
93450.51449.50C15H27N7O7Sarginyl-glycyl-aspartyl-cysteinepeptides[4,11]
94453.49452.50C21H36N6O5alanyl-leucyl-leucyl-histidinepeptides[4,11]
95463.61462.60C27H42O6shidasteronephytoecdysteroids[62,63]
96465.39464.40C21H20O12hyperosideflavonoids[9]
97469.51468.50C24H32N6O4phenylalanyl-arginyl-phenylalaninepeptides[4,11]
98481.59480.60C27H44O7ecdysteroneterpenoids[2,3,9,64]
99482.61481.60C16H31N7O6S2arginyl-threonyl-cysteinyl-cysteinepeptides[4,11]
100494.51493.50C21H31N7O7arginyl-glycyl-aspartyl-phenylalaninepeptides[4,11]
101521.59520.60C24H40N8O5H-Lys-Ala-Phe-Arg-OHpeptides[4,11]
102563.59562.60C30H42O10degluco-hellebrinterpenoids[60,64]
103572.69571.70C27H37N7O5Sarginyl-phenylalanyl-phenylalanyl-cysteinepeptides[4,11]
104611.49610.50C27H30O16rutinflavonoids[3,9]
105725.81724.80C36H52O15hellebrinterpenoids[2,3]
Table 3. The characteristic absorption bands attributed to the secondary metabolites identified in Helleborus purpurasces.
Table 3. The characteristic absorption bands attributed to the secondary metabolites identified in Helleborus purpurasces.
Secondary
Metabolites
Wavenumber (cm−1)Ref.
bufadienolides3560, 3035, 1740–1760, 1718–1736[110,111]
butenolide 1770, 1605, 830[83,112]
phytoecdysteroids3393, 3446, 1652 [113,114]
peptides3300, 3100, 1695–1610, 1575–1480,1320–1220, 800–640, 765–625, 605–535[115,116]
terpenoids2939, 1740, 1651, 810[117]
fatty acids3020–3010, 2924–2915, 2855–2847, 2800–2900, 1746, 1710, 1250, 720[118,119]
flavonoids4000–3125, 3140–3000, 1670–1620, 1650–1600, 1600–1500, 1450–1490[120,121]
phenolic acids1800–1650, 1734, 1720, 1627, 1522, 1440, 1410, 1420–1300, 1367, 1315, 1255, 1170–1100[117]
phytosterols3427, 2937, 1466, 1383, 1192, 1063, 740.5[122]
carbohydrates3524, 3324, 2943, 2917, 2874, 1129, 1089, 1048, 1016, 991, 782, 671, 602[123]
amino acids3400; 3330–3130; 2530–2760; 2130; 1724–1754, 1687, 1675, 1663, 1652, 1644, 1632, 1621, 1611, 1500–1600[124]
alkaloids3362, 1645.9, 1515.7, 1283, 1250.5 [125]
nucleoside3351, 3104, 2925, 2800, 1670, 1470, 1396, 1269, 1209, 1137. 1095, 1053, 980, 906, 830, 766, 572, 451[126]
glycosides3401, 1711, 1450, 1073, 991, 914, 863.5; 769, 743, 684, 630, 432[127]
stilbenoids1605, 1583, 1380, 960[128,129]
Table 4. The antioxidant assay results for hellebore, chitosan, hellebore-AgNPs system and HC and HAgC phyto-nanocarriers.
Table 4. The antioxidant assay results for hellebore, chitosan, hellebore-AgNPs system and HC and HAgC phyto-nanocarriers.
Sample NameIC50 (mg/mL)Total Phenolic Content (mg GAE/mL)mg AAE/mLmmol Fe/g
hellebore3.86 ± 0.165.16 ± 0.1891.41 ± 0.01234.91 ± 0.024
chitosan17.21± 0.091.82 ± −1.2415.16 ± 0.0149.16 ± 0.021
HC phyto-nanocarrier4.17 ± 0.124.85 ± 0.14100.48 ± 0.01727.75 ± 0.013
hellebore-AgNPs system2.35 ± 0.159.41 ± 0.17156.87 ± 0.01948.08 ± 0.027
HAgC phyto-nanocarrier2.56 ± 0.148.93 ± 0.15186.98 ± 0.02141.34 ± 0.018
Table 5. Parameters of the encapsulation process for the newly prepared phyto-nanocarriers.
Table 5. Parameters of the encapsulation process for the newly prepared phyto-nanocarriers.
Sample NameEE (%)EC (%)EY (%)
CN--70.34 ± 0.23
HC82.45 ± 0. 2181.65 ± 0.3587.45 ± 0.35
HAgC91.05 ± 0.1892.18 ± 0.2895.34 ± 0.48
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Segneanu, A.-E.; Vlase, G.; Vlase, T.; Sicoe, C.A.; Ciocalteu, M.V.; Herea, D.D.; Ghirlea, O.-F.; Grozescu, I.; Nanescu, V. Wild-Grown Romanian Helleborus purpurascens Approach to Novel Chitosan Phyto-Nanocarriers—Metabolite Profile and Antioxidant Properties. Plants 2023, 12, 3479. https://doi.org/10.3390/plants12193479

AMA Style

Segneanu A-E, Vlase G, Vlase T, Sicoe CA, Ciocalteu MV, Herea DD, Ghirlea O-F, Grozescu I, Nanescu V. Wild-Grown Romanian Helleborus purpurascens Approach to Novel Chitosan Phyto-Nanocarriers—Metabolite Profile and Antioxidant Properties. Plants. 2023; 12(19):3479. https://doi.org/10.3390/plants12193479

Chicago/Turabian Style

Segneanu, Adina-Elena, Gabriela Vlase, Titus Vlase, Crina Andreea Sicoe, Maria Viorica Ciocalteu, Dumitru Daniel Herea, Ovidiu-Florin Ghirlea, Ioan Grozescu, and Valentin Nanescu. 2023. "Wild-Grown Romanian Helleborus purpurascens Approach to Novel Chitosan Phyto-Nanocarriers—Metabolite Profile and Antioxidant Properties" Plants 12, no. 19: 3479. https://doi.org/10.3390/plants12193479

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop