Next Article in Journal
DNA Methylation Machinery in Gastropod Mollusks
Previous Article in Journal
Aflatoxins in Wheat Grains: Detection and Detoxification through Chemical, Physical, and Biological Means
Previous Article in Special Issue
Plant Production Protocols from Seeds of Threatened Atropa baetica and Widespread A. belladonna, Both Rich in Alkaloids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Anti-Cholinesterase Potential of Fifteen Different Species of Narcissus L. (Amaryllidaceae) Collected in Spain

by
Luciana R. Tallini
1,2,*,
Giulia Manfredini
1,3,
María Lenny Rodríguez-Escobar
1,
Segundo Ríos
4,
Vanessa Martínez-Francés
5,
Gabriela E. Feresin
6,
Warley de Souza Borges
7,
Jaume Bastida
1,
Francesc Viladomat
1 and
Laura Torras-Claveria
1,*
1
Departament de Biologia, Sanitat i Medi Ambient, Facultat de Farmàcia i Ciències de l’Alimentació, Universitat de Barcelona, Av. Joan XXIII 27-31, 08028 Barcelona, Spain
2
Faculdade de Farmácia, Universidade Federal do Rio Grande do Sul, Av. Ipiranga 2752, Porto Alegre 90610-000, RS, Brazil
3
Dipartimento di Scienze della Vita, Universita’ Degli Studi di Modena e Reggio Emilia, 41121 Modena, Italy
4
Estación Biológica Torretes-Jardín Botánico de la UA, Universidad de Alicante, Crtra. Sant Vicent del Raspeig, s/n, 03690 Alicante, Spain
5
Department Biología Aplicada, Area de Botánica, Universidad Miguel Hernández, Av. Universidad, s/n, 03202 Elche, Spain
6
Instituto de Biotecnología, Facultad de Ingeniería, Universidad Nacional de San Juan, Av. Libertador General San Martin 1109 Oeste, San Juan 5400, Argentina
7
Laboratory of Natural Products, Department of Chemistry, Federal University of Espirito Santo, Vitória 29075-910, ES, Brazil
*
Authors to whom correspondence should be addressed.
Life 2024, 14(4), 536; https://doi.org/10.3390/life14040536
Submission received: 1 March 2024 / Revised: 16 April 2024 / Accepted: 16 April 2024 / Published: 22 April 2024
(This article belongs to the Special Issue Alkaloids: Extraction, Analysis and Function Research)

Abstract

:
Narcissus L. is a renowned plant genus with a notable center of diversity and is primarily located in the Mediterranean region. These plants are widely recognized for their ornamental value, owing to the beauty of their flowers; nonetheless, they also hold pharmacological importance. In Europe, pharmaceutical companies usually use the bulbs of Narcissus pseudonarcissus cv. Carlton to extract galanthamine, which is one of the few medications approved by the FDA for the palliative treatment of mild-to-moderate symptoms of Alzheimer’s disease. The purpose of this study was to evaluate the potential of these plants in Alzheimer’s disease. The alkaloid extract from the leaves of different species of Narcissus was obtained by an acid-base extraction work-up -procedure. The biological potential of the samples was carried out by evaluating their ability to inhibit the enzymes acetyl- and butyrylcholinesterase (AChE and BuChE, respectively). The species N. jacetanus exhibited the best inhibition values against AChE, with IC50 values of 0.75 ± 0.03 µg·mL−1, while N. jonquilla was the most active against BuChE, with IC50 values of 11.72 ± 1.15 µg·mL−1.

1. Introduction

Narcissus L. is a well-known plant genus that belongs to the Amaryllidaceae family, specifically within the Amaryllidoideae subfamily [1]. These plants, commonly referred to as daffodils, are highly favored in gardens and serve as a significant commercial crop [2]. This genus encompasses around 100 wild species, primarily concentrated in southwestern Europe, with a significant center of diversity in the Iberian Peninsula—where 90% of all species are present, especially in Spain, and extend throughout North Africa [2,3,4].
The native habitats of the Narcissus species exhibit remarkable diversity, encompassing a broad spectrum of landscapes that span from lowland to mountainous regions. This genus includes a rich tapestry of environments, including verdant grasslands, rugged scrublands, serene woodlands, meandering riverbanks, and secluded rocky crevices [2]. The presentation of blooms exhibited by the majority of these species is a characteristic feature observed during late winter and spring. However, there are a limited number of species that deviate from this pattern and blossom during the autumnal season [5,6].
The hybridization of some of these species has led to the development of commercial Narcissus cultivars, which, in most instances, are larger and more robust than their wild ancestors [2,7]. This genus stands out as the main choice among commercial bulb planters, showcasing remarkable importance in various horticultural and commercial contexts [8] in Western Europe, being cultivated since the 16th century in the Netherlands. Currently, the United Kingdom, the Netherlands, and the United States are the main producing countries of Narcissus bulbs. Until the late nineties, Narcissus bulb production was oriented only towards ornamental use, but since 1999, the purpose of extracting galanthamine from these bulbs started to play an important and growing role [9]. Therefore, although the cultivation of Narcissus for ornamental purposes has had a long tradition, leading to extensive knowledge of its growing and breeding, as well as the production of large stocks of raw materials, only a few of the numerous cultivars developed could be cultivated in significant amounts and quality for chemical extraction [2]. This question has been successfully solved by in vitro cultivation [10].
The medicinal properties of the Narcissus species have been documented in the literature for many years. Hippocrates of Kos (460-370 BCE), the father of modern medicine, advocated the use of oil from the Narcissus species to alleviate symptoms that today would be recognized as cancer [11]. Dioscorides, in the 1st century, mentions some of their medicinal properties due to the presence of alkaloids, as well as the sedative and narcotic effects of their aroma [12]. In recent decades, the genus Narcissus has provided a range of useful or potentially valuable compounds, of which galanthamine has been extensively studied. Galanthamine hydrobromide is one of the few medicines available used for the palliative treatment of symptoms derived from Alzheimer’s disease, owing to its ability to inhibit the enzyme acetylcholinesterase. This alkaloid—specifically Amaryllidaceae alkaloid—was first isolated in 1952 from the species Galanthus woronowii and received the FDA’s approval in 2001 [13,14]. Pharmaceutical companies concentrated their search for galanthamine on the Amaryllidoideae subfamily, as it is the exclusive natural source of this metabolite [2,15]. For Narcissus cultivars, different studies have been carried out on variations in galanthamine content in bulbs, depending on their geographical origin or changes due to the addition of fungicides [16,17]. Torras-Claveria and co-workers [18] investigated the galanthamine content and bioactivity of more than one hundred Narcissus cultivars, highlighting those with higher galanthamine content and higher acetylcholinesterase inhibitory activity. In Central and Western Europe, the extraction of galanthamine is mainly derived from Narcissus pseudonarcissus cv. Carlton. Meanwhile, in Eastern Europe, Leucojum aestivum, known as snowflakes, serves as a key source, and it has been used in the past, although its collection for this purpose is currently not possible, as this species is in decline. In China, the red-tubed lily, Lycoris radiata, contributes to galanthamine production as well [9].
Amaryllidaceae alkaloids are unique structures originating from the Amaryllidoideae subfamily, covering a diverse group of compounds that exhibit interesting biological properties. Currently, there are more than 650 structures of Amaryllidaceae alkaloids reported in the literature [19]. While significant progress has been made in recent years, the comprehensive exploration of the diversity of structures, reactions, and genes within the Amaryllidaceae family remains ongoing. All of these metabolites originate from the aromatic amino acids L-phenylalanine (L-Phe) and L-tyrosine (L-Tyr), which, through distinct enzymatic reactions, ultimately lead to the formation of norbelladine [20,21]. The enzyme norbelladine 4′-O-methyltransferase catalyzes the production of 4′-O-methylnorbelladine, which is considered the main common precursor to Amaryllidaceae alkaloids [21]. The oxidative phenolic coupling of 4′-O-methylnorbelladine leads to the three main skeleton types that form the bases for the extensive structural diversity of Amaryllidaceae alkaloids. The ortho-para’ coupling of 4′-O-methylnorbelladine results in the formation of the lycorine- and homolycorine-type skeletons, para-para’ originates the crinine-, haemanthamine-, tazettine-, narciclasine- and montanine-type structures, and para-ortho’ gives the galanthamine-type skeleton [22]. Since the initial isolation of the alkaloid lycorine from Narcissus pseudonarcissus in 1877, significant advancements have been achieved in the study of Amaryllidaceae plants. Nevertheless, they continue to represent a relatively underexplored source of phytochemicals [3]. Generally, within each plant, a variety of related alkaloids is present, consisting of a few dominant metabolites and multiple minor compounds with varying substituent positions [3].
The historical bond between Amaryllidaceae and medicine stands as a testament to the enduring allure and profound significance of these remarkable compounds within the realm of human health and well-being. In the present day, continued research into Amaryllidaceae alkaloids promises to unveil further insights into their potential applications and therapeutic benefits. In this way, the aim of this study was to describe the potential different wild species of Narcissus L. collected in Spain against Alzheimer’s disease through cholinesterase inhibition assays.

2. Materials and Methods

2.1. Plant Material

Fifteen different species of Narcissus L. were provided from the Iberian Narcissus Collection of the Torretes Biological Research Station—Botanical Garden of the UA, in Alicante, Spain. All the samples were collected during the flowering season, April 2023. The origins of the species are different localities of the Iberian Peninsula, as described in Table 1 and Figure 1. All the species were authenticated by botanists Dr. Segundo Ríos and Dr. Vanessa Martínez-Francés. Most wild daffodils are distributed in small, scattered populations, in very specific and fragile microhabitats such as wetlands, rocky areas, and deciduous forests [2,23,24,25,26]. These populations with very small numbers (less than 1000 individuals) have a small surface area (often less than 1 ha), making them extremely vulnerable to changes due to natural (climate change, predation, etc.) or anthropogenic causes (land use, grazing, collecting, etc.), which endanger their survival [27,28].
The ecology of many species from the Pseudonarcissus DC. section, as N. asturiensis, N, jacetanus, N. minor, and N. vasconicus are perennial grasslands of external fringe woodlands. Other species, such as N. bujei and N. genesii-lopezii, are present in perennial mesophytic grasslands, and N. confusus is present in the secondary mesoforests of deciduous oaks (Table 1).
More ecological diversity is observed in the Nevadensis Zonn. section. N. nevadensis grows in the secondary mesoforests of deciduous oaks and the perennial grasslands of external fringe woodlands. The existing populations of N. alcaracensis develop on lacustrine and helophyte communities dominated by Carex hispida, and N. yepesii is present in meadows and chionophilous grassland vegetation (Table 1).
N. jonquilla, from section Jonquillae DC., grows in lacustrine and riverine helophyte graminoid communities, while N. assoanus grows in perennial succulent grassland in rocky, open communities. The representative of the Bulbocodium section, N. hedraeanthus, is present in meadows and chionophilous grassland vegetation. From the Ganymedes section, N. pallidulus grows in perennial mesophytic grasslands and bare sandy soils, and N. tazetta of section Tazettae DC. has been collected from gardens where it was cultivated (Table 1). Although ornamental domestication of the genus Narcissus by the United Kingdom and Holland began in the 16th century [7,24], today there is still continuous extraction of wild bulbs in Spain, Portugal, and North Africa to satisfy the global private collecting market, to obtain new disease-resistant genes in commercial bulbs, and to satisfy the demand for galanthamine by pharmaceutical companies. Despite the scarce data on the populations and conservation status of most of the species, the information collected for some of them has allowed for their inclusion in the IUCN Red List, considered a critical indicator of the health of biodiversity. Two species of section Nevadensis Zonn. Analyzed in this work, N. alcaracensis and N. nevadensis, and another of section Pseudonarcissus DC., N. bujei, have been assessed as Endangered (EN) in the IUCN list [30,31,32]. Another species of this first mentioned section, N. yepesii, has been evaluated and classified as Vulnerable (VU) [33].
N. asturiensis and N. hedraeanthus are listed under Least Concern (LC), with no other major conservation measures recommended [34,35]. N. pallidulus, although not mentioned, is included in N. triandrus, which is assessed as Least Concern (LC), considering its wide distribution, with large and stable populations; it is unlikely that existing threats will cause them to seriously decrease in the near future [36]. However, it should be added that the different species of Section Ganymedes (Salisb.) Schultes f. studied in [37] have not been taken into consideration in the preparation of this manuscript, and different problems are present regarding each of them, both with respect to population size and natural and anthropogenic issues, requiring, therefore, a new review.
Only the leaves were collected for biological assays. The use of leaves instead of bulbs for biological activity studies ensures the maintenance of the Iberian Daffodil Collection, allowing for its reproduction. Moreover, depending on the phenological stage of the plant, the alkaloid content and, therefore, the biological activity, may vary from bulb tissue to leaf tissue, and sometimes leaves may contain more alkaloids and be more active than bulbs [18,38].

2.2. Alkaloid Extracts Preparation

For the evaluation of the bioactivity, purified alkaloid extracts were performed. The species were meticulously processed as follows: First, they were cut into pieces and dried at a controlled temperature of 40 °C. Afterward, the dried plant material was finely powdered using a rotary blade mill, specifically a stainless-steel grinder (Taurus, Oliana, Spain). To extract the desired compounds, 1 g of the resulting powder was subjected to a maceration process with methanol at 25° for three days. During this period, the solvent was replaced each day using 3 × 50 mL aliquots and submitted to an ultrasonic bath (20 min, 4 daily intervals). Following methanolic extraction, the mixture was meticulously strained, and the solvent was carefully evaporated under reduced pressure, leaving behind crude extracts. These crude extracts were subsequently acidified using 30 mL of a 2% (v/v) sulfuric acid solution, lowering the pH to 2. After acidification, an ethyl acetate treatment (using 3 × 50 mL) was employed to eliminate neutral materials. The next step involved adjusting the pH of the remaining aqueous solution to a range of 9–10, achieved by the addition of a 25% (v/v) ammonium hydroxide solution. The volume of the ammonium hydroxide solution employed was the quantity required to achieve the aforementioned pH. Finally, the alkaloids were extracted with ethyl acetate (using 3 × 50 mL). After evaporation of the solvent, the dried alkaloid extract (AE) was obtained.

2.3. Cholinesterase Inhibitory Activity

Levels of ACh and BuCh decrease in patients with Alzheimer’s disease. Inhibition of AChE and BuChE has been shown to maintain ACh and BuCh levels in the brain, reducing disease progression [13,39]. The activity of Narcissus extracts inhibiting these enzymes can be assessed with colorimetric AChE and BuCHE inhibition assays, based on the formation of thiobenzoate anion (yellow) following the reaction of thiocholine and 2,2′-dinitro-5′-dinitro-5′-dithiobenzoic acid when the enzymes are active.
The inhibition activity of AChE and BuChE was assessed using the method described by Ellman and co-workers [40], with certain adaptations, as outlined by López and co-workers [41]. Enzyme stock solutions containing 518U of AChE from Electrophorus electricus (Merck, Darmstadt, Germany) and BuChE from equine serum (Merck, Darmstadt, Germany) were prepared and subsequently stored at −20 °C. 5,5-Dithiobis (2-nitrobenzoic acid) (DTNB), S-butyrylthiocholine iodide (BTCI), and acetylthiocholine iodide (ATCI) were supplied by Merck (Darmstadt, Germany). The reaction was initiated by mixing 50 µL of AChE or BuChE (both enzymes were employed at a concentration of 6.24 U in phosphate buffer (8 mM K2HPO4, 2.3 mM NaH2PO4, 0.15 NaCl, pH 7.5)) and 50 µL of the alkaloid extract dissolved in the same buffer solution. Immediately, the plates were incubated for 30 min at 25°. Finally, 100 µL of the substrate solution (comprising 0.1 M Na2HPO4, 0.5 M DTNB, and 0.6 mM ATCI or 0.24 mM BTCI in Millipore water, adjusted to pH 7.5) were introduced. Ten minutes later, the absorbance was measured at 405 nm using a Labsystem microplate reader (Helsinki, Finland). The activity of the enzymes was estimated as percentages in relation to a control (which consisted of a buffer without any inhibitor). Galanthamine served as a positive control. The galanthamine concentrations used for this positive control were the following: 0.1, 0.2, 0.3, 0.4, 0.5, 1.0, and 2.0 µg·mL−1 for AChE; and 1, 4, 6, 8, 10, 12, and 15 µg·mL−1 for BuChE. The calibration curves of samples A (0.1, 1.0, 3.0, 5.0, 7.0, 10, and 15 µg·mL−1), B, E and L (0.05, 0.1, 0.5, 1.0, 3.0, 5.0, and 10 µg·mL−1), C (0.1, 0.5, 1.0, 2.0, 3.0, 4.0, and 5.0 µg·mL−1), D and G (1.0, 2.5, 5.0, 7.5, 10, 15, and 25 µg·mL−1), F (5.0, 10, 15, 25, 50, 75, and 100 µg·mL−1), H (5.0, 10, 15, 20, 25, 50, and 75 µg·mL−1), I (1.0, 2.0, 3.0, 4.0, 5.0, 7.0, and 10.0 µg·mL−1), K (1.0, 10, 25, 50, 75, 100, and 125 µg·mL−1), M and O (0.1, 0.5, 1.0, 3.0, 5.0, 7.0, and 10.0 µg·mL−1), and N (5.0, 7.5, 10, 15, 25, 50, and 75 µg·mL−1) were applied to obtain the IC50 values against the AChE enzyme. To obtain the IC50 values against BuChE, the following curves were used: C and K (10, 25, 50, 75, 100, 125, and 150 µg·mL−1), E (1, 5, 10, 15, 25, 50, and 75 µg·mL−1), G, H, I, L, M and O (1, 10, 35, 50, 75, 100, and 125 µg·mL−1), and N (10, 25, 50, 100, 125, and 150 µg·mL−1). Analysis of the cholinesterase’s inhibitory data was conducted using Prism 10 software.

2.4. Statistical Evaluation

The inhibition of the cholinesterase activity of the Narcissus species was assessed using three separate assays. The PRISM software was used to analyze the results. The data are presented as the average ± standard deviation (SD). The significance of the results is represented versus the control (Gal), and it is indicated as follows: **** p < 0.0001, *** p < 0.001, ** p < 0.01, and ns (not significant). A one-way ANOVA test was performed following Dunnet’s multiple comparison test, comparing the differences with respect to the outcome of galanthamine with both AChE and BuChE.

3. Results and Discussion

The alkaloid extracts from all the Narcissus species collected in Spain were obtained through acid-base extraction, as described in Section 2.2. The yield of each species is available in Table 2. The average income value was 1.32%, with the highest value at 6.14% (sample N) and the lowest at 0.47% (sample B), which represent the species N. yepesii and N. jacetanus, respectively.
In vitro assessments were carried out to examine the inhibitory potential of the fifteen alkaloid extracts from Narcissus leaves against the enzymes AChE and BuChE. Among these plant samples, fourteen demonstrated activity against AChE, while only nine exhibited activity against BuChE (see Figure 2 and Figure 3). As was expected, all the species evaluated herein presented better results against AChE than BuChE. The species N. jacetanus (sample B) showed the best inhibition values against AchE, with IC50 values of 0.75 ± 0.03 µg·mL−1, while N. jonquilla (sample L) was the most active against BuChE, with IC50 values of 11.72 ± 1.15 µg·mL−1. The species N. jacetanus and N. jonquilla are illustrated in Figure 4 and Figure 5, respectively. The samples N. assoanus, N. minor, N. confuses, and N. jonquilla also presented noteworthy activity against AChE, with values of IC50 of 0.99 ± 0.06, 0.81 ± 0.10, 1.04 ± 0.07, and 1.88 ± 0.05 µg·mL−1, respectively. Regarding BuChE inhibition activity, N. confusus and N. genesii-lopezii also presented high and relevant activity, with IC50 of 12.83 ± 0.87 and 11.98 ± 0.93 µg·mL−1, respectively. Thus, N. jonquilla stands out for being the species of Narcissus with the highest global cholinesterase activity inhibition, followed by N. confusus and N. jonquilla.
Considering their biological potential, it is important to account for the potential synergistic interactions among Amaryllidaceae alkaloids in plant extracts [42]. These interactions, previously reported to contribute to acetylcholinesterase (AChE) inhibition, should be taken into consideration when explaining the anticholinesterase potential of certain Amaryllidaceae species [42].
Among other samples, Havlasová and co-authors [43] evaluated the inhibitory potential of N. jonquilla var. henriquesii against AChE and obtained IC50 values of 32.6 ± 4.3 µg·mL−1, while for galanthamine, it was 1.7 ± 0.06 µg·mL−1 [43]. As documented in the literature, the galanthamine-type skeleton, specifically the alkaloids galanthamine and sanguinine, are commonly active against AChE and BuChE [41]. In the literature, a great amount of galanthamine-type structures is reported, such as galanthamine, lycoramine, and narwedine in the species N. jonquilla, representing about 65% of its alkaloid profiling [4,44]. Furthermore, haemanthamine, tazettine, jonquailine, and narciclasine- and lycorine-type structures were also described in this plant species [4,44,45]. Furthermore, the majority of the reported alkaloids of N. jonquilla have been studied in terms of molecular modelling in front of AChE and BuChE, and data reported in the literature support the fact that they could be responsible for the especially high activity of this species inhibiting BuChE and AChE. Galanthamine has been reported to have good docking scores for BuCHE and ACHE [46,47,48,49,50]. Lycoramine has been reported to have molecular docking values of −8.84, −9.08, −8.87, −8.64, and −8.41 kcal·mol−1 for the human acetylcholinesterase X-ray crystals 4EY5, 4EY6, 4Ey7, 4M0E, and 4M0F, respectively, while values for galanthamine were −8.59, −8.75, −9.83, −7.90, and −8.74 kcal·mol−1, according to Tallini et al., 2022 [50]. Narwedine has been reported to have better scores (−9.15, −9.70, −10.41, −8.69, and −9.72 kcal·mol−1) than lycoramine and galanthamine [50].
The AChE inhibition activity of the species N. assoanus, N. jacetanus, N. bujei, N. vasconicus, and N. pallidulus have also been described by López and co-authors. However, no activity has been described for the latter two species [41]. Previous publications indicate that the alkaloid composition of the species N. assoanus includes assoanine, oxoassoanine, pseudolycorine, 1-O-acetylpseudolycorine, and 2-O-acetylpseudolycorine [51,52]. The species N. jacetanus has been previously documented to contain the compounds assoanine, oxoassoanine, pseudolycorine, and lycorine [53]. According to the literature, the compounds assoanine and oxoassoanine, which belong to a lycorine-type skeleton, are active against AChE, with IC50 values of 3.87 ± 0.24 and 47.21 ± 1.13 µM, respectively, while pseudolycorine exhibits poor activity [41]. According to prior publications, the alkaloid lycorine exhibits a very weak in vitro activity against AChE and BuChE, with IC50 values higher than 200 µM [41,54]. However, lycorine showed good energy values concerning docking studies with enzymes, with scores of 60.9444 for AChE (1EVE) and 52.7924 for BuChE (homology model) (scores corresponding to galanthamine were 65.4656 and 53.0089) [46].
Cortes et al., 2018 [47] obtained scores of −8.99, −8.87, and −8.94 kcal·mol−1 (energy of protein ligand interaction between lycorine and 1DX6 and 4EY7 from AChE, and 4BDS from BuCHE, respectively), while values corresponding to galanthamine were −10.10, −10.20, and −8.23 kcal·mol−1, respectively. León et al., 2021 [48] reported binding values for lycorine to AChE (1DX6) and BuChE (4BDS) of −8.82 and −8.94 kcal·mol−1, respectively, while values corresponding to galanthamine were −10.10 and −8.23 kcal·mol−1. Rojas-Vera et al., 2021 [49], estimated lycorine binding values of −8.89 and −8.38 kcal·mol−1 for AChE (4EY7 and 5HF5) and −7.74 kcal·mol−1 for BuChE (1P0I), while values corresponding to galanthamine were −9.92, −8.97, and −7.40 kcal·mol−1, respectively. Tallini et al., 2018 [54] reported lycorine binding values of −8.82 and −8.94 kcal·mol−1 for AchE (1DX6) and BuChE (4BDS), respectively, while scores corresponding to galanthamine were −9.55 and −8.23 kcal·mol−1, respectively.
Regarding the molecular modelling of assoanine and oxoassoanine vs. AchE, they did not show great interaction with AchE in terms of electrophilicity, with values of 0.0026 and 0.0036 eV, respectively, (galanthamine electrophilicity values corresponded to 8.5725 eV). In relation to the molecular electrostatic potential (MEP) of assoanine and oxoassoanine, values of negative (−0.08705 and −0.06995 u.a.) and positive (0.03474 and 0.05755 u.a.) regions (which represent the probability to conduct and suffer, respectively, nucleophilic attacks) are not as high as other Amaryllidaceae alkaloids with AChE inhibition properties such as galanthamine (−0.06995 and 0.05755 u.a. negative and positive regions, respectively), hydroxygalanthamine (−0.06822 and 0.06648 u.a.), or sanguinine (−0.07504 and 0.06400 u.a.) [55]. However, it must be considered that no other molecular modeling studies have been found in the literature regarding these two compounds; therefore, more studies should be performed to obtain a global idea of the characterization of the interaction of these compounds with AChE and BuChE.
The alkaloids homolycorine, lycorenine, haemanthamine, 8-O-demethylhomolycorine, O-methyllycorenine, crinamine, masonine, tazettine, O-methyloduline, 11-O-acetylhaemanthamine, and bujeine have been documented as constituents of the species N. bujei [56]. Additionally, four alkaloids have been reported in the species N. vasconicus, which are vasconine, lycorine, homolycorine, and 8-O-acetylhomolycorine [57].
As shown in Table 3, the alkaloid extract of N. pallidulus (sample J) was the only extract inactive against both cholinesterases. Previous studies have documented the presence of different structures from the Sceletium type in the section Ganymedes (Salisb.) Schultes f. [37,58]. This scaffold is the only group of alkaloids that is not exclusive to the monocotyledon subfamily Amaryllidoideae, being typical structures of the genus Sceletium that belongs to the dicotyledonous family Aizoaceae [4,37,59]. According to the literature, the following alkaloids have been described in this plant species: haemanthamine, lycorine, lycorenine, homolycorine, galanthamine, tazettine, mesembrine, mesembrenol, mesembrenone, 2-oxomesembrenone, 7,7a-dehydromesembrenone, 2-oxoepimesembrenol, 6-epimesembrenol, 6-epimesembranol, and 4′-O-demethylmesembrenone [58,60].
Considering the inhibitory potential of the samples against both enzymes (Figure 2 and Figure 3), the species N. confusus, N. jonquilla, and N. genesii-lopezii showed remarkable results, with IC50 values of 1.04 ± 0.07, 1.88 ± 0.05, and 5.28 ± 0.64 µg·mL−1, respectively, against AChE, and 12.83 ± 0.87, 11.72 ± 1.15, and 11.98 ± 0.93 µg·mL−1, respectively, against BuChE. López and co-authors [41] evaluated the activity of the alkaloid extract of twenty-six species of Narcissus and reported that the best results were obtained for N. confusus, which presented a high amount of galanthamine [41]. The literature documents the existence of various structures within the alkaloid composition of the species N. confusus, which are ismine, 11,12-dehydroanhydrolycorine, galanthamine, 3-O-acetylgalanthamine, N-demethylgalanthamine, N-formylgalanthamine, narwedine, 8-O-methylleucotamine, haemanthamine, haemanthidine, tazettine, pretazettine, 6-O-methylpretazettine, epimacronine, homolycorine, 8-O-demethylhomolycorine, 9-O-demethylhomolycorine, and also narciclasine-, and lycorine-type alkaloids [4,61,62,63]. As indicated in the literature, the chemical diversity of this species presents challenges to its viability as a consistent source of galanthamine on an industrial scale [15].
Various alkaloids have been documented in the alkaloid composition of N. tazetta, including galanthamine, sanguinine, narwedine, demethylmaritidine, anhydrolycorine, O-methylnorbelladine, pancratinine C, lycorine, 9-O-methylpseudolycorine, pseudolycorine, 1-O-acetyl-3-O-methylnarcissidine, 11, hydroxygalanthine, narcissidine, 9-O-demethyl-2alfa-hydroxyhomolycorine, ismine, tazettine, lycorenine, lycorine, masonine, 3-epimacronine, 1,2-dihydroclidanthine, assoanine, hippeastrine, and 4,5-ethylene-8,9-dimethoxy-6-phenanthridone [64,65,66]. Karakoyun and colleagues detailed an analysis of the anticholinesterase potential of alkaloids isolated from the species N. tazetta [65]. Their findings revealed 11-hydroxygalanthamine and narcissidine as important compounds with inhibitory effects against AChE, with IC50 values of 0.67 and 1.85 µM, respectively, while galanthamine showed IC50 values of 0.14 µM [65].
Viladomat and co-authors [67] identified haemanthamine, haemanthidine, tazettine, 3-epimacronine, ismine, and risperidona in the alkaloid composition of N. asturiensis [67]. The alkaloid type composition of N. alcaracensis, N. yepesii, N. genesii-lopezii, N. hedraeanthus, and N. jonquilla was determined by [4]. The alkaloid composition of N. alcaracensis and N. yepesii is dominated by alkaloid lycorine-type compounds. These species also contain haemanthamine- and homolycorine-type compounds. Pancracine- and galanthamine-type alkaloids have also been reported in N. yepesii [4]. N. genesi-lopezii composition is dominated by homolycorine-type alkaloids, although the presence of galanthamine-, haemanthamine-, and lycorine-type alkaloids is also reported. The alkaloid composition of N. hedraeanthus is dominated by haemanthamine-type alkaloids. However, this species also contains narciclasine-, galanthamine-, tazettine-, and lycorine-type alkaloids [4].
No information about the alkaloid profile of the species N. minor and N. nevadensis has been found in the literature. Regarding the alkaloid composition and anti-cholinesterase activity of certain species not listed in Table 3, Lisa-Molina and co-workers [6] evaluated the alkaloid profiling of ten samples of Narcissus, totaling nine different species (N. obsoletus, N. deficiens, N. serotinus, N. cavanillesii, N. viridiflorus, N. elegans, N. papyraceus, N. bulbocodium, N. blancoi) collected in Spain [6]. The authors detected thirty alkaloids among these samples, with two of them not being identified [6]. According to their results, the lycorine-type skeleton was the most diverse group detected among the samples, with these structures being observed in all of the species, except in N. viridiflorus and N.bulbocodium [6]. The authors also evaluated the AChE inhibitory activity of ten extracts of Narcissus, with seven of them able to act against this enzyme [6]. The best results were observed for the species N. obsoletus, which exhibited a substantial concentration of galanthamine in its alkaloid profile and showed IC50 values of 0.92 ± 0.06 µg·mL−1 [6]. Among all the species evaluated by them, N. obsoletus and N. blancoi were the only species that showed the presence of alkaloids from the galanthamine-type scaffold [6]. Furthermore, thirteen known and three new Amaryllidaceae alkaloids were isolated from Narcissus pseudonarcissus cv. Carlton by Mamun and co-authors [68]. Two of them were named carltonine A and carltonine B. The authors evaluated these new alkaloids in vitro against both cholinesterases, which showed a significant and selective inhibitory activity against BuChE, displaying IC50 values of 0.91 ± 0.02 and 0.031 ± 0.001 µM, correspondingly [68]. Twenty-one Amaryllidaceae alkaloids of various structural types and one new alkaloid, named narcimatuline, were obtained from the bulbs of Narcissus pseudonarcissus L. cv. Dutch Master by Hulcová and co-workers [69]. According to their results, narcimatuline showed interesting multipotent biological profiling, presenting properties against BuChE, prolyl oligopeptidase (POP), and glycogen synthase kinase-3β (GSK-3β) enzymes, with respective IC50 values of 5.9 ± 0.2, 29.2 ± 1.0, and 20.7 ± 2.4 µM [69].

4. Conclusions

In Europe, the species Narcissus plays a very important role for pharmaceutical companies as a source of galantamine. However, many species of this genus have not yet been studied. In this work, we describe the anticholinesterase potential of fifteen species of Narcissus collected in Spain, with seven of them being the first report of their biological activity (N. minor, N. asturiensis, N. hedraeanthus, N. alcaracensis, N. genesii-lopezii, N. yepesii and N. nevadensis). Some of these species showed interesting activity against AChE (N. jacetamus, N. assoanus, N. minor, N. confusus and N. jonquilla) and BuChE (N. jonquilla, N. confusus, N. genesii-lopezii). The species N. jonquilla and N. confusus show remarkable activity against both enzymes; therefore, they stand out as possible candidates for further studies and for determining their alkaloid composition. This study contributes to underlining the importance of Amaryllidaceae species as a source of important bioactive molecules.

Author Contributions

Conceptualization, L.T.-C., L.R.T., J.B., S.R. and V.M.-F.; methodology, L.R.T., G.M. and M.L.R.-E.; software, L.R.T., G.M. and M.L.R.-E.; validation, L.R.T. and L.T.-C.; formal analysis, L.R.T., G.M. and M.L.R.-E.; investigation, G.E.F., W.d.S.B., L.R.T. and L.T.-C.; resources, S.R., J.B. and V.M.-F.; writing—original draft preparation, L.R.T., S.R., V.M.-F. and L.T.-C.; writing—review and editing, L.R.T., S.R., V.M.-F. and L.T.-C.; visualization, J.B., L.R.T. and L.T.-C.; supervision, J.B., L.R.T., F.V. and L.T.-C.; project administration, J.B.; funding acquisition, J.B. All authors have read and agreed to the published version of the manuscript.

Funding

Programa Iberoamericano de Ciencia y Tecnologia para el Desarrollo (CYTED, 223RT0140).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors thank CCiTUB, from the University of Barcelona for technical support. L.R.T., F.V., J.B. and L.T-C., (Research Group Natural Products and Plant Biotechnology 2021-SGR-00693 from the University of Barcelona) thank CCiTUB for technical support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bremer, B.; Bremer, K.; Chase, M.W.; Fay, M.F.; Reveal, J.L.; Bailey, L.H.; Soltis, D.E.; Soltis, P.S.; Stevens, P.F.; Anderberg, A.A.; et al. An update of the Angiosperm Phylogeny Group classification for the orders and families of flowering plants: APG III. Bot. J. Linn. Soc. 2009, 161, 105–121. [Google Scholar] [CrossRef]
  2. Hanks, G.R. (Ed.) Narcissus and Daffodil: The Genus Narcissus; vol. 21 in the series: ‘‘Medicinal and Aromatic Plants—Industrial Profiles; Taylor & Francis: London, UK, 2002; pp. 1–428. [Google Scholar]
  3. Bastida, J.; Lavilla, R.; Viladomat, F. Chemical and Biological Aspects of Narcissus Alkaloids. In The Alkaloids: Chemistry and Biology; Cordell, G.A., Ed.; Elsevier: San Diego, CA, USA, 2006; Volume 63, pp. 87–179. [Google Scholar]
  4. Berkov, S.; Martínez-Francés, V.; Bastida, J.; Codina, C.; Ríos, S. Evolution of alkaloid biosynthesis in the genus Narcissus. Phytochemistry 2014, 99, 95–106. [Google Scholar] [CrossRef] [PubMed]
  5. Graham, S.W.; Barrett, S.C.H. Phylogenetic reconstruction of the evolution of stylar polymorphisms in Narcissus (Amaryllidaceae). Am. J. Bot. 2004, 91, 1007–1021. [Google Scholar] [CrossRef] [PubMed]
  6. Lisa-Molina, J.; Gómez-Murillo, P.; Arrelano-Martín, I.; Jiménez, C.; Rodríguez-Escobar, M.L.; Tallini, L.R.; Viladomat, F.; Torras-Claveria, L.; Bastida, J. Alkaloid profile in wild autumn-flowering daffodils and their acetylcholinesterase inhibitory activity. Molecules 2023, 28, 1239. [Google Scholar] [CrossRef] [PubMed]
  7. Rivera Núñez, D.; Obón de Castro, C.; Ríos Ruiz, S.; Alcaraz Ariza, F. The origin of cultivation and wild ancestors of daffodils (Narcissus subgenus Ajax) from an analysis of early illustrations. Sci. Hortic. 2003, 98, 307–330. [Google Scholar] [CrossRef]
  8. Marques, I.; Aguilar, J.F.; Martins-Louçao, M.A.; Moharrek, F.; Feliner, G.N. A three-genome five-gene comprehensive phylogeny of the bulbous genus Narcissus (Amaryllidaceae) challenges current classifications and reveals multiple hybridization events. Taxon 2017, 66, 832–854. [Google Scholar] [CrossRef]
  9. Berkov, S.; Georgieva, L.; Kondakova, V.; Atanassov, A.; Viladomat, F.; Bastida, J.; Codina, C. Plant sources of galanthamine: Phytochemical and biotechnological aspects. Biotechnol. Biotec. Eq. 2014, 23, 1170–1176. [Google Scholar] [CrossRef]
  10. Berkov, S.; Ivanov, I.; Georgiev, V.; Codina, C.; Pavlov, A. Galanthamine biosynthesis in plant in vitro systems. Eng. Life Sci. 2014, 14, 643–650. [Google Scholar] [CrossRef]
  11. van Goietsenoven, G.; Mathieu, V.; Lefranc, F.; Kornienko, A.; Evidente, A.; Kiss, R. Narciclasine as well as other Amaryllidaceae isocarbostyrils are Promising GTP-ase targeting agents against brain cancers. Med. Res. Rev. 2013, 33, 439–455. [Google Scholar] [CrossRef]
  12. Dioscorides, P. Acerca de la Materia Medicinal, y de los Venenos Mortiferos, Traducido de la Lengua Griega, en la Vulgar Castellana, & Illustrado con Claras y Substantiales Annotationes, y con las Figuras de Innumeras Plantas Exquisitas y Raras/Ror el Doctor Andres de Laguna, Medico de Iulio III. Pont. Maxi; Gast, M., Ed.; Real Jardín Botánico-CSIC: Salamanca, Spain, 1563; Available online: https://bibdigital.rjb.csic.es/records/item/13529-redirection (accessed on 7 February 2024).
  13. Maelicke, A.; Samochocki, M.; Jostock, R.; Fehrenbacher, A.; Ludwig, J.; Albuquerque, E.X.; Zerlin, M. Allosteric sensitization of nicotinic receptors by galantamine, a new treatment strategy for Alzheimer’s disease. Biol. Psychiatry 2001, 49, 279–288. [Google Scholar] [CrossRef]
  14. Heinrich, M.; Teoh, H.L. Galanthamine from snowdrop—The development of a modern drug against Alzheimer’s disease fromlocal Caucasian knowledge. J. Ethnopharmacol. 2004, 92, 147–162. [Google Scholar] [CrossRef]
  15. Berkov, S.; Georgieva, L.; Sidjimova, B.; Bastida, J. Evaluation of Hippeastrum papilio (Ravenna) Van Scheepen potencial as a new industrial source of galanthamine. Ind. Crop. Prod. 2022, 178, 114619. [Google Scholar] [CrossRef]
  16. Lubbe, A.; Pomahačová, B.; Choi, Y.H.; Verpoorte, R. Analysis of metabolic variation and galanthamine content in Narcissus bulbs by 1H NMR. Phytochem. Anal. 2010, 21, 66–72. [Google Scholar] [CrossRef]
  17. Lubbe, A.; Verpoorte, R.; Choi, Y.H. Effects of fungicides on galanthamine and metabolite profiles in Narcissus bulbs. Plant Physiol. Biochem. 2012, 58, 116–123. [Google Scholar] [CrossRef]
  18. Torras-Claveria, L.; Berkov, S.; Codina, C.; Viladomat, F.; Bastida, J. Daffodils as potential crops of galanthamine. Assessment of more than 100 ornamental varieties for their alkaloid content and acetylcholinesterase inhibitory activity. Ind. Crop. Prod. 2013, 43, 237–244. [Google Scholar] [CrossRef]
  19. Berkov, S.; Osorio, E.; Viladomat, F.; Bastida, J. Chemodiversity, chemotaxonomy and chemoecology of Amaryllidaceae alkaloids. In The Alkaloids: Chemistry and Biology; Knölker, H.-J., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; Volume 83, pp. 113–185. [Google Scholar] [CrossRef]
  20. Battersby, A.R.; Fales, H.M.; Wildman, W.C. Biosynthesis in the Amaryllidaceae. Tyrosine and norbelladine as precursors of haemanthamine. J. Am. Chem. Soc. 1961, 83, 4098–4099. [Google Scholar] [CrossRef]
  21. Kilgore, M.B.; Augustin, M.M.; Starks, C.M.; O’Neil-Johnson, M.; May, G.D.; Crow, J.A.; Kutchan, T.M. Cloning and characterization of a norbelladine 4′-O-methyltransferase involved in the biosynthesis of the Alzheimer’s drug galanthamine in Narcissus sp. aff. pseudonarcissus. PLoS ONE 2014, 9, e103223. [Google Scholar] [CrossRef] [PubMed]
  22. Barton, D.H.R.; Cohen, T. Some biogenic aspects of phenol oxidation. In Festschrift Arthur Stoll; Birkhauser, A.G., Ed.; Birkhauser: Basel, Switzerland, 1956; pp. 117–143. [Google Scholar]
  23. Rivera, D.; Ríos, S.; Obón, C.; Alcaraz, F. The image of daffodils in art and botanical illustration: Clues to the History of domestication and selection of Narcissus subgenus Ajax (Amaryllidaceae). In Floriculture, Ornamental and Plant Biotechnology; Teixeira, J., Ed.; Global Science Books: Ikenobe, Japan, 2006; Volume IV, Chapter 39; pp. 360–369. [Google Scholar]
  24. Rivera, D.; Ríos, S.; Alcaraz, F.; Obón, C. The biogeographical patterns of floral form in wild daffodils and their contribution to the cultivar groups of Narcissus L. Subgenus Ajax Spach (Amaryllidaceae). In Floriculture, Ornamental and Plant Biotechnology; Teixeira, J., Ed.; Global Science Books: Ikenobe, Japan, 2006; Volume IV, Chapter 32; pp. 300–313. [Google Scholar]
  25. Rivera, D.; Ríos, S.; Alcaraz, F.; Obón, C.; Verde, A.; Fajardo, J. The biogeography and habitat selection of Ajax daffodils (Narcissus L. subgenus Ajax Spach, Amaryllidaceae) in the Iberian Peninsula. In Frontiers of Vegetation Science—An Evolutionary Angle; Mucina, L., Kalwij, J., Smith, V.R., Chytrỳ, M., White, P.S., Ciliers, S.S., Pillar, V.D., Zobel, M., Sun, I.-F., Eds.; Keith Phillips Images: Somerset West, South Africa, 2008; pp. 152–153. [Google Scholar]
  26. Ríos, S.; Juan, J.; Martínez-Francés, V.; Laguna, E.; Rivera, D.; Alcaraz, F.; Verde, A.; Fajardo, J.; Carreño, E.; Casas, J.L.; et al. Endemic species of Narcissus in Central Spain: Biodiversity and conservation under grazing pressure by wild and domestic herbivorous. In Options Mediterranéenes Series A; Porqueddu, S., Ríos, S., Eds.; CIHEAM: Zaragoza, Spain, 2010; Volume 92, pp. 79–83. [Google Scholar]
  27. Blanca, G.; López, M.R.; Lorite, J.; Martínez, M.J.; Molero-Mesa, J.; Quintas, S.; Ruíz, M.; Varo, M.A.; Vidal, S. Flora amenazada y endémica de Sierra Nevada; Universidad de Granada: Granada, Spain, 2002; p. 407. [Google Scholar]
  28. Gutiérrez, L.; Lorite, J.; Ruiz, M.; López, M.; Donaire, F. Narcissus nevadensis Pugsley subsp. nevadensis. In Atlas y Libro Rojo de la Flora Vascular Amenazada de España; Bañares, A., Blanca, G., Güemes, J., Moreno, J.C., Ortiz, S., Eds.; Dirección General de Conservación de la Naturaleza: Madrid, Spain, 2003; pp. 796–797. [Google Scholar]
  29. Moreno Saiz, J.C.; Sainz ollero, H. Atlas corológico de las monocotiledóneas endémicas de la Península Ibérica e Islas Baleares. Bases para una política de conservación. Colección Técnica. ICONA; Ministerio de Agricultura, Pesca y Alimentación: Madrid, Spain, 1992; p. 354. [Google Scholar]
  30. Hernández-Bermejo, E.; Prados, J.; Herrera-Molina, F.; Narcissus bugei. The IUCN Red List of Threatened Species 2006: E.T61601A12519064. Available online: https://www.iucnredlist.org/species/61601/12519064 (accessed on 7 February 2024).
  31. García Murillo, P.G.; Ríos Ruiz, S.; Gutiérrez Carretero, L.; Lorite, J.; Ruiz Girela, M.; López, M.; Donaire Sánchez, F.J.; Sánchez Gómez, P.; Carrión Vilches, M.Á.; Navarro, F.B.; et al. Narcissus nevadensis. The IUCN Red List of Threatened Species 2018: E.T61682A47061845. Available online: https://www.iucnredlist.org/species/61682/47061845 (accessed on 7 February 2024).
  32. Juan Vicedo, J.; Ríos Ruiz, S.; García Murillo, P.G. Narcissus alcaracensis. The IUCN Red List of Threatened Species 2018: E.T61600A57488304. Available online: https://www.iucnredlist.org/species/61600/57488304 (accessed on 7 February 2024).
  33. Juan Vicedo, J.; Narcissus yepesii. The IUCN Red List of Threatened Species 2018: E.T57456144A57456147. Available online: https://www.iucnredlist.org/species/57456144/57456147 (accessed on 7 February 2024).
  34. Caldas, F.B.; Moreno Saiz, J.C.; Narcissus asturiensis. The IUCN Red List of Threatened Species 2013: E.T162098A5536148. Available online: https://www.iucnredlist.org/species/162098/5536148 (accessed on 7 February 2024).
  35. Juan Vicedo, J.; Narcissus hedraeanthus. The IUCN Red List of Threatened Species 2018: E.T57455245A57455282. Available online: https://www.iucnredlist.org/species/57455245/57455282 (accessed on 7 February 2024).
  36. Caldas, F.B.; Moreno Saiz, J.C.; Buord, S.; Narcissus triandrus. The IUCN Red List of Threatened Species 2013: E.T162050A5542491. Available online: https://www.iucnredlist.org/species/162050/5542491 (accessed on 7 February 2024).
  37. Pigni, N.B.; Ríos-Ruiz, S.; Martínez-Francés, V.; Nair, J.J.; Viladomat, F.; Codina, C.; Bastida, J. Alkaloids from Narcissus serotinus. J. Nat. Prod. 2012, 75, 1643–1647. [Google Scholar] [CrossRef]
  38. Torras-Claveria, L.; Berkov, S.; Codina, C.; Viladomat, F.; Bastida, J. Metabolomic analysis of bioactive Amaryllidaceae alkaloids of ornamental varieties of Narcissus by GC-MS combined with k-means cluster analysis. Ind. Crop. Prod. 2014, 56, 211–222. [Google Scholar] [CrossRef]
  39. Takos, A.M.; Rook, F. Towards a molecular understanding of the biosynthesis of Amaryllidaceae alkaloids in support of their expanding medical use. Int. J. Mol. Sci. 2013, 14, 11713. [Google Scholar] [CrossRef]
  40. Ellman, G.L.; Courtney, K.D.; Andres, V.; Featherstone, R.M. A new and rapid colorimetric determination of acetylcholinesterase activity. Biochem. Pharmacol. 1961, 7, 88–90. [Google Scholar] [CrossRef] [PubMed]
  41. López, S.; Bastida, J.; Viladomat, F.; Codina, C. Acetylcholinesterase inhibitory activity of some Amaryllidaceae alkaloids and Narcissus extracts. Life Sci. 2002, 71, 2521–2529. [Google Scholar] [CrossRef] [PubMed]
  42. Orhan, I.; Sener, B. Bioactivity-directed fractionation of alkaloids from some Amaryllidaceae plants and their anticholinesterase activity. Chem. Nat. Compd. 2003, 39, 383–386. [Google Scholar] [CrossRef]
  43. Havlasová, J.; Šafratová, M.; Siatka, T.; Štěpánková, S.; Novák, Z.; Ločárek, M.; Opletal, L.; Hrabinová, M.; Jun, D.; Benešová, N.; et al. Chemical Composition of Bioactive Alkaloid Extracts from Some Narcissus Species and Varieties and their Biological Activity. Nat. Prod. Commun. 2014, 9, 1151–1155. [Google Scholar] [CrossRef]
  44. Gotti, R.; Fiori, J.; Bartolini, M.; Cabrini, V. Analysis of Amaryllidaceae alkaloids from Narcissus by GC-MS and capillary electrophoresis. J. Pharmaceut. Biomed. 2006, 42, 17–24. [Google Scholar] [CrossRef] [PubMed]
  45. Masi, M.; Frolova, L.; Yu, X.; Mathieu, V.; Cimmino, A.; De Carvalho, A.; Kiss, R.; Rogelj, S.; Pertsemlidis, A.; Kornienko, A.; et al. Jonquailine, a new pretazettine-type alkaloid isolated from Narcissus jonquilla quail, with activity against drug-resistant cancer. Fitoterapia 2015, 102, 41–48. [Google Scholar] [CrossRef]
  46. Bozkurt, B.; Coban, G.; Kaya, G.I.; Onur, M.A.; Unver-Somer, N. Alkaloid profiling, anticholinesterase activity and molecular docking of Galanthus elwesii. S. Afr. J. Bot. 2017, 113, 119–127. [Google Scholar] [CrossRef]
  47. Cortes, N.; Sierra, K.; Alzate, F.; Osorio, E.H.; Osorio, E. Alkaloids of Amaryllidaceae as Inhibittors of Cholinesterases (AChE and BuChEs): An Integrated Bioguide Study. Phytochem. Anal. 2017, 29, 217–227. [Google Scholar] [CrossRef] [PubMed]
  48. León, K.A.; Inca, A.; Tallini, L.R.; Osorio, E.H.; Robles, J.; Bastida, J.; Oleas, N.H. Alkaloids of Phaedranassa dubia (Kunth) J.F. Macbr. and Phaedranassa brevifolia Meerow (Amaryllidaceae) from Ecuador and its cholinesterase-inhibitory activitty. S. Afr. J. Bot. 2021, 136, 91–99. [Google Scholar] [CrossRef]
  49. Rojas-Vera, J.d.C.; Buitrago-Díaz, A.A.; Possamai, L.M.; Timmers, L.F.S.M.; Tallini, L.R.; Bastida, J. Alkaloid profile and cholinesterase inhibition activity of five species of Amaryllidaceae family collected from Mérida state-Venezuela. S. Afr. J. Bot. 2021, 136, 126–136. [Google Scholar] [CrossRef]
  50. Tallini, L.R.; Osorio, E.H.; Berkov, S.; Torras-Claveria, L.; Rodríguez-Escobar, M.L.; Viladomat, F.; Meerow, A.; Bastida, J. Chemical Survey of Three Species of the Genus Rauhia Traub (Amaryllidaceae). Plants 2022, 11, 3549. [Google Scholar] [CrossRef] [PubMed]
  51. Llabrés, J.M.; Viladomat, F.; Bastida, J.; Codina, C.; Rubiralta, M. Phenanthridine alkaloids from Narcissus assoanus. Phytochemistry 1986, 25, 2637–2638. [Google Scholar] [CrossRef]
  52. Viladomat, F.; Llabrés, J.M.; Bastida, J.; Cusidó, R.M.; Codina, C. Ontogenic variations in the alkaloids of Narcissus assoanus. Physiol. Plant. 1986, 68, 657–661. [Google Scholar] [CrossRef]
  53. Bastida, J.; Viladomat, F.; Llabrés, J.M.; Codina, C.; Rubiralta, M. Alkaloids from Narcissus jacetanus. Planta Med. 1988, 54, 362. [Google Scholar] [CrossRef]
  54. Tallini, L.R.; Bastida, J.; Cortes, N.; Osorio, E.H.; Theoduloz, C.; Schmeda-Hirschmann, G. Cholinesterase inhibition activity, alkaloid profiling and molecular docking of Chilean Rhodophiala (Amaryllidaceae). Molecules 2018, 23, 1532. [Google Scholar] [CrossRef] [PubMed]
  55. Brito, M.F.B.; Ferreira, J.V.; de Souza, L.R.; Gemaque, L.R.P.; Sousa, K.P.A.; dos Santos, C.F.; Braga, F.S.; Pernomian, L.; da Silva, C.H.T.P.; Santos, C.B.R.; et al. Computational Molecular Modeling of Compounds from Amaryllidaceae Family as Potential Acetylcholinesterase Inhibitors. Curr. Bioact. Compd. 2017, 13, 121–129. [Google Scholar] [CrossRef]
  56. Labraña, J.; Choy, G.; Solans, X.; Font-Bardia, M.; de la Fuente, G.; Viladomat, F.; Codina, C.; Bastida, J. Alkaloids from Narcissus bujei (Amaryllidaceae). Phytochemistry 1999, 50, 183–188. [Google Scholar] [CrossRef]
  57. Bastida, J.; Codina, C.; Viladomat, F. Narcissus alkaloids, XIV. (+)-8-O-Acetylhomolycorine and vasconine, two novel alkaloids from Narcissus vasconicus. J. Nat. Prod. 1992, 55, 122–125. [Google Scholar] [CrossRef]
  58. Pigni, N.B.; Ríos-Ruiz, S.; Luque, L.; Viladomat, F.; Codina, C.; Bastida, J. Wild daffodils of the section Ganymedes from the Iberian Peninsula as a source of mesembrane alkaloids. Phytochemistry 2013, 95, 384–393. [Google Scholar] [CrossRef]
  59. Evidente, A. Advances on the Amaryllidaceae alkaloids collected in South Africa, Andean South America and the Mediterranean basin. Molecules 2023, 28, 4055. [Google Scholar] [CrossRef]
  60. Seijas, J.A.; Vázquez-Tato, M.P.; Linares, M.T.; Ramil-Rego, P.; Buján, M.I. Mesembrine alkaloids from Narcissus triandrus L. In Proceedings of ECSOC-8, The Eight International Electronic Conference on Synthetic Organic Chemistry; Seijas, J.A., Vázquez-Tato, M.P., Eds.; MDPI: Basel, Switzerland,, 2004; Available online: https://www.mdpi.org/ecsoc/ecsoc-8/BOCNP/008/index.htm (accessed on 7 February 2024).
  61. Bastida, J.; Viladomat, F.; LLabres, J.M.; Codina, C.; Feliz, M.; Rubiralta, M. Alkaloids from Narcissus confusus. Phytochemistry 1987, 26, 1519–1524. [Google Scholar] [CrossRef]
  62. Bastida, J.; Llabrés, J.M.; Viladomat, F.; Codina, C. Narcissus alkaloids, III. 9-O-demethylhomolycorine from Narcissus confusus. J. Nat. Prod. 1987, 50, 199–202. [Google Scholar] [CrossRef]
  63. Berkov, S.; Bastida, J.; Viladomat, F.; Codina, C. Development and validation of a GC-MS method for rapid determination of galanthamine in Leucojum aestivum and Narcissus ssp.: A metabolic approach. Talanta 2011, 83, 1455–1465. [Google Scholar] [CrossRef]
  64. Karakoyun, Ç.; Ünver-Somer, N. Simultaneous quantitative analysis of biologically important Amaryllidaceae alkaloids in Narcissus tazetta L. subsp. tazetta by HPLC/PDA. J. Res. Pharm. 2019, 23, 498–505. [Google Scholar] [CrossRef]
  65. Karakoyun, Ç.; Bozkurt, B.; Çoban, G.; Masi, M.; Cimmino, A.; Evidente, A.; Somer, N.U. A comprehensive study on Narcissus tazetta subsp. tazetta L.: Chemo-profiling, isolation, anticholinesterase activity and molecular docking of Amaryllidaceae alkaloids. S. Afr. J. Bot. 2020; 130, 148–154. [Google Scholar] [CrossRef]
  66. Tarakemeh, A.; Azizi, M.; Rowshan, V.; Salehi, H.; Spina, R.; Dupire, F.; Arouie, H.; Laurain-Mattar, D. Screening of Amaryllidaceae alkaloids in bulbs and tissue cultures of Narcissus papyraceus and four varieties of N. tazetta. J. Pharmaceut. Biomed. 2019, 172, 230–237. [Google Scholar] [CrossRef]
  67. Viladomat, F.; Sellés, M.; Codina, C.; Bastida, J. Alkaloids from Narcissus asturiensis. Planta Med. 1997, 63, 583. [Google Scholar] [CrossRef]
  68. Mamun, A.A.; Maríková, J.; Hulcová, D.; Janoušek, J.; Šafratová, M.; Nováková, L.; Kucera, T.; Hrabinová, M.; Kuneš, J.; Korábecny, J.; et al. Amaryllidaceae Alkaloids of belladine-type from Narcissus pseudonarcissus cv. Carlton as new selective inhibitors of butyrylcholinesterase. Biomolecules 2020, 10, 800. [Google Scholar] [CrossRef]
  69. Hulcová, D.; Mariková, J.; Korábecny, J.; Hostalková, A.; Jun, D.; Kuneš, J.; Chlebek, J.; Opletal, L.; De Simone, A.; Nováková, L.; et al. Amaryllidaceae alkaloids from Narcissus pseudonarcissus L. cv. Dutch Master as potential drugs in treatment of Alzheimer’s disease. Phytochemistry 2019, 165, 112055. [Google Scholar] [CrossRef]
Figure 1. The local collection of the different species of Narcissus in Spain according to Table 1 codes. Source: Google Earth. A = N. assoanus; B = N. jacetanus; C = N. vasconicus; D = N. minor; E = N. confusus; F = N. asturiensis; G = N. hedraeanthus; H = N. alcaracencis; I = N. bujei; J = N. pallidulus; K = N. tazetta; L = N. jonquilla; M = N. genesii-lopezii; N = N. yepesii; O = N. nevadensis. species name when we refer to.
Figure 1. The local collection of the different species of Narcissus in Spain according to Table 1 codes. Source: Google Earth. A = N. assoanus; B = N. jacetanus; C = N. vasconicus; D = N. minor; E = N. confusus; F = N. asturiensis; G = N. hedraeanthus; H = N. alcaracencis; I = N. bujei; J = N. pallidulus; K = N. tazetta; L = N. jonquilla; M = N. genesii-lopezii; N = N. yepesii; O = N. nevadensis. species name when we refer to.
Life 14 00536 g001
Figure 2. Graph of AChE inhibitory activity of the different samples of alkaloid extracts of Narcissus leaves collected in Spain. A = N. assoanus; B = N. jacetanus; C = N. vasconicus; D = N. minor; E = N. confusus; F = N. asturiensis; G = N. hedraeanthus; H = N. alcaracencis; I = N. bujei; K = N. tazetta; L = N. jonquilla; M = N. genesii-lopezii; N = N. yepesii; O = N. nevadensis; Gal = galanthamine; **** p < 0.0001, ** p < 0.01, ns—not significant.
Figure 2. Graph of AChE inhibitory activity of the different samples of alkaloid extracts of Narcissus leaves collected in Spain. A = N. assoanus; B = N. jacetanus; C = N. vasconicus; D = N. minor; E = N. confusus; F = N. asturiensis; G = N. hedraeanthus; H = N. alcaracencis; I = N. bujei; K = N. tazetta; L = N. jonquilla; M = N. genesii-lopezii; N = N. yepesii; O = N. nevadensis; Gal = galanthamine; **** p < 0.0001, ** p < 0.01, ns—not significant.
Life 14 00536 g002
Figure 3. Graph of BuChE inhibitory activity of the different samples of alkaloid extracts of Narcissus leaves collected in Spain. C = N. vasconicus; E = N. confusus; G = N. hedraeanthus; H = N. alcaracencis; I = N. bujei; L = N. jonquilla; M = N. genesii-lopezii; N = N. yepesii; O = N. nevadensis; Gal = galanthamine; **** p < 0.0001, *** p < 0.001, ** p < 0.01.
Figure 3. Graph of BuChE inhibitory activity of the different samples of alkaloid extracts of Narcissus leaves collected in Spain. C = N. vasconicus; E = N. confusus; G = N. hedraeanthus; H = N. alcaracencis; I = N. bujei; L = N. jonquilla; M = N. genesii-lopezii; N = N. yepesii; O = N. nevadensis; Gal = galanthamine; **** p < 0.0001, *** p < 0.001, ** p < 0.01.
Life 14 00536 g003
Figure 4. Photo of Narcissus jacetanus, the sample with the best results for AChE inhibition.
Figure 4. Photo of Narcissus jacetanus, the sample with the best results for AChE inhibition.
Life 14 00536 g004
Figure 5. Photo of Narcissus jonquilla, the sample with the best results for BuChE inhibition (also with noteworthy activity against AChE).
Figure 5. Photo of Narcissus jonquilla, the sample with the best results for BuChE inhibition (also with noteworthy activity against AChE).
Life 14 00536 g005
Table 1. Narcissus sample coding, biogeographical distribution, and province of collection. Biogeographical units: Baet: Baetic, CircMed: Circummediterranean, Iber: Iberic, Iber/LAlp: Iberic/Latealpine, Iber/NAfric: Iberic/Northafrican, Med Occ: Mediterranean occidental; political regions of origin: AND: Andalucía, AR: Aragón, CL: Castilla León, CLM: Castilla-La Mancha, VAL: Valencian region, BAS: Basque region. Biogeographical units according to [29].
Table 1. Narcissus sample coding, biogeographical distribution, and province of collection. Biogeographical units: Baet: Baetic, CircMed: Circummediterranean, Iber: Iberic, Iber/LAlp: Iberic/Latealpine, Iber/NAfric: Iberic/Northafrican, Med Occ: Mediterranean occidental; political regions of origin: AND: Andalucía, AR: Aragón, CL: Castilla León, CLM: Castilla-La Mancha, VAL: Valencian region, BAS: Basque region. Biogeographical units according to [29].
CodeSpeciesSectionBiogeographical
Distribution
Origin
AN. assoanus Dufour ex Schult. and Schult.f.Jonquillae DC.Med OccHuesca (AR)
BN. jacetanus Fern.CasasPseudonarcissus DC.Iber/LAlpHuesca (AR)
CN. vasconicus (Fern.Casas) Fern.CasasPseudonarcissus DC.Iber/LAlpVitoria (BAS)
DN. minor L.Pseudonarcissus DC.Iber/LAlpLeón (CL)
EN. confusus PugsleyPseudonarcissus DC.Iber/LAlpSalamanca (CL)
FN. asturiensis (Jord.) PugsleyPseudonarcissus DC.Iber/LAlpSalamanca (CL)
GN. hedraeanthus (Webb and Heldr.) ColmeiroBulbocodium (Salisb.) DC.Iber/NAfricAlbacete (CLM)
HN. alcaracensis S.Ríos, D.Rivera, Alcaraz and ObónNevadensis Zonn.BaetAlbacete (CLM)
IN. bujei (Fern.Casas) Fern.CasasPseudonarcissus DC.Baet Albacete (CLM)
JN. pallidulus GraellsGanymedes (Salisb.) Schultes f.IberAlbacete (CLM)
KN. tazetta L.Tazettae DC.CircMedAlicante (VAL)
LN. jonquilla L.Jonquillae DC.IberToledo (CLM)
MN. genesii-lopezii Fern.CasasPseudonarcissus DC.IberCuenca (CLM)
NN. yepesii S.Ríos, D.Rivera, Alcaraz and ObónNevadensis Zonn.BaetJaén (AND)
ON. nevadensis PugsleyNevadensis Zonn.BaetGranada (AND)
Table 2. Yield of alkaloid extracted, obtained from the Narcissus species.
Table 2. Yield of alkaloid extracted, obtained from the Narcissus species.
SamplesDry Weight (g)Alkaloid Extract (mg)Yield (%)
A1.000808.930.89
B1.001144.690.47
C1.000466.050.60
D1.000504.950.49
E1.0003214.541.45
F1.000359.790.98
G0.882536.830.77
H1.0006910.661.07
I1.0012518.831.88
J1.0013710.171.02
K1.001038.380.84
L1.0009415.321.53
M1.000996.850.68
N1.0003761.436.14
O1.000669.270.93
A = N. assoanus; B = N. jacetanus; C = N. vasconicus; D = N. minor; E = N. confusus; F = N. asturiensis; G = N. hedraeanthus; H = N. alcaracencis; I = N. bujei; J = N. pallidulus; K = N. tazetta; L = N. jonquilla; M = N. genesii-lopezii; N = N. yepesii; O = N. nevadensis.
Table 3. IC50 values of AChE and BuChE inhibitory activity of the Narcissus species collected in Spain. Values expressed in µg·mL−1. A = N. assoanus; B = N. jacetanus; C = N. vasconicus; D = N. minor; E = N. confusus; F = N. asturiensis; G = N. hedraeanthus; H = N. alcaracencis; I = N. bujei; K = N. tazetta; L = N. jonquilla; M = N. genesii-lopezii; N = N. yepesii; O = N. nevadensis; Gal = galanthamine.
Table 3. IC50 values of AChE and BuChE inhibitory activity of the Narcissus species collected in Spain. Values expressed in µg·mL−1. A = N. assoanus; B = N. jacetanus; C = N. vasconicus; D = N. minor; E = N. confusus; F = N. asturiensis; G = N. hedraeanthus; H = N. alcaracencis; I = N. bujei; K = N. tazetta; L = N. jonquilla; M = N. genesii-lopezii; N = N. yepesii; O = N. nevadensis; Gal = galanthamine.
SamplesSpeciesAChEBuChE
AN. assoanus0.99 ± 0.06>100
BN. jacetanus0.75 ± 0.03>100
CN. vasconicus2.98 ± 0.5123.14 ± 0.58
DN. minor0.81 ± 0.10>100
EN. confusus1.04 ± 0.0712.83 ± 0.87
FN. asturiensis34.28 ± 2.06>100
GN. hedraeanthus5.73 ± 0.3629.23 ± 3.84
HN. alcaracencis9.54 ± 1.2659.60 ± 2.61
IN. bujei6.14 ± 0.6834.86 ± 1.75
JN. pallidulus>100>100
KN. tazetta3.68 ± 0.36>100
LN. jonquilla1.88 ± 0.0511.72 ± 1.15
MN. genesii-lopezii5.28 ± 0.6411.98 ± 0.93
NN. yepesii10.63 ± 0.0887.20 ± 3.01
ON. nevadensis7.03 ± 1.4931.68 ± 0.49
Gal 0.46 ± 0.035.13 ± 0.48
Gal: galanthamine.
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

Tallini, L.R.; Manfredini, G.; Rodríguez-Escobar, M.L.; Ríos, S.; Martínez-Francés, V.; Feresin, G.E.; Borges, W.d.S.; Bastida, J.; Viladomat, F.; Torras-Claveria, L. The Anti-Cholinesterase Potential of Fifteen Different Species of Narcissus L. (Amaryllidaceae) Collected in Spain. Life 2024, 14, 536. https://doi.org/10.3390/life14040536

AMA Style

Tallini LR, Manfredini G, Rodríguez-Escobar ML, Ríos S, Martínez-Francés V, Feresin GE, Borges WdS, Bastida J, Viladomat F, Torras-Claveria L. The Anti-Cholinesterase Potential of Fifteen Different Species of Narcissus L. (Amaryllidaceae) Collected in Spain. Life. 2024; 14(4):536. https://doi.org/10.3390/life14040536

Chicago/Turabian Style

Tallini, Luciana R., Giulia Manfredini, María Lenny Rodríguez-Escobar, Segundo Ríos, Vanessa Martínez-Francés, Gabriela E. Feresin, Warley de Souza Borges, Jaume Bastida, Francesc Viladomat, and Laura Torras-Claveria. 2024. "The Anti-Cholinesterase Potential of Fifteen Different Species of Narcissus L. (Amaryllidaceae) Collected in Spain" Life 14, no. 4: 536. https://doi.org/10.3390/life14040536

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