Next Article in Journal
Identification of QTL under Brassinosteroid-Combined Cold Treatment at Seedling Stage in Rice Using Genotyping-by-Sequencing (GBS)
Next Article in Special Issue
Investigation of the Genotoxicological Profile of Aqueous Betula pendula Extracts
Previous Article in Journal
Response Surface Methodology (RSM) Optimization of the Physicochemical Quality Attributes of Ultraviolet (UV-C)-Treated Barhi Dates
Previous Article in Special Issue
Cytotoxic Activities and the Allantoinase Inhibitory Effect of the Leaf Extract of the Carnivorous Pitcher Plant Nepenthes miranda
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Extraction Techniques and Analytical Methods for Isolation and Characterization of Lignans

by
Andrzej Patyra
1,2,3,*,
Małgorzata Kołtun-Jasion
1,
Oktawia Jakubiak
1 and
Anna Karolina Kiss
1,*
1
Department of Pharmacognosy and Molecular Basis of Phytotherapy, Medical University of Warsaw, 02-097 Warsaw, Poland
2
Doctoral School, Medical University of Warsaw, 02-091 Warsaw, Poland
3
Institut des Biomolécules Max Mousseron, Université de Montpellier, CNRS, ENSCM, 34293 Montpellier, France
*
Authors to whom correspondence should be addressed.
Plants 2022, 11(17), 2323; https://doi.org/10.3390/plants11172323
Submission received: 3 August 2022 / Revised: 30 August 2022 / Accepted: 31 August 2022 / Published: 5 September 2022
(This article belongs to the Special Issue Plant Extracts and Their Cytotoxic Activities)

Abstract

:
Lignans are a group of natural polyphenols present in medicinal plants and in plants which are a part of the human diet for which more and more pharmacological activities, such as antimicrobial, anti-inflammatory, hypoglycemic, and cytoprotective, are being reported. However, it is their cytotoxic activities that are best understood and which have shed light on this group. Two anticancer drugs, etoposide, and teniposide, were derived from a potent cytotoxic agent—podophyllotoxin from the roots of Podophyllum peltatum. The evidence from clinical and observational studies suggests that human microbiota metabolites (enterolactone, enterodiol) of dietary lignans (secoisolariciresinol, pinoresinol, lariciresinol, matairesinol, syringaresinol, medioresinol, and sesamin) are associated with a reduced risk of some hormone-dependent cancers. The biological in vitro, pharmacological in vivo investigations, and clinical studies demand significant amounts of pure compounds, as well as the use of well-defined and standardized extracts. That is why proper extract preparation, optimization of lignan extraction, and identification are crucial steps in the development of lignan use in medicine. This review focuses on lignan extraction, purification, fractionation, separation, and isolation methods, as well as on chromatographic, spectrometric, and spectroscopic techniques for their qualitative and quantitative analysis.
Keywords:
lignans; extraction; HPLC; MS; TLC

1. Introduction

Lignans are a group of natural polyphenols (and one of the most lipophilic), located chiefly in the plant cell walls, many of which are present in medical plants and in plants that are a part of the human diet. According to the recent nomenclature, lignans are dimers of two coniferyl, sinapyl, paracoumaryl, or similar alcohol monomers. Some authors restrict the term lignan only to those molecules coupled by the central carbon of the side-chain (i.e., 8,8′ or β,β’ dimers). In this manner, there is a distinction between neolignans—similar compounds coupled differently (by another carbon or through oxygen bonds), norlignans—lignans which lost one of their carbon atoms from their skeleton, lignoids—compounds that have undergone skeletal rearrangement, and oligolignans—lignan condensation products—which we can divide into sequilignans (trimers) and dilignans (tetramers) [1]. Lignans were found in all of the plant parts, though their highest content was reported in plant wood—especially in wood knots—roots, bark, leaves, flowers, fruits, and seeds [2]. Lignans’ and neolignans’ formation are subject to stereoselective biosynthesis (Figure 1); thus, they are enantiomerically pure [1]. They mostly appear linked with carbohydrates, while more lipophilic aglycon-free forms can be found in wood and bark [3].
There is a rising interest in lignans and plants rich in this class of polyphenols as more and more pharmacological activities are being reported, such as antimicrobial, anti-inflammatory, hypoglycemic, and cytoprotective. However, it is their cytotoxic activities that are best understood, and which have shed light on this group. Two anticancer drugs, etoposide and teniposide, were derived from a potent cytotoxic agent—podophyllotoxin from the roots of Podophyllum peltatum L. This led to the search for other cytotoxic metabolites among these polyphenols, with interesting finds concerning lignans and neolignans from nearly every structural subtype [12]. Interestingly, the evidence from clinical and observational studies suggests that the human microbiota metabolites (enterolactone, enterodiol) of dietary lignans (secoisolariciresinol, pinoresinol, lariciresinol, matairesinol, syringaresinol, medioresinol, and sesamin) are associated with reduced risk of some hormone-dependent cancers [13]. Primarily, the attention is focused on flaxseeds’ lignan—secoisolariciresinol diglucoside (SDG) and the effect on breast and colorectal cancer development [14,15].
Moreover, some in vitro and in vivo studies suggest that arctiin and its aglycon arctigenin, occurring in the Great Burdock fruits and Forsythia spp., are promising cytotoxic agents. Arctigenin was shown to induce tumor cell death, such as prostate, breast, lungs, liver, and colon, by inducing apoptotic signaling pathways, caspases, cell cycle arrest, and the modulating proteasome [16]. Other reports suggest the potential antiproliferative effect of lignans from Phyllanthus spp., Schisandra chinensis (Turcz.) Baill. or Magnolia officinalis Rehder & E.H. Wilson [17,18,19].
The biological in vitro, pharmacological in vivo investigations, and clinical studies demand significant amounts of pure compounds, as well as the use of well-defined and standardized extracts. That is why proper extract preparation, optimization of lignan extraction, and identification are crucial steps in the development of lignan use in medicine.
This review focuses on the lignan extraction, purification, fractionation, separation, and isolation methods, as well as on chromatographic, spectrometric, and spectroscopic techniques for their analysis. Considering that lignans and similar compounds are abundant in plants, and each year many new compounds are isolated, we decided to devote this paper to plant materials where the lignans are present in more significant amounts and adopt the strict definition of lignans mentioned before. This does not mean that the methods and techniques presented here do not apply to neolignans, norlignans, lignoids, or oligolignans’ isolation and analysis, as they may be present in the same plant material as true lignans. Still, our attention will not be directed to their physiochemical properties, and further considerations should be taken in relation to them.
A literature search for the present review was conducted using the following databases: PubMed, Scopus, and Google Scholar, with the following search terms: “lignans”, “extraction”, “isolation”, “detection”, “quantification”, and “analysis” in various combinations. The results were first screened for their relevance based on their abstract and afterwards the full texts were analyzed. Studies from the year 1998–2022 were considered, as this period was characterized by considerable interest in lignan analysis. The search was conducted on 4 July 2022.

2. Sample Preparation

The isolation of plant metabolites begins with the collection and handling of plant material. In this regard, the procedure is typical for isolating any plant metabolite and should include botanical or taxonomical identification and preparation and storage of voucher specimens [20]. The plant material that is otherwise acquired should have its origin noted, and certificates should be provided if possible. When collecting plant material, it should be considered in which plant parts the target compounds are accumulated. For instance, in wood, the lignans are primarily present in knot wood (up to 30% (w/w) of extract) than in heartwood (less than 0.1%), and so the separation of knot wood from other wood parts may prove beneficial [21,22]. Similarly, lignans are found in the hulls of seeds rather than in their embryos. Thus, isolation can be facilitated by hulling the seeds, i.e., separating the hulls from kernels [23].
Every plant material should be handled and stored in a way that will provide stability to plant metabolites; thus, they will be protected from transformation and degradation. In most cases, the material will be air-dried, oven-dried, or freeze-dried [24]. Lignans and their glycosides are relatively resistant to high temperatures, so the options are numerous. Both freeze-drying methods [25] and drying in temperatures as high as 60 °C [26] were successfully applied.
Higher temperatures have been applied to study the temperature stability of lignans in plant material and food products. Generally, lignans and their aglycones from flaxseed, sesame seed, and rye seem stable below the temperature of 100 °C, while applied heat also facilitated the extraction procedure [27]. The stir-frying of Great Burdock fruit at 150 °C led to changes in the lignan content, predominantly manifesting a higher aglycones’ and lower glycosides’ yield, which could be explained by the enhancement of hydrolysis processes [28,29]. Some aglycones (e.g., pinoresinol, sesamin) are stable in temperatures as high as 180–200 °C [30]. Similarly, the secoisolariciresinol diglucoside ester-linked complex with hydroxymethyl glutaryl (found in flaxseed) is stable even in bakery products [27]. The overall thermal stability of lignans depends on the compound structure and its interactions with the other compounds present in the plant matrix [27,30]. Although heating may positively impact on the extractability of lignans, the loss of some less temperature-stable compounds (i.e., neolignans linked by oxygen or sesquilignans), and the possible hydrolysis of glycosides, should always be brought into consideration.
Not many studies have focused on the photostability of lignans. For instance, Kawamura et al. [31] light-irradiated 7-hydroxymatairesinol in different media (chloroform, methanol, water, and directly in plant matrix) and identified oxidation products, such as α-conidendrin, 7-oxomatairesinol, allohydroxymatairesinol, 7-methoxymatairesinol, and vanillin. Despite the scarcity of data on the photostability of lignans, the protection of the plant material and isolated compounds from light is reasonable.
Lastly, sometimes lignans can be stored in plants in a macromolecular complex, such as secoisolariciresinol diglucoside in flaxseed. It is esterified with 3-hydroxymethylglutaric acid and other phenolic compounds, such as p-coumaric acid and ferulic acid glycosides [32]. Various methods have been implemented to facilitate the extraction, such as acidic, alkaline, and enzymatic hydrolysis [33]. Although acidic hydrolysis can efficiently break the ester and glycoside linkages, liberating secoisolariciresinol, the product is not stable in these conditions and can transform to anhydrosecoisolariciresinol [34].

3. Extraction

The extraction process is a crucial and often overlooked step in isolating secondary metabolites. Several parameters influencing the extraction yield should be considered for the optimum recovery of lignans from the plant matrix. These include the choice of extraction solvent or solvents, method, time, temperature, solvent-to-sample ratio, and the number of repeat extractions of a sample [24].

4. Solvents

The choice of extraction solvent should be based on its physicochemical properties (polarity, solubility, lipophilicity), safety, ease of use, the potential for artifact formation, grade and purity, selectivity, and cost. The solvent should be matched with the target compound for the physicochemical properties. In this case, lignans are fairly lipophilic polyphenols (logP = 3.3 ± 1.0) with a limited water solubility (logS = −4.2 ± 1.4), whilst lignans glycosides tend to be more hydrophilic (logP = 0.4 ± 1.1; logS = −2.5 ± 1.3) [35]. Thus, for plant material containing aglycones, medium polarity solvents, such as ethyl acetate, may be used, as well as polar solvents such as ethanol, methanol, and their aqueous mixtures. Although some less polar lignans may be extracted with less polar solvents, e.g., dichloromethane, chloroform, or even n-hexane, they seem to be rarely used in the primary extraction from the plant matrix and are rather found in further steps, i.e., partitioning, separation, and isolation [26,36].
Most of the lignan glycosides will not be extracted by medium polarity solvents, and the ones that are more polar may be hard to obtain using pure ethanol or methanol. In this case, aqueous mixtures of ethanol and methanol give the best results. In the case of very polar lignan glycosides, pure water may be used [37,38,39,40].
It is generally advised to select a solvent by assessing the lipophilicity of the target lignan. When planning to isolate an unknown compound or obtain more than one compound (e.g., both aglycones and glycosides), choosing aqueous mixtures of ethanol or methanol is advisable. Aqueous alcohols have the advantage over pure solvents of greatly facilitating the penetration of the solvents into the plant matrix with even a small addition of water (5–10%). The mixtures of water and alcohols have an advantage over sequential extraction with those solvents separately of lowering the water surface tension, reducing the polarity of water, and increasing its density [41].
Most commonly, concentrations of 70–100% of either aqueous ethanol or methanol are used for lignans and lignans glycosides extraction (Table 1). As shown in the table, the concentration of ethanol or methanol used does not depend on the lignan form in the plant matrix. We observed both pure alcohols’ use in glycosides’ extraction [42,43,44,45] and water’s use in free aglycones’ extraction [39].
The choice between methanol and ethanol seems to depend mainly on the practice of a particular laboratory. It is of course worth noting that methanol has certain advantages, such as higher polarity, lower boiling point (which facilitates evaporation of solvent), and lower water impurity compared to ethanol. On the other hand, it is much more toxic than ethanol, and thus ethanol is more commonly used for the isolation of compounds for biological studies and when traditionally prepared extracts are studied [92].
The presence of other substances in the plant matrix is not without influence on the extraction of lignans. In particular, some lipophilic components, such as resins, terpenoids, and fatty acids, may cause some difficulties. Thus, it is advised to use sequential extraction with a non-polar solvent for lignan sources rich in lipophilic contents, e.g., flaxseeds, sesame seeds, Burdock fruit, or conifer wood. Petroleum ether and n-hexane are the most commonly used solvents for defatting extracts, although pentane and dichloromethane may also be used for this purpose. Caution is advised when using dichloromethane or similar medium polar solvents, as some more lipophilic lignans, e.g., the lignans from Magnolia spp. L. flowers, can be extracted in those solvents [93]. Other methods may also be applied, such as lipid precipitation in 10% acetone in −40 °C [26]. The removal of fatty contents can be executed directly on the plant material or crude extract.
Nowadays, an essential factor during solvent and method consideration is its environmental impact; thus, much attention is paid to reducing organic solvent use. Recently green extraction methods had been applied in lignan isolation from Eleutherococcus senticosus (Rupr. & Maxim.) Maxim. root. The authors reported using choline chloride and a lactic acid mixture in a 1:2 ratio with a 20% addition of water, successfully extracting lignan glycosides from plant matrix [94]. Although it is the first and only use of deep eutectic solvent in lignan extraction, the development of this method in the future should be expected.
In laboratory practice, solid to liquid ratios of 1:5 to 1:20 are generally considered appropriate. Higher ratios are not cost-efficient and should be used only to extract very valuable compounds. Thus, microscale experiments, which can sometimes offer ratios as high as 1:200, are not applicable on a larger scale [95]. Table 1 shows the solvent-to-sample ratios used in lignans extraction.

5. Methods

Lignans’ extraction can be executed using a variety of techniques. Commonly, the conventional procedures such as Soxhlet, maceration, and digestion, were applied. However, the use of other methods, such as ultrasound-assisted extraction (UAE), accelerated solvent extraction (ASE), microwave-assisted extraction (MAE), and supercritical fluid extraction (SFE) is increasing. The use of these methods is reviewed in Table 1.
The conventional methods are often techniques of choice because they do not require complex apparatus and because of their simplicity while producing satisfactory extraction rates. On the other hand, these methods carry essential disadvantages, such as long extraction times and the use of large volumes of solvents [96]. Contrary to other conventional methods, maceration is performed at room temperature, which is why it is mainly used to extract thermolabile compounds. Although macerations as short as 1 h have been performed in lignans’ extraction [70], they usually require much longer extraction times, reaching up to 7 days [60], with 24 h commonly applied [57,59,72]. Moreover, the extraction of phenolics (and by that lignans) by maceration is relatively weak [79]. As we have discussed before, lignans are not thermolabile compounds, and thus heat may be applied to shorten the extraction procedure. For instance, Kwon et al. [83] applied gentle heat of 50 °C in the maceration of Schisandra chinensis (Turcz.) Baill. fruit (a technique called digestion). Similar digestion procedures were used by Sicilia et al. [34] and Frishe et al. [72] in flaxseed lignans’ extraction.
Soxhlet and other heated reflux techniques are the most common methods used in lignans’ extraction. In their papers, the authors did not include the information on the temperature they used in their methods. However, as Zhang et al. pointed out, we can assume that it usually is performed at 80–100 °C—a temperature capable of boiling ethanol, methanol, and their dilutions, as well as other organic solvents and water [96]. The main advantage of this method is limiting the use of large volumes of solvents, as one batch of solvent is constantly recycled [92]. The time of extraction under this method varies from the 30 min used by Yang et al. in the extraction of eleutheroside E (syringaresinol diglucoside) from Eleutherococcus senticosus (Rupr. & Maxim.) Maxim. root [37], up to 10 h which was applied by Kumar et al. in the extraction of sesamin and sesamolin from the sesame seeds [26].
Sequential extraction at elevated temperature and pressure in short time periods, called accelerated solvent extraction (ASE), was applied several times in lignans’ extraction from wood knots. Willfor et al. were probably the first ones to utilize this technique in the extraction of lignans from Picea abies (L.) H.Karst. [97] and Pinus sylvestris L. wood knots [80]. The use of ASE followed studies on other conifer species, such as Picea spp. Mill. [98], Abies spp. Mill. [99], Larix spp. Mill., Thuja spp. L., Tsuga spp. Carrière, and Pseudotsuga spp. Carrière [100]. It was not seen outside this application.
Alternative techniques were implemented to overcome the disadvantages of conventional methods, such as solvent and time consumption, and environmental impact. Their common characteristic is to increase the efficiency of extraction using auxiliary energies (microwaves, ultrasounds) in place of conventional heating, stirring, shaking, or vortexing [101]. These methods are more complex than conventional ones, as the extraction yield is not only the result of solvent choice, time, temperature, solvent-to-sample ratio, and the number of repeat extractions of a sample; it is also reliant on specific parameters of sonification (wave frequency and distribution, probe or bath system) or microwave techniques (microwave power) [24]. While the advantages of these techniques are numerous, there is a risk that ultrasounds and microwaves may alter lignans’ and other metabolites’ structures [101].
Ultrasound-assisted extraction (UAE) was used in the recovery of lignans from Linum usitatissimum L. seeds, Sesamum indicum L. seeds, Arctium lappa L. fruit, and Forsythia spp. Vahl fruits, stems, and roots [29,30,42,62,67,68,74,75,76,88]. The extraction times were much shorter than using conventional methods, ranging from 5 min for Sesamum indicum L. seeds [30], to 90 min for Forsythia viridissima Lindl. roots [68]. However, for other parameters of sonification, only the authors of 2 papers included information on wave frequency used in their experiments—40 kHz [88] and 70 kHz [67]. Similarly, in most works, little was said about the sonification system (bath or probe) and the apparatus used. Perhaps an important note should be made on the temperature effect in ultrasound-assisted extraction. Contrary to conventional methods, a higher temperature may lower the extraction yield, as ultrasound waves will not distribute properly in a boiling solvent. For instance, Mekky et al. [88] performed their sonification experiment for sesame seeds at room temperature, while Corbin et al. [75] studied the temperature range of 25 to 60 °C during the sonification of flaxseeds and showed the lowest temperature to be superior. UAE sequential extraction method developed by Mekky et al. for lignan extraction from sesame seed cake and flaxseed cake afforded the obtention of lignans without the hydrolysis of glycosides, thus up to four hexosides’ conjugates (sesaminol tetrahexosides) were observed in the sesame seed cake extract and up to three hexosides conjugates (secoisolariciresinol trihexoside dihydroxyethylglutaryl esters) in the flaxseed cake extract [76,88].
Another important method is microwave-assisted extraction (MAE), which was used for the first time in lignan extraction by Beejmohun et al. in their study of flaxseeds [71]. A valuable work on the optimization of this method’s conditions for the preparation of lignan-rich extract from Saraca asoca (Roxb.) Willd. bark was completed by Mishra and Aeri [82]. According to the authors, their work concludes that microwave-assisted extraction is superior to conventional methods for the isolation of lignans—in their case, it was lyoniside, a lignan glycoside. Regarding these conditions, 70% methanol dilution was chosen as the best extraction solvent, 10 min as the extraction time, and 1:30 as the sample-to-solvent ratio. A similar optimization was completed by Lu et al. for the isolation of arctiin from Arctium lappa L. fruit [49]. The authors considered five extraction parameters, i.e., methanol concentration, microwave power, solid-to-liquid ratio, extraction time, and extraction times. Based on their results, the optimum extraction conditions were 40% methanol, 500 W microwave power, 1:15 solid-to-liquid ratio, 200 s as the extraction time, and three times of the extraction. According to the authors, with these conditions, an extract containing 17.5% of arctiin and 2.2% of arctigenin was obtained, though the efficiency of the extraction is not quite clear as the authors did not include the quantity of plant material used for this experiment. A fine comparison of maceration, Soxhlet extraction, and microwave-assisted extraction was provided by Garg et al. [102]. In their paper, the authors showed that MAE is a superior method for phyllantin extraction to maceration and Soxhlet extraction in terms of lignan yield, shorter extraction time, and lower extraction temperature.
Lastly, supercritical fluid extraction (SFE) may be used as another environmentally friendly extraction technique. In the same way as other alternative methods, it can lower the use of extraction solvents, shorten extraction time, and increase the yield. However, the apparatus needed for the implementation of this technique is much higher than for the previously mentioned methods [24]. This method has been so far mostly used in the extraction of lignans from Schisandra chinensis (Turcz.) Baill. fruits, seeds, and leaves. The lignans present in this plant material are free aglycones, with most of them having little to no hydroxyl groups, making them rather lipophilic. Many papers optimizing the Schisandra lignans’ SFE extraction were published, with two of them recently [84,85]. Perhaps the most important point from these works is that temperature and pressure variations have a lesser effect on extractability than the addition of methanol or ethanol to carbon dioxide, which dramatically increased the lignan yield [103].

6. Artifacts

During the extraction of lignans, artifacts may be formed by unwanted chemical reactions, such as oxidation, polymerization, thermal degradation, and other chemical rearrangements. Their cause usually lies in exposure to light, high temperatures, acidic or alkaline conditions, and the presence of free radicals. Their occurrence is unfortunately rarely reported.
In terms of artifact formation, one of the more sensitive lignans is 7-hydroxymatairesinol. As mentioned before, light irradiation may oxidize 7-hydroxymatairesinol to α-conidendrin, 7-oxomatairesinol, allohydroxymatairesinol, 7-methoxymatairesinol, and vanillin, subsequently forming colored oligomers [31]. Similarly, 7-hydroxymatairesinol may be transformed by acidic and alkaline conditions to α-conidendrin and other oxidation products, which were described in detail by Eklund et al. [104]. Acidic media (extraction solvents, acidic hydrolysis, chromatographic systems, etc.) caused artifact formation from other lignans. Sicilia et al. [34] identified anhydrosecoisolariciresinol, lariciresinol, and cyclolariciresinol in flaxseeds, which were probably products of secoisolariciresinol rearrangements. Other reports confirm the intramolecular cyclization of lariciresinol to cyclolariciresinol [105]. The presence of acetonides and cyclization products of lignans were also reported to occur in some plant substances naturally; thus, it is not always clear whether they should be classified as artifacts [106].
Artifact formation due to temperature is relatively rare in lignans’ extraction. As discussed in previous paragraphs, the lignans are generally stable in temperatures up to 100 °C, with some of them even immune to much higher temperatures [27]. Considering that lignans’ extraction, isolation, and purification usually happen at temperatures below this limit, we should not expect degradation products. On the other hand, the temperature effect may enhance the hydrolysis processes and thus the isolation of free aglycones from substances containing only glycosylated forms [28,29].
The prevention of artifacts forming can be achieved by implementing simple measures, such as the absence of light, avoidance of high temperatures, control of pH, and storage of extracts in frozen/solid-state [2].

7. Isolation and Purification

The separation of individual structures is a crucial step in the chemical analysis of the raw material. The obtained pure compounds can be valuable as new standards for HPLC-MS analysis and for biological experiments to determine their in vitro and in vivo activity.
Flash chromatography on silica gel or Sephadex LH-20 columns has been used for years as the most popular preliminary method for the initial fractionation, and separation, of the samples, as well as for the purification and preparative isolation of natural compounds from raw materials [53]. In addition, various chromatographic methods have been mentioned in the literature, such as open column chromatography [107,108], medium performance liquid chromatography (MPLC) [109], and the most popular method—semi-preparative HPLC [110].
Liquid chromatography is a bidirectional technique, both as an analytical method and as a method of isolating and purifying a single or several compounds from a mixture of many substances, often having very similar molecular structures. Due to the amount of isolated product, preparative chromatography can be divided into several subgroups: semi-preparative chromatography involving the isolation of a few micrograms of analyte (usually for the assessment of initial physiological, toxicological, or pharmacological properties), laboratory-scale preparative chromatography involving the isolation of a few to several grams of analyte as an intermediate in a synthesis process, or for detailed studies of the pharmacological and pharmacokinetic mechanisms under in vivo conditions and process chromatography involving the production of large quantities (many kilograms to tons), commonly for commercialization [111].
Using a semi-preparative HPLC system equipped with a UV–VIS detector is the most versatile method to obtain pure structures from the lignan group with UV absorption monitored at 280 nm or 254 nm. Semi-preparative HPLC analysis can be conducted on a reverse-phase C18 column using an isocratic elution, as was presented during noreucol A, (+)-epiolivil, or (+)-olivil isolation from Eucomiae bark [112], lignans from Euphorbia hirta L. [113], and neolignans from Pouzolzia sanguinea (Blume) Merr. [114] using a mobile phase containing acetonitrile-H2O mixture in various proportions. Similarly, Su et al. performed the isolation of lignans from the seeds of Arctium lappa L. obtaining arctigenin, matairesinol, arctiin, and a mixture of two isomers containing lappaol A and isolappaol A, using preparative and semipreparative chromatography [43]. On the other hand, the lignans glucosides from the stems of Alibertia sessilis were isolated using the same method but with a methanolic-water mobile phase (30:70) [115]. Trachelogenin, nortrachelogenin, nortracheloside, and other lignans were obtained by Lee et al. from Trachelospermum asiaticum Nakai in amounts ranging from 94.6 mg, 25 mg, and 128.2 mg, respectively, using the same mobile phase but with the addition of 0.1% formic acid. The slight addition of acid in the mobile phase is found to be useful for the improvement of peak resolution for the structures containing hydroxyl groups [116]. In contrast, more hydrophilic deoxypodophyllotoxin, with potential anticancer properties, was isolated from Juniperus communis L. branches and needles using isocratic, but normal-phase HPLC with an n-hexane/EtOAc mixture elution (65:35) [117].
Isolation with preparative HPLC with gradient grade (85: 15 → 0: 100) proved to be an effective method to isolate epipinoresinoil, matairesinol, phylligenin, and arctigenin among the other lignans’ glucosides from Forsythia ×intermedia Zabel leaves and flowers (mobile phase consisted of 0.1% HCOOH in H2O (A) and 0.1% HCOOH in MeCN (B)) in gram quantities, as was presented by Michalak B et al. [58]. Similarly, gradient elution was used during lignans’ isolation from the seeds of Carthamus tinctorius L. The choice of water (A) and methanol (B) as the mobile phase components, at a flow rate of 1 mL/min, ultimately yielded 21.8 mg of trachelogenin, 4.2 mg of arctigenin, and 5.8 mg of matairesinol [53].
Another method used for the separation and purification of components from natural products (including lignans) is centrifugal partition chromatography (CPC), based on continuous liquid–liquid partitioning. This method has many advantages, such as the possibility of using a wide range of solvent systems, less solvent consumption, and an effective way of obtaining a large number of pure target compounds. In addition, this method allows higher mobile phase flow rates to be used, reducing analysis time compared to traditional chromatographic methods. Jeon J. et al. used CPC for the effective purification and isolation of lipid-soluble lignans, such as sesamin and sesamolin from defatted sesame seeds [118]. High-performance countercurrent chromatography (HSCCC) has also been used to purify sesamin and sesamolin from this raw material, but the process yielded product only on the milligram scale [119]. On the other hand, HSCCC proved to be an efficient method for the isolation of secoisolariciresinol diglucoside from Linum usitatissimum L., using a gradient method with a yield of 280 mg SDG per 800 mg of flaxseed [73].

8. Qualitative and Quantitative Analysis

8.1. Thin Layer Chromatography (TLC)

Thin Layer Chromatography has been applied as an inexpensive supporting method in lignan analysis, e.g., during the qualitative examination of plant extracts, fraction collection, monitoring isolation procedures, or the screening of many samples [2].
In lignan TLC analysis, silica gel is mostly used, with different eluents and detection methods. The choice of eluent, similar to the extraction solvent consideration, depends on the physicochemical properties of the lignans present in the sample. For instance, Willfor et al. [2] used dichloromethane: ethanol 93:7 (v/v) for samples containing aglycones from conifer wood. Kuehl et al. [53] analyzed samples from seeds of Carthamus tinctorius L., containing glucosides of trachelogenin, arctigenin, and matairesinol, with toluene:ethyl acetate:formic acid 3:1:0.2 (v/v/v) and dichloromethane:acetone:formic acid 7:2:0.1 (v/v/v). In another example, Lee et al. [93] studied the flower buds of Magnolia fargesii (Finet & Gagnep.) W.C. Cheng, rich in lipophilic lignan aglycones, with n-hexane:ethyl acetate 3:1 (v/v). Another example is the chloroform:methanol:water 70:30:4 (v/v/v) solvent system, which was used to analyze the roots of Eleutherococcus senticosus (Rupr. & Maxim.) Maxim. and weed of Viscum album L. [120].
Various detection techniques have been applied for lignans. As all lignans absorb UV light, plates without chemical treatment can be observed under 254 nm wavelength. Some authors also use 366 nm wavelength to check for the other compounds present in the samples or impurities. Other detection methods include spraying with coloring reagents, such as vanillin in phosphoric or sulfuric acid, and 5% sulfuric acid in ethanol [2,53,93].
Although TLC was previously also used in quantitative analysis (densitometry) as well as for isolation and purification procedures (preparative TLC), it has been nowadays replaced with more advanced methods and its use is rather limited. However, some authors, such as Zare et al. [121], used preparative TLC to purify the lignans isolated from Linum mucronatum Bertol. Similarly, Goels et al. [122] isolated pinoresinol from Picea abies (L.) H. Karst. balm. One important notion should be highlighted here, that is that the isolation using preparative TLC was executed on a microscale—the yield was 1.2 mg of pinoresinol.
One of perhaps the newer TLC methods used is thin layer chromatography—direct bio-autophagy (TLC-DB), which allows the direct on-plate identification of active components (mostly regarding antioxidative, antibacterial, and enzyme inhibition properties). Recently, Sobstyl et al. [123] performed the TLC-DB method on Schisandra chinensis (Turcz.) Baill. fruit for the analysis of plant metabolites’ acetylcholinesterase inhibition and antibacterial effects.
The High-Performance Thin Layer Chromatography (HPTLC) method was used to quantify sesamin in sesame oil without saponification, which confirms the validity of this method in lignans’ analysis.

8.2. High-Performance Liquid Chromatography (HPLC)

High-performance liquid chromatography (HPLC) seems to be the technique of choice applied for the detection, identification, separation, and quantification of lignans from plant matrixes or human fluids. The advantage of this method is, among others, that no complicated sample preparation is required. Based on the literature data, analysis of the lignans was performed by HPLC with various detection methods (Table 2), including UV detection with diode array detector (DAD), coulometric electrode array detection (CEAD), pulse amperometric detection (PAD), fluorescence detector (FLD), and mass spectrometry (MS) which has been considered the most sensitive method for lignan structures to determine their molecular mass and purity. The main advantages of combining liquid chromatography with mass spectrometry include high selectivity, resolution, speed of analyses, sensitivity, repeatability, and the possibility of quantitative analysis with analyte structure determination. In most cases, the analyses are performed with satisfying sensitivity in the UV detection at 280 nm or 254 nm.
Due to the medium polarity of lignans and their metabolites, HPLC is especially suitable for their analysis, on reverse-phase column using a gradient elution mode. The most common reversed-phase column is octadecyl—RP-18 and RP-8—for more hydrophilic structures. In their recent report, Shi et al. also point to the sporadic but successful use of normal-phase liquid chromatography to identify the lipophilic lignans (sesamin, sesamolin, and sesamol) from sesame oil [130] or lignans from the Podophyllum species [131]. On the other hand, the Podophyllum spp. lignans (podophyllotoxin, deoxypodophyllotoxin, β-peltatin, yatein, matairesinol, anhydropodorhizol) can also be analyzed in the reverse phase system [132].
The most common organic solvents used as the mobile phase in this method are acetonitrile, methanol, and water, with formic, acetic, and phosphoric acids used for pH adjustment. Some modifications, such as mixtures of solvents, e.g., methanol, acetonitrile, and DMSO are necessary for the separation of diastereoisomer combinations (methanol) and functional group derivates (acetonitrile, DMSO). Due to the chiral structure, some lignans occur in plants in enantiomeric forms or enantiomeric excess. HPLC chiral columns with the isocratic flow can be used in this case, both normal and reverse phase ones, with cellulose carbamate as the packing material [131,133].
Dar et al. described the summarized techniques employed for the separation and quantification of lignans (sesamin, sesamolin, secoisolariciresinol, secoisolariciresinol diglucoside, sesaminol triglucoside, pinoresinol, matairesinol) from flaxseed and sesame oil [134]. The main types of chromatography techniques used for quantification in sesame include the many forms of HPLC, such as LC-NMR-MS [72], methods using atmospheric pressure chemical ionization (APCI-MS) [135], and electrospray ionization mass spectrometry (ESI-MS). Hata et al. confirmed the possibility of sesamin detection from leaves using the technique of ultra-performance liquid chromatography–fluorescence detection (UPLC-FLD) [136,137]. In line with literature results, a high-performance liquid chromatography–atmospheric pressure chemical ionization–tandem mass spectrometry (HPLC–APCI–MS/MS) method in comparison to the ESI-MS model appears to be a more generic method in sesame lignans’ analysis, but slightly better for their aglycones than glucosides [136]. In contrast, the combination of an atmospheric pressure-heated nebulizer interface with chemical ionization (HN-APCI) combined with tandem mass spectrometry (HPLC-HN-APCI-MSn (MRM)) was probably the first assay to detect the essential enterolignans (enterodiol, enterolactone) in human serum and urine with a limit of detection in the low parts per billion range [138].
The dibenzocyclooctadiene structures of the lignans from Schisandra species were quantitively and qualitatively determined by high-performance liquid chromatography (HPLC) with different detection methods, such as capillary electrophoresis (CE), and mass spectrometry (MS), or by gas chromatography coupled with mass spectrometry (GC-MS). Moreover, due to the strong UV absorbance of the Schisandra lignans in the range of 230–255 nm, ultraviolet (UV) detectors can be used as a predominant method with high sensitivity and specificity [139]. Twenty-four lignans, e.g., schisanhenol, schisandrin A-C, and gomisin A-O, were also confirmed with the UHPLC–MS/MS method from fruits and leaves of Schisandra rubiflora Rehder & E.H. Wilson [128].
Another method allowing the simultaneous separation and structural identification of compounds in the mixtures or plant extracts is liquid chromatography coupled with nuclear magnetic resonance (LC-NMR). It is one of the most powerful means of compound identification, which solved problems with signal interferences by the eluents and sensitivity by solid-phase extraction cartridge (SPE) application. The reversed-phase HPLC-SPE-NMR method has found application in the identification of structurally similar lignans from Phyllanthus urinaria L. [140]. The use of LC-NMR coupled with mass spectrometer (LC-NMR-MS) was also noted in lignan analysis, for example in the structural elucidation of the isolated and purified secoisolariciresinol diglucoside diastereomers obtained from flaxseed [72].

8.3. Liquid Chromatography Mass Spectrometry

The development of hyphenated techniques related to liquid chromatography and electrospray ionization mass spectrometry (LC-ESI-MS) has provided a routine method for the identification of plant metabolites (including lignans) in complex plant matrixes. In most cases, the compounds are identified by comparing fragmentation patterns with available libraries or using reference compounds [141,142]. A valuable work on the identification of lignans by liquid chromatography–electrospray ionization ion-trap mass spectrometry was completed by Eklund et al. [141]. Table 3 presents the fragmentation data of the major lignans based on gathered reports.
The most common ionization methods used during MS/MS analysis of lignans include negative electrospray ionization (ESI), negative atmospheric pressure chemical ionization (APCI), and negative ionization in a heated nebulizer (HN).
LC-ESI-MSn (ion trap) method is a method of choice, because of its shorter run time, smaller injection, high sensitivity, and selectivity. This method is useful for detecting structures with multiple ingredients because of the lower limit of detection and ability to elucidate their structure. LC-ESI-MS can be recorded in the positive and negative ion mode, with the latter being used in most cases [153]. This is because the acid phenolic groups found in the lignan structures, have a good proton-donating capacity, and thus are easily deprotonated (e.g., secoisolariciresinol, enterodiol), apart from lignans with methylenedioxy-bridged furofuran structures which, due to the lack of phenolic hydroxyl groups, can be optimally detected in the positive ionization mode [136]. Deoxypodophyllotoxin and its precursors from Anthriscus sylvestris (L.) Hoffm. roots were analyzed using this method [154], as well as honokiol and magnolol from Magnolia officinalis Rehder & E.H. Wilson [155]. In contrast, a QTRAP LC-MS system with hybrid triple quadrupole/linear ion trap mass spectrometer has been successfully applied by Cui et al. to analyze bioactive constituents (including lignans) in different parts of Forsythia suspensa Vahl [144].
As a method of high specificity, HPLC-ESI-MS happens to be useful for the determination in the negative-ion scan mode of flaxseed lignans (secoisolariciresinol, secoisolariciresinol diglucoside) and enterolignans (enterodiol, enterolactone, and their glucuronides) from urine and blood samples, 12 and 24 h after lignan intake [156]. On the other hand, 11 lignans from Magnolia biondii Pamp. (furofuran structure) were identified with the same method using positive ion mode [126]. LC-ESI-MS also allowed for the differentiation of the metabolite profiles, including lignans (i.e., schisandrin, deoxyschisandrin, and schisandrin B), between two species: Schisandra chinensis (Turcz.) Baill. and Schisandra sphenanthera Rehder & E.H. Wilson [157]. Phyllanthin and related lignans were also analyzed by UPLC-ESI-MS, using 0.1% formic acid in methanol and 0.1% formic acid in water as the mobile phases. Five lignans (two new, and three known structures) isolated from Zanthoxylum armatum DC peels, as well as another five lignans (euphorhirtins A-D, 5-methoxyvirgatusin) from Euphorbia hirta L. were analyzed using high resolution mass spectrometry (HRMS). Two dibenzocyclooctadiene lignans, i.e., schisphenlignan M and N, from ethanol extract of the root bark of Schisandra sphenanthera Rehder & E.H. Wilson [158] and lignans from Ginkgo biloba L. roots were analyzed using the same method [159].
HPLC-TOF-MS and HPLC-DAD-ESI-QTOF-MS analysis allowed the identification of nudiposide, a lignan from avocado (Persea americana Mill.) seed and seed coat [160], as well as lignans (arctiin, arctigenin, matairesinol) and sesquilignans (lappaol C, isolappaol C, lappaol A, isolappaol A) from Cnicus benedictus L. fruit samples [161], and olivil-type lignans among other types from Eucommia ulmoides Oliv. bark [112]. UPLC-ESI-QTOF-MS proved to be a good option for confirming the presence of dibenzylbutyrolactone lignans from Great Burdock seeds (Arctium lappa L.) [162] and furofuran-type lignans (pinoresinol, epipinoresinol) from Carduus nutans L. fruits [163]. Doussot et al. for the first time successfully analyzed the content of lignans from wild flax species, such as Linum flavum L., as well as Juniperus and Callitris species for the content of lignan podophyllotoxin and its precursors by liquid chromatography coupled with high-resolution mass spectrometer (LC-HRMS-QTOF-MS) [145].

8.4. Gas Chromatography Mass Spectrometry

Gas chromatography (GC) is a less popular method for the identification of phenolic compounds from the lignan group due to their low volatility and the need for high temperatures, which can damage the analytes [164]. On the other hand, this method allows the identification of multicomponent mixtures (e.g., polyphenols), combined with a detection system (usually a mass spectrometer—GC-MS) giving unambiguous information about its composition, and thus can be an alternative to LC-MS.
Although gas chromatography coupled with mass spectrometry (GC-MS) is far from being a method of choice for lignans’ analyses, the attempts to identify some of them (e.g., pinoresinol, matairesinol, sesamin, asarinin, sesamolin, medioresinol secoisolariciresinol, isolariciresinol, anhydrosecoisolariciresinol, lariciresinol, isolariciresinol, and syringaresinol among others) by GC-MS can be found in the literature [165,166,167,168,169].
The biggest obstacle in non-volatile lignans’ analysis by GC-MS is the need for sample derivatization to make the functional groups (i.e., hydroxyl groups) detectable. Furthermore, the derivatization process guarantees an increase in the thermal stability of the analytes to prevent their degradation under analytical conditions, thereby increasing the sensitivity and specificity of the assay. The most common techniques for the derivatization of lignan compounds include silylation and acylation or combinations of both methods. As an example of the reagents used to derivatize labile hydroxyl groups, for example in extracts from pomegranate by-products [168] and juices or flaxseed extracts [166], is BSTFA reagent (N,O-Bis(trimethylsilyl)trifluoroacetamide), which among others requires the longest time for the derivatization process [166]. Other common silylation mixtures include pyridine with BSTFA containing 1% trimethylchlorosilane (TMCS) [166], hexamethyldisilazane/trimethylchlorosilane in pyridine 2:1:10 (Tri-Sil reagent), N,O-bis(trimethylsilyl)acetamide (BSA), deuterated BSA [170], or pentafluoropropionic anhydride (PFPA) [171]. During the pinoresinol, matairesinol, and secoisolariciresinol analysis in various food samples, a Tri-Sil reagent was used in the derivatization process [165]. Some specific lignans, e.g., dibenzocyclooctadiene type from the fruit of the Schisandrae spp. and podophyllotoxin-type lignans, can be detected directly without the need for derivatization because of the lack of hydroxyl groups [172,173].
The accurate quantification of the key human lignan metabolites, such as enterodiol and enterolactone, is drawing attention due to their beneficial protective effects in various diseases. The successful detection of enterolignans in human plasma has been accomplished primarily using liquid chromatography–mass spectrometry (LC-MS) or gas chromatography–mass spectrometry (GC-MS) operated in full scan mode [174,175]. Edel et al. proposed a successful method for the precise quantification of enterolignans after consumption of flaxseed using supported liquid extraction (SLE) with quantification using gas chromatography and mass spectrometry in the micro-selected ion storage mode (GC-MS-μSIS) [176].
During GC-MS lignans’ analyses, helium is commonly used as the carrier gas at a constant flow rate of 1 mL/min [165,176]. According to the literature data, analyses are carried out with a gradual increase in the detector temperature from 270–280 °C [165,166,176], up to 320–340 °C [175], in total ion current (TIC) mode and/or selected ion monitoring (SIM) mode. The most common choice of a chromatography column is standard capillary non-polar columns (DB-17, HP-1) [177] or columns with mid polarity i.e., HP-5 [166], DB-17 [172], and VF-5 [176]. The electron impact (EI) is the most frequently chosen ionization technique in the GC-MS analysis of lignans and enterolignans [176].
An interesting addition to the methods of identifying the presence of specific compounds in the plant is the direct visualization of their distribution in individual parts of the raw material. Yu et al. first performed visualization of lignans’ and neolignans’ compounds (olivil, pinoresinol, lariciresinol, secoisolariciresinol, and their glucosides) in Gingko spp. stems using cryo-temporal secondary ion mass spectrometry coupled with scanning electron microscopy (cryo-TOF-SIMS/SEM). This method appears to be an essential element necessary to understand their biological functions by allowing visualization of their distribution with submicron transverse resolution in two-dimensional images [178].

9. Conclusions

This review provided a summary of both the conventional and alternative methods used to extract lignans from plant material. As isolation of pure compounds may be time-consuming, complicated, expensive, and impact the environment, an appropriate technique should be selected. Among the presented methods, alternative or advanced methods such as UAE, MAE, and SFE are rapid, efficient, consume fewer extraction solvents, and are environmentally friendly. Analytical methods for detecting lignans have also been discussed, especially emphasizing the advantages of hyphenated methods, such as LC-MS and GC-MS. Since most lignans have free hydroxyl groups, LC-MS should be regarded as a superior technique in most cases. Nevertheless, TLC and other conventional techniques still have a role as auxiliary methods in lignans’ analysis.
In summary, many previous reports have shown that lignans may interfere with cancer cells or even offer a means of preventing carcinogenic diseases. These compounds are present in the everyday human diet and thus should be more comprehensively studied to develop new therapies. A properly selected method for extracting and isolating lignans from the plant material may facilitate this process.

Author Contributions

Conceptualization: A.P.; methodology, A.P. and A.K.K.; resources, A.P., M.K.-J., and O.J.; writing—original draft preparation, A.P. and M.K.-J.; writing—review and editing, A.K.K. and A.P.; visualization, A.P. and O.J.; supervision, A.K.K.; project administration, A.P.; funding acquisition, A.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Science Centre, Poland, 2021/41/N/NZ7/00313. For the purpose of Open Access, the authors have applied a CC-BY public copyright license to any Author Accepted Manuscript (AAM) version arising from this submission.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Marvin was used for drawing, displaying, and characterizing chemical structures, substructures, and reactions, Marvin 17.21.0, Chemaxon (https://www.chemaxon.com), accessed on 4 July 2022.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rodriguez-Garcia, C.; Sanchez-Quesada, C.; Toledo, E.; Delgado-Rodriguez, M.; Gaforio, J.J. Naturally Lignan-Rich Foods: A Dietary Tool for Health Promotion? Molecules 2019, 24, 917. [Google Scholar] [CrossRef] [PubMed]
  2. Willfor, S.M.; Smeds, A.I.; Holmbom, B.R. Chromatographic analysis of lignans. J. Chromatogr. A 2006, 1112, 64–77. [Google Scholar] [CrossRef] [PubMed]
  3. Ward, R.S. Recent Advances in the Chemistry of Lignans. In Studies in Natural Products Chemistry; Attaur, R., Ed.; Elsevier: Amsterdam, The Netherlands, 2000; Volume 24, pp. 739–798. [Google Scholar]
  4. Ferrer, J.L.; Austin, M.B.; Stewart, C., Jr.; Noel, J.P. Structure and function of enzymes involved in the biosynthesis of phenylpropanoids. Plant Physiol. Biochem. 2008, 46, 356–370. [Google Scholar] [CrossRef] [PubMed]
  5. Kato, M.J.; Chu, A.; Davin, L.B.; Lewis, N.G. Biosynthesis of antioxidant lignans in Sesamum indicum seeds. Phytochemistry 1998, 47, 583–591. [Google Scholar] [CrossRef]
  6. Lewis, N.G.; Davin, L.B.; Sarkanen, S. Lignin and Lignan Biosynthesis: Distinctions and Reconciliations. In Lignin and Lignan Biosynthesis; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 1998; Volume 697, pp. 1–27. [Google Scholar]
  7. Nakatsubo, T.; Mizutani, M.; Suzuki, S.; Hattori, T.; Umezawa, T. Characterization of Arabidopsis thaliana pinoresinol reductase, a new type of enzyme involved in lignan biosynthesis. J. Biol. Chem. 2008, 283, 15550–15557. [Google Scholar] [CrossRef]
  8. Noguchi, A.; Fukui, Y.; Iuchi-Okada, A.; Kakutani, S.; Satake, H.; Iwashita, T.; Nakao, M.; Umezawa, T.; Ono, E. Sequential glucosylation of a furofuran lignan, (+)-sesaminol, by Sesamum indicum UGT71A9 and UGT94D1 glucosyltransferases. Plant J. 2008, 54, 415–427. [Google Scholar] [CrossRef]
  9. Ono, E.; Nakai, M.; Fukui, Y.; Tomimori, N.; Fukuchi-Mizutani, M.; Saito, M.; Satake, H.; Tanaka, T.; Katsuta, M.; Umezawa, T.; et al. Formation of two methylenedioxy bridges by a Sesamum CYP81Q protein yielding a furofuran lignan, (+)-sesamin. Proc. Natl. Acad. Sci. USA 2006, 103, 10116–10121. [Google Scholar] [CrossRef]
  10. Umezawa, T. Diversity in lignan biosynthesis. Phytochem. Rev. 2003, 2, 371–390. [Google Scholar] [CrossRef]
  11. Dewick, P.M. Medicinal Natural Products: A Biosynthetic Approach; Wiley: Chichester, UK, 2012. [Google Scholar]
  12. Zalesak, F.; Bon, D.J.D.; Pospisil, J. Lignans and Neolignans: Plant secondary metabolites as a reservoir of biologically active substances. Pharm. Res 2019, 146, 104284. [Google Scholar] [CrossRef]
  13. Adlercreutz, H. Lignans and Human Health. Crit. Rev. Clin. Lab. Sci. 2007, 44, 483–525. [Google Scholar] [CrossRef]
  14. Flower, G.; Fritz, H.; Balneaves, L.G.; Verma, S.; Skidmore, B.; Fernandes, R.; Kennedy, D.; Cooley, K.; Wong, R.; Sagar, S.; et al. Flax and Breast Cancer:A Systematic Review. Integr. Cancer Ther. 2014, 13, 181–192. [Google Scholar] [CrossRef] [PubMed]
  15. DeLuca, J.A.A.; Garcia-Villatoro, E.L.; Allred, C.D. Flaxseed Bioactive Compounds and Colorectal Cancer Prevention. Curr. Oncol. Rep. 2018, 20, 59. [Google Scholar] [CrossRef] [PubMed]
  16. Shabgah, A.G.; Suksatan, W.; Achmad, M.H.; Bokov, D.O.; Abdelbasset, W.K.; Ezzatifar, F.; Hemmati, S.; Mohammadi, H.; Soleimani, D.; Jadidi-Niaragh, F.; et al. Arctigenin, an anti-tumor agent; a cutting-edge topic and up-to-the-minute approach in cancer treatment. Eur. J. Pharmacol. 2021, 909, 174419. [Google Scholar] [CrossRef] [PubMed]
  17. Paul, S.; Patra, D.; Kundu, R. Lignan enriched fraction (LRF) of Phyllanthus amarus promotes apoptotic cell death in human cervical cancer cells in vitro. Sci. Rep. 2019, 9, 14950. [Google Scholar] [CrossRef] [PubMed]
  18. Hwang, D.; Shin, S.Y.; Lee, Y.; Hyun, J.; Yong, Y.; Park, J.C.; Lee, Y.H.; Lim, Y. A compound isolated from Schisandra chinensis induces apoptosis. Bioorg. Med. Chem. Lett. 2011, 21, 6054–6057. [Google Scholar] [CrossRef]
  19. Shen, L.; Zhang, F.; Huang, R.; Yan, J.; Shen, B. Honokiol inhibits bladder cancer cell invasion through repressing SRC-3 expression and epithelial-mesenchymal transition. Oncol. Lett. 2017, 14, 4294–4300. [Google Scholar] [CrossRef]
  20. Zidorn, C. Guidelines for consistent characterisation and documentation of plant source materials for studies in phytochemistry and phytopharmacology. Phytochemistry 2017, 139, 56–59. [Google Scholar] [CrossRef]
  21. Holmbom, B.; Willfoer, S.; Hemming, J.; Pietarinen, S.; Nisula, L.; Eklund, P.; Sjoeholm, R. Knots in Trees: A Rich Source of Bioactive Polyphenols. In Materials, Chemicals, and Energy from Forest Biomass; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 2007; pp. 350–362. [Google Scholar]
  22. Mansikkala, T.; Patanen, M.; Kärkönen, A.; Korpinen, R.; Pranovich, A.; Ohigashi, T.; Swaraj, S.; Seitsonen, J.; Ruokolainen, J.; Huttula, M.; et al. Lignans in Knotwood of Norway Spruce: Localisation with Soft X-ray Microscopy and Scanning Transmission Electron Microscopy with Energy Dispersive X-ray Spectroscopy. Molecules 2020, 25, 2997. [Google Scholar] [CrossRef]
  23. Hosseinian, F.S.; Beta, T. Patented techniques for the extraction and isolation of secoisolari-ciresinol diglucoside from flaxseed. Recent Pat. Food Nutr. Agric. 2009, 1, 25–31. [Google Scholar] [CrossRef]
  24. Khoddami, A.; Wilkes, M.A.; Roberts, T.H. Techniques for analysis of plant phenolic compounds. Molecules 2013, 18, 2328–2375. [Google Scholar] [CrossRef]
  25. Liu, J.; Cai, Y.Z.; Wong, R.N.; Lee, C.K.; Tang, S.C.; Sze, S.C.; Tong, Y.; Zhang, Y. Comparative analysis of caffeoylquinic acids and lignans in roots and seeds among various burdock (Arctium lappa) genotypes with high antioxidant activity. J. Agric. Food Chem. 2012, 60, 4067–4075. [Google Scholar] [CrossRef] [PubMed]
  26. Mahendra Kumar, C.; Singh, S.A. Bioactive lignans from sesame (Sesamum indicum L.): Evaluation of their antioxidant and antibacterial effects for food applications. J. Food Sci. Technol. 2015, 52, 2934–2941. [Google Scholar] [CrossRef] [PubMed]
  27. Gerstenmeyer, E.; Reimer, S.; Berghofer, E.; Schwartz, H.; Sontag, G. Effect of thermal heating on some lignans in flax seeds, sesame seeds and rye. Food Chem. 2013, 138, 1847–1855. [Google Scholar] [CrossRef] [PubMed]
  28. Hu, J.; Shi, Y.; Yang, B.; Dong, Z.; Si, X.; Qin, K. Changes in chemical components and antitumor activity during the heating process of Fructus Arctii. Pharm. Biol. 2019, 57, 363–368. [Google Scholar] [CrossRef]
  29. Qin, K.; Liu, Q.; Cai, H.; Cao, G.; Lu, T.; Shen, B.; Shu, Y.; Cai, B. Chemical analysis of raw and processed Fructus arctii by high-performance liquid chromatography/diode array detection-electrospray ionization-mass spectrometry. Pharm. Mag. 2014, 10, 541–546. [Google Scholar] [CrossRef]
  30. Chen, Y.; Lin, H.; Lin, M.; Zheng, Y.; Chen, J. Effect of roasting and in vitro digestion on phenolic profiles and antioxidant activity of water-soluble extracts from sesame. Food Chem. Toxicol. 2020, 139, 111239. [Google Scholar] [CrossRef]
  31. Kawamura, F.; Miyachi, M.; Kawai, S.; Ohashi, H. Photodiscoloration of western hemlock (Tsuga heterophylla) sapwood III Early stage of photodiscoloration reaction with lignans. J. Wood Sci. 1998, 44, 47–55. [Google Scholar] [CrossRef]
  32. Yuan, J.-P.; Li, X.; Xu, S.-P.; Wang, J.-H.; Liu, X. Hydrolysis Kinetics of Secoisolariciresinol Diglucoside Oligomers from Flaxseed. J. Agric. Food Chem. 2008, 56, 10041–10047. [Google Scholar] [CrossRef]
  33. Renouard, S.; Hano, C.; Corbin, C.; Fliniaux, O.; Lopez, T.; Montguillon, J.; Barakzoy, E.; Mesnard, F.; Lamblin, F.; Lainé, E. Cellulase-assisted release of secoisolariciresinol from extracts of flax (Linum usitatissimum) hulls and whole seeds. Food Chem. 2010, 122, 679–687. [Google Scholar] [CrossRef]
  34. Sicilia, T.; Niemeyer, H.B.; Honig, D.M.; Metzler, M. Identification and stereochemical characterization of lignans in flaxseed and pumpkin seeds. J. Agric. Food Chem. 2003, 51, 1181–1188. [Google Scholar] [CrossRef]
  35. Pilkington, L.I. Lignans: A Chemometric Analysis. Molecules 2018, 23, 1666. [Google Scholar] [CrossRef] [PubMed]
  36. Grougnet, R.; Magiatis, P.; Laborie, H.; Lazarou, D.; Papadopoulos, A.; Skaltsounis, A.L. Sesamolinol glucoside, disaminyl ether, and other lignans from sesame seeds. J. Agric. Food Chem. 2012, 60, 108–111. [Google Scholar] [CrossRef]
  37. Yang, F.; Yang, L.; Wang, W.; Liu, Y.; Zhao, C.; Zu, Y. Enrichment and purification of syringin, eleutheroside E and isofraxidin from Acanthopanax senticosus by macroporous resin. Int. J. Mol. Sci. 2012, 13, 8970–8986. [Google Scholar] [CrossRef]
  38. Wang, Z.; Zhang, L.; Sun, Y. Semipreparative separation and determination of eleutheroside E in Acanthopanax giraldii Harms by high-performance liquid chromatography. J. Chromatogr. Sci. 2005, 43, 249–252. [Google Scholar] [CrossRef]
  39. Benković, E.T.; Grohar, T.; Žigon, D.; Švajger, U.; Janeš, D.; Kreft, S.; Štrukelj, B. Chemical composition of the silver fir (Abies alba) bark extract Abigenol® and its antioxidant activity. Ind. Crops Prod. 2014, 52, 23–28. [Google Scholar] [CrossRef]
  40. Jahagirdar, A.; Usharani, D.; Srinivasan, M.; Rajasekharan, R. Sesaminol diglucoside, a water-soluble lignan from sesame seeds induces brown fat thermogenesis in mice. Biochem. Biophys. Res. Commun. 2018, 507, 155–160. [Google Scholar] [CrossRef] [PubMed]
  41. Fang, W.; Hemming, J.; Reunanen, M.; Eklund, P.; Pineiro, E.C.; Poljanšek, I.; Oven, P.; Willför, S. Evaluation of selective extraction methods for recovery of polyphenols from pine. Holzforschung 2013, 67, 843–851. [Google Scholar] [CrossRef]
  42. Liu, H.; Zhang, Y.; Sun, Y.; Wang, X.; Zhai, Y.; Sun, Y.; Sun, S.; Yu, A.; Zhang, H.; Wang, Y. Determination of the major constituents in fruit of Arctium lappa L. by matrix solid-phase dispersion extraction coupled with HPLC separation and fluorescence detection. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2010, 878, 2707–2711. [Google Scholar] [CrossRef]
  43. Su, S.; Wink, M. Natural lignans from Arctium lappa as antiaging agents in Caenorhabditis elegans. Phytochemistry 2015, 117, 340–350. [Google Scholar] [CrossRef]
  44. Tokar, M.; Klimek, B. The content of lignan glycosides in Forsythia flowers and leaves. Acta Pol. Pharm. 2004, 61, 273–278. [Google Scholar]
  45. Kuo, P.C.; Chen, G.F.; Yang, M.L.; Lin, Y.H.; Peng, C.C. Chemical constituents from the fruits of Forsythia suspensa and their antimicrobial activity. Biomed. Res. Int. 2014, 2014, 304830. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Li, Y.-L.; Yang, X.-W.; Zhang, W.-D. Stilbenes, lignans, and phenols from Abies chensiensis. Biochem. Syst. Ecol. 2008, 36, 932–934. [Google Scholar] [CrossRef]
  47. Li, Y.L.; Gao, Y.X.; Jin, H.Z.; Shan, L.; Liang, X.S.; Xu, X.K.; Yang, X.W.; Wang, N.; Steinmetz, A.; Chen, Z.; et al. Chemical constituents of Abies nukiangensis. Phytochemistry 2014, 106, 116–123. [Google Scholar] [CrossRef]
  48. Zhang, H.; Gao, Y.; Zhang, J.; Wang, K.; Jin, T.; Wang, H.; Ruan, K.; Wu, F.; Xu, Z. The effect of total lignans from Fructus Arctii on Streptozotocin-induced diabetic retinopathy in Wistar rats. J. Ethnopharmacol. 2020, 255, 112773. [Google Scholar] [CrossRef]
  49. Lu, H.; Sun, Z.; Shan, H.; Song, J. Microwave-Assisted Extraction and Purification of Arctiin and Arctigenin from Fructus Arctii by High-Speed Countercurrent Chromatography. J. Chromatogr. Sci. 2016, 54, 472–478. [Google Scholar] [CrossRef] [PubMed]
  50. He, J.; Huang, X.Y.; Yang, Y.N.; Feng, Z.M.; Jiang, J.S.; Zhang, P.C. Two new compounds from the fruits of Arctium lappa. J. Asian Nat. Prod. Res. 2016, 18, 423–428. [Google Scholar] [CrossRef] [PubMed]
  51. Yang, Y.N.; Huang, X.Y.; Feng, Z.M.; Jiang, J.S.; Zhang, P.C. New Butyrolactone Type Lignans from Arctii Fructus and Their Anti-inflammatory Activities. J. Agric. Food Chem. 2015, 63, 7958–7966. [Google Scholar] [CrossRef]
  52. Park, S.Y.; Hong, S.S.; Han, X.H.; Hwang, J.S.; Lee, D.; Ro, J.S.; Hwang, B.Y. Lignans from Arctium lappa and their inhibition of LPS-induced nitric oxide production. Chem. Pharm. Bull. 2007, 55, 150–152. [Google Scholar] [CrossRef]
  53. Kuehnl, S.; Schroecksnadel, S.; Temml, V.; Gostner, J.M.; Schennach, H.; Schuster, D.; Schwaiger, S.; Rollinger, J.M.; Fuchs, D.; Stuppner, H. Lignans from Carthamus tinctorius suppress tryptophan breakdown via indoleamine 2,3-dioxygenase. Phytomedicine 2013, 20, 1190–1195. [Google Scholar] [CrossRef]
  54. Lee, S.; Ban, H.S.; Kim, Y.P.; Kim, B.K.; Cho, S.H.; Ohuchi, K.; Shin, K.H. Lignans from Acanthopanax chiisanensis having an inhibitory activity on prostaglandin E2 production. Phytother. Res. 2005, 19, 103–106. [Google Scholar] [CrossRef]
  55. Jin, L.; Schmiech, M.; El Gaafary, M.; Zhang, X.; Syrovets, T.; Simmet, T. A comparative study on root and bark extracts of Eleutherococcus senticosus and their effects on human macrophages. Phytomedicine 2020, 68, 153181. [Google Scholar] [CrossRef] [PubMed]
  56. Li, T.; Ferns, K.; Yan, Z.Q.; Yin, S.Y.; Kou, J.J.; Li, D.; Zeng, Z.; Yin, L.; Wang, X.; Bao, H.X.; et al. Acanthopanax senticosus: Photochemistry and Anticancer Potential. Am. J. Chin. Med. 2016, 44, 1543–1558. [Google Scholar] [CrossRef] [PubMed]
  57. Lee, D.Y.; Seo, K.H.; Jeong, R.H.; Lee, S.M.; Kim, G.S.; Noh, H.J.; Kim, S.Y.; Kim, G.W.; Kim, J.Y.; Baek, N.I. Anti-inflammatory lignans from the fruits of Acanthopanax sessiliflorus. Molecules 2012, 18, 41–49. [Google Scholar] [CrossRef] [PubMed]
  58. Michalak, B.; Filipek, A.; Chomicki, P.; Pyza, M.; Wozniak, M.; Zyzynska-Granica, B.; Piwowarski, J.P.; Kicel, A.; Olszewska, M.A.; Kiss, A.K. Lignans From Forsythia x Intermedia Leaves and Flowers Attenuate the Pro-inflammatory Function of Leukocytes and Their Interaction with Endothelial Cells. Front. Pharm. 2018, 9, 401. [Google Scholar] [CrossRef]
  59. Lee, Y.G.; Jang, S.A.; Seo, K.H.; Gwag, J.E.; Kim, H.G.; Ko, J.H.; Ji, S.A.; Kang, S.C.; Lee, D.Y.; Baek, N.I. New Lignans from the Flower of Forsythia koreana and Their Suppression Effect on VCAM-1 Expression in MOVAS Cells. Chem. Biodivers. 2018, 15, e1800026. [Google Scholar] [CrossRef]
  60. Lim, H.; Lee, J.G.; Lee, S.H.; Kim, Y.S.; Kim, H.P. Anti-inflammatory activity of phylligenin, a lignan from the fruits of Forsythia koreana, and its cellular mechanism of action. J. Ethnopharmacol. 2008, 118, 113–117. [Google Scholar] [CrossRef]
  61. Kang, H.S.; Lee, J.Y.; Kim, C.J. Anti-inflammatory activity of arctigenin from Forsythiae Fructus. J. Ethnopharmacol. 2008, 116, 305–312. [Google Scholar] [CrossRef]
  62. Kim, C.Y.; Ahn, M.J.; Kim, J. A preparative isolation and purification of arctigenin and matairesinol from Forsythia koreana by centrifugal partition chromatography. J. Sep. Sci. 2006, 29, 656–659. [Google Scholar] [CrossRef]
  63. Chang, M.J.; Hung, T.M.; Min, B.S.; Kim, J.C.; Woo, M.H.; Choi, J.S.; Lee, H.K.; Bae, K. Lignans from the Fruits of Forsythia suspensa (Thunb.) Vahl Protect High-Density Lipoprotein during Oxidative Stress. Biosci. Biotechnol. Biochem. 2008, 72, 2750–2755. [Google Scholar] [CrossRef]
  64. Guo, H.; Liu, A.H.; Ye, M.; Yang, M.; Guo, D.A. Characterization of phenolic compounds in the fruits of Forsythia suspensa by high-performance liquid chromatography coupled with electrospray ionization tandem mass spectrometry. Rapid Commun. Mass Spectrom. 2007, 21, 715–729. [Google Scholar] [CrossRef]
  65. Li, C.; Dai, Y.; Duan, Y.-H.; Liu, M.-L.; Yao, X.-S. A new lignan glycoside from Forsythia suspensa. Chin. J. Nat. Med. 2014, 12, 697–699. [Google Scholar] [CrossRef]
  66. Kuo, P.C.; Hung, H.Y.; Nian, C.W.; Hwang, T.L.; Cheng, J.C.; Kuo, D.H.; Lee, E.J.; Tai, S.H.; Wu, T.S. Chemical Constituents and Anti-inflammatory Principles from the Fruits of Forsythia suspensa. J. Nat. Prod. 2017, 80, 1055–1064. [Google Scholar] [CrossRef] [PubMed]
  67. Liang, J.; Gong, F.Q.; Sun, H.M. Simultaneous Separation of Eight Lignans in Forsythia suspensa by beta-Cyclodextrin-Modified Capillary Zone Electrophoresis. Molecules 2018, 23, 514. [Google Scholar] [CrossRef] [PubMed]
  68. Huh, J.; Song, J.H.; Kim, S.R.; Cho, H.M.; Ko, H.J.; Yang, H.; Sung, S.H. Lignan Dimers from Forsythia viridissima Roots and Their Antiviral Effects. J. Nat. Prod. 2019, 82, 232–238. [Google Scholar] [CrossRef]
  69. Shang, N.; Guerrero-Analco, J.A.; Musallam, L.; Saleem, A.; Muhammad, A.; Walshe-Roussel, B.; Cuerrier, A.; Arnason, J.T.; Haddad, P.S. Adipogenic constituents from the bark of Larix laricina du Roi (K. Koch; Pinaceae), an important medicinal plant used traditionally by the Cree of Eeyou Istchee (Quebec, Canada) for the treatment of type 2 diabetes symptoms. J. Ethnopharmacol. 2012, 141, 1051–1057. [Google Scholar] [CrossRef] [PubMed]
  70. Waszkowiak, K.; Gliszczynska-Swiglo, A.; Barthet, V.; Skrety, J. Effect of Extraction Method on the Phenolic and Cyanogenic Glucoside Profile of Flaxseed Extracts and their Antioxidant Capacity. J. Am. Oil Chem. Soc. 2015, 92, 1609–1619. [Google Scholar] [CrossRef]
  71. Beejmohun, V.; Fliniaux, O.; Grand, E.; Lamblin, F.; Bensaddek, L.; Christen, P.; Kovensky, J.; Fliniaux, M.A.; Mesnard, F. Microwave-assisted extraction of the main phenolic compounds in flaxseed. Phytochem. Anal. 2007, 18, 275–282. [Google Scholar] [CrossRef]
  72. Fritsche, J.; Angoelal, R.; Dachtler, M. On-line liquid-chromatography-nuclear magnetic resonance spectroscopy-mass spectrometry coupling for the separation and characterization of secoisolariciresinol diglucoside isomers in flaxseed. J. Chromatogr. A 2002, 972, 195–203. [Google Scholar] [CrossRef]
  73. Degenhardt, A.; Habben, S.; Winterhalter, P. Isolation of the lignan secoisolariciresinol diglucoside from flaxseed (Linum usitatissimum L.) by high-speed counter-current chromatography. J. Chromatogr. A 2002, 943, 299–302. [Google Scholar] [CrossRef]
  74. Li, X.; Yuan, J.P.; Xu, S.P.; Wang, J.H.; Liu, X. Separation and determination of secoisolariciresinol diglucoside oligomers and their hydrolysates in the flaxseed extract by high-performance liquid chromatography. J. Chromatogr. A 2008, 1185, 223–232. [Google Scholar] [CrossRef]
  75. Corbin, C.; Fidel, T.; Leclerc, E.A.; Barakzoy, E.; Sagot, N.; Falguieres, A.; Renouard, S.; Blondeau, J.P.; Ferroud, C.; Doussot, J.; et al. Development and validation of an efficient ultrasound assisted extraction of phenolic compounds from flax (Linum usitatissimum L.) seeds. Ultrason. Sonochem. 2015, 26, 176–185. [Google Scholar] [CrossRef] [PubMed]
  76. Hassan Mekky, R.; Abdel-Sattar, E.; Segura-Carretero, A.; Contreras, M.d.M. A comparative study on the metabolites profiling of linseed cakes from Egyptian cultivars and antioxidant activity applying mass spectrometry-based analysis and chemometrics. Food Chem. 2022, 395, 133524. [Google Scholar] [CrossRef] [PubMed]
  77. Lee, D.H.; Kwon, S.Y.; Woo, M.H.; Lee, J.-H.; Son, K.H. Phytochemical Studies on Magnoliae Flos (I) Isolation of Lignans from the Flower Buds of Magnolia biondii. Nat. Prod. Sci. 2013, 19, 160–165. [Google Scholar]
  78. Nguyen, T.T.M.; Lee, H.S.; Nguyen, T.T.; Ngo, T.Q.M.; Jun, C.D.; Min, B.S.; Kim, J.A. Four New Lignans and IL-2 Inhibitors from Magnoliae Flos. Chem. Pharm. Bull. 2017, 65, 840–847. [Google Scholar] [CrossRef] [PubMed]
  79. Aspé, E.; Fernández, K. The effect of different extraction techniques on extraction yield, total phenolic, and anti-radical capacity of extracts from Pinus radiata Bark. Ind. Crops Prod. 2011, 34, 838–844. [Google Scholar] [CrossRef]
  80. Willför, S.; Hemming, J.; Reunanen, M.; Holmbom, B. Phenolic and Lipophilic Extractives in Scots Pine Knots and Stemwood. Wood Res. Technol. 2003, 57, 359–372. [Google Scholar] [CrossRef]
  81. Ekman, R.; Willför, S.; Sjöholm, R.; Reunanen, M.; Mäki, J.; Lehtilä, R.; Eckerman, C. Identification of the Lignan Nortrachelogenin in Knot and Branch Heartwood of Scots Pine (Pinus sylvestris L.). Wood Res. Technol. 2002, 56, 253–256. [Google Scholar] [CrossRef]
  82. Mishra, S.; Aeri, V. Optimization of microwave-assisted extraction conditions for preparing lignan-rich extract from Saraca asoca bark using Box-Behnken design. Pharm. Biol. 2016, 54, 1255–1262. [Google Scholar] [CrossRef]
  83. Kwon, D.Y.; Kim, D.S.; Yang, H.J.; Park, S. The lignan-rich fractions of Fructus Schisandrae improve insulin sensitivity via the PPAR-gamma pathways in in vitro and in vivo studies. J. Ethnopharmacol. 2011, 135, 455–462. [Google Scholar] [CrossRef]
  84. Dai, Z.; Xin, H.; Fu, Q.; Hao, H.; Li, Q.; Liu, Q.; Jin, Y. Exploration and optimization of conditions for quantitative analysis of lignans in Schisandra chinensis by an online supercritical fluid extraction with supercritical fluid chromatography system. J. Sep. Sci. 2019, 42, 2444–2454. [Google Scholar] [CrossRef]
  85. Razgonova, M.; Zakharenko, A.; Pikula, K.; Kim, E.; Chernyshev, V.; Ercisli, S.; Cravotto, G.; Golokhvast, K. Rapid Mass Spectrometric Study of a Supercritical CO(2)-extract from Woody Liana Schisandra chinensis by HPLC-SPD-ESI-MS/MS. Molecules 2020, 25, 2689. [Google Scholar] [CrossRef]
  86. Zhu, X.; Zhang, X.; Sun, Y.; Su, D.; Sun, Y.; Hu, B.; Zeng, X. Purification and fermentation in vitro of sesaminol triglucoside from sesame cake by human intestinal microbiota. J. Agric. Food Chem. 2013, 61, 1868–1877. [Google Scholar] [CrossRef] [PubMed]
  87. Kuo, P.C.; Lin, M.C.; Chen, G.F.; Yiu, T.J.; Tzen, J.T. Identification of methanol-soluble compounds in sesame and evaluation of antioxidant potential of its lignans. J. Agric. Food Chem. 2011, 59, 3214–3219. [Google Scholar] [CrossRef] [PubMed]
  88. Mekky, R.H.; Abdel-Sattar, E.; Segura-Carretero, A.; Contreras, M.D.M. Phenolic Compounds from Sesame Cake and Antioxidant Activity: A New Insight for Agri-Food Residues’ Significance for Sustainable Development. Foods 2019, 8, 432. [Google Scholar] [CrossRef] [PubMed]
  89. Feng, X.; Su, G.; Ye, Y.; Zhang, R.; Yang, X.; Du, B.; Peng, B.; Tu, P.; Chai, X. Alashinols F and G, two lignans from stem bark of Syringa pinnatifolia. Nat. Prod. Res. 2017, 31, 1555–1560. [Google Scholar] [CrossRef]
  90. Shao, L.W.; Wang, C.H.; Li, G.Q.; Huang, X.J.; Li, Z.; Wang, G.C. A new lignan from the roots of Syringa pinnatifolia. Nat. Prod. Res. 2014, 28, 1894–1899. [Google Scholar] [CrossRef]
  91. Wozniak, M.; Michalak, B.; Wyszomierska, J.; Dudek, M.K.; Kiss, A.K. Effects of Phytochemically Characterized Extracts From Syringa vulgaris and Isolated Secoiridoids on Mediators of Inflammation in a Human Neutrophil Model. Front. Pharm. 2018, 9, 349. [Google Scholar] [CrossRef]
  92. Tiwari, P.; Kaur, M.; Kaur, H. Phytochemical screening and Extraction: A Review. Int. Pharm. Sci. 2011, 1, 98–106. [Google Scholar]
  93. Lee, S.U.; Ryu, H.W.; Lee, S.; Shin, I.S.; Choi, J.H.; Lee, J.W.; Lee, J.; Kim, M.O.; Lee, H.J.; Ahn, K.S.; et al. Lignans Isolated from Flower Buds of Magnolia fargesii Attenuate Airway Inflammation Induced by Cigarette Smoke in vitro and in vivo. Front. Pharm. 2018, 9, 970. [Google Scholar] [CrossRef]
  94. Shi, X.; Yang, Y.; Ren, H.; Sun, S.; Mu, L.t.; Chen, X.; Wang, Y.; Zhang, Y.; Wang, L.h.; Sun, C. Identification of multiple components in deep eutectic solvent extract of Acanthopanax senticosus root by ultra-high-performance liquid chromatography with quadrupole orbitrap mass spectrometry. Phytochem. Lett. 2020, 35, 175–185. [Google Scholar] [CrossRef]
  95. Vinatoru, M. Ultrasonically assisted extraction (UAE) of natural products some guidelines for good practice and reporting. Ultrason. Sonochem. 2015, 25, 94–95. [Google Scholar] [CrossRef] [PubMed]
  96. Zhang, Q.W.; Lin, L.G.; Ye, W.C. Techniques for extraction and isolation of natural products: A comprehensive review. Chin. Med. 2018, 13, 20. [Google Scholar] [CrossRef]
  97. Willför, S.; Hemming, J.; Reunanen, M.; Eckerman, C.; Holmbom, B. Lignans and Lipophilic Extractives in Norway Spruce Knots and Stemwood. Wood Res. Technol. 2003, 57, 27–36. [Google Scholar] [CrossRef]
  98. Willför, S.; Nisula, L.; Hemming, J.; Reunanen, M.; Holmbom, B. Bioactive phenolic substances in industrially important tree species. Part 1: Knots and stemwood of different spruce species. Wood Res. Technol. 2004, 58, 335–344. [Google Scholar] [CrossRef]
  99. Willför, S.; Nisula, L.; Hemming, J.; Reunanen, M.; Holmbom, B. Bioactive phenolic substances in industrially important tree species. Part 2: Knots and stemwood of fir species. Holzforschung 2004, 58, 650–659. [Google Scholar] [CrossRef]
  100. Pietarinen, S.P.; Willför, S.M.; Ahotupa, M.O.; Hemming, J.E.; Holmbom, B.R. Knotwood and bark extracts: Strong antioxidants from waste materials. J. Wood Sci. 2006, 52, 436–444. [Google Scholar] [CrossRef]
  101. Kalyniukova, A.; Holuša, J.; Musiolek, D.; Sedlakova-Kadukova, J.; Płotka-Wasylka, J.; Andruch, V. Application of deep eutectic solvents for separation and determination of bioactive compounds in medicinal plants. Ind. Crops Prod. 2021, 172, 114047. [Google Scholar] [CrossRef]
  102. Garg, C.; Verma, S.; Satija, S.; Mehta, M.; Dureja, H.; Garg, M. Microwave assisted extraction of bioactive compound phyllanthin from Phyllanthus amarus and optimization using central composite design. Int. J. Pharm. Sci. Res. 2016, 1, 30–35. [Google Scholar]
  103. Slanina, J.; Glatz, Z. Separation procedures applicable to lignan analysis. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2004, 812, 215–229. [Google Scholar] [CrossRef]
  104. Eklund, P.C.; Sundell, F.J.; Smeds, A.I.; Sjoholm, R.E. Reactions of the natural lignan hydroxymatairesinol in basic and acidic nucleophilic media: Formation and reactivity of a quinone methide intermediate. Org. Biomol. Chem. 2004, 2, 2229–2235. [Google Scholar] [CrossRef]
  105. Eklund, P.; Sillanpää, R.; Sjöholm, R. Synthetic transformation of hydroxymatairesinol from Norway spruce (Picea abies) to 7-hydroxysecoisolariciresinol, (+)-lariciresinol and (+)-cyclolariciresinol. J. Chem. Soc. Perkin Trans. 2002, 1, 1906–1910. [Google Scholar] [CrossRef]
  106. Smeds, A.I.; Eklund, P.C.; Sjöholm, R.E.; Willför, S.M.; Nishibe, S.; Deyama, T.; Holmbom, B.R. Quantification of a broad spectrum of lignans in cereals, oilseeds, and nuts. J. Agric. Food Chem. 2007, 55, 1337–1346. [Google Scholar] [CrossRef]
  107. Macías-Villamizar, V.; Cuca-Suárez, L. Structural modification of lignan compounds isolated from nectandra species (lauraceaE). J. Chil. Chem. Soc. 2017, 62, 3427–3431. [Google Scholar] [CrossRef]
  108. Eklund, P.C.; Willför, S.M.; Smeds, A.I.; Sundell, F.J.; Sjöholm, R.E.; Holmbom, B.R. A new lariciresinol-type butyrolactone lignan derived from hydroxymatairesinol and its identification in spruce wood. J. Nat. Prod. 2004, 67, 927–931. [Google Scholar] [CrossRef] [PubMed]
  109. Wang, P.; Liu, Y.; Chen, T.; Xu, W.; You, J.; Liu, Y.; Li, Y. One-step Separation and Purification of Three Lignans and One Flavonol from Sinopodophyllum emodi by Medium-pressure Liquid Chromatography and High-speed Counter-current Chromatography. Phytochem. Anal. 2013, 24, 603–607. [Google Scholar] [CrossRef]
  110. Sok, D.E.; Cui, H.S.; Kim, M.R. Isolation and bioactivities of furfuran type lignan compounds from edible plants. Recent Pat. Food Nutr. Agric. 2009, 1, 87–95. [Google Scholar] [CrossRef] [PubMed]
  111. Schmidt-Traub, H.; Schulte, M.; Seidel-Morgenstern, A. Preparative Chromatography; Wiley: Hoboken, NJ, USA, 2020. [Google Scholar]
  112. Huang, Q.; Tan, J.B.; Zeng, X.C.; Wang, Y.Q.; Zou, Z.X.; Ouyang, D.S. Lignans and phenolic constituents from Eucommia ulmoides Oliver. Nat. Prod. Res. 2021, 35, 3376–3383. [Google Scholar] [CrossRef]
  113. Zhang, L.; Wang, X.L.; Wang, B.; Zhang, L.T.; Gao, H.M.; Shen, T.; Lou, H.X.; Ren, D.M.; Wang, X.N. Lignans from Euphorbia hirta L. Nat. Prod. Res. 2022, 36, 26–36. [Google Scholar] [CrossRef]
  114. Nhung, L.T.H.; Anh, N.T.H.; Tai, B.H.; Kiem, P.V. Isolation of lignans and neolignans from Pouzolzia sanguinea with their cytotoxic activity. Vietnam J. Chem. 2021, 59, 146–152. [Google Scholar]
  115. Silva, V.C.d.; Bolzani, V.d.S.; Lopes, M.N.; Silva, G.H. Isolation of lignans glycosides from Alibertia sessilis (Vell) K Schum (Rubiaceae) by preparative high-performance liquid chromatography. Eclet. Quim. 2006, 31, 55–58. [Google Scholar] [CrossRef]
  116. Lee, J.; Yang, H.S.; Jeong, H.; Kim, J.-H.; Yang, H. Targeted Isolation of Lignans from Trachelospermum asiaticum Using Molecular Networking and Hierarchical Clustering Analysis. Biomolecules 2020, 10, 378. [Google Scholar] [CrossRef] [PubMed]
  117. Benzina, S.; Harquail, J.; Jean, S.; Beauregard, A.P.; Colquhoun, C.D.; Carroll, M.; Bos, A.; Gray, C.A.; Robichaud, G.A. Deoxypodophyllotoxin isolated from Juniperus communis induces apoptosis in breast cancer cells. Anticancer Agents Med. Chem. 2015, 15, 79–88. [Google Scholar] [CrossRef]
  118. Takahashi, M.; Nishizaki, Y.; Sugimoto, N.; Takeuchi, H.; Nakagawa, K.; Akiyama, H.; Sato, K.; Inoue, K. Determination and purification of sesamin and sesamolin in sesame seed oil unsaponified matter using reversed-phase liquid chromatography coupled with photodiode array and tandem mass spectrometry and high-speed countercurrent chromatography. J. Sep. Sci. 2016, 39, 3898–3905. [Google Scholar] [CrossRef]
  119. Wang, X.; Lin, Y.; Geng, Y.; Li, F.; Wang, D. Preparative separation and purification of sesamin and sesamolin from sesame seeds by high-speed counter-current chromatography. Cereal Chem. 2009, 86, 23. [Google Scholar] [CrossRef]
  120. Wagner, H.; Bladt, S. Plant Drug Analysis: A Thin Layer Chromatography Atlas; Springer: Berlin/Heidelberg, Germany, 2009. [Google Scholar]
  121. Zare, K.; Movafeghi, A.; Mohammadi, S.A.; Asnaashari, S.; Nazemiyeh, H. New Phenolics from Linum mucronatum subsp. orientale. Bioimpacts 2014, 4, 117–122. [Google Scholar] [CrossRef]
  122. Goels, T.; Eichenauer, E.; Langeder, J.; Hoeller, F.; Sykora, C.; Tahir, A.; Urban, E.; Heiss, E.H.; Saukel, J.; Glasl, S. Norway Spruce Balm: Phytochemical Composition and Ability to Enhance Re-epithelialization In Vitro. Planta Med. 2020, 86, 1080–1088. [Google Scholar] [CrossRef] [PubMed]
  123. Sobstyl, E.; Szopa, A.; Ekiert, H.; Gnat, S.; Typek, R.; Choma, I.M. Effect directed analysis and TLC screening of Schisandra chinensis fruits. J. Chromatogr. A 2020, 1618, 460942. [Google Scholar] [CrossRef] [PubMed]
  124. Pi, J.J.; Wu, X.; Rui, W.; Feng, Y.F.; Guo, J. Identification and Fragmentation Mechanisms of Two Kinds of Chemical Compositions in Eucommia ulmoides By UPLC-ESI-Q-TOF-MS/MS. Chem. Nat. Compd. 2016, 52, 144–148. [Google Scholar] [CrossRef]
  125. Kraushofer, T.; Sontag, G. Determination of matairesinol in flax seed by HPLC with coulometric electrode array detection. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2002, 777, 61–66. [Google Scholar] [CrossRef]
  126. Li, J.; Wen, J.; Tang, G.; Li, R.; Guo, H.; Weng, W.; Wang, D.; Ji, S. Development of a comprehensive quality control method for the quantitative analysis of volatiles and lignans in Magnolia biondii Pamp. by near infrared spectroscopy. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2020, 230, 118080. [Google Scholar] [CrossRef]
  127. Zhang, Q.; Zhu, W.; Guan, H.; Liu, H.; Yang, W.; Wang, H.; Cai, D. Development of a matrix solid-phase dispersion extraction combined with high-performance liquid chromatography for determination of five lignans from the Schisandra chinensis. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2016, 1011, 151–157. [Google Scholar] [CrossRef] [PubMed]
  128. Sobstyl, E.; Szopa, A.; Dziurka, M.; Ekiert, H.; Nikolaichuk, H.; Choma, I.M. Schisandra rubriflora Fruit and Leaves as Promising New Materials of High Biological Potential: Lignan Profiling and Effect-Directed Analysis. Molecules 2022, 27, 2116. [Google Scholar] [CrossRef]
  129. Liu, H.; Zhang, J.; Li, X.; Qi, Y.; Peng, Y.; Zhang, B.; Xiao, P. Chemical analysis of twelve lignans in the fruit of Schisandra sphenanthera by HPLC–PAD-MS. Phytomedicine 2012, 19, 1234–1241. [Google Scholar] [CrossRef] [PubMed]
  130. Shi, L.; Zheng, L.; Xiang, Y.; Liu, R.; Chang, M.; Jin, Q.; Wang, X. A Rapid Method for Simultaneous Analysis of Lignan and γ-Tocopherol in Sesame Oil by Using Normal-Phase Liquid Chromatography. J. Am. Oil Chem. Soc. 2018, 95, 13–19. [Google Scholar] [CrossRef]
  131. Zhao, L.; Tian, X.; Fan, P.C.; Zhan, Y.J.; Shen, D.W.; Jin, Y. Separation, determination and identification of the diastereoisomers of podophyllotoxin and its esters by high-performance liquid chromatography/tandem mass spectrometry. J. Chromatogr. A 2008, 1210, 168–177. [Google Scholar] [CrossRef]
  132. Avula, B.; Wang, Y.-H.; Moraes, R.M.; Khan, I.A. Rapid analysis of lignans from leaves of Podophyllum peltatum L. samples using UPLC-UV-MS. Biomed. Chromatogr. 2011, 25, 1230–1236. [Google Scholar] [CrossRef]
  133. Lu, L.W.; Le, Z.; Hou, Z.L.; Jie, W.; Yao, G.D.; Lin, B.; Huang, X.X.; Song, S.J. Chiral-phase resolution of sesquilignans from raspberries (Rubus idaeus L.) and their neuroprotective effects. Fitoterapia 2020, 146, 104655. [Google Scholar] [CrossRef]
  134. Dar, A.A.; Arumugam, N. Lignans of sesame: Purification methods, biological activities and biosynthesis—A review. Bioorg. Chem. 2013, 50, 1–10. [Google Scholar] [CrossRef]
  135. Olmo-García, L.; Kessler, N.; Neuweger, H.; Wendt, K.; Olmo-Peinado, J.M.; Fernández-Gutiérrez, A.; Baessmann, C.; Carrasco-Pancorbo, A. Unravelling the Distribution of Secondary Metabolites in Olea europaea L.: Exhaustive Characterization of Eight Olive-Tree Derived Matrices by Complementary Platforms (LC-ESI/APCI-MS and GC-APCI-MS). Molecules 2018, 23, 2419. [Google Scholar] [CrossRef]
  136. Struijs, K.; Vincken, J.P.; Gruppen, H. Comparison of atmospheric pressure chemical ionization and electrospray ionization mass spectrometry for the detection of lignans from sesame seeds. Rapid Commun. Mass Spectrom. 2008, 22, 3615–3623. [Google Scholar] [CrossRef]
  137. Hata, N.; Hayashi, Y.; Okazawa, A.; Ono, E.; Satake, H.; Kobayashi, A. Comparison of sesamin contents and CYP81Q1 gene expressions in aboveground vegetative organs between two Japanese sesame (Sesamum indicum L.) varieties differing in seed sesamin contents. Plant Sci. 2010, 178, 510–516. [Google Scholar] [CrossRef]
  138. Valentín-Blasini, L.; Blount, B.C.; Rogers, H.S.; Needham, L.L. HPLC-MS/MS method for the measurement of seven phytoestrogens in human serum and urine. J. Expo. Anal. Environ. Epidemiol. 2000, 10, 799–807. [Google Scholar] [CrossRef] [PubMed]
  139. Zhang, F.; Zhai, J.; Weng, N.; Gao, J.; Yin, J.; Chen, W. A Comprehensive Review of the Main Lignan Components of Schisandra chinensis (North Wu Wei Zi) and Schisandra sphenanthera (South Wu Wei Zi) and the Lignan-Induced Drug-Drug Interactions Based on the Inhibition of Cytochrome P450 and P-Glycoprotein Activities. Front. Pharm. 2022, 13, 816036. [Google Scholar] [CrossRef]
  140. Wang, C.Y.; Lee, S.S. Analysis and identification of lignans in Phyllanthus urinaria by HPLC-SPE-NMR. Phytochem. Anal. 2005, 16, 120–126. [Google Scholar] [CrossRef]
  141. Eklund, P.C.; Backman, M.J.; Kronberg, L.A.; Smeds, A.I.; Sjoholm, R.E. Identification of lignans by liquid chromatography-electrospray ionization ion-trap mass spectrometry. J. Mass Spectrom. 2008, 43, 97–107. [Google Scholar] [CrossRef]
  142. Wolfender, J.L.; Terreaux, C.; Hostettmann, K. The Importance Of LC-MS And LC-NMR In The Discovery Of New Lead Compounds from Plants. Pharm. Biol. 2000, 38 (Suppl. 1), 41–54. [Google Scholar] [CrossRef] [PubMed]
  143. Martini, S.; Cattivelli, A.; Conte, A.; Tagliazucchi, D. Black, green, and pink pepper affect differently lipid oxidation during cooking and in vitro digestion of meat. Food Chem. 2021, 350, 129246. [Google Scholar] [CrossRef]
  144. Cui, Y.; Wang, Q.; Shi, X.; Zhang, X.; Sheng, X.; Zhang, L. Simultaneous quantification of 14 bioactive constituents in Forsythia suspensa by liquid chromatography-electrospray ionisation-mass spectrometry. Phytochem. Anal. 2010, 21, 253–260. [Google Scholar] [CrossRef]
  145. Kiss, A.K.; Michalak, B.; Patyra, A.; Majdan, M. UHPLC-DAD-ESI-MS/MS and HPTLC profiling of ash leaf samples from different commercial and natural sources and their in vitro effects on mediators of inflammation. Phytochem. Anal. 2020, 31, 57–67. [Google Scholar] [CrossRef]
  146. Jiao, Q.S.; Xu, L.L.; Zhang, J.Y.; Wang, Z.J.; Jiang, Y.Y.; Liu, B. Rapid Characterization and Identification of Non-Diterpenoid Constituents in Tinospora sinensis by HPLC-LTQ-Orbitrap MS(n). Molecules 2018, 23, 274. [Google Scholar] [CrossRef]
  147. Huang, Q.; Zhang, F.; Liu, S.; Jiang, Y.; Ouyang, D. Systematic investigation of the pharmacological mechanism for renal protection by the leaves of Eucommia ulmoides Oliver using UPLC-Q-TOF/MS combined with network pharmacology analysis. Biomed. Pharmacother. 2021, 140, 111735. [Google Scholar] [CrossRef] [PubMed]
  148. Ling, X.-x.; Chen, H.; Fu, B.-b.; Ruan, C.-s.; Pana, M.; Zhou, K.; Fang, Z.-r.; Shao, J.-t.; Zhu, F.-q.; Gao, S. Xin-Ji-Er-Kang protects myocardial and renal injury in hypertensive heart failure in mice. Phytomedicine 2021, 91, 153675. [Google Scholar] [CrossRef] [PubMed]
  149. Stasevich, O.V.; Mikhalenok, S.G.; Kurchenko, V.P. Isolation of secoisolariciresinol diglucoside from lignan-containing extract of Linum usitatissimum seeds. Chem. Nat. Compd. 2009, 45, 21–23. [Google Scholar] [CrossRef]
  150. Yan, G.; Li, Q.; Tan, H.; Ge, T. Electrospray ionization ion-trap time-of-flight tandem mass spectrometry of two furofurans: Sesamin and gmelinol. Rapid Commun. Mass Spectrom. 2007, 21, 3613–3620. [Google Scholar] [CrossRef] [PubMed]
  151. Peng, Z.; Xu, Y.; Meng, Q.; Raza, H.; Zhao, X.; Liu, B.; Dong, C. Preparation of Sesaminol from Sesaminol Triglucoside by β-Glucosidase and Cellulase Hydrolysis. J. Am. Oil Chem. Soc. 2016, 93, 765–772. [Google Scholar] [CrossRef]
  152. Liu, X.T.; Wang, X.G.; Yang, Y.; Xu, R.; Meng, F.H.; Yu, N.J.; Zhao, Y.M. Qualitative and Quantitative Analysis of Lignan Constituents in Caulis Trachelospermi by HPLC-QTOF-MS and HPLC-UV. Molecules 2015, 20, 8107–8124. [Google Scholar] [CrossRef]
  153. Lehraiki, A.; Attoumbré, J.; Bienaimé, C.; Matifat, F.; Bensaddek, L.; Nava-Saucedo, E.; Fliniaux, M.A.; Ouadid-Ahidouch, H.; Baltora-Rosset, S. Extraction of lignans from flaxseed and evaluation of their biological effects on breast cancer MCF-7 and MDA-MB-231 cell lines. J. Med. Food 2010, 13, 834–841. [Google Scholar] [CrossRef]
  154. Hendrawati, O.; Woerdenbag, H.J.; Michiels, P.J.A.; Aantjes, H.G.; van Dam, A.; Kayser, O. Identification of lignans and related compounds in Anthriscus sylvestris by LC–ESI-MS/MS and LC-SPE–NMR. Phytochemistry 2011, 72, 2172–2179. [Google Scholar] [CrossRef]
  155. Lee, S.; Khoo, C.; Halstead, C.W.; Huynh, T.; Bensoussan, A. Liquid chromatographic determination of honokiol and magnolol in hou po (Magnolia officinalis) as the raw herb and dried aqueous extract. J. AOAC Int. 2007, 90, 1210–1218. [Google Scholar] [CrossRef]
  156. Knust, U.; Hull, W.E.; Spiegelhalder, B.; Bartsch, H.; Strowitzki, T.; Owen, R.W. Analysis of enterolignan glucuronides in serum and urine by HPLC-ESI-MS. Food Chem. Toxicol. 2006, 44, 1038–1049. [Google Scholar] [CrossRef]
  157. Guo, Z.; Zhao, A.; Chen, T.; Xie, G.; Zhou, M.; Qiu, M.; Jia, W. Differentiation of Schisandra chinensis and Schisandra sphenanthera using metabolite profiles based on UPLC-MS and GC-MS. Nat. Prod. Res. 2012, 26, 255–263. [Google Scholar] [CrossRef]
  158. Huang, S.; Liu, Y.; Li, Y.; Fan, H.; Huang, W.; Deng, C.; Song, X.; Zhang, D.; Wang, W. Dibenzocyclooctadiene lignans from the root bark of Schisandra sphenanthera. Phytochem. Lett. 2021, 45, 137–141. [Google Scholar] [CrossRef]
  159. Wei, X.-l.; Chen, Y.; Chen, X.-y.; Liang, J.-y.; Qu, W. A New Lignan from the Roots of Ginkgo biloba. Chem. Nat. Compd. 2015, 51, 819–821. [Google Scholar] [CrossRef]
  160. Figueroa, J.G.; Borrás-Linares, I.; Lozano-Sánchez, J.; Segura-Carretero, A. Comprehensive characterization of phenolic and other polar compounds in the seed and seed coat of avocado by HPLC-DAD-ESI-QTOF-MS. Food Res. Int. 2018, 105, 752–763. [Google Scholar] [CrossRef] [PubMed]
  161. Sólyomváry, A.; Tóth, G.; Kraszni, M.; Noszál, B.; Molnár-Perl, I.; Boldizsár, I. Identification and quantification of lignans and sesquilignans in the fruits of Cnicus benedictus L.: Quantitative chromatographic and spectroscopic approaches. Microchem. J. 2014, 114, 238–246. [Google Scholar] [CrossRef]
  162. Dias, M.M.; Zuza, O.; Riani, L.R.; de Faria Pinto, P.; Pinto, P.L.S.; Silva, M.P.; de Moraes, J.; Ataíde, A.C.Z.; de Oliveira Silva, F.; Cecílio, A.B.; et al. In vitro schistosomicidal and antiviral activities of Arctium lappa L. (Asteraceae) against Schistosoma mansoni and Herpes simplex virus-1. Biomed Pharm. 2017, 94, 489–498. [Google Scholar] [CrossRef]
  163. Sólyomváry, A.; Alberti, Á.; Darcsi, A.; Könye, R.; Tóth, G.; Noszál, B.; Molnár-Perl, I.; Lorántfy, L.; Dobos, J.; Őrfi, L.; et al. Optimized conversion of antiproliferative lignans pinoresinol and epipinoresinol: Their simultaneous isolation and identification by centrifugal partition chromatography and high performance liquid chromatography. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2017, 1052, 142–149. [Google Scholar] [CrossRef]
  164. Carrasco-Pancorbo, A.; Nevedomskaya, E.; Arthen-Engeland, T.; Zey, T.; Zurek, G.; Baessmann, C.; Deelder, A.M.; Mayboroda, O.A. Gas Chromatography/Atmospheric Pressure Chemical Ionization-Time of Flight Mass Spectrometry: Analytical Validation and Applicability to Metabolic Profiling. Anal. Chem. 2009, 81, 10071–10079. [Google Scholar] [CrossRef]
  165. Thompson, L.U.; Boucher, B.A.; Liu, Z.; Cotterchio, M.; Kreiger, N. Phytoestrogen content of foods consumed in Canada, including isoflavones, lignans, and coumestan. Nutr. Cancer 2006, 54, 184–201. [Google Scholar] [CrossRef]
  166. Sarajlija, H.; Čukelj Mustač, N.; Novotni, D.; Mršić, G.; Brncic, M.; Curic, D. Preparation of Flaxseed for Lignan Determination by Gas Chromatography-Mass Spectrometry Method. Czech J. Food Sci. 2012, 30, 45. [Google Scholar] [CrossRef]
  167. Popova, I.E.; Hall, C.; Kubatova, A. Determination of lignans in flaxseed using liquid chromatography with time-of-flight mass spectrometry. J. Chromatogr. A 2009, 1216, 217–229. [Google Scholar] [CrossRef] [PubMed]
  168. Bonzanini, F.; Bruni, R.; Palla, G.; Serlataite, N.; Caligiani, A. Identification and distribution of lignans in Punica granatum L. fruit endocarp, pulp, seeds, wood knots and commercial juices by GC–MS. Food Chem. 2009, 117, 745–749. [Google Scholar] [CrossRef]
  169. Załuski, D.; Mendyk, E.; Smolarz, H.D. Identification of MMP-1 and MMP-9 inhibitors from the roots of Eleutherococcus divaricatus, and the PAMPA test. Nat. Prod. Res. 2016, 30, 595–599. [Google Scholar] [CrossRef]
  170. Liu, Z.; Saarinen, N.M.; Thompson, L.U. Sesamin is one of the major precursors of mammalian lignans in sesame seed (Sesamum indicum) as observed in vitro and in rats. J. Nutr. 2006, 136, 906–912. [Google Scholar] [CrossRef] [PubMed]
  171. Čukelj, N.; Jakasa, I.; Sarajlija, H.; Novotni, D.; Ćurić, D. Identification and quantification of lignans in wheat bran by gas chromatography-electron capture detection. Talanta 2011, 84, 127–132. [Google Scholar] [CrossRef]
  172. Xia, Y.-G.; Yang, B.-Y.; Liang, J.; Yang, Q.; Wang, D.; Kuang, H.-X. Quantitative Analysis and Fingerprint Profiles for Quality Control of Fructus Schisandrae by Gas Chromatography: Mass Spectrometry. Sci. World J. 2014, 2014, 806759. [Google Scholar] [CrossRef]
  173. Koulman, A.; Bos, R.; Medarde, M.; Pras, N.; Quax, W.J. A fast and simple GC MS method for lignan profiling in Anthriscus sylvestris and biosynthetically related Plant species. Planta Med. 2001, 67, 858–862. [Google Scholar] [CrossRef]
  174. Saarinen, N.M.; Smeds, A.I.; Peñalvo, J.L.; Nurmi, T.; Adlercreutz, H.; Mäkelä, S. Flaxseed ingestion alters ratio of enterolactone enantiomers in human serum. J. Nutr. Metab. 2010, 2010, 403076. [Google Scholar] [CrossRef]
  175. Bommareddy, A.; Arasada, B.L.; Mathees, D.P.; Dwivedi, C. Determination of mammalian lignans in biological samples by modified gas chromatography/mass spectrometry. J. AOAC Int. 2007, 90, 641–646. [Google Scholar] [CrossRef]
  176. Edel, A.L.; Aliani, M.; Pierce, G.N. Supported liquid extraction in the quantitation of plasma enterolignans using isotope dilution GC/MS with application to flaxseed consumption in healthy adults. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2013, 912, 24–32. [Google Scholar] [CrossRef]
  177. Willför, S.M.; Ahotupa, M.O.; Hemming, J.E.; Reunanen, M.H.T.; Eklund, P.C.; Sjöholm, R.E.; Eckerman, C.S.E.; Pohjamo, S.P.; Holmbom, B.R. Antioxidant Activity of Knotwood Extractives and Phenolic Compounds of Selected Tree Species. J. Agric. Food Chem. 2003, 51, 7600–7606. [Google Scholar] [CrossRef] [PubMed]
  178. Yu, M.; Aoki, D.; Akita, T.; Fujiyasu, S.; Takada, S.; Matsushita, Y.; Yoshida, M.; Fukushima, K. Distribution of lignans and lignan mono/diglucosides within Ginkgo biloba L. stem. Phytochemistry 2022, 196, 113102. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Biosynthetic pathway of major lignans reconstructed from the published records [1,4,5,6,7,8,9,10,11]. DP (dirigent protein), PSS (piperitol/sesamin Synthase), PLR (pinoresinol/lariciresinol reductase), CPR1 (cytochrome P450 oxidoreductase 1), SID (secoisolariciresinol dehydrogenase), MMT (matairesinol O-methyltransferase).
Figure 1. Biosynthetic pathway of major lignans reconstructed from the published records [1,4,5,6,7,8,9,10,11]. DP (dirigent protein), PSS (piperitol/sesamin Synthase), PLR (pinoresinol/lariciresinol reductase), CPR1 (cytochrome P450 oxidoreductase 1), SID (secoisolariciresinol dehydrogenase), MMT (matairesinol O-methyltransferase).
Plants 11 02323 g001
Table 1. Comparison of methods used in lignans’ extraction.
Table 1. Comparison of methods used in lignans’ extraction.
Species Plant Part Method Solvent Extraction Times Extraction Time (min) Temperature (°C) Solvent to Sample Ratio Reference
Abies alba Mill. bark Digestion H2O 120 70 1:5 [39]
Abies chensiensis Tiegh. aerial parts Maceration 80% EtOH/H2O (v/v) 180 RT - [46]
Abies delavayi var. nukiangensis (W.C.Cheng & L.K.Fu) Farjon & Silba aerial parts Maceration 95% MeOH/H2O (v/v) 180 RT - [47]
Arctium lappa L. fructus Digestion 95% EtOH/H2O (v/v) 180 80 1:10 [48]
Arctium lappa L. fructus MAE 40% MeOH/H2O (v/v) 200 s - 1:15 [49]
Arctium lappa L. fructus Soxhlet/heated reflux 70% EtOH/H2O (v/v) 60 BP 1:8 [50]
Arctium lappa L. fructus Soxhlet/heated reflux 80% EtOH/H2O (v/v) - BP 1:6 [51]
Arctium lappa L. fructus UAE MeOH 20 - 1:90 [42]
Arctium lappa L. fructus UAE MeOH + H2O 30 (with MeOH) and 30 (with water) - 1:15 [29]
Arctium lappa L. roots Maceration 80% MeOH/H2O (v/v) 8 h RT 1:50 [25]
Arctium lappa L. seed Maceration MeOH - RT 1:2 [43]
Arctium lappa L. seed Maceration MeOH - RT 3:5 [52]
Arctium lappa L. seeds Maceration 80% MeOH/H2O (v/v) 8 h RT 1:50 [25]
Carthamus tinctorius L. seed Maceration MeOH 15 h RT 1:2 [53]
Eleutherococcus divaricatus (Siebold & Zucc.) S.Y.Hu root Soxhlet/heated reflux MeOH - BP 1:2 [54]
Eleutherococcus senticosus (Rupr. & Maxim.) Maxim. bark Soxhlet/heated reflux 70% MeOH/H2O (v/v) 180 BP 1:6 [55]
Eleutherococcus senticosus (Rupr. & Maxim.) Maxim. root Soxhlet/heated reflux H2O 30 BP 1:6 [37]
Eleutherococcus senticosus (Rupr. & Maxim.) Maxim. stem Maceration MeOH 2 weeks RT - [56]
Eleutherococcus sessiliflorus (Rupr. & Maxim.) S.Y.Hu fruit Maceration 70% EtOH/H2O (v/v) 24 h RT 5:18 [57]
Forsythia × intermedia Zabel flower Digestion 75% MeOH/H2O (v/v) 120 70 1:20 [58]
Forsythia × intermedia Zabel flower Soxhlet/heated reflux MeOH - BP 1:200 [44]
Forsythia × intermedia Zabel leaf Digestion 75% MeOH/H2O (v/v) 120 70 1:20 [58]
Forsythia × intermedia Zabel leaf Soxhlet/heated reflux MeOH - BP 1:200 [44]
Forsythia koreana (Rehder) Nakai flower Maceration 80% MeOH/H2O (v/v) 24 h RT 3:100 [59]
Forsythia koreana (Rehder) Nakai fruit Maceration MeOH 7 days RT - [60]
Forsythia koreana (Rehder) Nakai fruit Maceration MeOH - RT 1:1 [61]
Forsythia koreana (Rehder) Nakai stem UAE MeOH 60 - 1:2 [62]
Forsythia suspensa flower Soxhlet/heated reflux MeOH - BP 1:200 [44]
Forsythia suspensa fruit Maceration 70% EtOH/H2O (v/v) - RT 2:1 [63]
Forsythia suspensa fruit Soxhlet/heated reflux 50% MeOH/H2O (v/v) 60 BP 1:25 [64]
Forsythia suspensa (Thunb.) Vahl fruit Soxhlet/heated reflux 60% EtOH/H2O (v/v) 120 BP - [65]
Forsythia suspensa (Thunb.) Vahl fruit Soxhlet/heated reflux 95% EtOH/H2O (v/v) 120 BP 1:60 [66]
Forsythia suspensa (Thunb.) Vahl fruit Soxhlet/heated reflux MeOH - BP 3:10 [45]
Forsythia suspensa (Thunb.) Vahl fruit UAE 20% MeOH/H2O (v/v) 30 - 1:5 [67]
Forsythia suspensa (Thunb.) Vahl leaf Soxhlet/heated reflux MeOH - BP 1:200 [44]
Forsythia viridissima Lindl. flower Soxhlet/heated reflux MeOH - BP 1:200 [44]
Forsythia viridissima Lindl. leaf Soxhlet/heated reflux MeOH - BP 1:200 [44]
Forsythia viridissima Lindl. root UAE 80% MeOH/H2O (v/v) 90 RT 7:10 [68]
Larix laricina (Du Roi) K.Koch bark Maceration 80% EtOH/H2O (v/v) 24 h RT 1:10 [69]
Linum usitatissimum L. seed Maceration 60% EtOH/H2O (v/v) 60 RT 10:75 [70]
Linum usitatissimum L. seed MAE 70% MeOH/H2O (v/v) 180 60 25:1 [71]
Linum usitatissimum L. seed Digestion 75% MeOH/H2O (v/v) 24 h 65 - [72]
Linum usitatissimum L. seed Digestion 80% EtOH/H2O (v/v) 240 55 1:14 [34]
Linum usitatissimum L. seed Maceration 70% MeOH/H2O (v/v) 120 RT 1:3 [73]
Linum usitatissimum L. seed Maceration H2O 60 RT 1:15 [70]
Linum usitatissimum L. seed UAE 70% MeOH/H2O (v/v) 4h - 1:6 [74]
Linum usitatissimum L. seed UAE H2O + 0.2 N NaOH 60 25 - [75]
Linum usitatissimum L. seed UAE 50% MeOH/H2O (v/v) 10 + 60 RT 1:25 [76]
70% acetone/H2O (v/v) - RT 1:25
Magnolia biondii Pamp. flower buds Soxhlet/heated reflux MeOH 5 h BP - [77]
Magnolia officinalis Rehder & E.H. Wilson flower buds Soxhlet/heated reflux MeOH 180 BP 9:4 [78]
Pinus radiata D.Don bark Soxhlet/heated reflux 70% acetone/H2O (v/v) 180 BP 1:10 [79]
Pinus sylvestris L. wood ASE 95% acetone/H2O (v/v) 5 100 - [80]
Pinus sylvestris L. wood Soxhlet/heated reflux 90% acetone/H2O (v/v) 180 BP 3:250 [81]
Saraca asoca (Roxb.) Willd. bark MAE 70% MeOH/H2O (v/v) 10 - 1:30 [82]
Schisandra chinensis (Turcz.) Baill. fruit Digestion 70% MeOH/H2O (v/v) 5 h 50 - [83]
Schisandra chinensis (Turcz.) Baill. fruit SFE CO2 + MeOH 20 - 1:10 [84]
Schisandra chinensis (Turcz.) Baill. wood SFE CO2 + EtOH 6 h - - [85]
Sesamum indicum L. seed Maceration 80% EtOH/H2O (v/v) 8 h RT 1:10 [86]
Sesamum indicum L. seed Maceration cyclohexane + DCM + MeOH (1:1:1) - RT 1:1 [36]
Sesamum indicum L. seed Maceration H2O 24 h RT - [40]
Sesamum indicum L. seed Soxhlet/heated reflux n-hexane 10 h BP - [26]
Sesamum indicum L. seed Soxhlet/heated reflux MeOH 8 h BP 1:10 [87]
Sesamum indicum L. seed UAE 50% MeOH/H2O (v/v) 10 + 60 RT 1:25 [88]
70% acetone/ H2O (v/v) - RT 1:25
Sesamum indicum L. seed UAE 80% acetone/H2O (v/v) 5 - 1:10 [30]
Syringa pinnatifolia Hemsl. bark Soxhlet/heated reflux 95% EtOH + 80% EtOH 90 BP 1:2 [89]
Syringa pinnatifolia Hemsl. root Maceration 95% EtOH/H2O (v/v) - RT 1:5 [90]
Syringa vulgaris L. bark Digestion 60% EtOH/H2O (v/v) 60 70 1:20 [91]
ASE—accelerated solvent extraction; BP—boiling point; DCM—dichloromethane; EtOH—ethanol; MAE—microwave-assisted extraction; MeOH—methanol; RT—room temperature; SFE—supercritical fluid extraction; UAE—ultrasound-assisted extraction.
Table 2. HPLC methods used in the analysis of lignans.
Table 2. HPLC methods used in the analysis of lignans.
Species Plant Part Mobile Phase Column Elution Type Detection System Reference
Abies alba Mill. bark H2O and MeCN RP-C18 column (10 cm × 4.6 mm, 2.7 µm) gradient DAD [39]
Arctium lappa L. fruit H2O + 0.01% HCOOH and MeCN RP-C18 (250 mm × 4.6 mm, 5 µm) gradient DAD [28]
Arctium lappa L. fruit H2O and MeCN RP-C18 (150 mm × 4.6 mm, 5 μm) gradient FLD [42]
Arctium lappa L. fruit MeCN and H2O + 0.1% HCOOH RP-C18 (250 mm × 4.6 mm, 5 μm) gradient UV detection at 254 nm [29]
Arctium lappa L. fruit, root H2O + 0.1% HCOOH and MeOH + 0.1% HCOOH RP-C18 (250 mm × 2.0 mm, 5 µm) gradient DAD [25]
Carthamus tinctorius L. fruit H2O + MeOH RP-C12 (150 mm × 4.6 mm, 3.5 µm) gradient UV detection at 221/ESI-MS [53]
Eleutherococcus senticosus (Rupr. & Maxim.) Maxim. bark H2O + 0.2% CH3COOH and MeOH + 0.2% CH3COOH RP-C18 (125 mm × 3 mm, 3 µm) gradient UV detection at 210 nm, 254 nm and 280 nm [55]
Eleutherococcus senticosus (Rupr. & Maxim.) Maxim. root H2O + 15% MeCN + 0.1% HCOOH RP-C18 (250 mm × 4.6 mm, 5 µm) isocratic UV detection at 205 nm [37]
Eucommia ulmoides Oliv. bark H2O + 0.1% HCOOH and MeOH RP-C18 (50 mm × 2.1 mm, 1.7 μm) gradient ESI-MS [124]
Forsythia × intermedia Zabel flower, leaf H2O + 0.1% HCOOH and MeCN + 0.1% HCOOH RP-C18 (150 mm × 2.1 mm, 1.9 μm) gradient DAD, ESI-MS [58]
Forsythia koreana (Rehder) Nakai stem H2O + 32% MeCN RP-C18 (150 mm × 4.6 mm, 5 μm) isocratic UV detection at 254 nm [62]
Forsythia suspensa (Thunb.) Vahl fruit 25% MeCN + 0.1% HCOOH RP-C18 (250 mm × 4.6 mm, 5 μm) isocratic UV detection at 250 nm [65]
Forsythia suspensa (Thunb.) Vahl fruit H2O + 0.1% HCOOH and MeOH RP-C18 (150 mm × 4.6 mm, 5 μm) coupled with RP-C18 (12.5 mm × 4.6 mm, 5 μm) gradient DAD [64]
Linum usitatissimum L. seed H2O + 0.2% CH3COOH and MeCN RP-C18 (250 mm × 4.6 mm, 5 µm) gradient UV detection at 280 nm [73]
Linum usitatissimum L. seed H2O + 0.2% CH3COOH and MeCN RP-C18 (250 mm × 4.6 mm, 5 µm) gradient UV detection at 280 nm [71]
Linum usitatissimum L. seed H2O + 0.2% CH3COOH and MeOH RP-C18 (250 mm × 4.0 mm, 5 µm) gradient DAD [75]
Linum usitatissimum L. seed H2O + 16% MeOH + HCOOH and 100% MeOH RP-C18 (250 mm × 4.6 mm, 5 µm) gradient UV detection at 283 nm [34]
Linum usitatissimum L. seed H2O and MeCN + HCOOH RP-C8 (250 mm × 4 mm, 5 µm) isocratic CEAD [125]
Magnolia biondii Pamp. flower H2O + 0.2% HCOOH and MeCN RP-C18 (250 mm x 4.6 mm, 5 μm) gradient UV detection at 278 nm/ESI-MS [126]
Schisandra chinensis (Turcz.) Baill. fruit H2O + 0.1% H3PO4 and MeCN RP-C18 (150 mm × 2.0 mm, 5 µm) gradient UV detection at 280 nm [83]
Schisandra chinensis (Turcz.) Baill. fruit H2O and MeCN RP-C18 (250 mm × 4.6 mm, 5 µm) gradient UV detection at 215 nm [127]
Schisandra rubriflora Rehder & E.H.Wilson fruit, leaf H2O and MeOH + 0.1% HCOOH RP-C18 (150 mm × 4.6 mm, 2.7 µm) gradient ESI-MS [128]
Schisandra sphenanthera Rehder & E.H.Wilson fruit H2O and MeCN RP-C18 (250 mm × 4.6 mm, 5 µm) gradient PAD-MS, ESI-MS [129]
Sesamum indicum L. seed H2O + 0.5% CH3COOH and MeCN RP-C18 (150 mm × 4.6 mm, 2.7 µm) gradient DAD [88]
Sesamum indicum L. seed H2O and MeOH RP-C18 (250 mm × 4 mm, 10 µm) gradient UV detection at 280 nm [40]
Sesamum indicum L. seed H2O + 0.5% CH3COOH and MeCN RP-C18 (150 mm × 2.1 mm, 1.9 µm) gradient APCI-MS, ESI-MS [130]
Syringa vulgaris L. bark H2O + 0.1% HCOOH and MeCN + 0.1% HCOOH RP-C18 (150 mm × 2.1 mm, 1.9 µm) gradient DAD, ESI-MS [91]
APCI-MS—atmospheric pressure chemical ionization mass spectrometry; ESI-MS—electrospray ionization mass spectrometry; CEAD—coulometric electrode array detection; DAD—diode array detector; FLD—fluorescence detection; MeCN—acetonitrile; PAD-MS—photodiode array detection mass spectrometry; RP—reversed phase.
Table 3. MS fragmentation patterns of major lignans.
Table 3. MS fragmentation patterns of major lignans.
Compound ESI a Extracted Ion [m/z] Fragment Ions [m/z] Reference
Arctigenin [M−H] 371 356, 295, 209 [58]
Arctigenin glucoside [M+HCOO] 579 371 [58]
[M−H] 533 371
Aschantin [M+H]+ 401 365, 353, 261, 231, 219, 181, 151 [126]
[M+H−H2O]+ 383
Conidendrin [M−H] 355 340, 286, 147 [143]
Cyclolariciresinol [M−H] 359 344, 329, 313, 159 [141]
Epipinoresinol [M−H] 357 151 [144]
Epipinoresinol glucoside [M+HCOO] 565 357 [145]
[M−H] 519 357
[M−H−Glc] 357 -
Fargesin [M+H]+ 371 335, 323, 283, 231, 219, 151 [126]
[M+H−H2O]+ 353
Hinokinin [M+H]+ 355 337, 319, 261, 135 [141]
7-Hydroxylariciresinol [M+Na+]+ 399 384, 381, 369, 351, 219, 202 [146]
7-Hydroxymatairesinol [M−H] 373 355, 340, 311, 296, 231, 160 [141]
Lariciresinol [M−H] 359 344, 329, 208, 161 [143]
Magnolin [M+H]+ 417 381, 369, 329, 279, 249, 231, 219, 189 [126]
[M+H−H2O]+ 399
Matairesinol [M−H] 357 342, 313, 298, 281, 209 [58]
Matairesinol glucoside [M−H] 519 357 [58]
[M−H−Glc] 357 342, 313, 298, 281, 209
Medioresinol [M−H] 387 372, 181, 166, 151, 123 [141]
Medioresinol diglucoside [M−H] 711 548, 387 [124]
Nortrachelogenin [M−H] 373 355, 327, 235, 223 [141]
Olivil [M−H] 375 357, 345, 327, 195, 179, 164 [147]
Olivil glucoside [M−H] 537 375, 345, 327, 195, 179 [124]
[M−H−Glc] 375 -
Olivil diglucoside [M+HCOO] 745 - [124]
[M−H] 699 375, 345, 195, 179
[M−H−Glc] 537 327
7-Oxomatairesinol [M−H] 371 356, 327, 205 [141]
Phillygenin [M−H] 371 356 [58]
Phillygenin glucoside [M+HCOO] 579 371 [58]
[M−H] 533 371
Pinoresinol [M−H] 357 342, 311, 151, 136 [141]
Pinoresinol glucoside [M−H] 519 357 [145]
[M−H−Glc] 357 342, 311, 151, 136
Pinoresinol diglucoside [M+HCOO] 727 - [124]
[M−H] 681 357, 151
[M−H−Glc] 519 -
Schizandrin [M+H]+ 433 415, 384, 369 [148]
Secoisolariciresinol [M−H] 361 346, 331, 313, 179, 165 [141]
Secoisolariciresinol diglucoside [M+2Na+−H] 732 722, 686 [149]
Sesamin [M+H]+ 355 353, 337 [150]
[M+H−H2O]+ 337 319, 289, 261, 231, 203
[M+H−H2]+ 353 323, 135, 77
Sesaminol [M−H] 369 - [151]
Sesaminol glucoside [M−H] 531 - [151]
Sesaminol diglucoside [M−H] 693 - [151]
Sesaminol triglucoside [M−H] 855 - [151]
Syringaresinol [M−H] 417 402, 181, 166, 151 [141]
Syringaresinol glucoside [M−H] 579 417, 181 [124]
[M−H−Glc] 417 -
Syringaresinol diglucoside [M−H] 741 579, 417, 181 [124]
[M+HCOO] 787 579
[M−H−Glc] 579 -
Trachelogenin [M−H] 387 357, 339, 329, 249, 193 [141]
Trachelogenin glucoside [M+Na+]+ 573 389, 371, 343, 325, 247, 203, 151, 137 [152]
a cation or anion formed and type of ionization mode (positive or negative); [M−H]—deprotonated molecule; [M+H]+—protonated molecule; [M+HCOO]—formate adduct ion; [M−H−Glc]—deprotonated and deglycosylated molecule; [M+Na+]+—sodium adduct ion.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Patyra, A.; Kołtun-Jasion, M.; Jakubiak, O.; Kiss, A.K. Extraction Techniques and Analytical Methods for Isolation and Characterization of Lignans. Plants 2022, 11, 2323. https://doi.org/10.3390/plants11172323

AMA Style

Patyra A, Kołtun-Jasion M, Jakubiak O, Kiss AK. Extraction Techniques and Analytical Methods for Isolation and Characterization of Lignans. Plants. 2022; 11(17):2323. https://doi.org/10.3390/plants11172323

Chicago/Turabian Style

Patyra, Andrzej, Małgorzata Kołtun-Jasion, Oktawia Jakubiak, and Anna Karolina Kiss. 2022. "Extraction Techniques and Analytical Methods for Isolation and Characterization of Lignans" Plants 11, no. 17: 2323. https://doi.org/10.3390/plants11172323

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