Open access peer-reviewed chapter

Application of Liquid Chromatography in the Analysis of Flavonoid Metabolism in Plant

Written By

Ngoc Van Thi Nguyen

Reviewed: 18 August 2022 Published: 16 September 2022

DOI: 10.5772/intechopen.107182

From the Edited Volume

Flavonoid Metabolism - Recent Advances and Applications in Crop Breeding

Edited by Hafiz Muhammad Khalid Abbas and Aqeel Ahmad

Chapter metrics overview

161 Chapter Downloads

View Full Metrics

Abstract

Plants have evolved the capacity to create a wide range of chemicals during the process of their existence. In contrast to specialized metabolites that accumulate in a small number of plant species, flavonoids are broadly distributed across the plant kingdom. Therefore, a detailed analysis of flavonoid metabolism in genomics and metabolomics is an ideal way to investigate how plants have developed their unique metabolic pathways during the process of evolution. Among the analysis methods used for flavonoids, the coupling of liquid chromatography (LC) with ultraviolet (UV) and/or electrospray ionization (ESI) mass spectrometric detection has been demonstrated as a powerful tool for the identification and quantification of phenolics in plant extracts. This chapter mainly introduces of chemistry and metabolism of flavonoids and the application of liquid chromatography in the analysis of plant flavonoids.

Keywords

  • flavonoid metabolism
  • liquid chromatography
  • secondary metabolites
  • phenolics and phenylpropanoids
  • flavonoids

1. Introduction

For decades, it has been commonly accepted that plant compounds have a wide range of biological activities. They are secondary metabolites that have considerable pharmacological characteristics and play an important role in improving human health, and flavonoids are one of the substances that have been isolated. Flavonoids, which are responsible for the color and perfume of flowers, have long been known to be synthesized in specific locations, and there are presently around 6000 flavonoids that contribute to the colorful pigments of fruits, herbs, vegetables, and medicinal plants [1]. Flavonoids are hydroxylated phenolic substances known to be formed by plants in response to microbial infection and are a broad set of polyphenolic chemicals with a benzo-pyrone structure synthesized through the phenylpropanoid pathway [2, 3, 4]. Fruits, vegetables, cereals, bark, roots, stems, flowers, tea, and wine all contain it. The chemical properties of flavonoids are determined by their structural class, degree of hydroxylation, various substitutions and conjugations, and degree of polymerization [5].

Flavonoids and other phenylpropanoids are generated from phenylalanine in plants, including the subgroups of flavanones (e.g., flavanone, hesperetin, and naringenin), flavones (e.g., flavone, apigenin, and luteolin), isoflavones (e.g., daidzein, genistein, glycitein). The quantity of oxidation and pattern of substitution of the C ring change within flavonoid classes, whereas individual compounds within a class differ in the pattern of substitution of the A and B rings [6]. Flavonoids are regarded to have health-promoting effects as dietary components due to their strong antioxidant activity in both in vivo and in vitro systems [7, 8]. Flavonoids’ functional hydroxyl groups influence their antioxidant activities by scavenging free radicals and/or chelating metal ions [9, 10]. In addition to antioxidant capabilities, flavonoids have been shown to have antiviral, antibacterial, anti-inflammatory, vasodilatory, anticancer, and anti-ischemic activities. The metabolism of flavonoids is performed by one or more membrane-associated multienzyme complexes rather than free-floating “soluble” enzymes [11]. The primary enzymes involved in flavonoid metabolism are chalcone synthase (a key enzyme in the phenylpropanoid pathway).

Although the separation, identification, and quantification of constituents in complex plant extracts and most likely will be a challenging task, today a multiplicity of different separation techniques, specific stationary phases, and detectors are available, helping to achieve the desired selectivity, sensitivity, and speed for nearly any separation problem. The most prominent and popular technique in this area of research is liquid chromatography [12]. Chromatographic techniques contributed significantly to the area of natural products, especially regarding identification, separation, and characterization of bioactive compounds from plant sources [13]. Flavonoid metabolism is a strong supporter in disease treatment and prevention with chemicals and is an indispensable ingredient in many nutritional, pharmaceutical, and cosmetic applications. The extensive research of flavonoid metabolism in the genome and metabolism is a great technique to investigate how plants’ unique metabolic pathways originated during evolution. The coupling of liquid chromatography (LC) with ultraviolet (UV) and/or electrospray ionization (ESI) mass spectrometric detection is a potent instrument for the identification of phenolics in plant extracts. This chapter focuses on the chemistry and metabolism of flavonoids, as well as the use of liquid chromatography in the study of plant flavonoids.

Advertisement

2. Overview of chemistry and metabolism of flavonoid

Flavonoids are phytonutrients that belong to the polyphenol class. Polyphenols have been employed in Chinese and Ayurvedic medicine for centuries. A novel chemical was extracted from oranges in 1930. It was thought to be a member of a novel class of vitamins at the time and was labeled as vitamin P. Later, it was discovered that this chemical was a flavonoid (rutin), and over 4000 other flavonoids have been found [14].

2.1 Basic chemistry

2.1.1 General characteristics of the C15 unit

Flavonoids exist in the form of aglycones, glycosides, and methylated derivatives. Flavonoids have a diphenyl propane skeleton with 15 carbon atoms in their main nucleus: two six-membered rings coupled with a three-carbon unit that may or may not be part of a third ring. Two benzene rings (A and B in Figure 1) are primarily connected together by a third heterocyclic oxygen-containing pyrene ring (C) [15].

Figure 1.

General structure of flavonoid.

Isoflavones are flavonoids in which the B ring is connected in position 3 of the C ring. Those with the B ring linked in position 4, are referred to as neoflavonoids whereas those with the B ring linked in position 2 can be further classified into many subgroups based on the structural characteristics of the C ring. Flavones, flavonols, flavanones, flavanonols, flavanols or catechins, anthocyanins, and chalcones are the subclasses [1].

2.1.2 Hydroxylation patterns of A-, B-, and C-rings

Positions 3, 5, 7, 2, 3′, 4′, and 5′ are frequently hydroxylated in flavonoids. The most prevalent A-ring hydroxylation pattern is 5,7-hydroxylation; however, a 7-hydroxy ring (also known as a 5-deoxy ring) is seen in isoflavonoid subgroups and several proanthocyanidins. On rare occasions, a 5,7,8 or 5.6.7-hydroxylation pattern is discovered. The B-ring is often 4′-, 3,4′-, or 3′,4,5′-hydroxylation. Rare flavonoids do not have B-ring oxygenation. A 2′-hydroxylation pattern is present in isoflavonoids. In isoflavonoids, the C ring is commonly hydroxylated at the carbon 3 position and occasionally at the carbon 6a position. This six-membered ring can have a carbonyl group, a hydroxyl group, a double bond between positions 2 and 3, or it can be totally unsubstituted, as in unsubstituted flavans. The isoflavonoid pterocarpans have extra rings as a consequence of 2′-hydroxylation of the original B-ring or cyclization of the added prenyl groups (Figure 2) [16].

Figure 2.

Structure backbone of the main flavonoid group.

2.2 Basic substitution

2.2.1 Hydroxylation

There are just a few flavonoid structures with no hydroxyl groups in the A-ring or one hydroxyl group in position 6 [17]. Such unusual structures appear to occur most frequently in the Primulaceae, Rutaceae, and Thymelaeaceae groups. However, the mechanisms of their biochemical synthesis remains unclear. The great majority of flavonoids have a basic 5,7-hydroxylation pattern of the A-ring, which is formed from malonyl-CoA during chalcone synthesis.

The C6-C3 precursor employed by chalcone synthase determines the hydroxylation pattern of the B-ring first. The physiological standard precursor is typically p-coumaroyl-CoA (4-hydroxycinnamoyl-CoA). The resultant basic C15 chalcone intermediate, naringenin chalcone, has a hydroxyl group in position 4′, which is seen in typical flavonoid structures, and all derived flavonoid structures have the hydroxyl group in position 4′. Thus, cinnamate 4-hydroxylase, a cytochrome P450-dependent monooxygenase that catalyzes the production of p-coumaric acid [18, 19], performs a pre-flavonoid step by introducing the hydroxyl group in position 4′.

The majority of flavonoid families have a hydroxyl group in C-ring position 3. (Figure 1). The well-studied flavanone 3-hydroxylase, a 2-oxoglutarate, Fe(II), and ascorbate-dependent dioxygenase [20], introduces the hydroxyl group at the flavanone level. The soluble dioxygenase catalyzes the 3-hydroxylation of the flavanone C-ring to create 3-hydroxyflavanone (flavanol). The same dioxygenase has also been linked to the catalysis of flavone synthases in several plants, with a 2-hydroxylation of the flavanone C-ring proposed as an intermediary step [16].

2.2.2 Methylation

Methylated flavonoids are a form of natural flavonoid derivative with possibly many health advantages, including enhanced bioavailability when compared to flavonoid precursors [21]. According to studies, methylating these flavonoids might boost their promise as pharmacological agents, leading to innovative uses [22]. Flavonoids have been shown to have a wide range of bioactivities, including anticancer, immunomodulation, and antioxidant activities, which can be enhanced to some extent by methylation [21]. Methylation of flavonoids through their free hydroxyl groups or C atoms significantly boosts their metabolic stability and improves membrane transport, resulting in easier absorption and significantly enhanced oral bioavailability [22]. Although cinnamic acid derivatives are methylated at the phenylpropanoid level in certain circumstances, and feruloyl-CoA serves as a poor substrate for chalcone synthase in others, methylation mainly happens at the C15 level. Poulton presents an overview of plant transmethylation and demethylation processes [23]. Grisebach documented the substrate specificities of several O-methyltransferases from parsley and soybean cell cultures, as well as shoots of Chrysosplenium americanum [24], while Heller and Forkmann added further test characteristics [25]. The methyl donor in all of these reactions is S-adenosyl-L-methionine. Ibrahim’s group has now identified five different O-methyltransferases in the flavonol pathway from C. americanum [26].

2.2.3 Glycosylation

The stability of flavonoids under glycosylation reaction circumstances is an important element to consider. For some flavonoids, direct glycosylation might result in the degradation of the changed molecules. It should also be noted that glycosylation is more than just attaching the carbohydrate residue to the flavonoid component of the molecule; it also involves the removal of the protecting groups. Some flavonoids may be partially destroyed as a result of this procedure [27]. It is generally agreed that glycosylation is a late or terminal step in flavonoid glycoside biosynthesis, except for acylation and prenylation reactions. Since the glycosylation step converts the flavonoid into a more water-soluble constituent, a step necessary for the retention of some flavonoids in the vacuole, the site of the glycosylation might be expected to act at the tonoplast boundary during transfer via the cisternae of, or vesicles derived from the endoplasmic reticulum [16]. The process of direct glycosylation for some classes of flavonoids can lead to the destruction of the modified compounds. There are two major types of linkages that form either O-glycosides or C-glycosides. Parsley preparations contain both types. There is a strict specificity for the position of the hydroxyl group, generally at the 3, 5, and 7 positions of the C- and A-rings. Both 3′ and 4’ B-ring glycosides are known. Recently, rarer 2′ and 5′ glycosides of highly methylated flavonol glucosides have been identified in Ibrahim’s laboratory [26].

2.2.4 Acylation

Many flavonoid families include acylated sugars. They exhibit a variety of physicochemical characteristics and biological activity; however, they have limited solubility and stability. To make use of these features, various publications have indicated that enzymatic acylation of these molecules with fatty and aromatic acids by protease and lipase under varied working conditions is a potential strategy. However, it is critical to strike a balance between increasing stability and solubility while maintaining biological activity. In fact, the acylation site (regioselectivity) can significantly alter these features [28]. The acyl groups are often aromatic acids like hydroxycinnamic acids or aliphatic acids like malonic acid. They appear to be position-specific for glucoside. Malonyl glucosides, which are catalyzed by malonyl transferases, are found in isoflavonoids, flavones, flavonols, and potentially anthocyanins. O-Malonyltransferases were isolated from parsley, which included malonated flavones and flavonols. Aromatic acylation, particularly of anthocyanins, has been observed in Silene and Matthiola sp. In both cases, the acyl groups transferred were either 4-coumaroyl or caffeoyl. Acylation has been observed to promote flavonoid absorption into parsley vacuoles; alterations in the molecular symmetry of the malonylglucosides are thought to be responsible for flavonoid vacuolar entrapment inside the vacuole [29].

2.2.5 Prenylation

Prenylflavonoids are useful natural compounds found in a wide range of plants. They frequently have diverse biological features, such as phytoestrogenic, antibacterial, antitumor, and antidiabetic qualities [28]. Prenyl groups are frequently found in phytoalexins and stress-induced isoflavonoids. They are occasionally cyclized. Pterocarpans having a 2′-oxy function and a phenyl group linked to the B-ring were the most active insect feeding deterrents [30]. Elicitor-challenged bean cell cultures include a prenyltransferase in a microsomal fraction that adds a prenyl group at position 10 on the “B-ring” (also known as the D-ring) of 3,9-dihydroxypterocarpan to create phaseollidin, which is then cyclized to phaseollin. Dimethylallylpyrophosphate was the prenyl donor. The identical preparations were capable of introducing prenyl groups into medicarpin and coumestrol, but the products were not recognized. Previously, two distinct flavonoid-specific prenyltransferases that need Mn2+ for full activity were discovered in soybean cotyledons and cell suspension cultures [31].

2.2.6 Sulfonation

Nature mostly employs sulfation of endogenous and external substances to minimize possible harm. Sulfonated flavonoids have recently been found in substantial numbers. The majority of them are sulfate esters of common flavones and flavonols. Flavone sulfates are mostly composed of apigenin, luteolin, or its 6- and 8-hydroxy derivatives. Flavonol glycosides were sulfated through the sugar or a separate hydroxyl group. They are found in both dicots and monocots, primarily in herbaceous species or advanced morphological groupings. They are exclusively seen in ferns on rare occasions and have not been identified in bryophytes or gymnosperms [32].

2.3 Stereochemistry

Flavanones have a unique structural property known as chirality that separates them from all other groups of flavonoids (Figure 3). The chemical structure of all flavanones is based on a C6–C3–C6 configuration consisting of two aromatic rings connected by a three-carbon bond [33]. Almost all flavanones have one chiral carbon atom in position 2 (Figure 3). Except for a subgroup of flavanones known as 3-hydroxyflavanones or dihydroflavonols, which have two chiral carbon atoms in positions 2 and 3 (Figure 4). In the C7 position of ring A, certain flavanones include an extra d-configured mono or disaccharide sugar. These flavanone-7-O-glycosides occur as diastereoisomers or epimers with opposing configurations at just one of two or more tetrahedral stereogenic centers in the corresponding chemical entities [34].

Figure 3.

Spatial disposition of the enantiomers of chiral flavanones.

Figure 4.

Chemical structure of the chiral 3-hydroxyflavanones or dihydroflavonols.

Most natural flavonoids now only have one stereoisomer at C-2. The RS nomenclature identifies the R or S that changes at carbon 2 without any change in stereochemistry, depending on the choice of the change adjacent groups should lead to confusion for flavonoid metabolism. It is not sufficient to utilize (+) - or 2,3-cis or -trans alone to define the four potential isoforms of dihydroquercetin or catechin; consequently, consideration should be given to alternate terminology. For mirror pictures, the ent-prefix is utilized. (+)-Catechin (2,3-trans isomer) with 2R, 3S absolute stereochemistry is simply known as catechin, whilst its mirror counterpart (−)-catechin (2,3-trans) with 2R, 3R stereochemistry is simply known as ent-catechin. Similarly, the (−)-epicatechin (2,3-cis) isomer (2R, 3R) and its mirror image (2S,3S) are known as epicatechin and ent-epicatechin [35].

There are other structures designated for hydroxylation patterns and inter-liquid bonding. To minimize ambiguity in the RS system of the configuration of the interflavanoid bond at C-4, Porter and Hemingway provided sugar chemistry terminology, particularly when defining proanthocyanidin isomers. The words are also used to characterize the stereochemistry of the added hydroxyl group at the C-3 position, which results in the 2,3-cis (a-OH) and more prevalent 2,3-trans (B-OH) forms. However, such language does not accurately describe the metabolic route [16].

2.4 Overall pathways metabolism of flavonoid

Flavonoids, which include chalcones, flavones, flavonols, anthocyanins, and proanthocyanidins, are abundant in plants and have been extensively researched using biochemical and molecular biology approaches. Until recently, liverworts and mosses were thought to be the earliest flavonoid-producing plants. Genes encoding enzymes in the phenylpropanoid biosynthetic pathway, including the first two enzymes for flavonoid biosynthesis (chalcone synthase and chalcone isomerase), have not been found in the algal genera Chlamydomonas, Micromonas, Ostreococcus, and Klebsormidium, whereas genes encoding enzymes in the shikimate pathway have been found in algae, liverwort [36].

The overall route to main flavonoid groups via 5,7-hydroxy A-rings. The key intermediates in the production of flavonoids are 4-coumaroyl-coA and 3-malonyl-coA. Synthesis of 4-coumaroyl-CoA and malonyl-CoA naringenin chalcone is synthesized by chalcone synthase, an enzyme involved in the phenylpropanoid pathway. Naringenin chalcone has the ability to spontaneously cyclize to naringenin. Furthermore, naringenin chalcone synthesizes a variety of chemicals such as chalcones, aurones, and biflavonoids (it can synthesize from flavanones). Naringenin has three routes for drug synthesis. The first is the direct synthesis of isoflavonoids, followed by the addition of 3’ OH to produce eriodictyol, followed by the addition of 5’OH to produce 5’OH-erio flavones. Eriodictyol may be combined with 3-OH to get DHQ (dihydroquercetin), then with 5’OH to form DHM (dihydromyricetin) to form flavonols and flavan-3,4-diols (it can synthesize flavan-3-ols, proanthocyanidins, and anthocyanidins). Finally, adding 3-OH to naringenin results in DHK (dihydrokaempferol) (Figure 5) [16].

Figure 5.

Overall pathway to major flavonoid groups with 5,7-hydroxyl a rings (DHK = dihydrokaempferol; DHQ = dihydroquercetin; DHM = dihydromyricetin).

Advertisement

3. Liquid chromatography in the analysis of flavonoid metabolism

To show the chemical variety of flavonoids, chromatographic methods have been utilized to examine their structures. Previously, the major methods used to analyze flavonoids were paper chromatography, thin layer chromatography, column chromatography, and liquid chromatography (LC) [36]. Efficient screening of plant extracts may be accomplished using biological assays as well as chromatographic techniques such as high-performance liquid chromatography (HPLC) in conjunction with different detection modalities [37]. Because it permits systematic profiling of complex plant samples and especially focuses on their identification and consistent assessment of the found compounds, HPLC is a potent tool for the quick investigation of bioactive ingredients. Modern HPLC separation of flavonoids nearly entirely uses reversed-phase liquid chromatography (RP-LC), with significant exceptions being normal phase liquid chromatography (NP-LC) for oligomeric proanthocyanins [38] and the recent rising use of hydrophilic interaction chromatography (HILIC). Other flavonoid separation methods include mixed-mode ion-exchange-reversed phase separation of anthocyanins [39, 40, 41], size exclusions chromatography (SEC) analysis of flavonol glycosides [42], and theaflavins and proanthocyanidins [43]. However, because of the infrequent usage of the later modes, this chapter will concentrate mostly on RP-LC in line with the extent and predominance of this method in flavonoid literature.

RP-LC has proved its applicability for the separation of flavonoids depending on the nature of the aglycone (including the oxidation state, substitution patterns, and stereochemistry), the type and degree of glycosylation, and the nature and degree of acylation. The vast majority of RP-LC separations are accomplished using C18 octadecyl-silica (ODS) phases, however, C8 [44], C12 [45], phenyl or phenyl-hexyl [46, 47, 48, 49, 50], pentafluorophenyl (PFP) [51, 52, 53] and polar embedded RP phases [54, 55, 56] as well as polymeric RP-LC phases were still widely used [57]. Aqueous/organic phases including methanol, acetonitrile, and less commonly tetrahydrofuran [58], isopropanol [59] or ethanol [57], and acidic modifiers such as acetic acid, formic acid, ammonium acetate, or trifluoroacetic acid (TFA) [59] are typical mobile phases (phosphoric, citric, or perchloric acids have also been used in combination with UV detection, although these are not suited to hyphenation with MS). Highly acidic mobile phases (>4–10% formic acid, 0.1–0.6% TFA) [60, 61, 62, 63] are utilized for anthocyanins to assure the presence of flavylium cationic species in solution and therefore increase chromatographic efficiency. To detect and/or identify flavonoids, a variety of detectors may and have been used in conjunction with HPLC separation. These include electrochemical detection (ED) [64, 65], fluorescence (FL) [66], UV–Vis, diode array [59, 67], NMR [68, 69], and of course MS detectors [70, 71]. The most prevalent currently are diode array and MS detectors, which will be explored briefly in this and the next sections.

3.1 Liquid chromatography (LC) with ultraviolet (UV) detector

Particularly in early flavonoid research, the conjugated aromatic nature of flavonoids proved to be a significant advantage: absorption at relatively long wavelengths increases the selectivity of qualitative and quantitative spectrophotometric methods, and the distinctive spectra of various classes of flavonoids allow differentiation between them. These qualities are similarly helpful when HPLC separation is combined with UV–Vis detection. Flavonoids exhibit two UV–Vis absorption maxima: Band II (Band II), which is attributed to the A-ring, and Band III (Band III), which is attributed to the B-ring (Band 274 I). Due to the fact that it offers more specialized information and since all flavonoids absorb between 240 and 285 nm, the latter of these is more useful. Due to the absence of conjugation between the A and B rings, flavanols, flavanones, dihydroflavonols, and isoflavones only display Band II absorption (269–279 nm). Anthocyanidins may be easily identified by their Band I absorption between 460 and 550 nm in the visible range, in contrast to flavonols and flavones, which exhibit Band I absorption between 300 and 380 nm [72]. Figure 6 provides typical illustrations of the UV–Vis absorbance spectra of the major groups of flavonoids.

Figure 6.

UV–vis absorbance spectra of the principal classes of flavonoids: (a) Luteolin [73]; (b) quercetin [74].

In research by Mohamed A. Farag et al., an integrated approach utilizing HPLC–UV was used for the large-scale and systematic identification of polyphenols in Medicago truncatula root and cell culture. UV spectra (200–600 nm) were recorded for different flavonoid sub-classes including 26 isoflavones (peaks 1–6, 8, 10, 12, 14, 19–29, and 31–35), 4 flavones (peaks 9, 13, 18, and 31), 2 flavanones (peaks 21 and 30), 2 aurones (peaks 7 and 11), and a chalcone (peak 32) with isoflavone representing the major sub-class. For example, isoflavones typically have a maximum absorbance near 255 nm with a second maximum between 300 and 330 nm (peak 34 in Figure 7), whereas aurones have the first maximum near 250 nm and the second peak around 390 nm (peak 7 in Figure 7) [75].

Figure 7.

HPLC–UV at 260 nm (a), 1 and 2 represent UV spectra of peak 7 (hispidol 40-O-glucoside) and peak 34 (afrormosin), respectively [75].

Kim-Ngan Huynh Nguyen et al.’s study quantified seven major compounds, including phenolic acids (chlorogenic acid, caffeic acid, and p-coumaric acid) and flavonoids (rutin, quercitrin, quercetin, and kaempferol) in three aerial parts of Physalis angulata, that is, leaves, calyces, and fruits. Chromatographic separation was carried out on a Kromasil C18 column (150 mm × 4.6 mm i.d., 5 μm) with a gradient elution of 0.1% formic acid in acetonitrile, 0.2% ammonium acetate/0.1% formic acid in water and methanol at a flow rate of 1.0 mL/min; detection was at 250 and 300 nm. The applications of liquid chromatography (LC) with ultraviolet (UV) for the analysis of flavonoid metabolism of plants that show in Table 1 (Figure 8) [81].

PlantFlavonoidDetection wavelengthStationary phases (column dimensions)Mobile phaseRef.
Medicago truncatulaisoflavone afrormosin, irilone260-321 nm, 260-325 nmC18, 5 μm, 4.6 × 250 mm columnWater (0.1% acetic acid) and acetonitrile using gradient mode[75]
Brassica rapa L. Ssp. chinensis L. (Hanelt.)Quercetin, kaempferol, isorhamnetin204 nm, 254 nm, 352 nmAlltima HP C18 column0.5% orthophosphoric acid (v/v) in 30% methanol (v/v) and 0.5% orthophosphoric acid (v/v) in methanol[76]
Lupinus albus, Lupinus angustifoliusGenistein (isoflavone), 2′-hydroxygenistein glycosides259 nmRP C-18 silica gelA (95% acetonitrile, 4.5% H2O, 0.5% acetic acid; v/v/v)[77]
Brazilian Vernonieae (Asteraceae)flavones, flavonols, flavone C-glycosides, flavonol O-glycosides200–600 nmKinetex 1.7 mm XB-C18Water and acetonitrile, both with formic acid 0.1% (v/v) using gradient mode[78]
Sutherlandia frutescens (L.)Four flavonoids260 nmC18 columnWater (0.1% acetic acid) and acetonitrile (0.1% acetic
acid) using gradient mode
[79]
Orange juiceFlavanones, flavones, flavonols280 nm, 265 nm, 265 nmC18 standard-bore columnAqueous formic acid (pH 2.4)/and acetonitrile (80:20, v/v)[80]

Table 1.

Applications of liquid chromatography (LC) with ultraviolet (UV) for the analysis of flavonoid metabolism.

Figure 8.

HPLC chromatogram of the mixed standards solution (a), P. angulata leaves (b). P1–P13: Phenolic acids, including P4: Chlorogenic acid, P6: Caffeic acid, and P8: p-coumaric acid. F1–F9: Flavonoids, including F4: Rutin and F9: Quercetin [81].

3.2 Liquid chromatography (LC) with mass spectrometry (MS)

In contrast to 30 years ago, routine separation and preliminary identification of complex mixtures of flavonoids ranging over many orders of magnitude in concentration are now achievable because of the combination of chromatographic resolution offered by HPLC and structural data offered by MS. Furthermore, during the past 10 years, significant advancements in LC technology have been realized. UHPLC (ultra-high pressure liquid chromatography), alternative stationary phases including monoliths and superficially porous phases, high-temperature HPLC [82, 83], and multidimensional HPLC [84, 85, 86, 87] are a few noteworthy advancements.

Cheminformatics methods combined with LC-MS/MS provide a potent tool for high-throughput surveys of flavonoid variety [88, 89]. Utilizing straightforward solvent combinations and LC columns, glycosylated, acylated, and prenylated flavonoid molecules and their aglycones may be separated. For MS/MS analysis, the isolated molecules are ionized next. In order to analyze flavonoids, LC-tandem mass spectrometry (LC-MS/MS) has emerged as the method of choice. Algae had previously been thought to have no flavonoids. But using an LC-MS/MS technique, flavonoids were identified as intermediates and end products, demonstrating the occurrence of flavonoid production in microalgae [90]. It implies that the undiscovered flavonoids in every plant species can be discovered using cutting-edge metabolomics technology.

With an emphasis on general ionization and fragmentation processes, a brief overview of the underlying knowledge pertinent to the MS detection and MS/MS structural elucidation of flavonoids will be provided in this part. Additionally, specialized research reports provide much more in-depth information on particular classes of flavonoids, including dihydroflavonols [91], isoflavones [92], flavone-di-C-glycosides [93], flavonoid-aglycones [94], flavonoid-O-glycosides [95], and flavonoid glycosides [96]. The discussion that follows will be restricted to the API sources electrospray ionization (ESI), atmospheric pressure chemical ionization (APCI), and atmospheric pressure chemical ionization, which are presently the most pertinent LC-MS ionization sources (APPI). The applications of liquid chromatography (LC) with mass spectrometry (MS) to the analysis of flavonoid metabolism are shown in Table 2.

PlantFlavonoidsStationary phase (column dimensions)Mobile phase(s)MS analyzerRef.
Arabidopsis thalianaChrysoeriol (3′ -O-methylluteolin), kaempferol, myricetin, and quercetinA Superspher 100 RP-18 columnA (95% acetonitrile, 4.5% H2O, 0.5% formic acid, v/v/v) and B (95% H2O, 4.5% acetonitrile, 0.5% formic acid v/v/v)ESI-QqQ[97]
Citrus fruitsNaringin, apigenin, eriodictyol, homoeriodictyol, and hesperetinWaters BEH C18 column, Phenomenex Kinetex C18 column, and Agilent Poroshell 120 EC-C18 column. (2.1 mm × 50 mm, 1.7 μm)Water and methanol, both with 0.1% formic acidESI-QqQ[98]
Chenopodium hybridumRutin, 3-kaempferolEclipse XDB-C18 (150 × 4.6 mm, 5 μmA: 0.1% Acid formic, B: Acetonitrile with 0.1% Acid formicESI QqQ-IT[99]
Orange peelKaempferol, neohesperidin, luteolin, homoorientin, tangeretin, diosmetin formononetin quercetin, hesperidin, apigenin, naringenin, naringin, and oleuropeinMediterranea Sea C18 (150 × 0.46 mm, 3 μm);A: 0.1% Acid formic, B: Acetonitrile with 0.1% Acid formicESI-QqQ[100]
Vitex negundo var. cannabifoliaVitexin, hyperoside, luteoloside, liquiritin, and albiflorinACQUITY UPLC Cortest C18 (100 × 2.1 mm, 1.6 μm);A: Acetonitrile, B: 5% MeOH:H2O with 0.1% acid formicESI-QqQ[101]

Table 2.

Applications of liquid chromatography (LC) with mass spectrometry (MS) for the analysis of flavonoid metabolism.

*ESI: electrospray ionization; QqQ: triple quadrupole MS, IT: ion trap MS.

Shoucuang Wang et al. (2017) researched comprehensive profiling of metabolites in citrus fruits. Non-targeted high-performance liquid chromatography with diode array detection and electrospray ionization mass spectrometry (HPLC-DAD-ESI-MS/MS) was used to profile the metabolites in fruit tissues. As a result, 7416 metabolic signals were detected. In addition to those reported metabolites, seven O-glycosylpolymethoxylated flavonoids were newly annotated in the study. To better characterize these flavonoids, the 3′,4′,5,6,7,8-hexamethoxyflavone standard (m70, RT 15.3 min, m/z 403.1389, error − 0.5 ppm) was analyzed first. The precursor ions of the standard compound lost one to four methyl radicals in the MS/MS spectrum to form the base peaks of [M + H - 15]+, [M + H - 30]+, [M + H - 45]+, or [M + H - 60] + (Figure 9A). The characteristic loss of 162 Da was observed in the MS/MS spectra corresponding to the dissociation of a hexose moiety and a series of methyl loss of the diagnostic fragments of 15 and 30 Da (Figure 9BD) [102].

Figure 9.

Mass spectra and structures of polymethoxylated flavonoids glycosides in citrus. (A) 3, 4, 5, 6, 7,8-hexamethoxyflavone (m070). (B) Dihydroxy-trimethoxyflavone -O-hexoside (m117). (C) Hydroxytetramethoxyflavone-O-hexoside (m119). (D) Monohydroxy-hexamethoxyflavone-O-hexoside (m133). PMFs, DFI, diagnostic fragment ions of polymethoxylated flavonoids [102].

Paola Montoro et al. (2012) researched the metabolic profiles of different extracts (obtained by petals, stamens, and flowers) by LC-ESI-IT MS (liquid chromatography coupled with electrospray mass spectrometry equipped with an ion trap analyzer). MS/MS experiments were diagnostic for the identification of specific fragmentation patterns, that is, sugar loss for flavonoid O-glycosides or the loss of specific esterification units. Interpretation of the ESI-MS/MS experiment obtained by the analysis of Crocus sativus petals extracts allowed us to tentatively identify 31 flavon derivatives in the extracts under investigation, mainly glycosidated and metoxilated derivatives of kaempferol, quercetin, isorhamnetin, and tamaryxetin (Figure 10) [103].

Figure 10.

LC-ESI-MS. TIC and reconstructed ion chromatograms for Crocins qualitative analysis in H2O/EtOH extract of Crocus sativus petals [103].

3.3 High-performance liquid chromatography in chiral flavonoid

Enantiomer separation, resolution, and analysis have traditionally been achieved by the transitory or covalent synthesis of diastereoisomers. Diastereoisomers can be separated on an achiral chromatographic column by differential contact and retention because they have distinct physicochemical characteristics in an achiral environment. On a chemically bonded chiral stationary phase (CSP) with an achiral mobile phase, racemic flavonoid resolution has typically been achieved through chromatographic enantiospecific resolution through transient production of diastereoisomers.

In 1980, one of the earliest publications on flavanone glycoside HPLC separation appeared. Both naringin and narirutin may be acetylated using an equal mixture of pyridine and acetic anhydride and then resolved at low temperatures (between 0 and 5 OC) [104]. Prunus callus (sweet cherries), oranges, and grapefruit’s prunin (naringenin-7-O-glucoside) epimers were initially separated in the middle of the 1980s using benzoylated derivatives [105]. On Cyclobond I columns, the separation of prunin benzoate and naringin benzoate has also been shown. Naringenin derivatization to naringenin tribenzoate and separation on a Chiralcel OD column are also mentioned in the literature. Naringenin’s hydroxyl groups may prevent chiral identification in this stationary phase as the enantiomers could not be resolved [34].

The main advantage of chiral separation methods over achiral methods is a better understanding of the pharmacokinetics of flavanones and the development of effective dosing regimens. In the case of racemic flavanones or stereochemically pure flavanones, this requires knowledge of the in vivo behavior of the enantiomers and epimers. During the drug development process, understanding and comprehending the conformational stability of chiral compounds may have a significant influence on the pharmacological, pharmacokinetic, and pharmacodynamic data. Using stereospecific analytical techniques, racemization and enantiomerization/epimerization may be studied. Rapid interconversion in vivo would eliminate any potential distinctions in the enantiomers’ medicinal or harmful effects, making the synthesis of stereochemically pure enantiomers useless. Chirality must be taken into account from the beginning of the development process for stereochemically pure compounds and racemates [34].

In a study by Gaggeri R et al. (2011), the HPLC enantioselective separation of (R/S)-naringenin (Figure 11), a chiral flavonoid found in several fruits juices and well-known for its beneficial health-related properties, including antioxidant, anti-inflammatory, cancer chemopreventive, immunomodulating and antimicrobial activities, has been performed on both analytical and (semi)-preparative scale using amylose-derived Chiralpak AD chiral stationary phase (CSP). A standard screening protocol for cellulose and amylose-based CSPs was firstly applied to analytical Chiralcel OD-H and Chiralpak AD-H, as well as to Lux Cellulose-1, Lux Cellulose-2, and Lux Amylose-2 in order to identify the best experimental condition for the subsequent scaling-up. Using Chiralpak AD-H and eluting with pure methanol (without acidic or basic additives), relatively short retention times, high enantioselectivity, and good resolution (Rs = 3.48) were observed. Therefore, these experimental conditions were properly scaled up to (semi)-preparative scale using both a prepacked Regispack column and a Chiralpak AD column packed in-house with bulk CSP [106].

Figure 11.

(R/S)-Naringenin.

Advertisement

4. Conclusion

Many beneficial health effects have been attributed to flavonoids, which are popular in the plant. The study of metabolism and bioavailability is very important in defining the pharmacological and toxicological profile of these flavonoid compounds. Due to great structural diversity among flavonoids, these profiles differ greatly from one compound to another, so the most abundant polyphenols in our diet are not necessarily the ones that reach target tissues. Therefore, careful analysis of flavonoids and their metabolites in biological systems is critical. Several hundred papers on the HPLC of flavonoids have been published in the past 20 or so years, yet HPLC methods can detect flavonoids across one, two, or perhaps three subclasses in one run. The improvements in HPLC flavonoid analysis closely resemble and, to a certain extent, build on those in domains like proteomics and metabolomics, which are supported by important breakthroughs.

Advertisement

Acknowledgments

The authors would like to express their hearty gratitude to Can Tho University of Medicine and Pharmacy. The authors also thank all of their colleagues for their excellent assistance.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Panche AN, Diwan AD, Chandra SR. Flavonoids: An overview. Journal of Nutritions Science. 2016;5:e47. DOI: 10.1017/jns.2016.41
  2. 2. Mahomoodally MF, Gurib-Fakim A, Subratty AH. Screening for alternative antibiotics: An investigation into the antimicrobial activities of medicinal food plants of Mauritius. Journal of Food Science. 2010;75(3):M173-M177. DOI: 10.1111/j.1750-3841.2010.01555.x
  3. 3. Kumar S, Pandey S, Pandey AK. In vitro antibacterial, antioxidant, and cytotoxic activities of Parthenium hysterophorus and characterization of extracts by LC-MS analysis. BioMed Research International. 2014;2014:495154. DOI: 10.1155/2014/495154
  4. 4. Dixon RA, Dey PM, Lamb CJ. Phytoalexins: Enzymology and molecular biology. Advances in Enzymology and Related Areas of Molecular Biology. 1983;55:1-136. DOI: 10.1002/9780470123010.ch1
  5. 5. Heim KE, Tagliaferro AR, Bobilya DJ. Flavonoid antioxidants: Chemistry, metabolism and structure-activity relationships. The Journal of Nutritional Biochemistry. 2002;13(10):572-584. DOI: 10.1016/s0955-2863(02)00208-5
  6. 6. Middleton E Jr. Effect of plant flavonoids on immune and inflammatory cell function. Advances in Experimental Medicine and Biology. 1998;439:175-182. DOI: 10.1007/978-1-4615-5335-9_13
  7. 7. Rice-Evans CA, Miller NJ, Bolwell PG, Bramley PM, Pridham JB. The relative antioxidant activities of plant-derived polyphenolic flavonoids. Free Radical Research. 1995;22(4):375-383. DOI: 10.3109/10715769509145649
  8. 8. Cook NC, Samman S. Review: Flavonoids-chemistry, metabolism, cardioprotective effects and dietary sources. Journal of Nutritional Biochemistry. 1996;7(2):66-76
  9. 9. Kumar S, Mishra A, Pandey AK. Antioxidant mediated protective effect of Parthenium hysterophorus against oxidative damage using in vitro models. BMC Complementary and Alternative Medicine. 2013;13:120. DOI: 10.1186/1472-6882-13-120
  10. 10. Kumar S, Pandey AK. Phenolic content, reducing power and membrane protective activities of solanum xanthocarpum root extracts. Vegetos. 2013;26(1):301-307. DOI: 10.5958/j.2229-4473.26.1.043
  11. 11. Winkel-Shirley B. Evidence for enzyme complexes in the phenylpropanoid and flavonoid pathways. Physiologia Plantarum. 1999;107(1):142-149. DOI: 10.1034/j.1399-3054.1999.100119.x
  12. 12. Ganzera M, Sturm S. Recent advances on HPLC/MS in medicinal plant analysis-an update covering 2011-2016. Journal of Pharmaceutical and Biomedical Analysis. 2018;147:211-233. DOI: 10.1016/j.jpba.2017.07.038
  13. 13. Costa DC, Costa HS, Albuquerque TG, et al. Advances in phenolic compounds analysis of aromatic plants and their potential applications. Trends in Food Science and Technology. 2015;45:336-354
  14. 14. Kumar S, Pandey AK. Chemistry and biological activities of flavonoids: An overview. The Scientific World Journal. 2013;2013:162750. DOI: 10.1155/2013/162750
  15. 15. Karak P. Biological activities of flavonoids: An overview. IJPSR. 2019;3:1567-1574. DOI: 10.13040/IJPSR.0975-8232.10(4).1567-74
  16. 16. Stafford H. Flavonoid Metabolism. Taylor and Francis, Florida: CRC Press. 1990;360
  17. 17. Harborne JB, Baxter H. A Handbook of the Natural Flavonoids. Chichester, UK: Wiley; 1999
  18. 18. Hrazdina G, Wagner GJ. Metabolic pathways as enzyme complexes: Evidence for the synthesis of phenylpropanoids and flavonoids on membrane associated enzyme complexes. Archives of Biochemistry and Biophysics. 1985;237(1):88-100. DOI: 10.1016/0003-9861(85)90257-7
  19. 19. Werck-Reichhart D. Cytochromes P450 in phenylpropanoid metabolism. Drug Metabolism and Drug Interactions. 1995;12(3-4):221-243. DOI: 10.1515/dmdi.1995.12.3-4.221
  20. 20. Halbwirth H. The creation and physiological relevance of divergent hydroxylation patterns in the flavonoid pathway. International Journal of Molecular Sciences. 2010;11(2):595-621. DOI: 10.3390/ijms11020595
  21. 21. Wen L, Jiang Y, Yang J, Zhao Y, Tian M, Yang B. Structure, bioactivity, and synthesis of methylated flavonoids. Annals of the New York Academy of Sciences. 2017;1398(1):120-129. DOI: 10.1111/nyas.13350
  22. 22. Koirala N, Thuan NH, Ghimire GP, Thang DV, Sohng JK. Methylation of flavonoids: Chemical structures, bioactivities, progress and perspectives for biotechnological production. Enzyme and Microbial Technology. 2016;86:103-116. DOI: 10.1016/j.enzmictec.2016.02.003
  23. 23. Poulton JE. Transmethylation and demethylation reac tions in the metabolism of secondary plant products. In: Conn EE, editor. The Biochemistry of Plants, Vol. 7: Secondary Plant Products. New York: Academic Press; 1981. pp. 667-723
  24. 24. De Luca V, Ibrahim RK. Enzymatic synthesis of polymethylated flavonols in Chrysosplenium americanum. I. Partial purification and some properties of S-adenosyl-L-methionine: Flavonol 3-, 6-, 7-, and 4'-O-methyltransferases. Archives of Biochemistry and Biophysics. 1985;238(2):596-605. DOI: 10.1016/0003-9861(85)90205-x
  25. 25. Hahlbrock K, Grisebach H. Biosynthesis of Flavonoids. Boston, MA: In The flavonoids Springer; 1975. pp. 866-915
  26. 26. Ibrahim RK, De Luca V, Khouri H, Latchinian L, Brisson L, Charest PM. Enzymology and compartmentation of polymethylated flavonol glucosides in Chrysosplenium americanum. Phytochemistry. 1987;26(5):1237-1245
  27. 27. Khodzhaieva RS, Gladkov ES, Kyrychenko A, Roshal AD. Progress and achievements in glycosylation of flavonoids. Frontiers in Chemistry. 2021;9:637994. Published 2021 Mar 31. DOI: 10.3389/fchem.2021.637994
  28. 28. Yang X, Yang J, Jiang Y, et al. Regiospecific synthesis of prenylated flavonoids by a prenyltransferase cloned from fusarium oxysporum. Scientific Reports. 2016;6:24819. Published 2016 Apr 21. DOI: 10.1038/srep24819
  29. 29. Matern U, Reichenbach C, Heller W. Efficient uptake of flavonoids into parsley (Petroselinum hortense) vacuoles requires acylated glycosides. Planta. 1986;167(2):183-189. DOI: 10.1007/BF00391413
  30. 30. Yoneyama K, Akashi T, Aoki T. Molecular characterization of soybean Pterocarpan 2-Dimethylallyltransferase in Glyceollin biosynthesis: Local gene and Whole-genome duplications of Prenyltransferase genes led to the structural diversity of soybean Prenylated Isoflavonoids. Plant & Cell Physiology. 2016;57(12):2497-2509. DOI: 10.1093/pcp/pcw178
  31. 31. Zähringer U, Schaller E, Grisebach H. Induction of phytoalexin synthesis in soybean. Structure and reactions of naturally occurring and enzymatically prepared prenylated pterocarpans from elicitor-treated cotyledons and cell cultures of soybean. Zeitschrift für Naturforschung C. 1981;36(3-4):234-241
  32. 32. Correia-da-Silva M, Sousa E, Pinto MM. Emerging sulfated flavonoids and other polyphenols as drugs: Nature as an inspiration. Medicinal Research Reviews. 2014;34(2):223-279. DOI: 10.1002/med.21282
  33. 33. Shahidi F, Wanasundara PK. Phenolic antioxidants. Critical Reviews in Food Science and Nutrition. 1992;32(1):67-103. DOI: 10.1080/10408399209527581
  34. 34. Yáñez JA, Andrews PK, Davies NM. Methods of analysis and separation of chiral flavonoids. Journal of Chromatography. B, Analytical Technologies in the Biomedical and Life Sciences. 2007;848(2):159-181. DOI: 10.1016/j.jchromb.2006.10.052
  35. 35. Hemingway RW, Foo LY, Porter LJ. Linkage isomerism in trimeric and polymeric 2,3-cis-procyanidins. Journal of the Chemical Society, Perkin Transactions. 1982;1:1209
  36. 36. Yonekura-Sakakibara K, Higashi Y, Nakabayashi R. The origin and evolution of plant flavonoid metabolism. Frontiers in Plant Science. 2019;10:943. Published 2019 Aug 2. DOI: 10.3389/fpls.2019.00943
  37. 37. Wolfender JL, Rodriguez S, Hostettmann K. Liquid chromatography coupled to mass spectrometry and nuclear magnetic resonance spectroscopy for the screening of plant constituents. Journal of Chromatography. A. 1998;794(1-2):299-316. doi:10.1016/S0021-9673(97)00939-4
  38. 38. LEA, Andrew GH. High performance liquid chromatography of cider procyanidins. Journal of the Science of Food and Agriculture. 1979;30(8):833-838
  39. 39. McCallum JL, Yang R, Young JC, Strommer JN, Tsao R. Improved high performance liquid chromatographic separation of anthocyanin compounds from grapes using a novel mixed-mode ion-exchange reversed-phase column. Journal of Chromatography. A. 2007;1148(1):38-45. DOI: 10.1016/j.chroma.2007.02.088
  40. 40. Vergara C, Mardones C, Hermosín-Gutiérrez I, von Baer D. Comparison of high-performance liquid chromatography separation of red wine anthocyanins on a mixed-mode ion-exchange reversed-phase and on a reversed-phase column. Journal of Chromatography. A. 2010;1217:5710-5717
  41. 41. Jin H, Liu Y, Guo Z, Yang F, Wang J, Li X, et al. High-performance liquid chromatography separation of cis-trans anthocyanin isomers from wild Lycium ruthenicum Murr. Employing a mixed-mode reversed-phase/strong anion-exchange stationary phase. Journal of Agricultural and Food Chemistry. 2015;63(2):500-508. DOI: 10.1021/jf504525w
  42. 42. de Souza LM, Cipriani TR, Sant'ana CF, Iacomini M, Gorin PA, Sassaki GL. Heart-cutting two-dimensional (size exclusion x reversed phase) liquid chromatography-mass spectrometry analysis of flavonol glycosides from leaves of Maytenus ilicifolia. Journal of Chromatography. A. 2009;1216(1):99-105. DOI: 10.1016/j.chroma.2008.11.032
  43. 43. Scoparo CT, de Souza LM, Dartora N, Sassaki GL, Gorin PA, Iacomini M. Analysis of Camellia sinensis green and black teas via ultra high performance liquid chromatography assisted by liquid-liquid partition and two-dimensional liquid chromatography (size exclusion × reversed phase). Journal of Chromatography. A. 2012;1222:29-37. DOI: 10.1016/j.chroma.2011.11.038
  44. 44. Harborne JB, Boardley M. Use of high-performance liquid chromatography in the separation of flavonol glycosides and flavonol sulphates. Journal of Chromatography A. 1984;299:377-385. DOI: 10.1016/s0021-9673(01)97853-7
  45. 45. Omar MH, Mullen W, Crozier A. Identification of proanthocyanidin dimers and trimers, flavone C-glycosides, and antioxidants in Ficus deltoidea, a Malaysian herbal tea. Journal of Agricultural and Food Chemistry. 2011;59(4):1363-1369
  46. 46. Dugo P, Cacciola F, Herrero M, Donato P, Mondello L. Use of partially porous column as second dimension in comprehensive two-dimensional system for analysis of polyphenolic antioxidants. Journal of Separation Science. 2008;31(19):3297-3308. DOI: 10.1002/jssc.200800281
  47. 47. Tanaka Y, Yanagida A, Komeya S, Kawana M, Honma D, Tagashira M, et al. Comprehensive separation and structural analyses of polyphenols and related compounds from bracts of hops (Humulus lupulus L.). Journal of Agricultural and Food Chemistry. 2014;62(10):2198-2206. DOI: 10.1021/jf405544n
  48. 48. Cesla P, Fischer J, Jandera P. Separation of phenolic acids and flavone natural antioxidants by two-dimensional method combining liquid chromatography and micellar electrokinetic capillary chromatography. Electrophoresis. 2010;31(13):2200-2210. DOI: 10.1002/elps.200900689
  49. 49. Guillarme D, Casetta C, Bicchi C, Veuthey JL. High throughput qualitative analysis of polyphenols in tea samples by ultra-high pressure liquid chromatography coupled to UV and mass spectrometry detectors. Journal of Chromatography. A. 2010;1217(44):6882-6890. DOI: 10.1016/j.chroma.2010.08.060
  50. 50. Wu LZ, Zhang XP, Xu XD, Zheng QX, Yang JS, Ding WL. Characterization of aromatic glycosides in the extracts of Trollius species by ultra high-performance liquid chromatography coupled with electrospray ionization quadrupole time-of-flight tandem mass spectrometry. Journal of Pharmaceutical and Biomedical Analysis. 2013;75:55-63
  51. 51. Jandera P, Hájek T, Vyňuchalová K. Retention and bandwidths prediction in fast gradient liquid chromatography. Part 2-Core-shell columns. Journal of Chromatography. A. 2014;1337:57-66
  52. 52. Mirali M, Ambrose SJ, Wood SA, Vandenberg A, Purves RW. Development of a fast extraction method and optimization of liquid chromatography-mass spectrometry for the analysis of phenolic compounds in lentil seed coats. Journal of Chromatography. B, Analytical Technologies in the Biomedical and Life Sciences. 2014;969:149-161. DOI: 10.1016/j.jchromb.2014.08.007
  53. 53. Gavina JM, Priem J, Wood CM, Xiao CW, Feng YL. Determination of isoflavones in rat serum using liquid chromatography-tandem mass spectrometry with a highly efficient core-shell column. Analytical and Bioanalytical Chemistry. 2013;405(8):2643-2651. DOI: 10.1007/s00216-012-6688-x
  54. 54. Feuereisen MM, Hoppe J, Zimmermann BF, Weber F, Schulze-Kaysers N, Schieber A. Characterization of phenolic compounds in Brazilian pepper (Schinus terebinthifolius Raddi) exocarp. Journal of Agricultural and Food Chemistry. 2014;62(26):6219-6226. DOI: 10.1021/jf500977d
  55. 55. Li C, Zhao Y, Guo Z, Zhang X, Xue X, Liang X. Effective 2D-RPLC/RPLC enrichment and separation of micro-components from Hedyotis diffusa Willd. And characterization by using ultra-performance liquid chromatography/quadrupole time-of-flight mass spectrometry. Journal of Pharmaceutical and Biomedical Analysis. 2014;99:35-44. DOI: 10.1016/j.jpba.2014.06.020
  56. 56. Šatínský D, Jägerová K, Havlíková L, et al. A new and fast HPLC method for determination of Rutin, Troxerutin, Diosmin and hesperidin in food supplements using fused-Core column technology. Food Analytical Methods. 2013;6:1353-1360. DOI: 10.1007/s12161-012-9551-y
  57. 57. Reichelt KV, Peter R, Paetz S, Roloff M, Ley JP, Krammer GE, et al. Characterization of flavor modulating effects in complex mixtures via high temperature liquid chromatography. Journal of Agricultural and Food Chemistry. 2010;58(1):458-464. DOI: 10.1021/jf9027552
  58. 58. Olszewska MA. New validated high-performance liquid chromatographic method for simultaneous analysis of ten flavonoid aglycones in plant extracts using a C18 fused-core column and acetonitrile-tetrahydrofuran gradient. Journal of Separation Science. 2012;35(17):2174-2183. DOI: 10.1002/jssc.201200287
  59. 59. Santos-Buelga C, García-Viguera C, Tomas-Barberan FA. On-line identification of flavonoids by HPLC coupled to diode Array detection, in: C. Santos-Buelga, G. Williamson (Eds.). Methods Polyphen. Anal., The Royal Society of Chemistry, Cambridge, 2003: pp. 92-127.
  60. 60. Lopes-da-Silva F, de Pascual-Teresa S, Rivas-Gonzalo J, Santos-Buelga C. Identification of anthocyanin pigments in strawberry (cv Camarosa) by LC using DAD and ESI-MS detection. European Food Research and Technology. 2002;214(3):248-253
  61. 61. de Villiers A, Cabooter D, Lynen F, Desmet G, Sandra P. High performance liquid chromatography analysis of wine anthocyanins revisited: Effect of particle size and temperature. Journal of Chromatography. A. 2009;1216(15):3270-3279. DOI: 10.1016/j.chroma.2009.02.038
  62. 62. Xu Y, Simon JE, Ferruzzi MG, Ho L, Pasinetti GM, Wu Q. Quantification of anthocyanidins in the grapes and grape juice products with acid assisted hydrolysis using LC/MS. Journal of Functional Foods. 2012;4:710-717
  63. 63. Díaz-García MC, Obón JM, Castellar MR, Collado J, Alacid M. Quantification by UHPLC of total individual polyphenols in fruit juices. Food Chemistry. 2013;138(2-3):938-949. DOI: 10.1016/j.foodchem.2012.11.061
  64. 64. Manach C. The use of HPLC with Coulometric detection in the analysis of flavonoids in complex matrices. In: Santos-Buelga C, Williamson G, editors. Methods Polyphen. Anal. Cambridge: The Royal Society of Chemistry; 2003. pp. 63-91
  65. 65. Milbury PE. Analysis of complex mixtures of flavonoids and polyphenols by high-performance liquid chromatography electrochemical detection methods. Methods in Enzymology. 2001;335:15-26. DOI: 10.1016/s0076-6879(01)35228-x
  66. 66. Wolfender JL. HPLC in natural product analysis: The detection issue. Planta Medica. 2009;75(7):719-734. DOI: 10.1055/s-0028-1088393
  67. 67. Lin LZ, Harnly JM. Quantitation of flavanols, proanthocyanidins, isoflavones, flavanones, dihydrochalcones, stilbenes, benzoic acid derivatives using ultraviolet absorbance after identification by liquid chromatography-mass spectrometry. Journal of Agricultural and Food Chemistry. 2012;60(23):5832-5840. DOI: 10.1021/jf3006905
  68. 68. Rupasinghe HPV. Application of NMR spectroscopy in plant polyphenols associated with human health. Applications of NMR spectroscopy. Netherlands: Bentham Science Publishers. 2015:3-92.
  69. 69. Wolfender JL, Marti G, Thomas A, Bertrand S. Current approaches and challenges for the metabolite profiling of complex natural extracts. Journal of Chromatography. A. 2015;1382:136-164. DOI: 10.1016/j.chroma.2014.10.091
  70. 70. de Rijke E, Out P, Niessen WM, Ariese F, Gooijer C, Brinkman UA. Analytical separation and detection methods for flavonoids. Journal of Chromatography. A. 2006;1112(1-2):31-63. DOI: 10.1016/j.chroma.2006.01.019
  71. 71. de Pascual-Teresa S. Rivas-Gonzalo JC. Application of LC-MS for the identification of polyphenols. In: Santos-Buelga C, Williamson G, editors. Methods Polyphen. Anal. Cambridge: The Royal Society of Chemistry; 2003. pp. 48-62
  72. 72. de Villiers A, Venter P, Pasch H. Recent advances and trends in the liquid-chromatography-mass spectrometry analysis of flavonoids. Journal of Chromatography. A. 2016;1430:16-78. DOI: 10.1016/j.chroma.2015.11.077
  73. 73. da Silva JB, Temponi Vdos S, Gasparetto CM, Fabri RL, Aragão DM, Pinto Nde C, et al. Vernonia condensata baker (Asteraceae): A promising source of antioxidants. Oxidative Medicine and Cellular Longevity. 2013;2013:698018. DOI: 10.1155/2013/698018
  74. 74. Golonka I, Wilk S, Musiał W. The influence of UV radiation on the degradation of pharmaceutical formulations containing quercetin. Molecules. 2020;25(22):5454. DOI: 10.3390/molecules25225454
  75. 75. Farag MA, Huhman DV, Lei Z, Sumner LW. Metabolic profiling and systematic identification of flavonoids and isoflavonoids in roots and cell suspension cultures of Medicago truncatula using HPLC-UV-ESI-MS and GC-MS. Phytochemistry. 2007;68(3):342-354
  76. 76. Simone J. Rochfort, Michael Imsic, Rod Jones, V. Craige Trenerry, And Bruce Tomkins (2006) .Characterization of Flavonol conjugates in immature leaves of Pak Choi [Brassica rapa L. Ssp. chinensis L. (Hanelt.)] by HPLC-DAD and LC-MS/MS. Journal of Agricultural and Food Chemistry, 54, pp. 4855-4860.
  77. 77. Bednarek P, Kerhoas L, Einhorn J, Frański R, Wojtaszek P, Rybus-Zajac M, et al. Profiling of flavonoid conjugates in Lupinus albus and Lupinus angustifolius responding to biotic and abiotic stimuli. Journal of Chemical Ecology. 2003;29(5):1127-1142. DOI: 10.1023/a:1023877422403
  78. 78. Gallon ME, Jaiyesimi OA, Gobbo-Neto L. LC-UV-HRMS dereplication of secondary metabolites from Brazilian Vernonieae (Asteraceae) species supported through in-house database. Biochemical Systematics and Ecology. 2018;78:5-16
  79. 79. Avula B, Wang YH, Smillie TJ, Fu X, Li XC, Mabusela W, et al. Quantitative determination of flavonoids and cycloartanol glycosides from aerial parts of Sutherlandia frutescens (L.) R. BR. By using LC-UV/ELSD methods and confirmation by using LC-MS method. Journal of Pharmaceutical and Biomedical Analysis. 2010;52(2):173-180. DOI: 10.1016/j.jpba.2010.01.010
  80. 80. Careri M, Elviri L, Mangia A, Musci M. Spectrophotometric and coulometric detection in the high-performance liquid chromatography of flavonoids and optimization of sample treatment for the determination of quercetin in orange juice. Journal of Chromatography. A. 2000;881(1-2):449-460. DOI: 10.1016/s0021-9673(00)00256-9
  81. 81. Nguyen K-NH, Nguyen N-VT, Kim KH. Determination of phenolic acids and flavonoids in leaves, calyces, and fruits of Physalis angulata L. in Viet Nam. Pharmacia. 2021;68(2):501-509. DOI: 10.3897/pharmacia.68.e66044
  82. 82. Guillarme D, Ruta J, Rudaz S, Veuthey JL. New trends in fast and high-resolution liquid chromatography: A critical comparison of existing approaches. Analytical and Bioanalytical Chemistry. 2010;397(3):1069-1082. DOI: 10.1007/s00216-009-3305-8
  83. 83. Fekete S, Veuthey JL, Guillarme D. Comparison of the most recent chromatographic approaches applied for fast and high resolution separations: Theory and practice. Journal of Chromatography. A. 2015;1408:1-14. DOI: 10.1016/j.chroma.2015.07.014
  84. 84. Stoll DR, Li X, Wang X, Carr PW, Porter SE, Rutan SC. Fast, Comprehensive two-dimensional liquid chromatography. Journal of Chromatography. A 2007;1168(1-2):3-43; discussion 2. DOI: 10.1016/j.chroma.2007.08.054
  85. 85. François I, Sandra K, Sandra P. Comprehensive liquid chromatography: Fundamental aspects and practical considerations--a review. Analytica Chimica Acta. 2009;641(1-2):14-31. DOI: 10.1016/j.aca.2009.03.041
  86. 86. Stoll DR. Recent progress in online, comprehensive two-dimensional high-performance liquid chromatography for non-proteomic applications. Analytical and Bioanalytical Chemistry. 2010;397(3):979-986. DOI: 10.1007/s00216-010-3659-y
  87. 87. Malerod H, Lundanes E, Greibrokk T. Recent advances in on-line multidimensional liquid chromatography. Analytical Methods. 2010;2:110-122
  88. 88. Akimoto N, Ara T, Nakajima D, Suda K, Ikeda C, Takahashi S, et al. Flavonoid search: A system for comprehensive flavonoid annotation by mass spectrometry. Scientific Reports. 2017;7(1):1243. DOI: 10.1038/s41598-017-01390-3
  89. 89. Tsugawa H, Kind T, Nakabayashi R, Yukihira D, Tanaka W, Cajka T, et al. Hydrogen rearrangement rules: Computational MS/MS fragmentation and structure elucidation using MS-FINDER software. Analytical Chemistry. 2016;88(16):7946-7958. DOI: 10.1021/acs.analchem.6b00770
  90. 90. Goiris K, Muylaert K, Voorspoels S, Noten B, De Paepe D, Baart E, et al. Detection of flavonoids in microalgae from different evolutionary lineages. Journal of Phycology. 2014;50(3):483-492. DOI: 10.1111/jpy.12180
  91. 91. Abad-García B, Garmón-Lobato S, Berrueta LA, Gallo B, Vicente F. A fragmentation study of dihydroquercetin using triple quadrupole mass spectrometry and its application for identification of dihydroflavonols in citrus juices. Rapid Communications in Mass Spectrometry. 2009;23(17):2785-2792. DOI: 10.1002/rcm.4182
  92. 92. Vacek J, Klejdus B, Lojková L, Kubán V. Current trends in isolation, separation, determination and identification of isoflavones: A review. Journal of Separation Science. 2008;31(11):2054-2067. DOI: 10.1002/jssc.200700569
  93. 93. Abad-García B, Garmón-Lobato S, Berrueta LA, Gallo B, Vicente F. New features on the fragmentation and differentiation of C-glycosidic flavone isomers by positive electrospray ionization and triple quadrupole mass spectrometry. Rapid Communications in Mass Spectrometry. 2008;22(12):1834-1842. DOI: 10.1002/rcm.3560
  94. 94. Hughes RJ, Croley TR, Metcalfe CD, March RE. A tandem mass spectrometric study of selected characteristic flavonoids. International Journal of Mass Spectrometry. 2001;211:371-385
  95. 95. Cuyckens F, Rozenberg R, de Hoffmann E, Claeys M. Structure characterization of flavonoid O-diglycosides by positive and negative nano-electrospray ionization ion trap mass spectrometry. Journal of Mass Spectrometry. 2001;36(11):1203-1210. DOI: 10.1002/jms.224
  96. 96. March RE, Lewars EG, Stadey CJ, Miao X-S, Zhao X, Metcalfe CD. A comparison of flavonoid glycosides by electrospray tandem mass spectrometry. International Journal of Mass Spectrometry. 2006;248:61-85
  97. 97. Kachlicki P, Einhorn J, Muth D, Kerhoas L, Stobiecki M. Evaluation of glycosylation and malonylation patterns in flavonoid glycosides during LC/MS/MS metabolite profiling. Journal of Mass Spectrometry. 2008;43(5):572-586. DOI: 10.1002/jms.1344
  98. 98. Zeng X, Zheng Y, He Y, Peng W, Su W. A rapid LC-MS/MS method for simultaneous determination of ten flavonoid metabolites of Naringin in rat urine and its application to an excretion study. Food. 2022;11:316. DOI: 10.3390/foods11030316
  99. 99. Podolak I, Olech M, Galanty A, Załuski D, Grabowska K, Sobolewska D, et al. Flavonoid and phenolic acid profile by LC-MS/MS and biological activity of crude extracts from Chenopodium hybridum aerial parts. Natural Product Research. 2016;30(15):1766-1770. DOI: 10.1080/14786419.2015.1136908
  100. 100. Molina-Calle M, Priego-Capote F, Luque de Castro MD. Development and application of a quantitative method for determination of flavonoids in orange peel: Influence of sample pretreatment on composition. Talanta. 2015;144:349-355. DOI: 10.1016/j.talanta.2015.05.054
  101. 101. Huang M, Zhang Y, Xu S, Xu W, Chu K, Xu W, et al. Identification and quantification of phenolic compounds in Vitex negundo L. var. cannabifolia (Siebold et Zucc.) hand.-Mazz. Using liquid chromatography combined with quadrupole time-of-flight and triple quadrupole mass spectrometers. Journal of Pharmaceutical and Biomedical Analysis. 2015;108:11-20. DOI: 10.1016/j.jpba.2015.01.049
  102. 102. Wang S, Yang C, Tu H, et al. Characterization and metabolic diversity of flavonoids in citrus species. Scientific Reports. 2017;7(1):10549. Published 2017. DOI: 10.1038/s41598-017-10970-2
  103. 103. Montoro P, Maldini M, Luciani L, Tuberoso CI, Congiu F, Pizza C. Radical scavenging activity and LC-MS metabolic profiling of petals, stamens, and flowers of Crocus sativus L. Journal of Food Science. 2012;77(8):C893-C900. DOI: 10.1111/j.1750-3841.2012.02803.x
  104. 104. Rudolf G, Herrmann K. Analysis of flavonoids by high-performance liquid chromatography. Journal of Chromatography A. 1980;189(2):217-224. DOI: 10.1016/s0021-9673(00)81521-6
  105. 105. Feucht W, Treutter D, Schmid P. Inhibition of growth and xylogenesis and promotion of vacuolation in prunus callus by the flavanone prunin. Plant Cell Reports. 1988;7(3):189-192. DOI: 10.1007/BF00269320
  106. 106. Gaggeri R, Rossi D, Collina S, Mannucci B, Baierl M, Juza M. Quick development of an analytical enantioselective high performance liquid chromatography separation and preparative scale-up for the flavonoid Naringenin. Journal of Chromatography. A. 2011;1218(32):5414-5422. DOI: 10.1016/j.chroma.2011.02.038

Written By

Ngoc Van Thi Nguyen

Reviewed: 18 August 2022 Published: 16 September 2022