Skip to main content

Functional identification of three regiospecific flavonoid O-methyltransferases in Rhododendron delavayi and their applications in the biotechnological production of methoxyflavonoids

Abstract

Rhododendrons produce a variety of methoxyflavonoids, including rarely found 3-methoxyflavonoids and 5-methoxyflavonoids. It was thus suggested that they have a series of regiospecific flavonoid O-methyltransferases (FOMTs). The 18 Class II O-methyltransferase (OMT) genes were retrieved from the Rhododendron delavayi genome, designating them as RdOMTs. A comprehensive biochemical characterization of RdOMTs was performed to identify functional FOMTs. The FOMT activity of recombinant RdOMTs was assayed with flavonoid substrates of different subclasses. Among the examined RdOMTs, RdOMT3, RdOMT10, and RdOMT12 showed FOMT activity for diverse flavonoids. In particular, RdOMT3 consumed only flavonols as a substrate. Structural analyses of the methylated products demonstrated that RdOMT3, RdOMT10, and RdOMT12 catalyze regiospecific methylation of flavonoids at the 3'/5'-, 3-, and 4'-hydroxyl groups, respectively. Their broad substrate spectrum and different regiospecificity suggest that these RdOMTs contribute to the formation of complex methoxyflavonoids in R. delavayi. Bioconversion of flavonoids using E. coli harboring each RdOMT demonstrated that RdOMT3, RdOMT10, and RdOMT12 are useful tools for the biotechnological production of valuable methoxyflavonoids, including the rarely found 3-methoxyflavonoids.

Introduction

Rhododendron is a large genus of flowering plants in the family Ericaceae and widely distributed in the Northern Hemisphere [1,2,3,4]. In Asian, North American, and European traditional medicine, some Rhododendron species have been used to treat various diseases, such as inflammation, pain, gastrointestinal disorders, colds, and asthma [1, 2, 5, 6]. The bioactive constituents of rhododendrons have been extensively isolated and identified. They are mainly flavonoids and diterpenoids [1, 2, 5, 6]. In addition to common flavonoids, such as kaempferol, quercetin, and myricetin, a variety of O-methylflavonoids and C-methylflavonoids were identified from rhododendron plants [1, 2, 5,6,7]. The C-methyl derivatives of naringenin and apigenin, such as farrerol (6,8-C-dimethylnaringenin), syzalterin (6,8-C-dimethylapigenin), and matteucinol (4'-O-methylfarrerol), were isolated in some rhododendron species, including R. dauricum, R. concinnum, and R. fortunei [5, 8,9,10].

Common O-methylations of flavonoids occur at their 7-, 3'-, and 4'-hydroxy groups [11, 12]. The 5- and 3-O-methylations of flavonoids rarely occur in higher plants [11,12,13]. Interestingly, 5- and/or 3-methoxyflavonoids are often found in rhododendrons. Azaleatin (5-O-methylquercetin) was first isolated from a white azalea (R. mucronatum G. Don) [14]. Since then, other 5- and 3-O-methylflavonids, such as 5-O-methylkaempferol, 5-O-methylmyricetin, 3-O-methylquercetin, and caryatin (3,5-O-dimethylquercetin), have been identified in many Rhododendron species, including R. delavayi, R. austrinum, R. yedoense, R. ellipticum, and R. seniavinii [1, 7, 13, 15,16,17,18].

In plants, the methylation of secondary metabolites is mainly accomplished by S-adenosyl-L-methionine (SAM)-dependent methyltransferases [19, 20]. They can be involved in the O-, C-, and N-methylations of plant secondary metabolites. Although there are a few exceptions, flavonoid O-methyltransferase (FOMT) catalyzes the O-methylation of flavonoids in a regiospecific manner [11, 12, 19,20,21]. The complexity of methylation patterns in the rhododendron flavonoids implies that rhododendrons have an array of FOMTs with different regiospecificities. In this regard, R. delavayi OMT (RdOMT) genes were retrieved from the Rhododendron Plant Genome Database (RPGD) [4]. Biochemical studies of RdOMTs led us to identify three regiospecific FOMTs: RdOMT3, RdOMT10, and RdOMT12. The application of the regiospecific RdOMTs in the biotechnological production of methoxyflavonoids was also evaluated.

Materials and methods

In silico analysis of class II OMTs

The OMT genes in the R. delavayi genome were retrieved from the RPGD (http://bioinfor.kib.ac.cn/RPGD/) using the InterPro ID search tool, designating them as RdOMTs. The amino acid sequences of RdOMTs and other OMTs were aligned with Clustal-W to identify the signature motifs conserved in Class II OMTs [22]. The phylogenetic relationships among RdOMTs and other OMTs were inferred by the Neighbor-joining method using MEGA X [23,24,25].

Preparation of recombinant RdOMTs

The codons of the RdOMT genes were optimized for heterologous expression in Escherichia coli (Table S1). The synthetic RdOMT genes individually cloned in the pET-28b vector were obtained from Bionics (Seoul, Korea). Each RdOMT/pET-28a construct was transformed into E. coli BL21 (DE3) cells. The resulting E. coli transformants were grown in Luria–Bertani (LB) medium supplemented with kanamycin (25 µg/mL) at 37 °C. When the cell population reached an OD600 of ~ 0.7, RdOMT expression was induced with either 0.1 mM or 0.5 mM isopropyl β-D-thiogalactopyranoside (IPTG) at different temperatures ranging from 14 to 25 °C overnight. Recombinant RdOMTs were individually purified from the induction cultures by the method described by Park et al., 2024 [26].

Biochemical analysis

The FOMT assay of recombinant RdOMTs and a high-performance liquid chromatography (HPLC) analysis were performed by the method described by Park et al., 2024 [26]. RdOMT3 reactions for kinetic analysis were carried out with eriodictyol, luteolin, quercetin, and rhamnetin. Concentrations of eriodictyol, luteolin, and rhamnetin for the RdOMT3 kinetics ranged from 0.5 µM to 50 µM. The quercetin concentrations were 1–50 µM. Kaempferol, quercetin, and isorhamnetin were used for the kinetic analysis of RdOMT10. Substrate concentrations for the RdOMT10 reactions were 0.7–10 µM. The kinetic analysis of RdOTM12 was performed with naringenin, apigenin, luteolin, and kaempferol. The concentrations of the substrates for the kinetic analysis of RdOMT12 ranged from 0.5 µM to 50 µM.

Structural analysis of the methylated products

Liquid chromatography-mass spectrometry (LC–MS) analysis of authentic 3'- and 4'-methoxyflavonoids and the reaction mixtures of RdOMT3 and RdOMT12 were carried out by the method described by Park et al., 2024 [26]. The methylated products of the RdOMT10 reactions were isolated by HPLC equipped with a preparative SunFire C18 column (10 mm × 150 mm, Waters, Milford, MA, USA) using a linear gradient of 25–60% acetonitrile in 3% acetic acid–water for 25 min at a flow rate of 3 mL/min with detection at 280 nm. The 1H, 13C, and heteronuclear multiple bond correlation (HMBC) NMR spectra of the isolated products were recorded with an Avance III HD spectrometer (Bruker, Billerica, MA, USA) operating at 600 MHz (for 1H) and 150 MHz (for 13C) in acetone-d6.

3-O-Methylkaempferol. 1H NMR δ (ppm) 6.25 (d, J = 1.9 Hz, H-6), 6.50 (d, J = 1.9 Hz, H-8), 8.03 (d, J = 8.9 Hz, H-2', 6'), 7.02 (d, J = 8.9 Hz, H-3', 5'), 3.87 (s, 3-OCH3); 13C NMR δ 156.92 (C-2), 139.25 (C-3), 179.60 (C-4), 163.31 (C-5), 99.47 (C-6), 164.98 (C-7), 94.56 (C-8), 157.94 (C-9), 105.98 (C-10), 122.75 (C-1'), 131.28 (C-2'), 116.48 (C-3'), 160.99 (C-4'), 116.48 (C-5'), 131.28 (C-6'), 60.28 (3-OCH3).

3-O-Methylquercetin. 1H NMR δ 6.25 (d, J = 1.9 Hz, H-6), 6.49 (d, J = 1.9 Hz, H-8), 7.70 (d, J = 2.1 Hz, H-2'), 6.99 (d, J = 8.5 Hz, H-5'), 7.58 (dd, J = 8.5, 2.1 Hz, H-6'), 3.86 (s, 3-OCH3); 13C NMR δ 156.82 (C-2), 139.36 (C-3), 179.60 (C-4), 163.29 (C-5), 99.44 (C-6), 164.94 (C-7), 94.49 (C-8), 157.89 (C-9), 105.97 (C-10), 123.11 (C-1'), 116.35 (C-2'), 145.94 (C-3'), 149.18 (C-4'), 116.42 (C-5'), 122.21 (C-6'), 60.24 (3-OCH3).

3-O-Methylisorhamnetin. 1H NMR δ 6.25 (d, J = 1.9 Hz, H-6), 6.51 (d, J = 1.9 Hz, H-8), 7.77 (d, J = 2.0 Hz, H-2'), 7.00 (d, J = 8.5 Hz, H-5'), 7.68 (dd, J = 8.5, 2.0 Hz, H-6'), 3.89 (s, 3-OCH3), 3.95 (s, 3'-OCH3); 13C NMR δ 156.71 (C-2), 139.35 (C-3), 179.56 (C-4), 163.26 (C-5), 99.52 (C-6), 165.20 (C-7), 94.67 (C-8), 157.92 (C-9), 105.88 (C-10), 122.92 (C-1'), 112.71 (C-2'), 148.34 (C-3'), 150.51 (C-4'), 116.15 (C-5'), 123.41 (C-6'), 60.31 (3-OCH3), 56.48 (3'-OCH3).

Regiospecific methylation of flavonoids using whole-E. coli cells

Expression of each RdOMT in E. coli cells was induced with 0.5 mM IPTG at 25 °C for 4 h. The cells were harvested and resuspended in the same volume of fresh LB medium. The bioconversion by RdOMT3 was carried out with eriodictyol, luteolin, and quercetin. Kaempferol, quercetin, and isorhamnetin were used for the RdOMT10 reactions. The substrates used in the bioconversion by RdOMT12 were naringenin, apigenin, and luteolin. Flavonoid substrates (50 µM) were individually added to the fresh cell suspension and further incubated at 25 °C. An aliquot of the culture was harvested at the selected time points and centrifuged to obtain a cell-free medium. The medium was extracted with ethyl acetate and analyzed by the method described by Park et al., 2024 [26].

Results and discussion

Class II OMT family in the R. delavayi genome

OMTs are a large group of enzymes that belong to the SAM-dependent methyltransferase superfamily. They are involved in the methylation of diverse secondary metabolites, including flavonoids, phenylpropanoids, stilbenes, and alkaloids [19, 20]. Two classes of SAM-dependent OMT have been known to catalyze the O-methylation of phenolic compounds. The cation-dependent Class I OMT participates in the methylation of hydroxycinnamoyl-CoAs during monolignol biosynthesis [20]. Class II OMTs catalyze the methylation of a variety of phenolic compounds, including caffeic acids, flavonoids, and isoflavonoids [19,20,21]. To identify biochemically active FOMTs, OMT genes were retrieved from the R. delavayi genome in the RPGD. It has been well known that both an O-methyltransferase domain (InterPro ID: IPR001077) and a plant methyltransferase dimerisation domain (InterPro ID: IPR012967) are conserved in Class II OMT that is involved in the O-methylation of flavonoids and isoflavonoids [19,20,21, 27,28,29]. A search for the InterPro ID IPR001077 in the RPGD demonstrated that there are 18 tentative Class II OMTs (RdOMT1-RdOMT18) in the R. delavayi genome (Fig. 1). Most RdOMTs also have the dimerisation domain, except for RdOMT16 and RdOMT17 (Fig. 1). The gene lengths of the functionally identified Class II OMTs are about 1100 nucleotides long, encoding proteins of approximately 40 kDa molecular mass [12, 19,20,21]. The open reading frames of RdOMT15, RdOMT16, RdOMT17, and RdOMT18 are significantly shorter than typical Class II OMTs and the other RdOMTs (Table S1). RdOMT15, RdOMT17, and RdOMT18 have abnormally short O-methyltransferase domains compared to the other RdOMTs (Fig. 1). The defects in the key domains in RdOMT15-RdOMT18 suggest that they are unlikely to encode normal Class II OMTs. Therefore, the biochemical functions of 14 RdOMTs (RdOMT1-RdOMT14) were examined in the present study.

Fig. 1
figure 1

Conserved protein domains in the Class II OMTs of R. delavayi. Gene IDs of Class II OMTs in the R. delavayi genome retrieved from the RPGD are in parentheses. The plant methyltransferase dimerisation domains and O-methyltransferase domains are indicated in red and blue boxes, respectively. The amino acid positions of the domains are included in the boxes. The conserved domain was searched in InterPro (https://www.ebi.ac.uk/interpro/). The polypeptide lengths of RdOMTs are on the right

In our previous study, a phylogenetic analysis categorized FOMTs into two groups, Group I and II [26]. Another small group of FOMTs was found in the present study and named Group III (Fig. S1). Group I includes seven RdOMTs (RdOMT1-RdOMT5, RdOMT9, and RdOMT14), caffeic acid OMTs (COMTs), and a majority of functionally identified FOMTs (Fig. S1). Some COMTs have been reported to have FOMT activity. Recombinant Arabidopsis thaliana COMT (AtCOMT) was shown to catalyze the 3'-O-methylation of flavonols [30]. Rice COMT (OsCOMT) was reported to have flavonoid 3'-OMT (F3'OMT) activity [31]. Most Group I FOMTs are flavonoid 3'- and/or 5'-OMTs, and a few of them are flavonoid 7-OMTs (Fig. S1) [26, 30,31,32,33,34,35]. RdOMT3 is closely related to COMTs and Group I FOMTs. Separated from Group I, RdOMT10 is categorized with alfalfa ChOMT and Solanum habrochaites MOMTs (ShMOMT1 and ShMOMT3), designating Group III (Fig. S1). Isoflavonoid OMTs (IFOMTs) and some FOMTs were early separated from Group I and III FOMTs. These FOMTs were designated as Group II (Fig. S1). The Group II members are flavonoid 4'- and/or 7-OMTs [36,37,38,39]. RdOMT6, RdOMT7, RdOMT8, RdOMT11, RdOMT12, and RdOMT13 were included in Group II (Fig. S1).

RdOMT3, RdOMT10, and RdOMT12 are regiospecific FOMTs

Three biochemically active RdOMTs among the class II OMTs of R. delavayi

It has been reported that only a few OMT genes encode a biochemically active OMT, although many OMTs exist in plant genomes [26, 32, 40, 41]. RdOMTs were heterologously expressed in E. coli, and their FOMT activity toward flavonoids was examined to identify functional OMTs. Of the examined RdOMTs, RdOMT2, RdOMT3, RdOMT4, RdOMT5, RdOMT6, RdOMT8, RdOMT9, RdOMT10, and RdOMT12 were expressed as a soluble form in E. coli BL21 cells at 18 or 25 °C of growth temperature by 0.5 mM IPTG (Fig. 2 and Fig. S2). The recombinant RdOMTs were successfully purified by Ni2+-affinity chromatography (Fig. 2 and Fig. S2). RdOMT1, RdOMT7, RdOMT11, RdOMT13, and RdOMT14 were expressed in only insoluble form in E. coli BL21 cells at all attempted induction conditions.

Fig. 2
figure 2

Purification of the recombinant RdOMT3 (a), RdOMT10 (b), and RdOMT12 (c). M, size marker; 1, total protein extract from the induction cultures; 2, purified recombinant RdOMT protein

FOMT activity of the purified recombinant RdOMTs was assayed with diverse flavonoids of different subclasses. The HPLC analysis of the RdOMT reaction mixtures demonstrated that RdOMT3, RdOMT10, and RdOMT12 used flavonoids and produced the respective methylated flavonoids (Figs. 3 and 4). This result exhibits that RdOMT3, RdOMT10, and RdOMT12 are biochemically active FOMTs among the Class II OMTs of R. delavayi. No products were detected in the reactions of RdOMT2, RdOMT4, RdOMT5, RdOMT6, RdOMT8, and RdOMT9 to all flavonoids examined.

Fig. 3
figure 3

FOMT activities of recombinant RdOMT3 (a), RdOMT10 (b), and RdOMT12 (c) for flavonoid substrates. Relative activity indicates the ratio of FOMT activity of RdOMTs for flavonoid substrates relative to that for the selected substrates. The FOMT activity of RdOMT3, RdOMT10, and RdOMT12 for luteolin, kaempferol, and apigenin was set at 100%, respectively. The FOMT activities of RdOMTs for flavonoids were measured at substrate concentrations of 50 µM. Duplicated experiments were performed. The results represent the mean relative activities and standard deviations

Fig. 4
figure 4

Representative HPLC chromatograms of the RdOMT reactions. (a-c) The RdOMT3 reactions with eriodictyol (a), luteolin (b), and quercetin (c). (d-f) FOMT activity of RdOMT10 for kaempferol (d), quercetin (e), and isorhamnetin (f). (g-i) The RdOMT12 reactions with naringenin (g), apigenin (h), and luteolin (i). The HPLC chromatograms of reaction mixtures and authentic methoxyflavonoids are in red and blue lines, respectively. An HPLC chromatogram of the 3’-O-methylflavonoid homoeriodictyol (HE), chrysoeriol (CE), and isorhamnetin (IR), the 3-O-methylflavonoid 3-O-methylkaempferol (3MK), 3-O-methylquercetin (3MQ), and 3-O-methylisorhamnetin (3MIR), and the 4’-O-methylflavonoid isosakuranetin (IS), acacetin (AC), and diosmetin (DM), was individually inserted. S, substrate; P, product

Structural studies of alfalfa COMT (MsCOMT) and ChOMT demonstrated that three residues (His, Glu/Asp, and Glu) participate in the catalysis of Class II OMT [28, 29]. The catalytic residues are conserved in RdOMT3 (His270, Glu298, and Glu330), RdOMT10 (His281, Glu309, and Glu340), and RdOMT12 (His263, Asp293, and Glu347) (Fig. S3). The SAM-binding motifs (SAM-A, B, and C) and COMT motifs (COMT-I, J, K, and L) were suggested to be conserved in the Class II OMT [21]. An alignment of the amino acid sequences of RdOMTs and other functionally identified FOMTs showed that RdOMT3, RdOMT10, and RdOMT12 have all the signature motifs (Fig. S3). The SAM binding motifs were found to be highly conserved in functional FOMTs, including RdOMT3, RdOMT10, and RdOMT12. Minor substitutions were observed in the SAM-B and C motifs of the Group II and III FOMTs (Fig. S3). A few insertions were found in the SAM-C motif of the Group II FOMT RdOMT12 and CrOMT6 (Fig. S3). This finding agrees with the dependence of FOMTs on SAM. The COMT motifs are well conserved in RdOMT3, AtCOMT, and CaFOMT (Fig. S3). It is consistent with the close relationship between COMTs and Group I FOMTs. In RdOMT10, ShMOMT3, RdOMT12, and CrOMT6, some substitutions were found in the COMT motifs (Fig. S3).

RdOMT3 is a flavonoid 3'/5'-OMT

The FOMT reactions have been known to be mainly regiospecific [11, 12, 32,33,34]. In the phylogenetic tree, RdOMT3 was included in Group I and closely related to flavonoid 3'/5'-OMTs (F3'/5'OMTs) (Fig. S1). Although RdOMT3 methylated diverse flavonoids examined, it did not use naringenin, apigenin, or kaempferol (Fig. 3a). Compared to other flavonoids, naringenin, apigenin, and kaempferol have no 3'-OH group. It thus implied that RdOMT3 requires the 3'-OH on the flavonoid backbones for methylation. As expected, the methylated products of eriodictyol, luteolin, and quercetin by RdOMT3 exhibited an identical retention time to the authentic 3'-methoxyflavonoids, homoeriodictyol (3'-O-methyleriodictyol), chrysoeriol (3'-O-methylluteolin), and isorhamnetin (3'-O-methylquercetin), respectively (Fig. 4a–c). The 3'-O-methylation of flavonoids by RdOMT3 was further confirmed by a comparison of the mass spectra for the reaction products and the authentic 3'-methoxyflavonoids. In the LC–MS analysis of the RdOMT3 reaction mixtures, the methylated products and the authentic 3'-methoxyflavonoids exhibited identical molecular masses and fragmentation patterns (Fig. S4). Because the 3'-OH and 5'-OH on flavonoids are chemically equivalent, most F3'OMTs can catalyze the 5'-O-methylation [27, 30]. In this regard, the FOMT activity of RdOMT3 for laricitrin (3'-O-methylmyricetin) was examined. An HPLC analysis exhibited that RdOMT3 used laricitrin and produced syringetin (3',5'-O-dimethylmyricetin) (Fig. S5). These findings led us to conclude that RdOMT3 is an F3'/5'OMT.

RdOMT10 catalyzes the 3-O-methylation of flavonoids

RdOMT10 is categorized into Group III together with ShMOMT1, ShMOMT3, and ChOMT (Fig. S1). The FOMTs in Group III exhibited diverse regiospecificities. ShMOMT1 and ShMOMT3 were reported to be an F3'/5'OMT and a flavonoid 3-OMT (F3OMT), respectively [38, 42]. ChOMT catalyzes the 2'-O-methylation of chalcone [43]. Of different subclasses of flavonoids, RdOMT10 consumed only flavonols as a substrate and yielded methylated products (Figs. 3b and 4d-f). The restricted consumption of flavonols suggests that RdOMT10 likely catalyzes the 3-O-methylation of flavonoids. Because of the requirement for the 3-OH group, F3OMT is rarely found in plants [12]. A few F3OMTs have been identified in plants. C. roseus OMT1 (CrOMT1) was reported to methylate phenylpropanoids, such as caffeate and 5-hydroxyferulate, and also catalyze the methylation of flavonols to a lesser extent [44]. The strict dependence on flavonols as a substrate suggests that CrOMT1 is likely a F3OMT. An OMT was partially purified from Serratula tinctoria leaves and characterized as a flavonol 3-OMT [45]. ShMOMT3 and SlMOMT3 from wild and cultivated tomatoes were found to catalyze the 3-O-methylation of flavonols and 7/3'/4'/5'-O-methylflavonols [42].

Because no authentic 3-methoxyflavonoids were available, the methylated products of kaempferol, quercetin, and isorhamnetin were isolated from the RdOMT10 reactions by a preparative HPLC to confirm the regiospecificity. The NMR analyses of the isolated methylflavonoids were performed to determine the position of methylation. The methoxy proton signals of the methylated products, methylkaempferol and methylquercetin, were observed at δ 3.87 and δ 3.86, respectively, in the 1H NMR spectra (see Method). The 1H NMR spectrum of the methylisorhamnetin showed two methoxy proton signals. One methoxy proton signal of 3'-OCH3 appeared at δ 3.95, and another signal was observed at δ 3.89. The methoxy signal at δ 3.95 was found to correlate with the C-3' signal at δ 148.34 in the HMBC spectrum (Fig. S6). The C-3 signals of the methylated products, methylkaempferol, methylquercetin, and methylisorhamnetin, were observed at δ 139.25, δ 139.36, and δ 139.35, respectively. As expected, the proton signals of the newly added methyl group were observed to correlate with the C-3 signals of the methylated products in the HMBC spectra (Fig. S6). This result demonstrated that the methylation products of kaempferol, quercetin, and isorhamnetin by RdOMT10 are isokaempferide (3-O-methylkaempferol), 3-O-methylquercetin, and 3-O-methylisorhamnetin (3, 3'-O-dimethylquercetin), respectively. The strict dependence on flavonol substrates and the HMBC correlations between the methoxy signals and the C-3 signals indicate that RdOMT10 catalyzes the 3-O-methylation of flavonols.

RdOMT12 is a flavonoid 4’-OMT

Known FOMTs that belonged to Group II were shown to have flavonoid 4'- and/or 7-OMT activities (Fig. S1) [36,37,38,39]. RdOMT12 is a member of Group II. The FOMT activity assay showed that RdOMT12 catalyzes the methylation of all flavonoid substrates examined (Fig. 3c). The 4'-, 5-, and 7-OH groups are common in flavonoids because they are synthesized by the condensation of p-coumaroyl-CoA and malonyl-CoAs, followed by an isomerization (Fig. 5) [46,47,48,49]. Therefore, these OH-groups were suggested to be the possible positions for the RdOMT12 methylation. The retention times of the methylated products from naringenin, apigenin, and luteolin by RdOMT12 were the same as those of the authentic 4'-methoxyflavonoids, isosakuranetin (4'-O-methylnaringenin), acacetin (4'-O-methylapigenin), and diosmetin (4'-O-methylluteolin), in the HPLC analysis, respectively (Fig. 4g-i). This result implies that RdOMT12 catalyzes the 4'-O-methylation of flavonoids. An LC–MS analysis also showed that the molecular masses and fragmentation patterns of the methylated products from naringenin, apigenin, and luteolin by RdOMT12 were identical with those of isosakuranetin, acacetin, and diosmetin, respectively (Fig. S7). It was therefore concluded that RdOMT12 is a flavonoid 4'-OMT (F4'OMT).

Fig. 5
figure 5

Possible involvement of regiospecific FOMTs in the formation of complex methoxyflavonoids in rhododendrons. Three regiospecific RdOMTs are in red. Methoxyflavonoids identified in rhododendrons are in bold. CHS, chalcone synthase; CHI, chalcone isomerase; F3H, flavanone 3-hydroxylase; FLS, flavonol synthase; F3’H, flavonoid 3'-hydroxylase; F5'H, flavonoid 5'-hydroxylase; F3'/5'H, flavonoid 3'/5'-hydroxylase; CMT, C-methyltransferase; F3OMT; flavonoid 3-O-methyltransferase; F3'/5'OMT, flavonoid 3'/5'-O-methyltransferase; F4'OMT, flavonoid 4'-O-methyltransferase; F5OMT, flavonoid 5-O-methyltransferase; F7OMT; flavonoid 7-O-methyltransferase

Substrate preference of RdOMT3, RdOMT10, and RdOMT12

It has been well established that FOMTs have a broad substrate spectrum for flavonoids [11, 12, 20]. In the present study, a variety of flavonoids and methoxyflavonoids were used in the FOMT assay of RdOMTs. Of the flavonoid substrates examined, RdOMT3 showed strong FOMT activity for luteolin and quercetin at concentrations of 50 μM (Fig. 3a). In the same condition, the activity of RdOMT3 for eriodictyol was about fivefold lower than that for luteolin. Of the mono-methoxyflavonoid substrates, RdOMT3 consumed rhamnetin and laricitrin (Fig. 3a). However, no methylated products from hesperetin (4'-O-methyleriodictyol) and diosmetin (4'-O-methylluteolin) were detected in the ROMT3 reactions, suggesting that the vicinal methyl group likely interferes with the 3'-O-methylation by RdOMT3. A kinetic analysis was carried out to further evaluate the efficiency of the RdOMT3 reaction for the selected substrates (eriodictyol, luteolin, quercetin, and rhamnetin). The Km values of RdOMT3 for eriodictyol, luteolin, and rhamnetin were 0.75 µM, 1.35 µM, and 0.87 µM, respectively (Table 1). This result suggests that RdOMT3 has a similar binding affinity to these substrates. The Km value of RdOMT3 for quercetin was quite higher than that for the other substrates. RdOMT3 showed the highest Vmax value for quercetin, followed by luteolin (Table 1). The Vmax values of RdOMT3 for eriodictyol and rhamnetin were lower than those for quercetin and luteolin. In parallel with its relative activity, RdOMT3 showed the highest enzymatic efficiency to luteolin, with a kcat/Km value of 9.53 × 103 M−1 s−1 among the examined substrates (Table 1). These findings indicated that luteolin is a preferable substrate for RdOMT3, although it can use diverse flavonoids.

Table 1 Kinetic constants of the RdOMT3, RdOMT10, and RdOMT12 reactionsa

Because of the requirement for the 3-OH group, substrates for RdOMT10 were limited to flavonols. RdOMT10 was shown to methylate kaempferol and quercetin to similar extents in the saturated conditions (Fig. 3b). It also methylated the mono-methoxyflavonol rhamnetin and isorhamnetin. In the same condition, the FOMT activity of RdOMT10 for isorhamnetin was comparable to that for quercetin. The relative activity of RdOMT10 for rhamnetin was quite weaker than that for the other flavonols (Fig. 3b). In parallel with its relative activity, RdOMT10 has comparable Vmax values for kaempferol, quercetin, and isorhamnetin (Table 1). The Km values of RdOMT10 for quercetin and isorhamnetin were 1 µM and 1.06 µM, respectively, indicating that it has a similar binding affinity for these substrates (Table 1). RdOMT10 showed a high Km value for kaempferol compared to that for quercetin and isorhamnetin (Table 1). Therefore, kcat/Km values for quercetin and isorhamnetin were about tenfold higher than those for kaempferol (Table 1). Taken together, RdOMT10 prefers quercetin and isorhamnetin as substrates, followed by kaempferol.

Although RdOMT12 catalyzed the methylation of all examined flavonoids, it showed different FOMT activities for substrates. RdOMT12 exhibited the strongest activity for apigenin under the saturated condition (Fig. 3c). In the same condition, RdOMT12 showed moderate FOMT activities for naringenin, luteolin, and kaempferol, with relative activities of 25.32, 32.25, and 27.06%, respectively, to the activity for apigenin (Fig. 3c). The activity of RdOMT12 on the other flavonoids was less than 10% of that of apigenin. In the kinetic study, RdOMT12 had the lowest Km value of 0.3 µM for apigenin (Table 1). The Km values for naringenin, luteolin, and kaempferol were 3.11 µM, 0.5 µM, and 1.04 µM, respectively. RdOMT12 exhibited comparable Vmax values for apigenin, naringenin, luteolin, and kaempferol, ranging from 1.07 pmol min−1 μg−1 to 3.21 pmol min−1 μg−1 (Table 1). The overall enzyme efficiency of RdOMT12 for apigenin was the highest among the examined flavonoids. The kinetic parameters and relative activity indicated that apigenin is the best substrate for RdOMT12.

Regiospecific FOMTs contribute to the formation of complex methoxyflavonoids in rhododendrons

Extensive phytochemical analyses of rhododendrons have shown that they produce a wide array of flavonoids, from the common flavonoid naringenin, kaempferol, and quercetin to C- and O-methylflavonoids [1, 5,6,7, 50]. The flavanone naringenin is an entry point of the flavonoid metabolism and is synthesized from a p-coumaroyl-CoA and three malonyl-CoAs by consecutive actions of chalcone synthase and chalcone isomerase (Fig. 5) [46,47,48,49]. Therefore, flavonoids generally contain the 5-, 7-, and 4'-OH groups. In addition, flavonoid backbones are often hydroxylated at the 3'-position [46, 47]. These OH-groups, except for 5-OH, are common sites of the flavonoid O-methylations [12, 13]. The 3'-methoxyflavonoid isorhamnetin and laricitrin and the 4'-methoxyflavonoid isosakuranetin and hesperetin were identified in some Rhododendron species [1, 8, 51,52,53,54,55,56]. In addition to these common O-methylations, 3- and 5-O-methylations have often been found in the rhododendron flavonoids. 3-O-Methylquercetin was isolated from R. delavayi and R. luteum [17, 57]. Azaleatin, 5-O-methylkaempferol, and 5-O-methylmyricetin are widely found in rhododendrons [5, 7, 13]. Besides mono-methoxyflavonoids, di-methoxyflavonoids (caryatin, 5,4'-O-dimethylquercetin, and 7,4'-O-dimethylfarrerol) and 5,7,3'-O-trimethylquercetin have been identified in Rhododendron species, including R. delavayi, R. austrinum, R. ellipticum, R. seniavinii, and R. hainanense [7, 8, 13, 15, 17, 18].

The complexity of methoxyflavonoids suggests that rhododendrons need to have a series of FOMTs because of their regiospecific nature (Fig. 5). In the present study, we comprehensively investigated the Class II OMTs in the R. delavayi genome and identified three biochemically functional RdOMTs. The functional RdOMTs include a rarely found F3OMT, RdOMT10, as well as the F3'/5'OMT RdOMT3 and the F4'OMT RdOMT12. Phytochemical studies have been reported to identify diverse methoxyflavonoids, such as 3-O-methylquercetin, 3,5-O-dimethylquercetin, and isorhamnetin, from R. delavayi [17, 53]. Its 3-O-methylation activity and broad substrate spectrum suggest that RdOMT10 likely participates in the formation of 3-methoxyflavonoids in R. delavayi. RdOMT3 and RdOMT12 were also likely involved in the 3'-, 4'-, and/or 5'-O-methylation of flavonoids, leading to the synthesis of common methoxyflavonoids, such as isorhamnetin, in R. delavayi. The possible roles of regiospecific FOMTs in the formation of complex methoxyflavonoids in rhododendrons are summarized in Fig. 5.

Applications of regiospecific RdOMTs in the biotechnological production of methoxyflavonoids

Bioconversion using whole-cell microorganisms bearing biosynthetic enzymes has been considered a promising method for the production of valuable natural products [11, 58]. E. coli cells harboring Plagiochasma appendiculatum F4'OMT (PaF4'OMT) were reported to be used in the regiospecific conversion of apigenin to acacetin [59]. The flavonoid 7-O-methyltransferase PaOMT-7 was applied in the production of the 7-methoxyflavonoids from luteolin, kaempferol, apigenin, and quercetin [60].

Although a few F3OMTs were identified, their applications in the biotechnological production of 3-methoxyflavonoids have not been reported yet. In this regard, the bioconversion of kaempferol, quercetin, and isorhamnetin using E. coli cells bearing RdOMT10 was attempted to produce the respective 3-methoxyflavonoids (Fig. 6d–f). The HPLC analysis of the reaction mixtures showed that the examined substrates were almost consumed within 2 h. RdOMT10-transformed E. coli cells converted isorhamnetin to an equivalent amount of 3-O-methylisorhamnetin after 4 h of bioconversion (Fig. 6f). The production of isokaempferide and 3-O-methylquercetin from kaempferol and quercetin reached the highest levels at 3 h and 2 h, with maximum yields of 84.5% and 64.3%, respectively (Fig. 6d, e). This result implies that the E. coli cells bearing RdOMT10 are a useful whole-cell system for the production of 3-methoxyflavonoids.

Fig. 6
figure 6

Regiospecific production of methoxyflavonoids using whole E. coli cells expressing RdOMT. (ac) The RdOMT3-transformed E. coli cells were used in the production of the 3'-methoxyflavonoids from eriodictyol (a), luteolin (b), and quercetin (c). (df) The 3-methoxyflavonoids were produced from kaempferol (d), quercetin (e), and isorhamnetin (f) by E. coli cells harboring RdOMT10. (gi) The production of the 4'-methoxyflavonoids from naringenin (g), apigenin (h), and luteolin (i) was carried out with the RdOMT12-transformed E. coli cells. Three independent experiments were performed, and the results were presented

Production of 3'-methoxyflavonoids and 4'-methoxyflavonoids was tried with RdOMT3 and RdOMT12, respectively. Eriodictyol, luteolin, and quercetin were individually added to the culture of E. coli cells harboring RdOMT3 to evaluate the 3'-methoxyflavonoid production. The RdOMT3-transformed cells rapidly consumed luteolin and quercetin within 2 h (Fig. 6b, c). Eriodictyol was gradually used by the E. coli cells through the biotransformation period (Fig. 6a). This result agrees well with the relative activity of RdOMT3 on the examined substrates. The amount of chrysoeriol and isorhamnetin reached maximum levels at 2 h of bioconversion and then decreased (Fig. 6b, c). The maximum yields of chrysoeriol and isorhamnetin from luteolin and quercetin were 77.9% and 68.7%, respectively. The yield of homoeriodictyol from eriodictyol was 87.1% at 6 h of bioconversion.

E. coli cells harboring RdOMT12 were used in the production of the 4'-methoxyflavonoids from naringenin, apigenin, and luteolin. Naringenin was successfully converted to isosakuranetin by RdOMT12-transformed E. coli cells without significant loss (Fig. 6g). RdOMT12-transformed E. coli cells showed the highest initial bioconversion rate to apigenin among the examined substrates (Fig. 6g–i). The level of acacetin from apigenin reached its maximum at 1 h of bioconversion and rapidly decreased (Fig. 6h). This result suggests that the bioconversion time needs to be carefully controlled for a better production yield of acacetin. The production of diosmetin from luteolin by RdOMT12-transformed E. coli cells reached its highest level at 3 h of biotransformation, with a yield of 78.8% (Fig. 6i).

The health-beneficial properties of methoxyflavonoids have been consistently reported [61,62,63,64,65,66,67,68,69]. 3-O-Methylquercetin was reported to have chemoprevention effects against skin and esophageal cancer by targeting the extracellular signal-related kinase and mitogen-activated protein kinase pathways [65, 66]. Several studies have demonstrated the anti-cancer and anti-inflammatory activities of isokaempferide [64, 69]. Methoxyflavonoids, such as 3-O-methylisorhamnetin, sakuranetin, ponciretin, and rhamnazin, have been reported to have antimicrobial activity against bacteria and/or fungi [61, 70,71,72]. 3'-Methoxyflavonoids and 4'-methoxyflavonoids are widely found in plants. They have been suggested as potent therapeutic agents to treat cancer, inflammation, infections, and other ailments [67, 68, 73,74,75,76]. In the present study, three regiospecific RdOMTs, RdOMT3, RdOMT10, and RdOMT12, were applied in the biotechnological production of methoxyflavonoids. E. coli cells harboring RdOMTs carried out the regiospecific methylation of isorhamnetin and naringenin and produced 3-O-methylisorhamnetin and isosakuranetin as much as the added substrates, respectively. The other examined flavonoids were also successfully bioconverted by RdOMT-transformed E. coli cells to the respective methoxyflavonoids, with a higher than 64% yield. These findings demonstrate that the E. coli cells harboring RdOMTs are promising tools for the regiospecific production of valuable methoxyflavonoids.

Availability of data and materials

All data generated or analyzed during this study are included in this published article [and its supplementary information files].

Abbreviations

OMT:

O-Methyltransferase

FOMT:

Flavonoid O-methyltransferase

SAM:

S-adenosyl L-methionine

RdOMT:

R. delavayi O-Methyltransferase

RPGD:

Rhododendron Plant Genome Database

LB:

Luria–Bertani

IPTG:

Isopropyl β-D-thiogalactopyranoside

HPLC:

High-performance liquid chromatography

LC–MS:

Liquid chromatography-mass spectrometry

HMBC:

Heteronuclear multiple bond correlation

COMT:

Caffeic acid O-methyltransferase

F3'OMT:

Flavonoid 3'-O-methyltransferase

F3’/5’OMT:

Flavonoid 3'/5'-O-methyltransferase

F3OMT:

Flavonoid 3-O-methyltransferase

F4’OMT:

Flavonoid 4'-O-methyltransferase

References

  1. Qiang Y, Zhou B, Gao K (2011) Chemical constituents of plants from the genus Rhododendron. Chem Biodivers 8:792–815

    Article  CAS  PubMed  Google Scholar 

  2. Popescu R, Kopp B (2013) The genus Rhododendron: an ethnopharmacological and toxicological review. J Ethnopharmacol 147:42–62

    Article  CAS  PubMed  Google Scholar 

  3. Zhang L, Xu P, Cai Y, Ma L, Li S, Li S, Xie W, Song J, Peng L, Yan H, Zou L, Ma Y, Zhang C, Gao Q, Wang J (2017) The draft genome assembly of Rhododendron delavayi Franch. var. delavayi. Gigascience 6:1–11

    Article  PubMed  PubMed Central  Google Scholar 

  4. Liu N, Zhang L, Zhou Y, Tu M, Wu Z, Gui D, Ma Y, Wang J, Zhang, (2021) The Rhododendron plant genome database (RPGD): a comprehensive online omics database for Rhododendron. BMC Genomics 22:376

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Cao Y, Chu Q, Ye J (2004) Chromatographic and electrophoretic methods for pharmaceutically active compounds in Rhododendron dauricum. J Chromatogr B 812:231–240

    Article  CAS  Google Scholar 

  6. Liang Q, Hu JX, Zhang XM, Xu WH (2023) Traditional uses, phytochemistry, pharmacology, toxicology, and quality control of Rhododendron dauricum L. leaves: a comprehensive review. J Ethnopharmacol. https://doi.org/10.1016/j.jep.2022.116085

    Article  PubMed  Google Scholar 

  7. King B (1977) Flavonoid analysis of hybridization in Rhododendron section Pentanthera (Ericaceae). Syst Bot 2:14–27

    Article  CAS  Google Scholar 

  8. Zhao J, Ding HX, Zhao DG, Wang CM, Gao K (2012) Isolation, modification and cytotoxic evaluation of flavonoids from Rhododendron hainanense. J Pharm Pharmacol 64:1785–1792

    Article  CAS  PubMed  Google Scholar 

  9. Zhao L, Wu D, Yu X, Zhang Y (2010) Dihydroflavones in Rhododendron concinnum. Zhongguo Zhong Yao Za Zhi 35:722–724

    CAS  PubMed  Google Scholar 

  10. Lai Y, Zeng H, He M, Qian H, Wu Z, Luo Z, Xue Y, Yao G, Zhang Y (2016) 6,8-Di-C-methyl-flavonols with neuroprotective activities from Rhododendron fortunei. Fitoterapia 112:237–243

    Article  PubMed  Google Scholar 

  11. Kim BG, Sung SH, Chong Y, Lim Y, Ahn JH (2010) Plant flavonoid O-methyltransferases: substrate specificity and application. J Plant Biol 53:321–329

    Article  CAS  Google Scholar 

  12. Liu Y, Fernie AR, Tohge T (2022) Diversification of chemical structures of methoxylated flavonoids and genes encoding flavonoid-O-methyltransferases. Plants 11:564

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Harborne JB (1969) Occurrence of flavonol 5-methyl ethers in higher plants and their systematic significance. Phytochemistry 8:419–423

    Article  CAS  Google Scholar 

  14. Wada E (1956) On a flavonol glycoside isolated from flowers of a white azalea (R. mucronatum G. Don). J Am Chem Soc 78:4725–4726

    Article  CAS  Google Scholar 

  15. Ho LK, Lin WN (1995) Quercetin 5,4’-dimethyl ether from Rhododendron ellipticum. Phytochemistry 39:463–464

    Article  CAS  Google Scholar 

  16. Jung SJ, Kim DH, Hong YH, Lee JH, Song HN, Rho YD, Baek NI (2007) Flavonoids from the flower of Rhododendron yedoense var. poukhanense and their antioxidant activities. Arch Pharm Res 30:146–150

    Article  CAS  PubMed  Google Scholar 

  17. Song HJ, Pan YY, Wang WG, Fu LZ, Li HZ, Li HM, Li RT (2009) Studies on the chemical constituents from Rhododendron delavayi. Zhong Yao Cai 32:1840–1843

    CAS  PubMed  Google Scholar 

  18. Wang QQ, Wu C, Zhang Y, Liu BL, Zhou GZ (2015) A new flavonoid glucoside from Rhododendron seniavinii. J Asian Nat Prod Res 17:778–782

    Article  CAS  PubMed  Google Scholar 

  19. Ibrahim RK, Bruneau A, Bantignies B (1998) Plant O-methyltransferases: molecular analysis, common signature and classification. Plant Mol Biol 36:1–10

    Article  CAS  PubMed  Google Scholar 

  20. Lam KC, Ibrahim RK, Behdad B, Dayanandan S (2007) Structure, function, and evolution of plant O-methyltransferases. Genome 50:1001–1013

    Article  CAS  PubMed  Google Scholar 

  21. Joshi C, Chiang VL (1998) Conserved sequence motifs in plant S-adenosyl-L-methionine-dependent methyltransferases. Plant Mol Biol 37:663–674

    Article  CAS  PubMed  Google Scholar 

  22. Thompson JD, Higgins DG, Gibson TJ (1994) CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res 22:4673–4680

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Saitou N, Nei M (1987) The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evol 4:406–425

    CAS  PubMed  Google Scholar 

  24. Jones DT, Taylor WR, Thornton JM (1992) The rapid generation of mutation data matrices from protein sequences. Comput Appl Biosci 8:275–282

    CAS  PubMed  Google Scholar 

  25. Kumar S, Stecher G, Li M, Knyaz C, Tamura K (2018) MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol 35:1547–1549

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Park HL, Bhoo SH, Lee SW, Cho MH (2024) Biochemical characterization of a regiospecific flavonoid 3’-O-methyltransferase from orange. Appl Biol Chem 67:4

    Article  CAS  Google Scholar 

  27. Keller NP, Dischinger HC Jr, Bhatnagar D, Cleveland TE, Ullah AH (1993) Purification of a 40-kilodalton methyltransferase active in the aflatoxin biosynthetic pathway. Appl Environ Microbiol 59:479–484

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Zubieta C, He XZ, Dixon RA, Noel JP (2001) Structures of two natural product methyltransferases reveal the basis for substrate specificity in plant O-methyltransferases. Nat Struct Biol 3:271–279

    Article  Google Scholar 

  29. Zubieta C, Kota P, Ferrer JL, Dixon RA, Noel JP (2022) Structural basis for the modulation of lignin monomer methylation by caffeic acid/5-hydroxyferulic acid 3/5-O-methyltransferase. Plant Cell 14:1265–1277

    Article  Google Scholar 

  30. Muzac I, Wang J, Anzellotti D, Zhang H, Ibrahim RK (2000) Functional expression of an Arabidopsis cDNA clone encoding a flavonol 3’-O-methyltransferase and characterization of the gene product. Arch Biochem Biophys 375:385–388

    Article  CAS  PubMed  Google Scholar 

  31. Kim BG, Lee Y, Hur HG, Lim Y, Ahn JH (2006) Flavonoid 3’-O-methyltransferase from rice: cDNA cloning, characterization and functional expression. Phytochemistry 67:387–394

    Article  CAS  PubMed  Google Scholar 

  32. Cho MH, Park HL, Park JH, Lee SW, Bhoo SH, Hahn TR (2012) Characterization of a regiospecific flavonoid 3’/5’-O-methyltransferase from tomato and its application in flavonoid biotransformation. J Korean Soc Appl Biol Chem 55:749–755

    Article  CAS  Google Scholar 

  33. Park HL, Lee JC, Lee K, Lee JM, Nam HJ, Bhoo SH, Lee TH, Lee SW, Cho MH (2020) Biochemical characterization of a flavonoid O-methyltransferase from perilla leaves and its application in 7-methoxyflavonoid production. Molecules 25:4455

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Shimizu T, Lin F, Hasegawa M, Okada K, Nojiri H, Yamane H (2012) Purification and identification of naringenin 7-O-methyltransferase, a key enzyme in biosynthesis of flavonoid phytoalexin sakuranetin in rice. J Biol Chem 287:19315–19325

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Gauthier A, Gulick PJ, Ibrahim RK (1998) Characterization of two cDNA clones which encode O-methyltransferases for the methylation of both flavonoid and phenylpropanoid compounds. Arch Biochem Biophys 351:243–249

    Article  CAS  PubMed  Google Scholar 

  36. Schröder G, Wehinger E, Lukačin R, Wellmann F, Seefelder W, Schwab W, Schröder J (2004) Flavonoid methylation: a novel 4’-O-methyltransferase from Catharanthus roseus, and evidence that partially methylated flavanones are substrates of four different flavonoid dioxygenases. Phytochemistry 65:1085–1094

    Article  PubMed  Google Scholar 

  37. Willits MG, Giovanni M, Prata RTN, Kramer CM, De Luca V, Steffens JC, Graser G (2004) Bio-fermentation of modified flavonoids: an example of in vivo diversification of secondary metabolites. Phytochemistry 65:31–41

    Article  CAS  PubMed  Google Scholar 

  38. Schmidt A, Li C, Shi F, Daniel Jones A, Pichersky E (2011) Polymethylated myricetin in trichomes of the wild tomato species Solanum habrochaites and characterization of trichome-specific 3’/5’- and 7/4’-myricetin O-methyltransferases. Plant Physiol 155:1999–2009

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Berim A, Hyatt DC, Gang DR (2012) A set of regioselective O-methyltransferases gives rise to the complex pattern of methoxylated flavones in sweet basil. Plant Physiol 160:1052–1069

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Itoh N, Iwata C, Toda H (2016) Molecular cloning and characterization of a flavonoid-O-methyltransferase with broad substrate specificity and regioselectivity from Citrus depressa. BMC Plant Biol 16:180

    Article  PubMed  PubMed Central  Google Scholar 

  41. Liu H, Xu RX, Gao S, Cheng AX (2017) The functional characterization of a site-specific apigenin 4’-O-methyltransferase synthesized by the liverwort species Plagiochasma appendiculatum. Molecules 22:769

    Google Scholar 

  42. Schmidt A, Li C, Shi F, Daniel Jones A, Pichersky E (2012) Characterization of a flavonol 3-O-methyltransferase in the trichomes of the wild tomato species Solanum habrochaites. Planta 236:839–849

    Article  CAS  PubMed  Google Scholar 

  43. Maxwell CA, Harrison MJ, Dixon RA (1993) Molecular characterization and expression of alfalfa isoliquiritigenin 2’-O-methyltransferase, an enzyme specifically involved in the biosynthesis of an inducer of Rhizobium meliloti nodulation genes. Plant J 4:971–981

    Article  CAS  PubMed  Google Scholar 

  44. Schröder G, Wehinger E, Schröder J (2002) Predicting the substrates of cloned plant O-methyltransferases. Phytochemistry 59:1–8

    Article  PubMed  Google Scholar 

  45. Huang TS, Anzellotti D, Dedaldechamp F, Ibrahim RK (2004) Partial purification, kinetic analysis, and amino acid sequence information of a flavonol 3-O-methyltransferase from Serratula tinctoria. Plant Physiol 134:1366–1376

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Winkel-Shirley B (2002) Biosynthesis of flavonoids and effects of stress. Curr Opin Plant Biol 5:18–223

    Article  Google Scholar 

  47. Buer CS, Imin N, Djordjevic MA (2010) Flavonoids: new roles for old molecules. J Integr Plant Biol 52:98–111

    Article  CAS  PubMed  Google Scholar 

  48. Park HL, Yoo Y, Bhoo SH, Lee TH, Lee SW, Cho MH (2020) Two chalcone synthase isozymes participate redundantly in UV-induced sakuranetin synthesis in rice. Int J Mol Sci 21:3777

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Park SI, Park HL, Bhoo SH, Lee SW, Cho MH (2021) Biochemical and molecular characterization of the rice chalcone isomerase family. Plant 10:2064

    Article  CAS  Google Scholar 

  50. Liu JY, Guo PJ, Wang XL, Chen HM, Chen LJ, Sang YL, Hao YJ, Lu J (2022) Study on phytochemical and pharmacological activities of four Rhododendron plants endemic to Northeast China. J Agric Food Res 7:100255

    CAS  Google Scholar 

  51. Dai SJ, Chen RY, Yu DQ (2004) Studies on the flavonoid compounds of Rhododendron anthopogonoides. Zhongguo Zhong Yao Za Zhi 29:44–47

    CAS  PubMed  Google Scholar 

  52. Zhao, (2008) Separation and quantification of flavonoid compounds in Rhododendron anthopogonoides maxim by high-performance liquid chromatography. Acta Chromatogr 20:135–146

    Article  CAS  Google Scholar 

  53. Xu JJ, Wang YH, Wang HS, Wang H, Huang QQ, Long CL (2012) Chemical constituents from stems of Rhododendron delavayi Franch. Nat Pro Res Dev 23:757–760

    Google Scholar 

  54. Li Y, Zhao J, Gao K (2016) Activity of flavanones isolated from Rhododendron hainanense against plant pathogenic fungi. Nat Prod Commun 11:611–612

    PubMed  Google Scholar 

  55. Grimbs A, Shrestha A, Rezk ASD, Grimbs S, Said IH, Schepker H, Hütt MT, Albach DC, Brix K, Kuhnert N, Ullrich MS (2017) Bioactivity in Rhododendron: a systemic analysis of antimicrobial and cytotoxic activities and their phylogenetic and phytochemical origins. Front Plant Sci 8:551

    Article  PubMed  PubMed Central  Google Scholar 

  56. Alkan S, Akgün M, Ertürk Ö, Ayvaz MҪ, Başkan C (2020) Properties of honey and pollen samples obtained from different Rhododendron species collected from black sea region of Turkey. J Apic Sci 64:321–334

    CAS  Google Scholar 

  57. Lyko L, Olech M, Noeak R (2022) LC-ESI-MS/MS characterization of concentrated polyphenolic fractions from Rhododendron luteum and their anti-inflammatory and antioxidant activities. Molecules 27:827

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Fowler ZL, Koffas MAG (2009) Biosynthesis and biotechnological production of flavanones: current state and perspectives. Appl Microbiol Biotechnol 83:799–808

    Article  CAS  PubMed  Google Scholar 

  59. Liu H, Xu RX, Gao S, Cheng AX (2017) The functional characterization of a site-specific apigenin 4’-O-methyltransferase synthesized by the liverwort species Plagiochasma appendiculatum. Molecules 22:759

    Article  PubMed  PubMed Central  Google Scholar 

  60. Kim BG, Kim H, Hur HG, Lim Y, Ahn JH (2006) Regioselectivity of 7-O-methyltransferase of poplar to flavones. J Biotechnol 126:241–247

    Article  CAS  PubMed  Google Scholar 

  61. Wang J, Lou J, Luo C, Zhou L, Wang M, Wang L (2012) Phenolic compounds from Halimodendron halodendron (Pall.) voss and their antimicrobial and antioxidant activities. Int J Mol Sci 13:11349–11364

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Wei BL, Lu CM, Tsao LT, Wang JP, Lin CN (2001) In vitro anti-inflammatory effects of quercetin 3-O-methyl ether and other constituents from Rhamnus species. Planta Med 67:745–747

    Article  CAS  PubMed  Google Scholar 

  63. Leal LKAM, Costa MF, Pitombeira M, Barroso VM, Silveira ER, Canuto KM, Viana GSB (2006) Mechanisms underlying the relaxation induced by isokaempferide from Amburana cearensis in the guinea-pig isolated trachea. Life Sci 79:98–104

    Article  CAS  PubMed  Google Scholar 

  64. Leal LKAM, Canuto KM, Costa KCS, Nobre-Júnior HV, Vasconcelos SM, Silveira ER, Ferreira MVP, Fontenele JB, Andrade GM, Viana GSB (2009) Effects of amburoside A and isokaempferide, polyphenols from Amburana cearensis, on rodent inflammatory processes and myeloperoxidase activity in human neutrophils. Basic Clin Pharmacol Toxicol 104:198–205

    Article  CAS  PubMed  Google Scholar 

  65. Li J, Mottamal M, Li H, Liu K, Zhu F, Cho YY, Sosa CP, Zhou K, Bowden GT, Bode AM, Dong Z (2012) Quercetin-3-methyl ether suppresses proliferation of mouse epidermal JB6 P+ cells by targeting ERKs. Carcinogenesis 33:459–465

    Article  PubMed  Google Scholar 

  66. Zhao S, Jiang Y, Zhao J, Li H, Yin X, Wang Y, Xie Y, Chen X, Lu J, Dong Z, Liu K (2018) Quercetin-3-methyl ether inhibits esophageal carcinogenesis by targeting the AKT/mTOR/p70S6K and MAPK pathways. Mol Carcinog 57:1540–1552

    Article  CAS  PubMed  Google Scholar 

  67. Gong G, Guan YY, Zhang ZL, Rahman K, Wang SJ, Zhou S, Luan X, Zhang H (2020) Isorhamnetin: a review of pharmacological effects. Biomed Pharmacother 128:110301

    Article  CAS  PubMed  Google Scholar 

  68. Aboulaghras S, Sahib N, Bakrim S, Benali T, Charfi S, Guaouguaou FE, El Omari N, Gallo M, Montesano D, Zengin G, Taghzouti K, Bouyahya A (2022) Health benefits and pharmacological aspects of chrysoeriol. Pharmaceuticals 15:973

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Yin Q, Zhang H, Huang T, Liu B, Negm S, El-kott AF (2024) Anti-collagenase, anti-elastase, anti-urease, and anti-cancer potentials of isokaempferide as natural compound: In vitro and in silico study. J Oleo Sci 73:18–199

    Article  Google Scholar 

  70. Zhang L, Kong Y, Wu D, Zhang H, Wu J, Chen J, Ding J, Hu L, Jiang H, Shen X (2008) Three flavonoids targeting the β-hydroxyacyl-acyl carrier protein dehydratase from Helicobacter pylori: crystal structure characterization with enzymatic inhibition assay. Protein Sci 17:1971–1978

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Kim DH, Bae EA, Han MJ (1999) Anti-Helicobacter pylori activity of the metabolites of poncirin from Poncirus trifoliata by human intestinal bacteria. Biol Pharm Bull 22:422–424

    Article  CAS  PubMed  Google Scholar 

  72. Martini ND, Katerere DRP, Eloff JN (2004) Biological activity of five antibacterial flavonoids from Combretum erythrophyllum (Combretaceae). J Ethnopharmacol 93:207–212

    Article  CAS  PubMed  Google Scholar 

  73. Nath LR, Gorantla JN, Joseph SM, Antony J, Thankachan S, Menon DB, Sankar S, Lankalapalli RS, Anto RJ (2015) Kaempferide, the most active among the four flavonoids isolated and characterized from Chromolaena odorata, induces apoptosis in cervical cancer cells while being pharmacologically safe. RSC Adv 5:100912–100922

    Article  CAS  Google Scholar 

  74. Koosha S, Mohamed Z, Sinniah A, Alshawsh MA (2019) Investigation into the molecular mechanisms underlying the anti-proliferative and anti-tumorigenesis activities of diosmetin against HCT-116 human colorectal cancer. Sci Rep 26:5148

    Article  Google Scholar 

  75. Hu S, Huang Y, Chen Y, Zhou R, Yang X, Zou y, Gao D, Huang H, Yu D, (2020) Diosmetin reduces bone loss and osteoclastogenesis by regulating the expression of TRPV1 in osteoporosis rats. Ann Transl Med 8:1312

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Singh S, Gupta P, Meena A, Luqman S (2020) Acacetin, a flavone with diverse therapeutic potential in cancer, inflammation, infections and other metabolic disorders. Food Chem Toxicol 145:111708

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by the National Research Foundation (NRF) of Korea funded by the Korean Ministry of Education (NRF-2022R1I1A1A01068808) and Ministry of Science and ICT (NRF-2023R1A2C1005706).

Author information

Authors and Affiliations

Authors

Contributions

KL and MHC performed the experiments and wrote the manuscript. SHB and SWL analyzed the data and reviewed the manuscript. SWL and MHC conceived the study. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Sang-Won Lee or Man-Ho Cho.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

13765_2024_918_MOESM1_ESM.pptx

Supplementary Material 1. Figure S1. Evolutionary analysis of RdOMTs and other functional OMTs. Figure S2. Purification of the recombinant RdOMT2 (a), RdOMT4 (b), RdOMT5 (c), RdOMT6 (d), RdOMT8 (e), and RdOMT9 (f). Figure S3. Multiple alignments of three active RdOMTs and other functional FOMTs. Figure S4. Identification of the methylated products of the RdOMT3 reactions with flavonoid substrates by LC-MS analysis. Figure S5. HPLC analysis of the RdOMT3 reaction with laricitrin. Figure S6. Some HMBC correlations observed in 3-O-methylkaempferol (a), 3-O-methylquercetin (b), and 3-O-methylisorhamnetin (c) produced from kaempferol, quercetin, and isorhamnetin by RdOMT10, respectively. Figure S7. Identification of the methylated products of the RdOMT12 reactions with flavonoid substrates by LC-MS analysis.

Supplementary Material 2. Table S1. Nucleotide sequences of RdOMTs.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lee, K., Bhoo, S.H., Lee, SW. et al. Functional identification of three regiospecific flavonoid O-methyltransferases in Rhododendron delavayi and their applications in the biotechnological production of methoxyflavonoids. Appl Biol Chem 67, 64 (2024). https://doi.org/10.1186/s13765-024-00918-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13765-024-00918-2

Keywords