Next Article in Journal
TRPV4 Role in Neuropathic Pain Mechanisms in Rodents
Previous Article in Journal
Effects of In Vitro Digestion on the Antioxidant Activity of Three Phenolic Extracts from Olive Mill Wastewaters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Chemistry and the Anti-Inflammatory Activity of Polymethoxyflavonoids from Citrus Genus

by
Gianfranco Fontana
1,
Maurizio Bruno
1,*,
Francesco Sottile
2,* and
Natale Badalamenti
1
1
Dipartimento di Scienze e Tecnologie Biologiche, Chimiche e Farmaceutiche (STEBICEF), Università Degli Studi di Palermo, Viale delle Scienze Ed. 17, 90128 Palermo, Italy
2
Dipartimento di Architettura, Università Degli Studi di Palermo, Centro di Conservazione della Biodiversità di Interesse Agrario, Viale delle Scienze Ed. 14, 90128 Palermo, Italy
*
Authors to whom correspondence should be addressed.
Antioxidants 2023, 12(1), 23; https://doi.org/10.3390/antiox12010023
Submission received: 5 December 2022 / Revised: 15 December 2022 / Accepted: 21 December 2022 / Published: 22 December 2022

Abstract

:
Polymethoxyflavonoids (PMFs) are a large group of compounds belonging to the more general class of flavonoids that possess a flavan carbon framework decorated with a variable number of methoxy groups. Hydroxylated polymethoxyflavonoids (HPMFs), instead, are characterized by the presence of both hydroxyl and methoxy groups in their structural unities. Some of these compounds are the aglycone part in a glycoside structure in which the glycosidic linkage can involve the −OH at various positions. These compounds are particular to Citrus genus plants, especially in fruits, and they are present mainly in the peel. A considerable number of PMFs and HPMFs have shown promising biological activities and they are considered to be important nutraceuticals, responsible for some of the known beneficial effects on health associated with a regular consumption of Citrus fruits. Among their several actions on human health, it is notable that the relevant contribution in controlling the intracellular redox imbalance is associated with the inflammation processes. In this work, we aim to describe the status concerning the chemical identification and the anti-inflammatory activity of both PMFs and HPMFs. In particular, all of the chemical entities unambiguously identified by isolation and complete NMR analysis, and for which a biochemical evaluation on the pure compound was performed, are included in this paper.

Graphical Abstract

1. Introduction

The large group referred to as the genus Citrus has been the subject of multiple studies to determine its origin and spread [1]. There is a general agreement that the primary area of origin is identified in the downstream area of the Himalayan mountainous complex, in an area between northeastern India and southwestern China. In the remote area of Meghalaya, an Indian area bordering China, there are still some neglected Citrus able to demonstrate the presence of citrus in the domestication phase of the species [2]. Much has been discussed about the phylogeny of citrus, and this has also resulted in a complex botanical classification that has been modified several times over time.
The contribution of molecular techniques has allowed for significant advances on the definition of the origin and distribution of Citrus by highlighting probable processes of hybridization and mutations that have fostered the current diversification of species and varieties [3]. Today, there is molecular-based scientific evidence that allows us to consider citron (Citrus medica), mandarin (Citrus reticulata), and pomelo (Citrus maxima) as the progenitor species from which all other cultivated citrus species then originated in different ways, forms and at different times [4].
The great plasticity of adaptation of the species has allowed its presence in many countries of the world with a great spread in cultivation, mainly as oranges, mandarins, lemons, and grapefruits [5]. That is because of their well-known sensory characteristics and widespread therapeutic and nutritional properties, owing to the presence of nutrients and biologically active compounds providing sugars, volatiles, organic acids (citric acid), dietary metabolites, amino acids, fibers, vitamin B6, vitamin C, and macro- and micronutrients, determining that citrus represent the most produced, consumed, and processed fruits globally [6].
The latest official data [7] report over 140 million tons with about 50% grown in the Asian continent and 14% in Brazil. Spain, Egypt, and Turkey are the main producing countries in the Mediterranean whereas on the African continent, South Africa contributes to about 2% of world production. The sweet orange (C. sinensis) is the world’s most produced citrus fruit and accounts for about half of global production (50%), followed by the tangerine (C. reticulata), lemon (C. limon), and grapefruit (C. paradisi). Approximately 75% of the production is usually consumed as fresh fruit, whereas the remaining 25% is processed to produce juices, jams, and other by-products from whose processing, moreover, several secondary products are obtained [8,9].
Furthermore, and not to be overlooked, is the contribution of the production of citrus plants for ornamental purposes, which has now assumed a significant role in the sector with substantial volumes moving from the Mediterranean Basin to within Europe and the Arabian Peninsula [10].
Phytochemically, the fruits of the genus Citrus are characterized by the massive presence of sugars, especially polysaccharides, vitamins, and mineral salts, and by smaller quantities of protein and lipid compounds. Chemically, the colour of the fruit depends on both carotenoids and different flavonoids, whereas the limonoids are responsible for the bitter taste of the juice. Terpenoids, on the other hand, are the class of chemical constituents responsible for the aroma [11,12].
In particular, flavonoids are a wide class of phytochemicals most often extracted from the seeds, roots, leaves, and fruits peels of Citrus plants. Table 1, Table 2 and Table 3 report the botanical origin of compounds 1–79. Regarding the basic structure of their carbon framework, it is possible to distinguish three main categories: flavones (2-aryl-benzopyran-4-ones) (e.g., luteolin; Figure 1), flavanones ((S)-2-aryl-chroman-4-ones) (e.g., naringenin; Figure 1), and chalcones (1,3-diphenyl-2E-propene-1-one) (e.g., butein; Figure 1). These fundamental skeletons are decorated with several phenolic hydroxyl groups. A variable number of the −OH groups can be converted in vivo to the corresponding methyl ethers or to glycosides in which sugar moieties are attached by a glycosyl linkage to one of the several hydroxy groups.
Although there are different reviews that have extensively investigated the biosynthetic aspects of flavonoids from the Citrus genus [13,14] and their generic biological aspects [15], with a focus on the antioxidant properties of pure flavones extracted [16], this review aims to cover the literature concerning the spectroscopical properties and the anti-inflammatory activity of Citrus polymethoxyflavonoids (PMFs) that are chemical entities bearing only methoxy groups as substituents, and hydroxylated polymethoxyflavonoids (HPMFs) in which more than the one hydroxyl function is converted to a methoxy one.
Table 1. Polymethoxyflavones identified in the genus Citrus.
Table 1. Polymethoxyflavones identified in the genus Citrus.
No.CompoundSpeciesReferences
12′,3′-DimethoxyflavoneBrazilian orange essential oil[17]
22′,4′-DimethoxyflavoneBrazilian orange essential oil[17]
33′,4′-DimethoxyflavoneC. platymamma[18]
45,4′-DimethoxyflavoneBrazilian orange essential oil[17]
55,7-Dimethoxyflavone
(Chrysin dimethyl ether)
Brazilian orange essential oil[17]
67,4′-DimethoxyflavoneBrazilian orange essential oil[17]
3′,4′-Dihydroxy-7,5′-dimethoxyflavoneC. reticulata; C. sinensis[19]
75,6-Dihydroxy-7,4′-dimethoxyflavone
(Ladanein)
Fructus Aurantii
(C. aurantium dried)
[20]
83,5,6-Trihydroxy-7,4′-dimethoxyflavoneC. medica[21]
95,7,4′-Trihydroxy-6,3′-dimethoxyflavone
(Jaceosidin)
C. hassaku[20]
105,7,4′-Trihydroxy-6,8-dimethoxyflavone
(Demethoxysudachitin)
C. sudachi[22]
3,5,7,4′-Tetrahydroxy-8,3′-dimethoxyflavoneC. unshiu[23]
5,7,3′,4′-Tetrahydroxy-8,3′-dimethoxyflavoneC. sudachi[24]
112′,3′,4′-TrimethoxyflavoneBrazilian orange essential oil[17]
123,5,7-Trimethoxyflavone
(Galangin trimethyl ether)
Brazilian orange essential oil[17]
133′,4′,5′-TrimethoxyflavoneBrazilian orange essential oil[17]
145,7,4′-Trimethoxyflavone
(Apigenin trimethyl ether)
C. rutaceae[25]
155,3′,4′-TrimethoxyflavoneBrazilian orange essential oil[17]
166,2′,3′-TrimethoxyflavoneBrazilian orange essential oil[17]
3-Hydroxy-5,7,8-trimethoxyflavoneC. reticulata; C. sinensis[19]
175-Hydroxy-6,7,4′-trimethoxyflavone
(Salvigenin)
C. sinensis[20]
5-Hydroxy-7,3′,4′-trimethoxyflavoneC. reticulata; C. sinensis[19]
185-Hydroxy-7,8,4′-trimethoxyflavone
(Isoscutellarein 7,8,4′-trimethyl ether)
C. reticulata[26]
198-Hydroxy-5,7,4′-trimethoxyflavoneC. reticulata Blanco[27]
4′-Hydroxy-5,6,7-trimethoxyflavoneC. aurantium[28]
205,4′-Dihydroxy-6,7,8-trimethoxyflavone
(Xanthomicrol)
C. reticulata[29]
213,5,6-Trihydroxy-7,3′,4′-trimethoxyflavoneC. medica[21]
225,7,4′-Trihydroxy-6,8,3′-
trimethoxyflavone
C. sudachi[24]
233,5,7,4′-Tetrahydroxy-6,8,3′-trimethoxyflavone
(Limocitrol)
C. limon[30]
243,6,7,4′-TetramethoxyflavoneC. hallabong (hybrid)[31]
255,6,7,4′-Tetramethoxyflavone
(Scutellarein tetramethyl ether)
C. sinensis; C. reticulata[20]
265,7,8,4′-Tetramethoxyflavone
(6-Demethoxytangeretin)
C. reticulata Blanco cv. Ponkan[20]
275,7,3′,4′-Tetramethoxyflavone
(Luteolin tetramethyl ether)
C. reticulata; C. sinensis[19]
286,7,8,4′-TetramethoxyflavoneC. unshiu[32]
7,8,3′,4′-TetramethoxyflavoneC. platymamma[18]
297,3′,4′,5′-TetramethoxyflavoneC. reticulata[33]
303-Hydroxy-5,6,7,4′-tetramethoxyflavone
(Eupatorin-5-methylether)
C. sinensis[34]
5-Hydroxy-3,6,7,8-tetramethoxyflavoneC. reticulata; C. sinensis[19]
315-Hydroxy-3,6,7,4′-tetramethoxyflavone
(Penduletin 4′-methyl ether)
C. reticulata[27]
325-Hydroxy-3,7,8,4′-tetramethoxyflavoneC. albanicus; C. parviflorus[35]
335-Hydroxy-3,7,3′,4′-tetramethoxyflavone
(Retusin)
C. miaray[36]
345-Hydroxy-6,7,8,4′-tetramethoxyflavone
(Gardenin B)
Fructus aurantii
(C. aurantium dried)
[37]
355-Hydroxy-6,7,3′,4′-tetramethoxyflavone
(5-Desmethylsinensetin)
C. aurantium[28]
5-Hydroxy-7,8,3′,4′-tetramethoxyflavoneC. reticulata; C. sinensis[19]
366-Hydroxy-5,7,8,4′-tetramethoxyflavoneC. reticulata; C. sinensis[19]
377-Hydroxy-5,6,8,4′-tetramethoxyflavoneC. reticulata; C. sinensis[19]
7-Hydroxy-5,6,3′,4′-tetramethoxyflavoneC. reticulata[38]
384′-Hydroxy-5,6,7,8-tetramethoxyflavoneC. reticulata[26]
395,7-Dihydroxy-6,8,3′,4′-tetramethoxyflavone
(Hymenoxin)
C. medica L.[21]
5,8-Dihydroxy-3,7,3′,4′-tetramethoxyflavoneC. miaray[36]
405,3′-Dihydroxy-3,7,4′,5′-tetramethoxyflavoneC. monspeliensis[39]
415,4′-Dihydroxy-6,7,8,3′-tetramethoxyflavone
(8-Methoxycirsilineol)
C. reticulata[29]
7,4′-Dihydroxy-5,6,8,3′-tetramethoxyflavoneC. deliciosa[40]
423,5,6,8,4′-PentamethoxyflavoneC. reticulata[41]
433,5,7,3′,4′-Pentamethoxyflavone
(Quercetin pentamethyl ether)
C. miaray[20]
3,6,7,8,4′-PentamethoxyflavoneC. sinensis[20]
445,6,7,8,4′-Pentamethoxyflavone
(Tangeretin)
C. sinensis[20]
455,6,7,3′,4′-Pentamethoxyflavone
(Sinensitin)
C. reticulata; C. sinensis[19]
465,7,8,3′,4′-Pentamethoxyflavone
(Isosinensetin)
C. sinensis[20]
475,7,2′,3′,4′-PentamethoxyflavoneC. reticulata Blanco;
C. reticulata Chachi
[42]
486,7,8,3′,4′-Pentamethoxyflavone
(Demethylnobiletin)
C. reticulata Blanco cv. Ponkan[20]
493-Hydroxy-5,6,7,8,4′-pentamethoxyflavone
(3-Demethylnobiletin)
C. sinensis Osbeck[34]
505-Hydroxy-3,6,7,8,4′-pentamethoxyflavone
(5-Hydroxyauranetin)
C. aurantium[43]
515-Hydroxy-3,7,8,3′,4′-pentamethoxyflavone
(Gossypetin pentamethylether)
C. sinensis Osbeck[20]
525-Hydroxy-6,7,8,3′,4′-pentamethoxyflavone
(5-Demethylnobiletin)
C. sinensis[34]
535-Hydroxy-6,7,3′,4′,5′-pentamethoxyflavone
(Umuhengerin)
C. reticulata; C. sinensis[19]
547-Hydroxy-3,5,6,3′,4′-pentamethoxyflavoneC. reticulata[44]
557-Hydroxy-5,6,8,3′,4′-pentamethoxyflavone
(7-Demethylnobiletin)
C. reticulata; C. sinensis[19]
568-Hydroxy-3,5,6,7,4′-pentamethoxyflavoneC. aurantifolia[45]
573′-Hydroxy-5,6,7,8,4′-pentamethoxyflavone
(3-Hydroxytangeretin)
C. changshan-huyou[46]
584′-Hydroxy-5,6,7,8,3′-pentamethoxyflavone
(4′-Demethylnobiletin)
C. reticulata; C. sinensis[19]
593,5,6,7,8,4′-Hexamethoxyflavone
(3-Methoxytangeretin)
C. sinensis[20]
3,5,6,7,3′,4′-HexamethoxyflavoneCommercially Citrus peels extract[47]
3,5,6,8,3′,4′-HexamethoxyflavoneC. hassaku[20]
603,5,7,8,2′,5′-HexamethoxyflavoneC. reticulata Blanco[48]
613,5,7,8,3′,4′-Hexamethoxyflavone
(Gossypetin hexamethyl ether)
C. hassaku[20]
623,6,7,8,2′,5′-HexamethoxyflavoneCitrus unshiu[32]
635,6,7,8,3′,4′-Hexamethoxyflavone
(Nobiletin)
C. reticulata[44]
645,6,7,3′,4′,5′-HexamethoxyflavoneC. reticulata; C. sinensis[19]
655,7,8,3′,4′,5′-Hexamethoxyflavone
(Bannamurpanisin)
C. reticulata; C. sinensis[19]
663-Hydroxy-5,6,7,8,3′,4′-hexamethoxyflavone
(Natsudaidain)
C. aurantium[28]
675-Hydroxy-3,6,7,8,3′,4′-hexamethoxyflavoneC. kinokuni[20]
685-Hydroxy-6,7,8,3′,4′,5′-hexamethoxyflavone
(Gardenin A)
C. reticulata; C. sinensis[19]
6-Hydroxy-3,5,7,8,3′,4′-hexamethoxyflavoneC. unshiu[23]
697-Hydroxy-3,5,6,8,3′,4′-hexamethoxyflavoneC. reticulata[44]
708-Hydroxy-3,5,6,7,3′,4′-hexamethoxyflavoneC. aurantifolia[45]
8-Hydroxy-5,6,7,3′,4′,5′-hexamethoxyflavoneC. reticulata; C. sinensis[19]
3,5,6,7,8,3′,4′-HeptamethoxyflavoneC. miaray[36]
715,6,7,8,3′,4′,5′-HeptamethoxyflavoneBrazilian orange essential oil[49]
Table 2. Polymethoxyflavanones identified in the genus Citrus.
Table 2. Polymethoxyflavanones identified in the genus Citrus.
No.CompoundSpeciesReferences
725,6,7,4′-TetramethoxyflavanoneC. sinensis[34]
5-Hydroxy-3,6,7,4′-tetramethoxyflavanoneC. reticula Blanco[50]
735,6,7,8,4′-PentamethoxyflavanoneC. reticula Blanco[50]
745,6,7,3′,4′-PentamethoxyflavanoneC. reticulata; C. sinensis[19]
755,7,8,3′,4′-PentamethoxyflavanoneC. reticula Blanco[27]
6,7,8,3′,4′-PentamethoxyflavanoneC. reticula Blanco[50]
765-Hydroxy-6,7,8,3′,4′-pentamethoxyflavanone
(5-Demethylcitromitine)
C. sinensis[34]
6-Hydroxy-5,7,8,3′,4′-pentamethoxyflavanoneC. jambhiri[51]
775,6,7,8,3′,4′-Hexamethoxyflavanone
(Citromitin)
C. miaray[36]
Table 3. Polymethoxychalcones identified in the Citrus genus.
Table 3. Polymethoxychalcones identified in the Citrus genus.
No.CompoundSpeciesReferences
782′-Hydroxy-3,4,4′,5′,6′-pentamethoxychalconeC. sinensis[20]
792′-Hydroxy-3,4,3′,4′,5′,6′-hexamethoxychalconeC. sinensis[34]
Properly speaking, compounds characterized by the presence of just one methoxy group do not belong to the general type of HPMFs, and so they will be considered out of the scope of this paper. Further, we intend to report compounds with a chemical structure that has been unambiguously characterized by spectroscopical methods and that have not been discussed in previous reviews covering the topic. An update regarding molecules that have already been reviewed in the recent past is included.
An important criterion of the selection of the references reviewed within this work is also the presence of a defined relationship between bioactivity and the individual compounds. Reports concerning the bioactivity of mixtures (e.g., plant extracts) have not been taken into consideration, even where these are accompanied by a chemical identification of the mixture. This is because they lack any cause–effect evidence.
The biological, pharmacological, chemo-preventive, and therapeutic activities, as well as the molecular details of the mechanism of action of PMFs and HPMFs, have been described in several works [52,53,54,55,56]. Many of these molecules have been considered as powerful chemo-preventive agents for various important diseases, including cancer and neurodegenerative pathologies, as well as other disorders related to inflammation. Some of these compounds are rather common and ubiquitous in nature, and they have received a great deal of attention from natural product researchers; these include tangeretin (44) [57,58], nobiletin (63) [57,59,60,61], sinensitin (45) [62], and jaceosidin (9) [63]. Papers concerning the compounds cited above that have been already taken into consideration in other reviews will be mentioned here only when it is necessary for the purpose of comparing their bioactivity with that of the other compounds.

2. Radical Scavenging Activity

The occurrence of a strict relationship between the level of cellular oxidative stress and the tissue inflammation status is well known, so it is very important to evaluate the antioxidant potential of any putative anti-inflammatory compound. The comparison of the radical scavenging activity measured both by the ORAC and CUPRAC assays (Table 4) clearly shows that the presence of the −OH at C-5 implies a significant increase in the reduction power of the molecule.
Oxygenated radical species (ROS) are mainly responsible for the chemical degradation of important molecular cellular components, such as chromatin, and their control is crucial. Moreover, the presence of –OH groups at C-5 and C-7 in the A ring in the flavone framework was found to be essential in the inhibition of ROS generation in rabbit neutrophils activated with serum-opsonized zymosan (OZ) and insoluble ICs (ICIgG) [73]. This study included two flavones from Citrus lacking the above-mentioned structural feature, e.g., compounds 33 and 43, which showed a moderate inhibitory activity (ca. 30 and 60%, respectively, for OZ, and 13 and 45%, respectively, for ICIgG).
The involvement of tartrate-resistant acid phosphatase (TRAP) in the intracellular generation of reactive oxygen species (ROS) is well known. Table 4 reports the inhibition activity of several PMFs and HPMFs toward this protein.

3. Inhibition of Enzymatic Activity

PMFs and HPMFs can be active toward the proteins involved in the very complex multi-pathway activation of tissue inflammation. Prostaglandins are well recognized mediators of the inflammatory process and the inhibition of the key leukotrienes cycloxygenation step involved in their biosynthesis is a very important goal. Some relevant data concerning the reduction of the production of prostaglandins are reported in Table 5. Many other cases are cited within the following discussion.
Some authors reported computational data concerning the relationship between the structure of different known nutraceuticals, including some PMFs and HPMFs, and the inhibition activity on cycloxygenase-2 (COX-2) determined by a docking procedure. Indeed, the statistical parameters obtained, in particular the correlation coefficient R2 (0.462 for AutoDock Vina affinities and 0.238 for Gold scores) between the predicted and the experimental values are quite low, even though the authors claimed the opposite [77]. Experimental determinations of COX-2 inhibition are available for different PMFs and HPMFs from Citrus taxa (Table 6).
The inhibiting power toward this enzyme was evaluated for some natural occurring and semisynthetic flavones [76] and some relevant structural features that can enhance the bioactivity include the presence of a methoxy group at C-8 and of free phenolic –OH at C-5 and C-7. The type of substitution in ring C seems to be less important. Hence, whereas compounds 18 and 26 were inactive, compound 46 was found to be moderately active and 65 was significantly active (Table 6). Other authors investigated the structure–activity relationship for the COX-2 inhibition for a variable number of PMFs and HPMFs [52] by 2D-QSAR methods and obtained a significant correlation coefficient between the predicted versus the experimental mRNA inhibition (R2 = 0.80). The importance of the presence of the C-2, C-3 double bond as well as the methoxy group at C-4′ was highlighted.
Interestingly, the presence of the above-mentioned alkenyl moiety was also recognized as advantageous in the β-glucuronidase release stimulated by f-MetLeuPhe in human neutrophils [85]. On the other hand, unlike for the case of COX-2, the glucuronidase release appears to be favored by the presence of a phenolic −OH at C-4′, even though it is not very clear from the data reported in this work, whether the main structural contribution to this bioactivity has to be attributed to the 4′ −OH or perhaps to the presence of –OH at C-5 and C-7. However, all the tested PMFs lacking any −OH were inactive (compounds 25, 44, 45, 63, and 71).
Several other enzymes involved in several biochemical pathways can be considered as valuable targets to control the inflammation status and both PMFs and HPMFs were proven to be active (Table 6). Compounds 46 and 14 were also proven to be ineffective as collagenase inhibitors [86], unlike other flavonoids not present in the Citrus genus.

4. Inhibition of Nitric Oxide (NO) Production and the Related Biochemical Effects In Vitro

Nitric oxide (NO) is involved in a number of important physiological mechanisms and its level can be raised up as a consequence of an overexpression of the inducible nitric oxide synthase (iNOS), the isoform of NOS produced as a response to the stimulus by pro-inflammatory cytokines. The expression of iNOS can be activated by the translocation of the κB nuclear factor (NF-κB) to the nucleus. It is worth noting that a number of PMFs and HPMFs are able to reduce the levels of both NO (Table 7) and iNOS mRNA, as well as regulate the activity of NF-κB, as described below.
Table 7 reports data taken from several references and, where possible, they are converted in homogeneous units for the purpose of a suitable comparison. Sometimes different values were obtained by different authors; these discrepancies may be due to different degrees of purity of the isolated compounds, as well as by the use of completely different determination methods. However, it is our opinion that the data reported can be considered on the whole to be quite coherent.
Compounds 5, 14, and 27 were able to inhibit NO production in LPS-activated RAW267.4 cells, showing remarkable IC50 values (Table 7) [83]. Two structural features shared by these compounds are the lack of both the −OH at C-5 and the −OMe at C-3. The three compounds were demonstrated to act at the transcriptional level by inhibiting the expression of the inducible NO Syntase (iNOS) mRNA. Compound 5 was significantly capable to limit the expressions of iNOS, p-IκBα and p-NF-κB in human HaCaT cells stimulated by propionibacterium acnes, thus confirming its potential as an anti-acne agent [107]. The same molecule, when administrated to rats at doses of 100 mg/Kg bw, was able to renormalize the NED induced overexpression of pro-inflammatory proteins (iNOS, COX-2) and NF-κB in HCC cells obtained from pre-neoplastic nodules [108].
Compound 27 showed also a strong effect in inhibiting NO and IL-6 production, both in Raw267.4 and in GES-1 cells. The effect being relevant at only 3 μM [102]; a similar activity was disclosed for compound 52. Further, a mild effect in the TNFα release was proven for the three compounds (IC50 ca. 300 μM). On the contrary, compound 27 displayed a strong reduction effect toward TNF-α release in human mast cells [109]. Furthermore, 27 was found to be effective in preventing inflammatory reactions in human mast cells stimulated by neuropeptides substance P (SP) and neurotensin (NT) [110]. This compound inhibited the mTOR-mediated production of TNF, CXCL8, and VEGF and downregulated the gene expression of these proteins. It also inhibited the degranulation of MCs. The beneficial effect of 27 was still confirmed in a study where the inflammatory reaction of human MCs was induced by substance P together with IL-33 [111]. The pre-treatment with 27 100 μM implied the inhibition of secretion and gene expression of IL-1β, procaspase-1, and pro-IL-1β.
The effect of 27 was higher than that of its non-methoxylated analogue luteolin, even though the very low solubility and oral availability of the highly lipophilic 27 can be an obstacle for its eventual pharmacological development.
A similar effect was reported [106] in the RAW274,5 cell system for the NO production inhibited by the tetramethoxy derivative 25. This compound showed also a remarkable inhibition on TNFα release and a potent action toward PGE2 and IL-6 production that were almost nullified upon treatment at 100 μM. IL-1β was instead reduced by a much lesser extent (ca. 40%) at the same concentration of the tested compound. A similar pattern of inhibition of the production of pro-inflammatory factors was established for compound 25 by other authors [93]. They described the downregulation of gene activity for the expression of iNOS, COX-2, IL-6, and TNF-α when murine LPS-activated Raw267.4 cells were treated with 25 at 50, 100 μM. This bioactivity was related to the modulation of the NF-κB pathway by 25 that was able to reduce the translocation of p50 and p65 subunits by decreasing the phosphorylation of IKKβ.
With the use of the same cellular system, Li et al. [95] found a significant inhibiting power in NO generation for compounds 57 and 58, compared to that of 63, which was much lesser (Table 7). This result was in accordance with a downregulation of both iNOS and COX-2 expressions for the two most active compounds.
Similar effects were investigated for compound 58 formed in the phase I metabolism of 63 [112]. This compound dramatically reduced the NO production (Table 7) and the proinflammatory cytokines PGE2, IL-1β, and IL-6, acting by the inhibition of the transcription of both COX-2 and iNOS. Furthermore, the anti-inflammatory activity of 58 was also related to the inhibition of translocation of both NF-κB and AP-1, as well as the activation of NRf2 and its dependent genes NQO-1 and OH-1.
An interesting comparison was made between the anti-inflammatory activity of compounds 20 and 36, structurally differing only for the presence of a phenolic OH in C-4′ of the former versus a methoxy group in the latter [94]. The substitution of −CH3 with a -H was responsible for a decrease in some inflammatory parameters in RAW264.7 cells, such as NO (Table 7) and IL-1β production, as a consequence of reduced expression of iNOS and IL-1β mRNA. On the other hand, some others factors remained substantially unchanged, i.e., PGE2 and COX-2 mRNA. Finally, the OH-1 mRNA expression was higher after treatment with 20 than with 34 at the same dose of 2 μM. These differences may be justified on the grounds of the difference in bioavailability related to the reduced hydrophilicity of 34, with respect to 20.
In another investigation aiming to evaluate the inhibition of TNF-α release in LPS activated RAW264.7 cells, compounds 44, 63, and 71 were all weakly active [113].
Compounds 5 and 33 were found to inhibit the expression of the enzymes involved in the inflammatory response of LPS-activated RAW267.4 cells as they significantly reduced the mRNA expression of IL-1β, TNF-α, IL-6, COX-2, and iNOS [114]; iNOS and COX-2 were deeply downregulated in the same cell type by compound 17 at 10 μM [115]. Further, IL-1β and TNF-α were also effectively reduced in LPS-activated renal epithelia HK-2 cells by compounds 3, 17, 25, and 45 [116]. The antiinflammatory activity of 17 was also proven for another biological model, LPS-activated THP-1 monocytic leukemia cells, where this compound inhibited IL-6 release [117]. Sinensetin (47) inhibited the release of several cytokines and chemokines, by reducing their mRNA expression (IL-6, IL-8, IP-10, MCP-1, and TNF-α), which are involved in the inflammation process induced by the H1N1 influenza-A virus on A549 cells [118].
Another interesting structural comparison was proposed by During et al. [119], which evaluated the effect of −OH methylation in the capacity to reduce pro-inflammatory factors in intestinal caco-2 cells, in which the inflammatory response was induced by IL-1β. For example, compound 5 was significantly active in limiting the release of IL-6, MCP-1, COX-1 and COX-2, unlike for chrysin, which is the non-methoxylated form of 5. The same compound was able to partially reduce the amount of IL-8 and the activation of NF-κB.
Furthermore, other PMFs were compared with their related hydroxylated counterparts (3 vs. 3,4-dihydroxyflavone, 14 vs. apigenin, 27 vs. luteolin, and 43 vs. quercetin) in their inhibiting power toward the production of IL-6, IL-8, MCP-1, and PGE2. The authors were able to summarize the results obtained with a number of structure-activity inferences, i.e., methylation of the 5- and 7-hydroxyl groups on the A-ring increases the bioactivity, as well as the lack of methylation at the 3′-hydroxyl group on the B-ring, and finally the methylation of the 3-hydroxyl group on the C-ring.
In another investigation, comparing quercetin with different derivatives [70], the per-methylation of this compound yielded 43, and dramatically reduced both the antioxidant and the anti-inflammatory activities measured by several tests: DPPH scavenging (IC50 μg/mL 60 vs. 1.3), FRAP (μg ascorbic acid 0 vs. 2), lipids peroxydation (IC50 μg/mL 85 vs. 5), TXB2 inhibition (IC50 μg/mL 80 vs. 55), PGE2 inhibition (IC50 μg/mL 58 vs. 15), 12-HHT (IC50 μg/mL 50 vs. 22), and 12-HETE (IC50 μg/mL 79 vs. 5). It is worthy to note how the inhibition in the transcription of some inflammatory factors in SW1353 cells was dependent upon the number of methoxy substituents in the flavone core [120]. The order of inhibition power toward the mRNA expression of IL-1β and IL-6 was pentamethoxyflavone 43 > trimethoxyflavone 14 > dimethoxyflavone 5. On the other hand, the effect on the mRNA expression of TNF-α was less differentiated.
Compound 45 was also found to be active in inhibiting the superoxide anion production in neutrophils activated with the chemotactic peptide N-formylmethionine-leucyl-phenylalanine (FMLP). This activity was associated with a remarkable reduction of the elastase release [121]. Other authors were able to show that the presence of the phenolic −OH at 4′ strongly increased the inhibiting power toward the NO production in LPS activated RAW 264.7 macrophages [105]. Indeed, compound 38 was the most active PMF with respect to the NO inhibition (Table 7). Furthermore, this compound at 40 mM completely suppressed the PGE2 production in this cell model.
Similar dose-depending effects on RAW 264.7 cells were envisaged for compound 67 that effectively reduced the production of NO (Table 7) and PGE2 (Table 5) [76]. Other pro-inflammatory factors, such as TNF-α and IL-1β, were restored to physiological levels in the presence of 67 al 40 μM. These effects were associated with the inhibition of NF-κB translocation and activation. Compound 67 was also investigated for its effect in the significant inhibition of the TPA-induced mouse skin inflammation [122] involved in skin tumor genesis. The mRNA expressions of both iNOS and COX-2 were reduced significantly (5-fold and 6-fold respectively) after the topical application of a solution 5-15 mM in acetone for 67. These effects were accompanied, also in this case, by the inhibition of NF-κB activity and the inhibition of translocation of this protein due to the stoppage of the phosphorylation of its inhibitor IkBα.
The relevant inhibition of the generation of inflammatory proteins on a transcriptional level on SW982 cells stressed with inflammatory cytokines was related to the number of methoxy groups in the flavone framework [123]. The mRNA expression of IL-6, IL-1β, and COX-2 followed the order pentamethoxyflavone 43, which was significantly higher than trimethoxyflavone 14, and slightly higher than dimethoxyflavone 5. The range of mRNA expression observed was once to twice with respect to the positive control (cells untreated with cytokines) for 5 and 14, whereas 43 was poorly active (mRNA expression 25 to 70 times higher). These results could be explained by the lower polarity, which means lower bioavailability because of the presence of a higher number of methoxy substituents. On the other side, the expression of TNF-α mRNA was much lesser dependent on the structure of the three compounds.
Gardenin A (68), a hexamethoxyflavone bearing a hydroxyl group at C-5, was shown to be a potent anti-inflammatory agent in PMA/ionomycin-induced EL-4 cells. This compound reduced IL-5 and ROS production and induced HO-1cexpression through the transcription factor PPARγ. Overall, this data accounts for a possible application of HPMF in controlling the oxidative stress associated with asthma [124].
Compounds 34, 35, and 52 were investigated for their anti-inflammatory potential by employing carrageenan-challenged peripheral blood mononuclear cells (PBMCs) as a biological model [125]. The oxidative status of the lysate cells was evaluated by determining the levels of MDA, GSH, SOD, and CAT. Whereas the pre-treatment with the three compounds significantly mitigated MDA accumulation and SOD reduction associated with the inflammation status, GSH and CAT remained unaffected. Furthermore, the levels of inflammation markers COX-2, iNOS, MCP-1, IL-1β, IL-6, and MPO were notably decreased in cells treated with compounds 1 and 2 at 10 mM, whereas compound 3 was not effective in modulating the ratio of COX-2, iNOS, and MCP-1.
The inflammatory pathway in rat basophilic leukemia (RBL-2H3) cells was inhibited on a transcriptional level by both compound 5 and 33, which reduced both the protein concentration of TNF-α, IL-4, and MCP-1, and the expression of their mRNA [126]. The effect of these compounds was higher than that of the known anti-inflammatory molecule nobiletin (65). The cells were more responsive when sensitized by a degranulation- inducing antigen rather than with a calcium ionophore. The same biological target was involved in a comparative study concerning compounds 63 together with its 3-hydroxyderivative, natsudaidain (66) [127]. Both compounds had a fable effect on the histamine release and they both inhibited the expression of COX-2 and TNF-α to a significant extent. The presence of an extra –OH group in 68 had a slightly beneficial effect.

5. Anti-Neuroinflammatory Activity of Citrus PMFs and HPMFs

The role of microglia in neurodegeneration is an expanding area of biomedical investigation that requires the involvement of suitable biological models, such as immortalized BV-2 cells that have been being employed in many studies in recent years. These macrophage-like cells are significantly present in CNS and are involved in some significant neurodegenerative pathologies such as Alzheimer’s, Huntington’s, and Parkinson’s diseases.
The influence of HPMF 67 in the NO production and regulation mechanism was investigated in microglia BV2 cells [98]. The relevant inhibition in the NO production in LPS-activated BV2 cells, treated with 67, was accompanied by a downregulation of iNOS expression and by the inhibition of NF-κB activity and nuclear translocation. Furthermore, the reduction of NO in these cells was also associated with an overexpression of heme oxygenase 1 (OH-1) enzyme induced by compound 67 and stimulated by the activation of the nuclear factor Nfr2. Interestingly, the activation of the nuclear factor by 67 was reversed by the knockdown of its activator gene obtained by a specific siRNA.
The protective effect toward the neuroinflammation process in BV2 microglia was also investigated for HPMFs 9 [97] that displayed a remarkable IC50 value for the NO inhibition (27 μM). In the same work, the inhibitory effect of compound 9 on NO production was evidenced for other cell types such as HAPI microglial, primary astrocytes, and RAW 254,7. Furthermore, compound 9 significantly reduced the levels of the pro-inflammatory factors IL-1β, COX-2, iNOS, and TNF-α in BV2 cells. On the other hand, compound 33 was ineffective in the inhibition of the production of TNF-α, whereas the same compound was established as an effective inhibitor for the production of PGE2 (IC50 16.3 μM) [88,128].
It is worthy to note as even minimal structural differences in the substitution pattern of the flavone framework can cause relevant differences in the mechanism of the anti-inflammatory action. For example, compounds 44, 52, and 63 differ only in the presence of one more methoxy group in 63, and one methyl less in 44 with respect to compound 52. The three compounds showed a comparable inhibition power for the NO production in BV2 cells and a similar ability in reducing the expression of the cytokines IL-1β, IL-6, and TNF-α [129]. However, whereas both 44 and 52 reduced the expression and the phosphorylation of JAK2, as well as the phosphorylation of STAT3, 62 was ineffective toward the expression and the phosphorylation of STAT3. The anti-inflammatory activity of compounds 52 and 65 in BV2 cells was confirmed in their ability to downregulate the transcription of iNOS and IL-1β mRNA [130]; similar effects were demonstrated for compound 71. The last compound (63) also showed mild activity (IC50 ca. 20 μM) in the inhibition of the NF-κB activation [131].
The anti-inflammatory activity of 44 was responsible for a notable neuroprotective effect in ischemia/reperfusion (I/R)-injured rat brains [124,132]. Compound 44 significantly decreased the brain tissue pathological parameters such as water content, brain edema, infarct volume, neurological score, and Evans blue leakage. Some relevant pro-inflammatory factors were also downregulated after administration of 5–20 mg/Kg bw of 44; these include IL-1β, IL-6, TNF-α, TLR-4, and IFNG-γ.
Compound 45 showed anti-inflammatory protection toward SH-SY5Y cells with amyloid β25-35-induced oxidative stress [133]. The compound was efficient in reducing MDA and increasing the SOD and CAT activities at 20–40 μM. Nitric oxide production, IL-1β, and TNF-α were also inhibited together with iNOS and COX-2 mRNA expressions. Further the translocation of the p65 subunit of NF-κB was downregulated as well as TLR-4 expression.

6. Anti-Inflammatory Activity In Vivo of Citrus PMFs and HPMFs

Compound 9 has been proven to act efficiently against the cartilage destruction in an osteoarthritis mouse model. The biochemical basis of this in vivo effect was investigated thoroughly in vitro, and it was shown that the mechanism of action of this compound is the inhibition of IκB degradation in NF-κB pathway [134]. The same compound appears to be partially effective in reducing ear edema in mice by 23.2% (1 mg/ear) [135]; this activity was related to a notable reduction of NF-κB activation (77%).
Compound 58 effectively reduced ear edema in mice at a dose from 2 to 4 μmol/ear [96]. This effect was accompanied with a strong downregulation of IL-1β, IL-6, and TNF-α, in addition to a reduction of mRNA expression of COX-2, iNOS, and MMP9. Similar effects were disclosed for compounds 5 and 27 that, at a dose of 0.5 mg per ear, were demonstrated to be effective in reducing the ear edema volume (ca. 53 and 50%, respectively), induced by 12-0-tetradecanoylphorbol acetate (TPA) [136]. The same biological model was applied to test retusine (35), which displayed an IC50 of 0.43 mg/ear [91]. The same group determined the reduction of the paw edema in mice, induced by either TPA and carrageenan, and the administration at 20 mg/Kg of 35 was able to reduce the edema volume by ca. 50%.
Compound 5 was also investigated using rat paw edema as a model in vivo, showing a notable volume reduction (from 45 to 67%) when administrated orally at doses from 300 to 1200 mg/Kg bw [137]. The same authors also proved a modest reduction effect on carrageenan-induced pleurisy and on cotton pellet-induced granuloma formation (4.5% inhib. at 300 mg/Kg bw). This compound was able to reduce the PGE2 production by about 70%. A different number of flavonoids were compared for their oral anti-inflammatory activity in the rat carrageenan-induced hind paw edema test, to gain information on the structure activity relation [138]. Three of the compounds included in these investigations are flavonoids from Citrus, i.e., compounds 5, 14, and 43. These compounds had volume inhibition rates of 43.7, 19.5, and 25.3%, respectively, at a dose of 300 mg/Kg bw. The authors concluded that the presence of the pyrene ring and the methoxy groups at C-5 and C-7 are essential for this kind of anti-inflammatory activity.
Compounds 35 and 71 had a similar effect on paw edema (about 30% for both compounds after 7 h) after an intraperitoneal administration [139]. On the other hand, the result in the cotton pellet-induced granuloma test (chronic inflammation) was 16 and 21% inhibition, respectively, higher than that reported for compound 5.
The relevant action of compound 6 against acute paw inflammation in rats was reported. This compound is characterized by the presence of two methoxy groups at C-7 and C-4′ [140]. The dose-dependent effect was maximized at 100 mg/Kg bw (90%). A reduction in the number of inflammation biochemical indexes accompanied the in vivo effect of this molecule, that is the reduction of both COX-1 and COX-2 activities (IC50 52 and 24 μM, respectively), as well as the significant decrease in the production of both TNF-α and IL-1β.
Apparently, a higher number of methoxy groups is favorable to obtain a better bioavailability for this type of chronic inflammation test. This generalization is enforced by the results of other groups [141] that proved a quasi-linear correlation between the paw edema volume reduction and the number of methoxy groups in the compounds given at 75 mg/Kg bw, the most active compound being the pentamethoxyflavone quercetine (43).
Inflammation is one of the physiological responses to the allergy reaction in tissues, such as delayed-type hypersensitivity (DTH). Demethylnobiletin (52) was recognized to be able to modulate DHT induced in mice with the allergen oxazolone, dinitrofluorobenzene (DNFB), and sheep red blood cells (SRBC) by reducing cell infiltration and by suppressing the release of inflammation mediators [90]. Significant reduction rates were observed for IL-2 (IC50 1.63 μM), IL-4 (IC50 2.67 μM), TNF-α (IC50 0.66 μM), IL-1β (46% inhib. at 2.5 μM), and interferon-γ (IC50 1.35 μM). In the same study, compound 52 was effective in reducing NO production (Table 7), even though iNOS expression remained surprisingly unaffected.
Furthermore, 52 was able to restore the physiological level of TNF-α in mice when a pathological increase of this factor was induced by benzo[a]pyrene and DSS [142].
The in vivo anti-inflammatory action of 52 was confirmed by investigating the hepatoprotective effect of this compound in BULB/c mice treated with CCl4. The remarkable protective action of 52 toward ROS-mediated apoptosis in HepG2 cells was explained with the downregulation of CYP2E1 and hepatic SOD expressions, as well as with the notable reduction of the hepatic levels of MDA and GSH [143].
Compounds 17 and 45 confirmed the anti-inflammatory potential disclosed in vitro (see above) in the in vivo xylene-induced ear edema test. They were able to remarkably reduce the auricle swelling at a dose of 50 mg/Kg [116].
Compound 44 was active in restoring the level of pro-inflammatory factors in mice with dextran sulfate sodium (DSS)-induced colitis [144]. The levels of TNF-α, IL-1β, and IL-10 were significantly and dose-dependently downregulated after dietary administration of 44 (0.04 and 0.08%). On the contrary, the moderate reduction of IL-6 level was not statistically significant. Further, this compound was effective in the control of glucose- induced oxidative stress in mice [145]. It was capable of reducing both the 8-hydroxy-2-deoxy guanosine release, indicative of an oxidative DNA damage, and the ROS generation in diabetic podocytes when administrated at a dose of 10 mg/Kg bw. A relevant inhibitory activity toward ROS production was also disclosed for compound 22 at 10 μM in osteoclast lineage cells. The ROS increase induced by sRANKL was blocked by the simultaneous treatment of cells with sRANKL and 22. This activity was part of the mechanism by which this compound was able to suppress the inflammatory bone destruction [146]. An inhibitory activity of this compound toward NO (Table 7) and TNF-α (Table 8) was also highlighted in mouse macrophage-Like RAW264.7 cells, which was related to the downregulation of the relative mRNA activity [101].
Jecosidin (9) inhibited the hind paw edema induced by carrageenan by a volume of 30% at a dose of 20 mg/Kg bw. This activity was accompanied by a significant reduction of the inflammatory markers TNF-α and IL-1β (ca. 50%); the activity on PGE2 was instead weak [147]. The same compound (9) was found to be significantly active in inhibiting the ear edema in mice (ID50 ca. 0.50 μmol/ear) [148].
Compound 20 had similar activity in vivo, showing a reduction of ca. 7% vs. 34% of the positive control indomethacin [149]. Compound 71 had a significantly higher effect after intraperitoneal administration to rats (100 mg/Kg bw), as it was able to reduce the edema by 56%. Oral administration implied a lower hematic concentration of the active compound that reduced the activity at 21% [150]. This in vivo activity was associated with a relevant serum TNF-α reduction (45%) in the treated animals. On the other hand, the molecular mechanism of the anti-inflammatory action of 20 is still unclear, even though an in vitro inhibition toward TNF-α production (IC50: 8 μM in THP-1 cells and 4 μM in B16–F10 cells) was also established for this compound [151].
Compound 63 significantly ameliorated the histopathological score and the gastric mucosa injury index in mice suffering from acute gastric lesions induced by ethanol [152]. This effect was accompanied by a significant, linear dose-dependent reduction of several inflammatory markers such as MPO, SOD, GSH, and MDA, as well as serum and the tissues TNF-α and IL-6. PGE2 linearly increased within the same dose range of 63. The same compound also prevented colon carcinogenesis in mice through an anti-inflammatory mechanism on colon cells, where this flavone inhibited the synthesis of IL-1β, IL-6, and TNF-α [96,112,153].
The relative efficacy of PMFs and HPMFs in vivo obviously depends upon their metabolism, bioavailability, and pharmacokinetics after administration. The available literature data concerning this issue are still limited; however, a few useful investigations have been reported for compounds 5 [154], 9 [155,156], and 27 [157,158].
Regardless, because of the narrowness of these studies, it is possible to infer a few general points concerning the kinetics and the metabolism of PMFs that seem to show an average serum cmax of 1–2 μg/mL and a half-life of ca. 3–4 h. The metabolism pathways generally include demethylation at ring A (and eventually C) followed by solfonylation or glycoconiugation at various –OH. Hydroxylation of ring C was observed in compound 5. Interestingly, the high lipophilicity of compound 71 makes it easy to permeate the blood–brain barrier and exert its neuroprotective effects directly within the CNS after oral or intraperitoneal administration to rats [159].

7. Spectroscopical Data of Polymethoxy-Flavones, -Flavanones, and -Chalcones Isolated in Citrus Genus

This review, in addition to reporting the data on the anti-inflammatory activity of PMFs and HPMFs extracted from different species of the Citrus genus, reports the spectroscopic data of the single isolated polymethoxy-flavones (Figure 2), -flavanones (Figure 3), and -chalcones (Figure 4). This is because, as described in the previous paragraphs, the position of the different functionalities (−OMe or/and −OH) within the flavone skeleton is relevant and basic in the investigated activity.
In particular, the chemical structures of polymethoxyflavones (Table 9) and the spectroscopic values (expressed in ppm) of the single carbon signals (Table 10) and of the related methoxy groups (Table 11) of the polymethoxyflavones (171), and the chemical structures of polymethoxyflavonones (7277) and polymethoxychalcones (7879) and their 1H- and 13C-NMR data (Table 12, Table 13, Table 14 and Table 15) are reported. For the identification, an extensive bibliographic search was carried out including only the compounds purely isolated from the different species of the Citrus genus. Reported compounds include polymethoxyflavones (71 compounds), polymethoxyflavanones (6 compounds), and chalcones (2 compounds).
Table 9, Table 12 and Table 14 contain all flavonoids with their semisystematic names (and common names) sorted by increasing structural complexity.
All spectroscopic data were obtained using results from articles reporting single isolated polymethoxyflavonoids from the Citrus genus, with the help of electronic databases such as SciFinder, Scopus, Google Scholarm, and Web of Science, considering scientific works published until 2022.

8. Spectroscopical Data of Glycosylated Flavonoids Isolated in Citrus Genus

Only a few methoxylated and glycosylated flavonoids (8091) (Figure 5) have been found and identified mainly from the peels of various Citrus trees. They are essentially flavones, characterized by the constant presence of −OMe functionality in position 8 and 5′, glycosylated through β-glycosidic bonds in position 3, or in 7 or in 4′. These compounds were always characterized by the presence of a simple glucose moiety, at most esterified by the 3-hydroxy-3-methylglutaric acid chain (Table 16, Table 17 and Table 18).

9. Conclusions

Many species of Citrus taxa are part of the daily ordinary diet in several countries and have also been used in traditional medicine because of their beneficial effects on health. In recent years, some of the chemo-preventive and therapeutic effects of Citrus were proven by rigorous scientific investigations. The prevention and reduction of inflammation is a truly relevant bioactivity exerted by the Citrus secondary metabolites, which is the basis for the prevention of important disorders such as metabolic syndrome, cancer, cardiovascular, and neurodegenerative pathologies. The molecular basis of this anti-inflammatory action is rather complex and involves the regulatory action of Citrus PMFs and HPMFs, both at post-transductional and at post-transcriptional levels, as it was described in this paper. However, most of the investigations are performed in vitro or in animal models, such as the murine model that is now considered to poorly mimic human inflammatory diseases [110,190].
Studies of the bioactivities on humans are limited and they normally concern epidemiological issues and the use of fruit parts or extracts, rather than the pure isolated compounds. Relevant examples of these s kinds of investigations include the prevention of the metabolic syndrome development and the protection against heart failure by the bergamot extract [191,192] and the antioxidant effect of red orange extract in elderly subjects, which ameliorated their wellbeing (tested by SF-36 QoL and the MRS questionnaires) by reducing the TNF-α level and increasing GSH/GSSG ratio after an 8-week intake [193]. Even though these kinds of extracts present a rather complex chemical composition, it is strongly presumable that PMFs and HPMFs can play an important role. Pure compounds were much less investigated clinically, except for a few molecules. For example, nobiletin (63) and tangeretin (44) were the object of clinical trials aimed to ascertain their beneficial effect on human diseases such as Rhinoconjunctivitis, cognitive disfunctions, and nocturia [194]. It is generally assumed that Citrus extracts and their main components are safe and non-toxic and that they are considered as a possible valuable eventual alternative to standard anti-inflammatory drugs such as NSAIDs. However, the acute oral toxicity of sinensitin (45) was reported [195]. Hence, deep attention must be paid in order to avoid unsuitable generalizations on presumed healthy effects of Citrus extracts and preparations containing PMFs and HPMFs based on scientifically uncontrolled claims. Moreover, there is no doubt that the way is open toward the recognition of PMFs and HPMFs as valuable nutraceuticals or even pharmaceutical leads, likewise for their more established polyphenolic precursor. However, more investigations are necessary in order to better assess the efficacy, the toxicity, and the mechanism of actions of many PMFs and HPMFs and to improve their bioavailability by means of available and suitable drug delivery systems.

Author Contributions

Conceptualization, G.F. and N.B.; methodology, G.F.; software, N.B., G.F. and F.S.; formal analysis, G.F. and N.B.; investigation, G.F., N.B. and F.S.; resources, G.F. and N.B.; writing—original draft preparation, G.F., N.B. and F.S.; writing—review and editing, M.B., N.B. and G.F.; visualization, M.B.; supervision, M.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Scora, R.W. On the history and origin of Citrus. Bull. Torrey Bot. Club 1975, 102, 369–375. Available online: https://www.jstor.org/stable/248476 (accessed on 25 September 2022). [CrossRef]
  2. Jena, S.N.; Kumar, S.; Nair, N.K. Molecular phylogeny in Indian Citrus L. (Rutaceae) inferred through PCR-RFLP and trnL-trnF sequence data of chloroplast DNA. Sci. Hortic. 2009, 119, 403–416. [Google Scholar] [CrossRef]
  3. Kalita, B.; Roy, A.; Annamalai, A.; Ptv, L. A molecular perspective on the taxonomy and journey of Citrus domestication. Perspect. Plant Ecol. Evol. Syst. 2021, 53, 403–416. [Google Scholar] [CrossRef]
  4. Nicolosi, E.; Deng, Z.N.; Gentile, A.; La Malfa, S.; Continella, G.; Tribulato, E. Citrus phylogeny and genetic origin of important species as investigated by molecular markers. Theor. Appl. Genet. 2000, 100, 1155–1166. [Google Scholar] [CrossRef]
  5. Abbas, F.; Fares, A. Best management practices in citrus production. Tree For. Sci. Biotechnol. 2009, 3, 1–11. [Google Scholar]
  6. Satari, B.; Karimi, K. Citrus processing wastes: Environmental impacts, recent advances, and future perspectives in total valorization. Resour. Conserv. Recycl. 2018, 129, 153–167. [Google Scholar] [CrossRef]
  7. Food and Agricultue Organization of the United Nations (FAO). Citrus Fruit Fresh and Processed; Statistical Bulletin: Rome, Italy, 2020; Available online: https://www.fao.org/3/cb6492en/cb6492en.pdf (accessed on 2 December 2022).
  8. Lucia, C.; Laudicina, V.A.; Badalucco, L.; Galati, A.; Palazzolo, E.; Torregrossa, M.; Viviani, G.; Corsino, S.F. Challenges and opportunities for citrus wastewater management and valorisation: A review. J. Environ. Manag. 2022, 321, 115924. [Google Scholar] [CrossRef]
  9. Fontana, G. The Orange Peel: An Outstanding Source of Chemical Resources. Citrus—Res. Dev. Biotechnol. 2021, 207, 1–22. [Google Scholar]
  10. Sottile, F.; Del Signore, M.B.; Barone, E. Ornacitrus: Citrus plants (Citrus spp.) as ornamentals. Folia Hortic. 2019, 31, 239–251. [Google Scholar] [CrossRef] [Green Version]
  11. Badalamenti, N.; Bruno, M.; Schicchi, R.; Geraci, A.; Leporini, M.; Tundis, R.; Loizzo, M.R. Reuse of food waste: The chemical composition and health properties of pomelo (Citrus maxima) cultivar essential oils. Molecules 2022, 27, 3273. [Google Scholar] [CrossRef]
  12. Badalamenti, N.; Bruno, M.; Schicchi, R.; Geraci, A.; Leporini, M.; Gervasi, L.; Tundis, R.; Loizzo, M.R. Chemical compositions and antioxidant activities of essential oils, and their combinations, obtained from flavedo by-product of seven cultivars of Sicilian Citrus aurantium L. Molecules 2022, 27, 1580. [Google Scholar] [CrossRef] [PubMed]
  13. Zhao, C.; Wang, F.; Lian, Y.; Xiao, H.; Zheng, J. Biosynthesis of citrus flavonoids and their health effects. Food Sci. Nutr. 2020, 60, 566–583. [Google Scholar] [CrossRef] [PubMed]
  14. Addi, M.; Elbouzidi, A.; Abid, M.; Tungmunnithum, D.; Elamrani, A.; Hano, C. An Overview of bioactive flavonoids from Citrus fruits. Appl. Sci. 2022, 12, 29. [Google Scholar] [CrossRef]
  15. Yi, L.; Ma, S.; Ren, D. Phytochemistry and bioactivity of Citrus flavonoids: A focus on antioxidant, anti-inflammatory, anticancer and cardiovascular protection activities. Phytochem. Rev. 2017, 16, 479–511. [Google Scholar] [CrossRef]
  16. Wang, Y.; Liu, X.J.; Chen, J.B.; Cao, J.P.; Li, X.; Sun, C.D. Citrus flavonoids and their antioxidant evaluation. Food Sci. Nutr. 2022, 62, 3833–3854. [Google Scholar] [CrossRef]
  17. Gaydou, E.M.; Berahia, T.; Wallet, J.C.; Bianchini, J.P. Gas chromatography of some polymethoxylated flavones and their determination in orange peel oils. J. Chromatogr. A 1991, 549, 440–445. [Google Scholar] [CrossRef]
  18. Kim, Y.D.; Ko, W.J.; Koh, K.S.; Jeon, Y.J.; Kim, S.H. Composition of flavonoids and antioxidative activity from juice of Jeju native Citrus fruits during maturation. Korean J. Nutr. 2009, 42, 278–290. [Google Scholar] [CrossRef] [Green Version]
  19. Xing, T.T.; Zhao, X.J.; Zhang, Y.D.; Li, Y.F. Fast separation and sensitive quantitation of polymethoxylated flavonoids in the peels of Citrus using UPLC-Q-TOF-MS. J. Agric. Food Chem. 2017, 65, 2615–2627. [Google Scholar] [CrossRef]
  20. Uckoo, R.M.; Jayaprakasha, G.K.; Patil, B.S. Chromatographic techniques for the separation of polymethoxyflavones from citrus. ACS Symp. Ser. 2012, 1093, 3–15. [Google Scholar] [CrossRef]
  21. He, H.; Ling, I. Chemical studies of a Chinese traditional drug fingered citron (Citrus medica L. var sarcodactylis (Noot.) Swingle. Acta Pharm. Sin. 1985, 20, 433–435. [Google Scholar]
  22. Tsukayama, M.; Sasaki, T.; Yamamoto, K.; Kawamura, Y.; Ichikawa, R. Microwave-Assisted extraction and methylation of useful flavones from waste peels of Citrus sudachi. Nippon. Shokuhin Kagaku Kogaku Kaishi 2010, 57, 427–433. [Google Scholar] [CrossRef] [Green Version]
  23. Chkhikvishvili, I.D.; Kurkin, V.A.; Pervykh, L.N. Flavonoids of the peel of the fruit of Citrus unshiu. Chem. Nat. Compd. 1994, 30, 765–766. [Google Scholar] [CrossRef]
  24. Nakasugi, T.; Nakashima, M.; Komai, K. Antimutagens in Citrus sudachi Hort. et Shirai. Nat. Med. 2000, 54, 26–28. [Google Scholar]
  25. Mizuno, M.; Iinuma, M.; Ohara, M.; Tanaka, T.; Iwamasa, M. Chemotaxonomy of the genus Citrus based on polymethoxyflavones. Chem. Pharm. Bull. 1991, 39, 945–949. [Google Scholar] [CrossRef]
  26. Iinuma, M.; Matsuura, S.; Kurogochi, K.; Tanaka, T. Studies on the constituents of useful plants. V. Multisubstituted flavones in the fruit peel of Citrus reticulata and their examination by Gas-Liquid Chromatography. Chem. Pharm. Bull. 1980, 28, 717–722. [Google Scholar] [CrossRef] [Green Version]
  27. Zhao, X.J.; Xing, T.T.; Li, Y.F.; Jiao, B.N. Analysis of phytochemical contributors to antioxidant capacity of the peel of Chinese mandarin and orange varieties. Int. J. Food Sci. Nutr. 2019, 70, 825–833. [Google Scholar] [CrossRef]
  28. Ding, Y.-Q.; Xiong, Y.; Zhou, B.; Deng, M.-Z.; Deng, K.-Z. Isolation and structural identification of flavonoids from Aurantii Fructus. Zhongguo Zhongyao Zazhi 2015, 40, 2352–2356. [Google Scholar] [CrossRef]
  29. Pinkas, J.; Lavie, D.; Chorin, M. Fungistatic constituents in citrus varieties resistant to the mal-secco disease. Phytochemistry 1968, 7, 169–174. [Google Scholar] [CrossRef]
  30. Gentili, B.; Horowitz, R.M. Flavonoids of Citrus—VII. Tetrahedron 1964, 20, 2313–2318. [Google Scholar] [CrossRef]
  31. Han, S.; Kim, H.M.; Lee, J.M.; Mok, S.Y.; Lee, S. Isolation and identification of polymethoxyflavones from the hybrid Citrus, hallabong. J. Agric. Food Chem. 2010, 58, 9488–9491. [Google Scholar] [CrossRef]
  32. Shin, H.J.; Nam, J.W.; Yoon, U.J.; Han, A.R.; Seo, E.K. Identification of three new flavonoids from the peels of Citrus unshiu. Helv. Chim. Acta 2012, 95, 240–245. [Google Scholar] [CrossRef]
  33. Rizzi, G.P.; Boeing, S.S. Mass spectral analysis of some naturally occurring polymethoxyflavones. J. Agric. Food Chem. 1984, 32, 551–555. [Google Scholar] [CrossRef]
  34. Li, S.; Lo, C.Y.; Ho, C.T. Hydroxylated polymethoxyflavones and methylated flavonoids in sweet orange (Citrus sinensis) peel. J. Agric. Food Chem. 2006, 54, 4176–4185. [Google Scholar] [CrossRef]
  35. Vogt, T.; Proksch, P.; Gülz, P.G.; Wollenweber, E. Rare 6- and 8-O-methylated epicuticular flavonols from two Cistus species. Phytochemistry 1987, 26, 1027–1030. [Google Scholar] [CrossRef]
  36. Uckoo, R.M.; Jayaprakasha, G.K.; Vikram, A.; Patil, B.S. Polymethoxyflavones isolated from the Peel of Miaray mandarin (Citrus miaray) have biofilm inhibitory activity in Vibrio harveyi. J. Agric. Food Chem. 2015, 63, 7180–7189. [Google Scholar] [CrossRef] [PubMed]
  37. Yang, W.; Chen, H.; Bao-Jin, Y.; Zhang, P.; Yan, Z. Study on the active constituents of fructus aurantii. J. Chin. Med. Mater. 2008, 32, 15–22. [Google Scholar]
  38. Mizuno, M.; Matoba, Y.; Tanaka, T.; Tachibana, H.; Iinuma, M.; Iwamasa, M. Two new flavones in Citrus reticulata. J. Nat. Prod. 1987, 50, 751–753. [Google Scholar] [CrossRef]
  39. Berti, G.; Livi, O.; Segnini, D.; Cavero, I. Determination of constitution and synthesis of a new flavone from Cistus monspeliensis L. Tetrahedron 1967, 23, 2295–2300. [Google Scholar] [CrossRef]
  40. Eldomiaty, M.; Abdelaal, M.; Elshafae, A. On the chemical, botanical, and chemotaxonomical evaluation of the genus Citrus. Part I. Thyroid. Off. J. Am. Thyroid. Assoc. 1997, 21, 469–470. [Google Scholar]
  41. Abdel Alim, M.A.; El-Hamouly, W.S.; Aboutabl, E.A.; Hetta, M.H. Polymethoxylated flavones from Citrus reticulata Blanco cuv. Baladi. Egypt. J. Pharm. Sci. 1998, 38, 71–78. [Google Scholar]
  42. Luo, Y.; Zeng, W.; Huang, K.-E.; Li, D.-X.; Chen, W.; Yu, X.-Q.; Ke, X.-H. Discrimination of Citrus reticulata Blanco and Citrus reticulata ‘Chachi’ as well as the Citrus reticulata ‘Chachi’ within different storage years using ultra high performance liquid chromatography quadrupole/time-of-flight mass spectrometry based metabol. J. Pharm. Biomed. Anal. 2019, 171, 218–231. [Google Scholar] [CrossRef] [PubMed]
  43. Sarin, P.S.; Eshars, T.R. New components from Citrus aurantium. Tetrahedron 1960, 8, 64–66. [Google Scholar] [CrossRef]
  44. Chen, J.; Montanari, A.M.; Widmer, W.W. Two new polymethoxylated flavones, a class of compounds with potential anticancer activity, isolated from cold pressed dancy tangerine peel oil solids. J. Agric. Food Chem. 1997, 45, 364–368. [Google Scholar] [CrossRef]
  45. Johann, S.; Smânia, A.; Pizzolatti, M.G.; Schripsema, J.; Braz-Filho, R.; Branco, A. Complete 1H and 13C NMR assignments and antifungal activity of two 8-hydroxy flavonoids in mixture. An. Acad. Bras. Cienc. 2007, 79, 215–222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Zhao, Z.; Wang, W.; Wang, F.; Zhao, K.; Han, Y.; Xu, W.; Tang, L. Effects of Astragaloside IV on heart failure in rats. Chin. Med. 2009, 4, 6. [Google Scholar] [CrossRef] [Green Version]
  47. Wang, J.; Zhu, W.; Yang, S.; Wang, W.; Zhou, Y. Catalytic wet air oxidation of phenol with pelletized ruthenium catalysts. Appl. Catal. B Environ. 2008, 78, 30–37. [Google Scholar] [CrossRef]
  48. Zhong, W.-J.; Luo, Y.-J.; Li, J.; Wu, Y.-P.; Gao, Y.-J.; Luo, H.-J.; Yang, Y.-T.; Jiang, L. Polymethoxylated flavonoids from Citrus reticulata Blanco. Biochem. Syst. Ecol. 2016, 68, 11–14. [Google Scholar] [CrossRef]
  49. Shuji, K.; Fumiak, K. Polymethoxyflavones, their manufacture by extraction, and aqueous organic acid solutions containing them. JP, 2009051738 A, 3 December 2009. [Google Scholar]
  50. Zhao, X.J.; Xing, T.T.; Li, Y.F.; Jiao, B.N.; Jiang, D. Efficient analysis of phytochemical constituents in the peel of Chinese wild citrus Mangshanju (Citrus reticulata Blanco) by ultra high performance liquid chromatography–quadrupole time-of-flight-mass spectrometry. J. Sep. Sci. 2018, 41, 1947–1959. [Google Scholar] [CrossRef]
  51. Hamdan, D.; El-Readi, M.Z.; Tahrani, A.; Herrmann, F.; Kaufmann, D.; Farrag, N.; El-Shazly, A.; Wink, M. Chemical composition and biological activity of Citrus jambhiri Lush. Food Chem. 2011, 127, 394–403. [Google Scholar] [CrossRef]
  52. Li, S.; Pan, M.H.; Lo, C.Y.; Tan, D.; Wang, Y.; Shahidi, F.; Ho, C.T. Chemistry and health effects of polymethoxyflavones and hydroxylated polymethoxyflavones. J. Funct. Foods 2009, 1, 2–12. [Google Scholar] [CrossRef]
  53. Lai, C.S.; Wu, J.C.; Ho, C.T.; Pan, M.H. Disease chemopreventive effects and molecular mechanisms of hydroxylated polymethoxyflavones. BioFactors 2015, 41, 301–313. [Google Scholar] [CrossRef] [PubMed]
  54. Gao, Z.; Gao, W.; Zeng, S.L.; Li, P.; Liu, E.H. Chemical structures, bioactivities and molecular mechanisms of citrus polymethoxyflavones. J. Funct. Foods 2018, 40, 498–509. [Google Scholar] [CrossRef]
  55. Manthey, J.A.; Grohmann, K.; Guthrie, N. Biological properties of citrus flavonoids pertaining to cancer and inflammation. Curr. Med. Chem. 2001, 8, 135–153. [Google Scholar] [CrossRef] [PubMed]
  56. Owis, A.I. Citrus polymethoxyflavones: Biofunctional molecules of therapeutic interest. Stud. Nat. Prod. Chem. 2018, 59, 509–530. [Google Scholar] [CrossRef]
  57. Braidy, N.; Behzad, S.; Habtemariam, S.; Ahmed, T.; Daglia, M.; Nabavi, S.M.; Sobarzo-Sanchez, E.; Nabavi, S.F. Neuroprotective effects of citrus fruit-derived flavonoids, nobiletin and tangeretin in Alzheimer’s and Parkinson’s disease. CNS Neurol. Disord. Drug Targets 2017, 16, 387–397. [Google Scholar] [CrossRef]
  58. Ashrafizadeh, M.; Ahmadi, Z.; Mohammadinejad, R.; Ghasemipour Afshar, E. Tangeretin: A mechanistic review of its pharmacological and therapeutic effects. J. Basic Clin. Physiol. Pharmacol. 2020, 31, 1–12. [Google Scholar] [CrossRef]
  59. Li, S.; Wang, H.; Guo, L.; Zhao, H.; Ho, C.T. Chemistry and bioactivity of nobiletin and its metabolites. J. Funct. Foods 2014, 6, 2–10. [Google Scholar] [CrossRef]
  60. Marzocchella, L.; Fantini, M.; Benvenuto, M.; Masuelli, L.; Tresoldi, I.; Modesti, A.; Bei, R. Dietary Flavonoids: Molecular mechanisms of action as anti- inflammatory agents. Recent Pat. Inflamm. Allergy Drug Discov. 2011, 5, 200–220. [Google Scholar] [CrossRef]
  61. Singh, A.P.; Kandpal, J.B.; Sharma, R.K.; Chitme, H. Nobiletin a biologically active phytoconstituent: Systematic review. J. Biol. Act. Prod. Nat. 2021, 11, 204–211. [Google Scholar] [CrossRef]
  62. Han Jie, L.; Jantan, I.; Yusoff, S.D.; Jalil, J.; Husain, K. Sinensetin: An insight on its pharmacological activities, mechanisms of action and toxicity. Front. Pharmacol. 2021, 11, 1–16. [Google Scholar] [CrossRef]
  63. Nageen, B.; Rasul, A.; Hussain, G.; Shah, M.A.; Anwar, H.; Syed, M.; Uddin, S.; Riaz, A.; Selamoglu, Z. Jaceosidin: A natural flavone with versatile pharmacological anf biological activities. Curr. Pharm. Des. 2021, 27, 456–466. [Google Scholar] [CrossRef] [PubMed]
  64. Shaikh, A.; Makhmoor, T.; Choudhary, M.I. Radical scavenging potential of compounds isolated from—Vitex agnus-castus. Turk. J. Chem. 2010, 34, 119–126. [Google Scholar] [CrossRef]
  65. García, B.F.; Torres, A.; Macías, F.A. Synergy and other interactions between polymethoxyflavones from citrus byproducts. Molecules 2015, 20, 20079–20106. [Google Scholar] [CrossRef] [Green Version]
  66. Karioti, A.; Milošević-Ifantis, T.; Pachopos, N.; Niryiannaki, N.; Hadjipavlou-Litina, D.; Skaltsa, H. Antioxidant, anti-inflammatory potential and chemical constituents of Origanum dubium Boiss., growing wild in Cyprus. J. Enzym. Inhib. Med. Chem. 2015, 30, 38–43. [Google Scholar] [CrossRef] [PubMed]
  67. Kim, M.J.; Han, J.M.; Jin, Y.Y.; Baek, N.I.; Bang, M.H.; Chung, H.G.; Choi, M.S.; Lee, K.T.; Sok, D.E.; Jeong, T.S. In vitro antioxidant and anti-inflammatory activities of jaceosidin from Artemisia princeps Pampanini cv. Sajabal. Arch. Pharm. Res. 2008, 31, 429–437. [Google Scholar] [CrossRef] [PubMed]
  68. Tapia, A.; Rodriguez, J.; Theoduloz, C.; Lopez, S.; Feresin, G.E.; Schmeda-Hirschmann, G. Free radical scavengers and antioxidants from Baccharis grisebachii. J. Ethnopharmacol. 2004, 95, 155–161. [Google Scholar] [CrossRef]
  69. Al-Dabbas, M.M.; Kitahara, K.; Suganuma, T.; Hashimoto, F.; Tadera, K. Antioxidant and α-amylase inhibitory compounds from aerial parts of Varthemia iphionoides Boiss. Biosci. Biotechnol. Biochem. 2006, 70, 2178–2184. [Google Scholar] [CrossRef]
  70. Lesjak, M.; Beara, I.; Simin, N.; Pintać, D.; Majkić, T.; Bekvalac, K.; Orčić, D.; Mimica-Dukić, N. Antioxidant and anti-inflammatory activities of quercetin and its derivatives. J. Funct. Foods 2018, 40, 68–75. [Google Scholar] [CrossRef]
  71. Thao, N.P.; Luyen, B.T.T.; Lee, S.H.; Jang, H.D.; Kim, Y.H. Anti-osteoporotic and antioxidant activities by rhizomes of Kaempferia parviflora Wall. Ex Baker. Nat. Prod. Sci. 2016, 22, 13–19. [Google Scholar] [CrossRef] [Green Version]
  72. Cheng, L.Y.; Liao, H.R.; Chen, L.C.; Wang, S.W.; Kuo, Y.H.; Chung, M.I.; Chen, J.J. Naphthofuranone derivatives and other constituents from Pachira aquatica with inhibitory activity on superoxide anion generation by neutrophils. Fitoterapia 2017, 117, 16–21. [Google Scholar] [CrossRef]
  73. Kanashiro, A.; Kabeya, L.M.; Polizello, A.C.M.; Lopes, N.P.; Lopes, J.L.C.; Lucisano-Valim, Y.M. Inhibitory activity of flavonoids from lychnophora sp. on generation of reactive oxygen species by neutrophils upon stimulation by immune complexes. Phyther. Res. 2004, 18, 61–65. [Google Scholar] [CrossRef] [PubMed]
  74. An, J.Y.; Lee, H.H.; Shin, J.S.; Yoo, H.S.; Park, J.S.; Son, S.H.; Kim, S.W.; Yu, J.; Lee, J.; Lee, K.T.; et al. Identification and structure activity relationship of novel flavone derivatives that inhibit the production of nitric oxide and PGE2 in LPS-induced RAW 264.7 cells. Bioorganic Med. Chem. Lett. 2017, 27, 2613–2616. [Google Scholar] [CrossRef] [PubMed]
  75. Tian, D.; Yang, Y.; Yu, M.; Han, Z.Z.; Wei, M.; Zhang, H.W.; Jia, H.M.; Zou, Z.M. Anti-inflammatory chemical constituents of: Flos Chrysanthemi Indici determined by UPLC-MS/MS integrated with network pharmacology. Food Funct. 2020, 11, 6340–6351. [Google Scholar] [CrossRef] [PubMed]
  76. Kim, M.J.; Lee, H.H.; Jeong, J.W.; Seo, M.J.; Kang, B.W.; Park, J.U.; Kim, K.S.; Cho, Y.S.; Seo, K.-I.; Kim, G.Y. Anti-inflammatory effects of 5-hydroxy-3,6,7,8,3′,4’- hexamethoxyflavone via NF-κB inactivation in lipopolysaccharide-stimulated RAW 264.7 macrophage. Mol. Med. Rep. 2014, 9, 1197–1203. [Google Scholar] [CrossRef] [Green Version]
  77. Maldonado-Rojas, W.; Olivero-Verbel, J. Potential interaction of natural dietary bioactive compounds with COX-2. J. Mol. Graph. Model. 2011, 30, 157–166. [Google Scholar] [CrossRef]
  78. Silva, D.H.S.; Zhang, Y.; Santos, L.A.; Bolzani, V.S.; Nair, M.G. Lipoperoxidation and cyclooxygenases 1 and 2 Inhibitory Compounds from Iryanthera juruensis. J. Agric. Food Chem. 2007, 55, 2569–2574. [Google Scholar] [CrossRef]
  79. Moscatelli, V.; Hnatyszyn, O.; Acevedo, C.; Megías, J.; Alcaraz, M.J.; Ferraro, G. Flavonoids from Artemisia copa with anti-inflammatory activity. Planta Med. 2006, 72, 72–74. [Google Scholar] [CrossRef]
  80. Jang, J.; Hyun, P.K.; Park, H. Structure and antiinflammatory activity relationships of wogonin derivatives. Arch. Pharm. Res. 2005, 28, 877–884. [Google Scholar] [CrossRef]
  81. Alcaraz, M.J.; Ferrandiz, M.L.; Villar, A. Flavonoid inhibition of soybean lipoxygenase. Pharmazie 1986, 41, 299–300. [Google Scholar]
  82. Chan, A.L.F.; Huang, H.L.; Chien, H.C.; Chen, C.M.; Lin, C.N.; Ko, W.C. Inhibitory effects of quercetin derivatives on phosphodiesterase isozymes and high-affinity [3H]-rolipram binding in guinea pig tissues. Investig. New Drugs 2008, 26, 417–424. [Google Scholar] [CrossRef]
  83. Sae-Wong, C.; Matsuda, H.; Tewtrakul, S.; Tansakul, P.; Nakamura, S.; Nomura, Y.; Yoshikawa, M. Suppressive effects of methoxyflavonoids isolated from Kaempferia parviflora on inducible nitric oxide synthase (iNOS) expression in RAW 264.7 cells. J. Ethnopharmacol. 2011, 136, 488–495. [Google Scholar] [CrossRef] [PubMed]
  84. Tsuji, P.A.; Stephenson, K.K.; Wade, K.L.; Liu, H.; Fahey, J.W. Structure-activity analysis of flavonoids: Direct and indirect antioxidant, and antiinflammatory potencies and toxicities. Nutr. Cancer 2013, 65, 1014–1025. [Google Scholar] [CrossRef] [PubMed]
  85. Middleton, E.; Drzewiecki, G.; Tatum, J. The effects of citrus flavonoids on human basophil and neutrophil function. Planta Med. 1987, 53, 325–328. [Google Scholar] [CrossRef] [PubMed]
  86. Park, H.I.; Kim, H.-P. Inhibition of collagenase by anti-inflammatory synthetic flavones. J. Appl. Pharmacol. 2006, 14, 36–39. [Google Scholar]
  87. Jin, M.R.; Xu, H.; Duan, C.H.; Chou, G.X. Two new flavones from Salvia plebeia. Nat. Prod. Res. 2015, 29, 1315–1322. [Google Scholar] [CrossRef]
  88. Tewtrakul, S.; Subhadhirasakul, S.; Karalai, C.; Ponglimanont, C.; Cheenpracha, S. Anti-inflammatory effects of compounds from Kaempferia parviflora and Boesenbergia pandurata. Food Chem. 2009, 115, 534–538. [Google Scholar] [CrossRef]
  89. Puripattanavong, J.; Tewtrakul, S. Anti-allergic and anti-inflammatory compounds from Aglaia andamanica leaves. Songklanakarin J. Sci. Technol. 2015, 37, 37–41. [Google Scholar]
  90. Bas, E.; Recio, M.C.; Giner, R.M.; Manez, S.; Lopez-Gines, C.; Gil-Benso, R.; Rios, J.L. Demethylnobiletin inhibits delayed-type hypersensitivity reactions, human lymphocyte proliferation and cytokine production. Br. J. Pharmacol. 2007, 152, 1272–1282. [Google Scholar] [CrossRef] [Green Version]
  91. Ramırez Apan, A.A.M.T.; Perez-Castorena, A.L.; Romo de Vivar, A. Anti-inflammatory constituents of Mortonia greggii Gray. Z. Naturforsch. 2004, 59, 237–243. [Google Scholar]
  92. Bai, N.; He, K.; Zhou, Z.; Lai, C.S.; Zhang, L.; Quan, Z.; Shao, X.; Pan, M.H.; Ho, C.T. Flavonoids from Rabdosia rubescens exert anti-inflammatory and growth inhibitory effect against human leukemia HL-60 cells. Food Chem. 2010, 122, 831–835. [Google Scholar] [CrossRef]
  93. Pandith, H.; Zhang, X.; Thongpraditchote, S.; Wongkrajang, Y.; Gritsanapan, W.; Baek, S.J. Effect of Siam weed extract and its bioactive component scutellarein tetramethyl ether on anti-inflammatory activity through NF-κB pathway. J. Ethnopharmacol. 2013, 147, 434–441. [Google Scholar] [CrossRef] [PubMed]
  94. Guo, S.; Wu, X.; Zheng, J.; Charoensinphon, N.; Dong, P.; Qiu, P.; Song, M.; Tang, Z.; Xiao, H. Anti-inflammatory effect of xanthomicrol, a major colonic metabolite of 5-demethyltangeretin. Food Funct. 2018, 9, 3104–3113. [Google Scholar] [CrossRef] [PubMed]
  95. Li, S.; Sang, S.; Pan, M.H.; Lai, C.S.; Lo, C.Y.; Yang, C.S.; Ho, C.T. Anti-inflammatory property of the urinary metabolites of nobiletin in mouse. Bioorganic Med. Chem. Lett. 2007, 17, 5177–5181. [Google Scholar] [CrossRef] [PubMed]
  96. Wu, X.; Song, M.; Rakariyatham, K.; Zheng, J.; Wang, M.; Xu, F.; Gao, Z.; Xiao, H. Inhibitory effects of 4′ -demethylnobiletin, a metabolite of nobiletin, on 12-O-tetradecanoylphorbol-13-acetate (TPA)-induced inflammation in mouse ears. J. Agric. Food Chem. 2015, 63, 10921–10927. [Google Scholar] [CrossRef]
  97. Nam, Y.; Choi, M.; Hwang, H.; Lee, M.G.; Kwon, B.M.; Lee, W.H.; Suk, K. Natural flavone jaceosidin is a neuroinflammation inhibitor. Phyther. Res. 2013, 27, 404–411. [Google Scholar] [CrossRef]
  98. Kang, C.H.; Kim, M.J.; Seo, M.J.; Choi, Y.H.; Jo, W.S.; Lee, K.T.; Jeong, Y.K.; Kim, G.Y. 5-Hydroxy-3,6,7,8,3’4’-hexamethoxyflavone inhibits nitric oxide production in lipopolysaccharide-stimulated BV2 microglia via NF-κB suppression and Nrf-2-dependent heme oxygenase-1 induction. Food Chem. Toxicol. 2013, 57, 119–125. [Google Scholar] [CrossRef]
  99. Soo, J.K.; Park, H.; Hyun, P.K. Inhibition of nitric oxide production from lipopolysaccharide-treated RAW 264.7 cells by synthetic flavones:structure-activity relationship and action mechanism. Arch. Pharm. Res. 2004, 27, 937–943. [Google Scholar] [CrossRef]
  100. Duan, L.; Dou, L.L.; Yu, K.Y.; Guo, L.; Bai-Zhong, C.; Li, P.; Liu, E.H. Polymethoxyflavones in peel of Citrus reticulata ‘Chachi’ and their biological activities. Food Chem. 2017, 234, 254–261. [Google Scholar] [CrossRef]
  101. Yuasa, K.; Tada, K.; Harita, G.; Fujimoto, T.; Tsukayama, M.; Tsuji, A. Sudachitin, a polymethoxyflavone from Citrus sudachi, suppresses lipopolysaccharide-induced inflammatory responses in mouse macrophage-like RAW264 cells. Biosci. Biotechnol. Biochem. 2012, 76, 598–600. [Google Scholar] [CrossRef]
  102. Wu, J.; Liu, K.; Shi, X. The anti-inflammatory activity of several flavonoids isolated from Murraya paniculata on murine macrophage cell line and gastric epithelial cell (GES-1). Pharm. Biol. 2016, 54, 868–881. [Google Scholar] [CrossRef] [Green Version]
  103. Zhao, H.Y.; Yang, L.; Wei, J.; Huang, M.; Jiang, J.G. Bioactivity evaluations of ingredients extracted from the flowers of Citrus aurantium L. varamara Engl. Food Chem. 2012, 135, 2175–2181. [Google Scholar] [CrossRef] [PubMed]
  104. Gao, H.Y.; Wang, H.Y.; Li, G.Y.; Du, X.W.; Zhang, X.T.; Han, Y.; Huang, J.; Li, X.X.; Wang, J.H. Constituents from Zhuyeqing Liquor and their inhibitory effects on nitric oxide production. Phytochem. Lett. 2014, 7, 150–155. [Google Scholar] [CrossRef]
  105. Guo, S.; Wu, X.; Zheng, J.; Smith, S.A.; Dong, P.; Xiao, H. Identification of 4′-demethyltangeretin as a Major urinary metabolite of tangeretin in mice and its anti-inflammatory activities. J. Agric. Food Chem. 2021, 69, 4381–4391. [Google Scholar] [CrossRef] [PubMed]
  106. Hyun, J.M.; Jo, Y.J.; Kim, J.E.; An, H.J.; Choi, Y.H.; Hyun, C.G.; Lee, N.H. Tetramethyl-O-scutellarin isolated from peels of immature shiranuhi fruit exhibits anti-inflammatory effects on LPS-induced RAW264.7 cells. Trop. J. Pharm. Res. 2017, 16, 2197–2205. [Google Scholar] [CrossRef] [Green Version]
  107. Jin, S.; Lee, M.Y. Kaempferia parviflora extract as a potential anti-acne agent with anti-inflammatory, sebostatic and anti-propionibacterium acnes activity. Int. J. Mol. Sci. 2018, 19, 3457. [Google Scholar] [CrossRef] [Green Version]
  108. Khan, M.S.; Halagowder, D.; Devaraj, S.N. Methylated chrysin induces co-ordinated attenuation of the canonical Wnt and NF-kB signaling pathway and upregulates apoptotic gene expression in the early hepatocarcinogenesis rat model. Chem. Biol. Interact. 2011, 193, 12–21. [Google Scholar] [CrossRef]
  109. Weng, Z.; Patel, A.B.; Panagiotidou, S.; Theoharides, T.C. The novel flavone tetramethoxyluteolin is a potent inhibitor of human mast cells. J. Allergy Clin. Immunol. 2015, 135, 1044–1052. [Google Scholar] [CrossRef] [Green Version]
  110. Patel, A.B.; Theoharides, T.C. Methoxyluteolin inhibits neuropeptide-stimulated proinflammatory mediator release via mTOR activation from human mast cells. J. Pharmacol. Exp. Ther. 2017, 361, 462–471. [Google Scholar] [CrossRef] [Green Version]
  111. Taracanova, A.; Tsilioni, I.; Conti, P.; Norwitz, E.R.; Leeman, S.E.; Theoharides, T.C. Substance P and IL-33 administered together stimulate a marked secretion of IL-1β from human mast cells, inhibited by methoxyluteolin. Proc. Natl. Acad. Sci. USA 2018, 115, E9381–E9390. [Google Scholar] [CrossRef] [Green Version]
  112. Wu, X.; Song, M.; Wang, M.; Zheng, J.; Gao, Z.; Xu, F.; Zhang, G.; Xiao, H. Chemopreventive effects of nobiletin and its colonic metabolites on colon carcinogenesis. Mol. Nutr. Food Res. 2015, 59, 2383–2394. [Google Scholar] [CrossRef] [Green Version]
  113. Hu, Y.H.; Liu, J.; Li, H.; Tang, W.; Li, X.W.; Guo, Y.W. Chemical constituents from Citrus changshan-huyou and their anti-inflammatory activities. Chem. Biodivers. 2020, 17, 11–14. [Google Scholar] [CrossRef] [PubMed]
  114. Horigome, S.; Yoshida, I.; Ito, S.; Inohana, S.; Fushimi, K.; Nagai, T.; Yamaguchi, A.; Fujita, K.; Satoyama, T.; Katsuda, S.; et al. Inhibitory effects of Kaempferia parviflora extract on monocyte adhesion and cellular reactive oxygen species production in human umbilical vein endothelial cells. Eur. J. Nutr. 2017, 56, 949–964. [Google Scholar] [CrossRef] [PubMed]
  115. Vinh, L.B.; Jo, S.J.; Nguyen Viet, P.; Gao, D.; Cho, K.W.; Koh, E.J.; Park, S.S.; Kim, Y.R.; Kim, Y.H.; Yang, S.Y. The chemical constituents of ethanolic extract from Stauntonia hexaphylla leaves and their anti-inflammatory effects. Nat. Prod. Res. 2021, 35, 1852–1855. [Google Scholar] [CrossRef]
  116. Chen, W.; Zhao, Y.; Zhou, Z.D.Z.; Zhu, P.; Luo, Y.L.L.Z.X. Bioassay-guided isolation of anti-inflammatory diterpenoids with highly oxygenated substituents from kidney tea (Clerodendranthus spicatus). J. Food Biochem. 2020, 44, 1–10. [Google Scholar] [CrossRef] [PubMed]
  117. Leu, W.J.; Chen, J.C.; Guh, J.H. Extract from Plectranthus amboinicus inhibit maturation and release of interleukin 1β through inhibition of NF-κB nuclear translocation and NLRP3 inflammasome activation. Front. Pharmacol. 2019, 10, 1–12. [Google Scholar] [CrossRef]
  118. Li, J.; Jie, X.; Liang, X.; Chen, Z.; Xie, P.; Pan, X.; Zhou, B.; Li, J. Sinensetin suppresses influenza a virus-triggered inflammation through inhibition of NF-κB and MAPKs signalings. BMC Complement. Med. Ther. 2020, 20, 1–10. [Google Scholar] [CrossRef]
  119. During, A.; Larondelle, Y. The O-methylation of chrysin markedly improves its intestinal anti-inflammatory properties: Structure-activity relationships of flavones. Biochem. Pharmacol. 2013, 86, 1739–1746. [Google Scholar] [CrossRef]
  120. Ongchai, S.; Chiranthanut, N.; Tangyuenyong, S.; Viriyakhasem, N.; Kongdang, P. Kaempferia parviflora extract alleviated rat arthritis, exerted chondroprotective properties in vitro, and reduced expression of genes associated with inflammatory arthritis. Molecules 2021, 26, 1527. [Google Scholar] [CrossRef]
  121. Lin, C.F.; Leu, Y.L.; Al-Suwayeh, S.A.; Ku, M.C.; Hwang, T.L.; Fang, J.Y. Anti-inflammatory activity and percutaneous absorption of quercetin and its polymethoxylated compound and glycosides: The relationships to chemical structures. Eur. J. Pharm. Sci. 2012, 47, 857–864. [Google Scholar] [CrossRef]
  122. Lai, C.S.; Li, S.; Chai, C.Y.; Lo, C.Y.; Ho, C.T.; Wang, Y.J.; Pan, M.H. Inhibitory effect of citrus 5-hydroxy-3,6,7,8,3′,4′-hexamethoxyflavone on 12-O-tetradecanoylphorbol 13-acetate-induced skin inflammation and tumor promotion in mice. Carcinogenesis 2007, 28, 2581–2588. [Google Scholar] [CrossRef] [Green Version]
  123. Kongdang, P.; Jaitham, R.; Thonghoi, S.; Kuensaen, C.; Pradit, W.; Ongchai, S. Ethanolic extract of Kaempferia parviflora interrupts the mechanisms-associated rheumatoid arthritis in SW982 culture model via p38/STAT1 and STAT3 pathways. Phytomedicine 2019, 59, 152755. [Google Scholar] [CrossRef] [PubMed]
  124. Yang, W.L.; Chen, S.Y.; Ho, C.Y.; Yen, G.C. Citrus flavonoids suppress IL-5 and ROS through distinct pathways in PMA/ionomycin-induced EL-4 cells. Food Funct. 2020, 11, 824–833. [Google Scholar] [CrossRef] [PubMed]
  125. Abdel-Lateff, A.; Alarif, W.M.; Algandaby, M.M.; Alburae, N.A.; Abdel-Naim, A.B. Euryops arabicus displays anti-inflammatory activities in experimental models. J. Ethnopharmacol. 2020, 247, 112278. [Google Scholar] [CrossRef]
  126. Horigome, S.; Yoshida, I.; Tsuda, A.; Harada, T.; Yamaguchi, A.; Yamazaki, K.; Inohana, S.; Isagawa, S.; Kibune, N.; Satoyama, T. Identification and evaluation of anti-inflammatory compounds from Kaempferia parviflora. Biosci. Biotechnol. Biochem. 2014, 78, 851–860. [Google Scholar] [CrossRef]
  127. Matsui, T.; Ito, C.; Itoigawa, M.; Okada, T.; Furukawa, H. Effect of natsudaidain isolated from Citrus plants on TNFα and cyclooxygenase-2 expression in RBL-2H3 cells. J. Pharm. Pharmacol. 2008, 61, 109–114. [Google Scholar] [CrossRef]
  128. Tewtrakul, S.; Subhadhirasakul, S. Effects of compounds from Kaempferia parviflora on nitric oxide, prostaglandin E2 and tumor necrosis factor-alpha productions in RAW264.7 macrophage cells. J. Ethnopharmacol. 2008, 120, 81–84. [Google Scholar] [CrossRef]
  129. Wang, Y.; Zang, W.; Ji, S.; Cao, J.; Sun, C. Three polymethoxyflavones purified from ougan (Citrus reticulata Cv. suavissima) inhibited LPS-induced NO elevation in the neuroglia BV-2 cell line via the JAK2/STAT3 pathway. Nutrients 2019, 11, 791. [Google Scholar] [CrossRef] [Green Version]
  130. Adhikari-Devkota, A.; Kurauchi, Y.; Yamada, T.; Katsuki, H.; Watanabe, T.; Devkota, H.P. Anti-neuroinflammatory activities of extract and polymethoxyflavonoids from immature fruit peels of Citrus ‘Hebesu’. J. Food Biochem. 2019, 43, 1–10. [Google Scholar] [CrossRef]
  131. Tran, T.V.A.; Malainer, C.; Schwaiger, S.; Atanasov, A.G.; Heiss, E.H.; Dirsch, V.M.; Stuppner, H. NF-κB inhibitors from Eurycoma Longifolia. J. Nat. Prod. 2014, 77, 483–488. [Google Scholar] [CrossRef]
  132. Yang, T.; Feng, C.; Wang, D.; Qu, Y.; Yang, Y.; Wang, Y.; Sun, Z. Neuroprotective and anti-inflammatory effect of tangeretin against cerebral ischemia-reperfusion injury in rats. Inflammation 2020, 43, 2332–2343. [Google Scholar] [CrossRef]
  133. Zhi, Z.; Tang, X.; Wang, Y.; Chen, R.; Ji, H. Sinensetin attenuates amyloid Beta 25-35-induced oxidative stress, inflammation, and apoptosis in SH-SY5Y cells through the TLR4/NF-κB signaling pathway. Neurochem. Res. 2021, 46, 3012–3024. [Google Scholar] [CrossRef] [PubMed]
  134. Lee, H.; Jang, D.; Jeon, J.; Cho, C.; Choi, S.; Han, S.J.; Oh, E.; Nam, J.; Hum, C.; Yu, P. Seomae mugwort and jaceosidin attenuate osteoarthritic cartilage damage by blocking IκB degradation in mice. J. Cell. Mol. Med. 2020, 46, 8126–8137. [Google Scholar] [CrossRef] [PubMed]
  135. Clavin, M.; Gorzalczany, S.; Macho, A.; Mu, E. Anti-inflammatory activity of flavonoids from Eupatorium arnottianum. J. Ethnopharmacol. 2007, 112, 585–589. [Google Scholar] [CrossRef] [PubMed]
  136. Ahmad, F.; Sirat, H.M.; Jamaludin, F.; Musa, N.; Ali, R.M.; Arbain, D.; Aboul-enein, H.Y. Antimicrobial and anti-inflammatory activities of Piper porphyrophyllum (Fam. Piperaceae). Arab. J. Chem. 2014, 7, 1031–1033. [Google Scholar] [CrossRef] [Green Version]
  137. Panthong, A.; Tassaneeyakul, W.; Kanjanapothi, D.; Tantiwac, P.; Reutrakul, V. Anti-inflammatory activity of 5, 7-dimethoxyflavone. Planta Med. 1989, 55, 133–136. [Google Scholar] [CrossRef] [PubMed]
  138. Panthong, A.; Kanjanapothp, D.; Tuntiwachwutiikul, P.; Pancharoen, O.; Reutrakul, V. Antiinflammatory activity of flavonoids. Phytomedicine 1994, 1, 141–144. [Google Scholar] [CrossRef]
  139. Pelzer, L.E.; Guardia, T.; Juarez, O.; Guerreiro, E. Acute and chronic antiinflammatory effects of plant flavonoids. Il Farm. 1998, 53, 421–424. [Google Scholar] [CrossRef]
  140. Pandurangan, K.; Krishnappan, V. Anti-inflammatory effect of certain dimethoxy flavones. Inflammopharmacology 2015, 23, 307–317. [Google Scholar] [CrossRef]
  141. Reddy, G.B.S.; Udupa, A.I.; Shirwaikar, A.; Srinivasa Aithai, K.; Srinivasa, K.K. Comparison of the anti-inflammatory activity of certain methoxyflavonoids. Fitoterapia 1990, 61, 460–461. [Google Scholar]
  142. Chou, Y.; Lin, Y.; Lin, P.; Tung, Y.; Ho, C.; Pan, M. Dietary 5-demethylnobiletin modulates xenobiotic-metabolizing enzymes and ameliorates colon carcinogenesis in benzo-[a]-pyrene-induced mice. Food Chem. Toxicol. 2021, 155, 112380. [Google Scholar] [CrossRef]
  143. Chang, S.N.; Kim, S.H.; Dey, D.K.; Park, S.M.; Nasif, O.; Bajpai, V.K.; Kang, S.C.; Lee, J.; Park, J.G. 5-O-Demethylnobiletin alleviates CCl4 -induced acute liver injury by equilibrating ROS-mediated apoptosis and autophagy induction. Mol. Sci. 2021, 22, 1083. [Google Scholar] [CrossRef]
  144. Chen, B.; Luo, J.; Han, Y.; Du, H.; Liu, J.; He, W.; Zhu, J.; Xiao, J.; Wang, J.; Cao, Y. Dietary tangeretin alleviated dextran sulfate sodium-induced colitis in mice via inhibiting in fl ammatory response, restoring intestinal barrier function, and modulating gut microbiota. J. Agric. Food Chem. 2021, 69, 7663–7674. [Google Scholar] [CrossRef]
  145. Kang, M.; Kim, S.; Oh, S.Y.; Na, W.; Kang, Y. Tangeretin ameliorates glucose-induced podocyte injury through blocking epithelial to mesenchymal transition caused by oxidative stress and hypoxia. J. Mol. Sci. 2020, 21, 8577. [Google Scholar] [CrossRef]
  146. Ohyama, Y.; Ito, J.; Kitano, V.J.; Shimada, J.; Hakeda, Y. The polymethoxy flavonoid sudachitin suppresses inflammatory bone destruction by directly inhibiting osteoclastogenesis due to reduced ROS production and MAPK activation in osteoclast precursors. PLoS ONE 2018, 13, 1–17. [Google Scholar] [CrossRef] [Green Version]
  147. Min, S.W.; Kim, N.J.; Baek, N.I.; Kim, D.H. Inhibitory effect of eupatilin and jaceosidin isolated from Artemisia princeps on carrageenan-induced inflammation in mice. J. Ethnopharmacol. 2009, 125, 497–500. [Google Scholar] [CrossRef]
  148. Schinella, G.R.; Giner, R.M.; Del Carmen Recio, M.; De Buschiazzo, P.M.; Ríos, J.L.; Máñez, S. Anti-inflammatory effects of South American Tanacetum vulgare. J. Pharm. Pharmacol. 1998, 50, 1069–1074. [Google Scholar] [CrossRef]
  149. Küpeli, E.; Şahin, F.P.; Yeşilada, E.; Çaliş, I.; Ezer, N. In vivo anti-inflammatory and antinociceptive activity evaluation of phenolic compounds from Sideritis stricta. Z. Naturforsch. Sect. C J. Biosci. 2007, 62, 519–525. [Google Scholar] [CrossRef] [Green Version]
  150. Manthey, J.A.; Bendele, P. Anti-inflammatory activity of an orange peel polymethoxylated flavone, 3′,4′,3,5,6,7,8-heptamethoxyflavone, in the rat carrageenan/paw edema and mouse lipopolysaccharide-challenge assays. J. Agric. Food Chem. 2008, 56, 9399–9403. [Google Scholar] [CrossRef]
  151. Apaza, T.L.; Serban, A.M.; Cabanillas, A.H.; Villacampa, A.; Rumbero, A. Flavonoids of Tripodanthus acutifolius inhibit TNF–α production in LPS–activated THP–1 and B16–F10 cells. J. Ethnopharmacol. 2019, 242, 112036. [Google Scholar] [CrossRef]
  152. Li, W.; Wang, X.; Zhi, W.; Zhang, H.; He, Z.; Wang, Y.; Liu, F.; Niu, X.; Zhang, X. The gastroprotective effect of nobiletin against ethanol-induced acute gastric lesions in mice: Impact on oxidative stress and inflammation. Immunopharmacol. Immunotoxicol. 2017, 39, 354–363. [Google Scholar] [CrossRef]
  153. Wu, X.; Song, M.; Rakariyatham, K.; Zheng, J.; Guo, S.; Tang, Z.; Zhou, S.; Xiao, H. Anti-inflammatory effects of 4’-demethylnobiletin, a major metabolite of nobiletin. J. Funct. Foods 2015, 19, 278–287. [Google Scholar] [CrossRef] [PubMed]
  154. Bei, D.; An, G. Pharmacokinetics and tissue distribution of 5,7-dimethoxyflavone in mice following single dose oral administration. J. Pharm. Biomed. Anal. 2016, 119, 65–70. [Google Scholar] [CrossRef]
  155. Ji, H.Y.; Kim, S.Y.; Kim, D.K.; Jeong, J.H.; Lee, H.S. Effects of eupatilin and jaceosidin on cytochrome P450 enzyme activities in human liver microsomes. Molecules 2010, 15, 6466–6474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Song, W.Y.; Ji, H.Y.; Baek, N.I.; Jeong, T.S.; Lee, H.S. In Vitro metabolism of jaceosidin and characterization of cytochrome P450 and UDP-glucuronosyltransferase enzymes in human liver microsomes. Arch. Pharm. Res. 2010, 33, 1985–1996. [Google Scholar] [CrossRef]
  157. Wei, G.J.; Hwang, L.S.; Tsai, C.L. Absolute bioavailability, pharmacokinetics and excretion of 5,7,3’,4’-tetramethoxyflavone in rats. J. Funct. Foods 2014, 7, 136–141. [Google Scholar] [CrossRef]
  158. Lu, W.; Sheen, J.; Hwang, L.S.; Wei, G. Identification of 5,7,3′,4′-tetramethoxy flavone metabolites in rat urine by the isotope-labeling method and ultrahigh-performance liquid chromatography—Electrospray ionization—Mass spectrometry. J. Agric. Food Chem. 2012, 60, 8123–8128. [Google Scholar] [CrossRef]
  159. Furukawa, Y.; Okuyama, S.; Amakura, Y.; Sawamoto, A.; Nakajima, M.; Yoshimura, M.; Igase, M.; Fukuda, N.; Tamai, T.; Yoshida, T. Isolation and characterization of neuroprotective components from citrus peel and their application as functional food. Chem. Pharm. Bull. 2021, 69, 2–10. [Google Scholar] [CrossRef]
  160. Moon, B.H.; Lee, Y.; Ahn, J.H.; Lim, Y. Complete assignment of 1H and 13C NMR data of dihydroxyflavone derivatives. Magn. Reson. Chem. 2006, 44, 99–101. [Google Scholar] [CrossRef]
  161. Lee, S.; Moon, B.H.; Park, Y.; Lee, E.; Hong, S.; Lim, Y. Methyl substitution effects on 1H and 13C NMR data of methoxyflavones. Bull. Korean Chem. Soc. 2008, 29, 1793–1796. [Google Scholar] [CrossRef] [Green Version]
  162. Mayer, R. A β-hydroxychalcone from Leptospermum scoparium. Planta Med. 1993, 59, 269–271. [Google Scholar] [CrossRef]
  163. Das, B.; Chakravarty, A.K. Three flavone glycosides from Gelonium multiflorum. Phytochemistry 1993, 33, 493–496. [Google Scholar] [CrossRef]
  164. Horie, T.; Ohtsuru, Y.; Shibata, K.; Yamashita, K.; Tsukayama, M.; Kawamura, Y. 13C NMR spectral assignment of the A-ring of polyoxygenated flavones. Phytochemistry 1998, 47, 865–874. [Google Scholar] [CrossRef]
  165. Hajdú, Z.; Martins, A.; Orbán-Gyapai, O.; Forgo, P.; Jedlinszki, N.; Máthé, I.; Hohmann, J. Xanthine oxidase-inhibitory activity and antioxidant properties of the methanol extract and flavonoids of Artemisia asiatica. Rec. Nat. Prod. 2014, 8, 299–302. [Google Scholar]
  166. Bovicelli, P.; D’Angelo, V.; Collalto, D.; Verzina, A.; D’Antona, N.; Lambusta, D. Efficient synthesis of polyoxygenated flavones from naturally occurring flavanones. J. Pharm. Pharmacol. 2007, 59, 1697–1701. [Google Scholar] [CrossRef]
  167. Wenkert, E.; Gottlieb, H.E. Carbon-13 nuclear magnetic resonance spectroscopy of flavonoid and isoflavonoid compounds. Phytochemistry 1977, 16, 1811–1816. [Google Scholar] [CrossRef]
  168. Chari, V.M.; Ilyas, M.; Wagner, H.; Neszmélyi, A.; Fa-Ching, C.; Li-Kuang, C.; Yu-Chin, L.; Yu-Meei, L. 13C-NMR spectroscopy of biflavanoids. Phytochemistry 1977, 16, 1273–1278. [Google Scholar] [CrossRef]
  169. Lee, J.I.; Son, H.S.; Park, H. An efficient synthesis of flavones from 2-hydroxybenzoic acids. Bull. Korean Chem. Soc. 2004, 25, 1945–1947. [Google Scholar] [CrossRef]
  170. Wollenweber, E.; Dörr, M.; Rivera, D.; Roitman, J.N. Externally accumulated flavonoids in three mediterranean ononis species. Z. Naturforsch. Sect. C J. Biosci. 2003, 58, 771–775. [Google Scholar] [CrossRef]
  171. Tominaga, H.; Horie, T. Studies of the selective o-alkylation and dealkylation of flavonoids. XV. A convenient synthesis of 3,5,6-trihydroxy-7-methoxyflavones and revised structures of two natural flavones. Bull. Chem. Soc. Jpn. 1993, 66, 2668–2675. [Google Scholar] [CrossRef]
  172. Facundo, V.A.; Morais, S.M.; Braz Filho, R. Constituintes químicos de Ottonia corcovadensis Miq. da floresta Amazônica: Atribuição dos deslocamentos químicos dos átomos de hidrogênio e carbono. Quim. Nova 2004, 27, 79–83. [Google Scholar] [CrossRef]
  173. Yoshida, M.; Fujino, Y.; Doi, T. Synthesis of γ-benzopyranone by TfOH-promoted regioselective cyclization of O-alkynoylphenols. Org. Lett. 2011, 13, 4526–4529. [Google Scholar] [CrossRef]
  174. Akkal, S.; Benayache, F.; Bentamene, A.; Medjroubi, K.; Seguin, E.; Tillequin, F. Flavonoid Aglycones from Centaurea napifolia. Chem. Nat. Compd. 2003, 39, 219–220. [Google Scholar] [CrossRef]
  175. Nawasreh, M.; Zarga, M.A.; Sabri, S.S. New anthranilic acid derivatives from Ononis natrix of Jordanian origin. Alex. J. Pharm. Sci. 2004, 18, 165–170. [Google Scholar]
  176. Datta, B.K.; Datta, S.K.; Rashid, M.A.; Nash, R.J.; Sarker, S.D. A sesquiterpene acid and flavonoids from Polygonum viscosum. Phytochemistry 2000, 54, 201–205. [Google Scholar] [CrossRef]
  177. Dong, H.; Gou, Y.L.; Cao, S.G.; Chen, S.X.; Sim, K.Y.; Goh, S.H.; Kini, R.M. Eicosenones and methylated flavonols from Amomum koenigii. Phytochemistry 1999, 50, 899–902. [Google Scholar] [CrossRef]
  178. Machida, K.; Osawa, K. On the flavonoid constituents from the peels of Citrus hassaku Hort. ex. Tanaka. Chem. Pharm. Bull. 1989, 37, 1092–1094. [Google Scholar] [CrossRef] [Green Version]
  179. Wallet, J.-C.; Gaydou, E.M. 13C NMR spectra of a series of 2’,3’,4’-trimetboxyíiavones. Reinvestigation of signal assignment in 13C NMR spectra of flavones. Ref. Data 1993, 31, 518–519. [Google Scholar]
  180. Rwangabo, P.C.; Claeys, M.; Pieters, L.; Corthout, J.; Vanden Berghe, D.A.; Vlietinck, A.J. Umuhengerin, a new antimicrobially active flavonoid from Lantana trifolia. J. Nat. Prod. 1988, 51, 966–968. [Google Scholar] [CrossRef]
  181. Sugiyama, S.; Umehara, K.; Kuroyanagi, M.; Ueno, A.; Taki, T. Studies on the differentiation inducers of myeloid leukemic cells from Citrus species. Chem. Pharm. Bull. 1993, 41, 714–719. [Google Scholar] [CrossRef] [Green Version]
  182. Okuno, Y.; Miyazawa, M. Biotransformation of nobiletin by Aspergillus niger and the antimutagenic activity of a metabolite, 4′-hydroxy-5,6,7,8,3′-pentamethoxyflavone. J. Nat. Prod. 2004, 67, 1876–1878. [Google Scholar] [CrossRef]
  183. Ferracin, R.J.; Silva, M.; Fernandes, J.B.; Vieira, P.C. Flavonoids from the fruits of Murraya paniculata. Phytochemistry 1998, 41, 393–396. [Google Scholar] [CrossRef]
  184. Sukari, M.A.; Azziz, S.S.S.A.; Rahmani, M.; Ali, A.M.; Aimi, N.; Kitajima, M. Polysubstituted flavonoids from the leaves of Murraya paniculata (Rutaceae). Nat. Prod. Sci. 2003, 9, 56–59. [Google Scholar]
  185. Wang, Q.; Wu, Z.; Liu, L.; Zou, L.; Luo, M. Synthesis of citrus bioactive polymethoxyflavonoids and flavonoid glucosides. Chin. J. Org. Chem. 2010, 30, 1682–1688. [Google Scholar]
  186. Chen, J.; Montanari, A.M. Isolation and identification of new polymethoxyflavonoids from Dancy tangerine leaves. J. Agric. Food Chem. 1998, 46, 1235–1238. [Google Scholar] [CrossRef]
  187. Baek, J.; Lee, H.L.; Kang, K.S.; Kim, K.H. Chemical constituents from the fruit of Citrus unshiu and their inhibitory effects on acetylcholinesterase. Maced. J. Chem. Chem. Eng. 2017, 36, 1–8. [Google Scholar] [CrossRef] [Green Version]
  188. Sawabe, A.; Matsubara, Y. Bioactive glycosides in citrus fruit peels. Stud. Plant Sci. 1999, 6, 261–274. [Google Scholar] [CrossRef]
  189. Horie, T.; Tsukayama, M.; Yamada, T.; Miura, I.; Nakayama, M. Three flavone glycosides from Citrus sudachi. Phytochemistry 1986, 25, 2621–2624. [Google Scholar] [CrossRef]
  190. Seok, J.; Warren, H.S.; Cuenca, A.G.; Mindrinos, M.N.; Baker, H.V.; Xu, W.; Richards, D.R.; McDonald-Smith, G.P.; Gao, H.; Hennessy, L. Genomic responses in mouse models poorly mimic human inflammatory diseases. Proc. Natl. Acad. Sci. USA 2013, 110, 3507–3512. [Google Scholar] [CrossRef] [Green Version]
  191. Carresi, C.; Gliozzi, M.; Musolino, V.; Scicchitano, M.; Scarano, F.; Bosco, F.; Nucera, S.; Maiuolo, J.; Macrì, R.; Ruga, S.; et al. The effect of natural antioxidants in the development of metabolic syndrome: Focus on bergamot polyphenolic fraction. Nutrients 2020, 12, 1504. [Google Scholar] [CrossRef]
  192. Mollace, V.; Rosano, G.M.C.; Anker, S.D.; Coats, A.J.S.; Seferovic, P.; Mollace, R.; Tavernese, A.; Gliozzi, M.; Musolino, V.; Carresi, C.; et al. Pathophysiological basis for nutraceutical supplementation in heart failure: A comprehensive review. Nutrients 2021, 13, 257. [Google Scholar] [CrossRef]
  193. Nobile, V.; Pisati, M.; Cestone, E.; Insolia, V.; Zaccaria, V.; Malfa, G.A. Antioxidant efficacy of a standardized red orange (Citrus sinensis (L.) Osbeck) extract in elderly subjects: A randomized, double blind, controlled study. Nutrients 2022, 14, 4235. [Google Scholar] [CrossRef]
  194. PubChem. Nobiletin. 2005. Available online: https://pubchem.ncbi.nlm.nih.gov/#query=nobiletin (accessed on 14 December 2022).
  195. European Chemicals Agency. Summary of Classification and Labelling. 2021. Available online: https://echa.europa.eu/information-on-chemicals/cl-inventory-database/-/discli/details/236961 (accessed on 13 December 2022).
Figure 1. Chemical structure of flavone (e.g., luteolin), flavanone (e.g., naringenin), and chalcone (e.g., butein).
Figure 1. Chemical structure of flavone (e.g., luteolin), flavanone (e.g., naringenin), and chalcone (e.g., butein).
Antioxidants 12 00023 g001
Figure 2. Chemical structure of polymethoxyflavones.
Figure 2. Chemical structure of polymethoxyflavones.
Antioxidants 12 00023 g002
Figure 3. Chemical structure of polymethoxyflavanones.
Figure 3. Chemical structure of polymethoxyflavanones.
Antioxidants 12 00023 g003
Figure 4. Chemical structure of polymethoxychalcones.
Figure 4. Chemical structure of polymethoxychalcones.
Antioxidants 12 00023 g004
Figure 5. Glycosylated flavones (8091) from Citrus trees.
Figure 5. Glycosylated flavones (8091) from Citrus trees.
Antioxidants 12 00023 g005aAntioxidants 12 00023 g005b
Table 4. Radical scavenging activity of PMFs and HPMFs from Citrus genus.
Table 4. Radical scavenging activity of PMFs and HPMFs from Citrus genus.
CompoundType of TestResult (μM)References
30DPPH (% inhib.)24.27 at 1000 μM[64]
3511.44 at 25 μM[65]
26, 44, 45, 48, 52, 63, 712–6 at 25 μM
203.5 at 100 μM[66]
920 at 100 μM[67]
1025 at 300 μM[68]
340 at 300 μM
5910 at 500 μM[69]
43DPPH (IC50 mg/mL)0.06[70]
670.736[51]
520.752
633.56
91.15[68]
9Inhibition % of lipoperoxidation in erythrocytes71 at 300 μM[68]
1064 at 300 μM
3428 at 300 μM
5Reducing capacity Copper(I) ions μM (CUPRAC)0.08 at 10 μM[71]
120.29 at 1 μM
140.17 at 10 μM
331.05 at 10 μM
430 at 10 μM
5Peroxyl radical-scavenging capacity (ORAC) µM of Trolox equivalent1.26 at 10 μM
121.11 at 10 μM
141.51 at 10 μM
332.66 at 10 μM
430.79 at 10 μM
10Superoxide anion scavenging (%)60 at 150 μM[68]
3429 at 150 μM
31Superoxide anion scavenging (IC50)>100 μM[72]
5028.8 μM
9% LDL-oxidation inhibition (TBARS)65 at 20 μM[67]
Table 5. Inhibition of PGE2 production.
Table 5. Inhibition of PGE2 production.
CompoundTest TypeResultReferences
5IC50 (μM) on RAW264.7 LPS-activated cells7.43[74]
273.13
9PGE2 production in RAW264.7 LPS-activated cells290 pg/mL at 20 μM[75]
671500 pg/mL at 40 μM[76]
43IC50 (μM) on human platelets200[70]
Table 6. Enzymatic activity of PMFs and HPMFs.
Table 6. Enzymatic activity of PMFs and HPMFs.
CompoundType of TestResultReferences
6COX-1 inhibition (%)63 at 350 μM[78]
935 at 10 μM[79]
6COX-2 inhibition (%)30 at 350 μM[78]
4612 at 10 μM[80]
6565 at 10 μM
5TRAP activity in RANKL-induced osteoclastic RAW 264.7 cells expressed as % respect to control (Trolox)200[71]
12290
14240
33400
43200
10Xanthine oxidase activity inhibition (%)28 at 150 μM[68]
340 at 150 μM
52Lipoxygenase activity inhibition (%)69.0 at 10 μM[81]
33
43
Phosphodiesterase 1 (IC50, μM)>100[82]
33
43
Phosphodiesterase 2 (IC50, μM)9.2
33
43
Phosphodiesterase 3 (IC50, μM)>100
33
43
Phosphodiesterase 4 (IC50, μM)7.8
33
43
Phosphodiesterase 5 (IC50, μM)22.8
14AKT1 (PKB) inhib. % with ATP 75 μM7 at 10 μM[83]
BTK inhib. % with ATP 36 μM7 at 10 μM
CHUK (IKKα). Inhib. % with ATP 9 μM15 at 10 μM
IGF1R Inhib. % with ATP 9 μM7 at 10 μM
IKBKB (IKKβ) Inhib. % with ATP 5 μM11 at 100 μM
IKBKE (IKKe) Inhib. % with ATP 16 μM2 at 10 μM
IRAK4 Inhib. % with ATP 34 μM3 at 10 μM
JAK1 Inhib. % with ATP 87 μM−10 at 10 μM
JAK2 Inhib. % with ATP 31 μM−7 at 10 μM
MAP4K5 (KHS1) Inhib. % with ATP 55 μM5 at 10 μM
MAPK1 (ERK2) Inhib. % with ATP 100 μM−5 at 100 μM
MAPK3 (ERK1) Inhib. % with ATP 45 μM3 at 10 μM
MAPK8 (JNK1) Inhib. % with ATP 100 μM−5 at 10 μM
MAPK9 (JNK2) Inhib. % with ATP 100 μM5 at 10 μM
MAPK10 (JNK3) Inhib. % with ATP 100 μM3 at 10 μM
MAPKAPK2 Inhib. % with ATP 3 μM2 at 10 μM
NEK1 Inhib. % with ATP 119 μM5 at 10 μM
NEK2 Inhib. % with ATP 150 μM1 at 10 μM
PDK 1 Direct Inhib. % with ATP 27 μM7 at 10 μM
PRKACA (PKA) Inhib. % with ATP 4 μM1 at 10 μM
SYK Inhib. % with ATP 25 μM35 at 10 μM
TBK1 Inhib. % with ATP 31 μM17 at 10 μM
ZAP70 Inhib. % with ATP 2 μM3 at 10 μM
3NAD(P)H:quinoneoxidoreductase 1 (NQO1): = Concentration required to double the specific activity (μM)18.0[84]
40.085
50.2
2712.5
4IC50 inhib. of inducible NO-syntase on RAW264.7 LPS-activated cells (μM)5[84]
52.5
Table 7. Inhibition of NO production.
Table 7. Inhibition of NO production.
CompoundTest TypeResultReferences
5IC50 (μM) on RAW264.7 LPS-activated cells33.2[74]
5.1[83]
914.3[87]
1288.5[88]
60[83]
144.6
278.7
32.5[74]
33>100[89]
16.1[88]
66[83]
4396
527.53[90]
33% Inhibition on murine Mϕ LPS-activated14.2 at 100 μM[91]
418.96 at 50 μM[92]
41Nitrite production (μM) in Raw 246.7 LPS-activated cells19 at 50 μM[92]
253.5 at 50 μM[93]
203 at 30 μM[94]
3411 at 2 μM
5727 at 30 μM[95]
589 at 30 μM
<2 at 30 μM[96]
6345 at 30 μM[95]
678 at 40 μM[76]
919 at 20 μM[67]
21 at 20 μM[75]
6 at 10 μM[79]
4 at 50 μM[97]
9Nitrite production (μM) in primary microglia1 at 50 μM[97]
Nitrite production (μM) in HAPI microglial cells3 at 50 μM
Nitrite production (μM) in primary astrocytes20 at 50 μM
9Nitrite production (μM) in BV2 microglia12 at 30 μM[97]
675 at 30 μM[98]
1% Inhibition on RAW264.7 LPS-activated cells23 at 50 μM[99]
20 at 50 μM
7440 at 50 μM
49 at 200 μM[100]
972.5 at 50 μM[87]
1133 at 50 μM[99]
1332 at50 μM
2250 at 30 μM[101]
2770 at 3 μM[102]
3590 at 30 μM[103]
4473 at 200 μM[100]
42 at 50 μM[104]
4575 at 200 μM[100]
5248 at 200 μM
55 at 3 μM[102]
40 at 50 μM[104]
5527.9 at 50 μM
6347 at 200 μM[100]
42 at 50 μM[104]
7149 at 200 μM[100]
3895 at 50 μM[105]
2550 at 50 μM[106]
Table 8. Effect of TNF-α release by PMFs and HPMFs.
Table 8. Effect of TNF-α release by PMFs and HPMFs.
CompoundTest TypeResult *References
14IC50 (μM) on RAW264.7 LPS-activated cells206[83]
27292
25% production in RAW264.7 LPS-activated cells65 at 100 μM[106]
2248 at 30 μM[101]
* Data not cited within in the text.
Table 9. Chemical structure of polymethoxyflavones (171) isolated in Citrus genus.
Table 9. Chemical structure of polymethoxyflavones (171) isolated in Citrus genus.
No.NameR1R2R3R4R5R6R7R8R9R10
12′,3′-DimethoxyflavoneHHHHHOMeOMeHHH
22′,4′-DimethoxyflavoneHHHHHOMeHOMeHH
33′,4′-DimethoxyflavoneHHHHHHOMeOMeHH
45,4′-DimethoxyflavoneHOMeHHHHHOMeHH
55,7-Dimethoxyflavone
(Chrysin dimethyl ether)
HOMeHOMeHHHHHH
67,4′-DimethoxyflavoneHHHOMeHHHOMeHH
75,6-Dihydroxy-7,4′-dimethoxyflavone
(Ladanein)
HOHOHOMeHHHOMeHH
83,5,6-Trihydroxy-7,4′-dimethoxyflavoneOHOHOHOMeHHHOMeHH
95,7,4′-Trihydroxy-6,3′-dimethoxyflavone
(Jaceosidin)
HOHOMeOHHHOMeOHHH
105,7,4′-Trihydroxy-6,8-dimethoxyflavone
(Demethoxysudachitin)
HOHOMeOHOMeHHOMeHH
112′,3′,4′-TrimethoxyflavoneHHHHHOMeOMeOMeHH
123,5,7-Trimethoxyflavone
(Galangin trimethyl ether)
OMeOMeHOMeHHHHHH
133′,4′,5′-TrimethoxyflavoneHHHHHHOMeOMeOMeH
145,7,4′-Trimethoxyflavone
(Apigenin trimethyl ether)
HOMeHOMeHHHOMeHH
155,3′,4′-TrimethoxyflavoneHOMeHHHHOMeOMeHH
166,2′,3′-TrimethoxyflavoneHHOMeHHOMeOMeHHH
175-Hydroxy-6,7,4′-trimethoxyflavone
(Salvigenin)
HOHOMeOMeHHHOMeHH
185-Hydroxy-7,8,4′-trimethoxyflavone
(Isoscutellarein 7,8,4′-trimethyl ether)
HOHHOMeOMeHHOMeHH
198-Hydroxy-5,7,4′-trimethoxyflavoneHOMeHOMeOHHHOMeHH
205,4′-Dihydroxy-6,7,8-trimethoxyflavone
(Xanthomicrol)
HOHOMeOMeOMeHHOHHH
213,5,6-Trihydroxy-7,3′,4′-trimethoxyflavoneOHOHOHOMeHHOMeOMeHH
225,7,4′-Trihydroxy-6,8,3′-
trimethoxyflavone
HOHOMeOHOMeHOMeOMeHH
233,5,7,4′-Tetrahydroxy-6,8,3′-trimethoxyflavone
(Limocitrol)
OHOHOMeOHOMeHOMeOHHH
243,6,7,4′-TetramethoxyflavoneOMeHOMeOMeHHHOMeHH
255,6,7,4′-Tetramethoxyflavone
(Scutellarein tetramethyl ether)
HOMeOMeOMeHHHOMeHH
265,7,8,4′-Tetramethoxyflavone
(6-Demethoxytangeretin)
HOMeHOMeOMeHHOMeHH
275,7,3′,4′-Tetramethoxyflavone
(Luteolin tetramethyl ether)
HOMeHOMeHHOMeOMeHH
286,7,8,4′-TetramethoxyflavoneHHOMeOMeOMeHHOMeHH
297,3′,4′,5′-TetramethoxyflavoneHHHOMeHHOMeOMeOMeH
303-Hydroxy-5,6,7,4′-tetramethoxyflavone
(Eupatorin-5-methylether)
OHOMeOMeOMeHHHOMeHH
315-Hydroxy-3,6,7,4′-tetramethoxyflavone
(Penduletin 4′-methyl ether)
OMeOHOMeOMeHHHOMeHH
325-Hydroxy-3,7,8,4′-tetramethoxyflavoneOMeOHHOMeOMeHHOMeHH
335-Hydroxy-3,7,3′,4′-tetramethoxyflavone
(Retusin)
OMeOHHOMeHHOMeOMeHH
345-Hydroxy-6,7,8,4′-tetramethoxyflavone
(Gardenin B)
HOHOMeOMeOMeHHOMeHH
355-Hydroxy-6,7,3′,4′-tetramethoxyflavone
(5-Desmethylsinensetin)
HOHOMeOMeHHOMeOMeHH
366-Hydroxy-5,7,8,4′-tetramethoxyflavoneHOMeOHOMeOMeHHOMeHH
377-Hydroxy-5,6,8,4′-tetramethoxyflavoneHOMeOMeOHOMeHHOMeHH
384′-Hydroxy-5,6,7,8-tetramethoxyflavoneHOMeOMeOMeOMeHHOHHH
395,7-Dihydroxy-6,8,3′,4′-tetramethoxyflavone
(Hymenoxin)
HOHOMeOHOMeHOMeOMeHH
405,3′-Dihydroxy-3,7,4′,5′-tetramethoxyflavoneOMeOHHOMeHHOHOMeOMeH
415,4′-Dihydroxy-6,7,8,3′-tetramethoxyflavone
(8-Methoxycirsilineol)
HOHOMeOMeOMeHOMeOHHH
423,5,6,8,4′-PentamethoxyflavoneOMeOMeOMeHOMeHHOMeHH
433,5,7,3′,4′-Pentamethoxyflavone
(Quercetin pentamethyl ether)
OMeOMeHOMeHHOMeOMeHH
445,6,7,8,4′-Pentamethoxyflavone
(Tangeretin)
HOMeOMeOMeOMeHHOMeHH
455,6,7,3′,4′-Pentamethoxyflavone
(Sinensitin)
HOMeOMeOMeHHOMeOMeHH
465,7,8,3′,4′-Pentamethoxyflavone
(Isosinensetin)
HOMeHOMeOMeHOMeOMeHH
475,7,2′,3′,4′-PentamethoxyflavoneHOMeHOMeHOMeOMeOMeHH
486,7,8,3′,4′-Pentamethoxyflavone
(Demethylnobiletin)
HHOMeOMeOMeHOMeOMeHH
493-Hydroxy-5,6,7,8,4′-pentamethoxyflavone
(3-Demethylnobiletin)
OHOMeOMeOMeOMeHHOMeHH
505-Hydroxy-3,6,7,8,4′-Pentamethoxyflavone
(5-Hydroxyauranetin)
OMeOHOMeOMeOMeHHOMeHH
515-Hydroxy-3,7,8,3′,4′-pentamethoxyflavone
(Gossypetin pentamethylether)
OMeOHHOMeOMeHOMeOMeHH
525-Hydroxy-6,7,8,3′,4′-pentamethoxyflavone
(5-Demethylnobiletin)
HOHOMeOMeOMeHOMeOMeOHH
535-Hydroxy-6,7,3′,4′,5′-pentamethoxyflavone
(Umuhengerin)
HOHOMeOMeHHOMeOMeOMeH
547-Hydroxy-3,5,6,3′,4′-pentamethoxyflavoneOMeOMeOMeOHHHOMeOMeHH
557-Hydroxy-5,6,8,3′,4′-pentamethoxyflavone
(7-Demethylnobiletin)
HOMeOMeOHOMeHOMeOMeHH
568-Hydroxy-3,5,6,7,4′-pentamethoxyflavoneOMeOMeOMeOMeOHHHOMeHH
573′-Hydroxy-5,6,7,8,4′-pentamethoxyflavone
(3-Hydroxytangeretin)
HOMeOMeOMeOMeHOHOMeHH
584′-Hydroxy-5,6,7,8,3′-pentamethoxyflavone
(4′-Demethylnobiletin)
HOMeOMeOMeOMeHOMeOHHH
593,5,6,7,8,4′-Hexamethoxyflavone
(3-Methoxytangeretin)
OMeOMeOMeOMeOMeHHOMeHH
603,5,7,8,2′,5′-HexamethoxyflavoneOMeOMeHOMeOMeOMeHHOMeH
613,5,7,8,3′,4′-Hexamethoxyflavone
(Gossypetin hexamethyl ether)
OMeOMeHOMeOMeHOMeOMeHH
623,6,7,8,2′,5′-HexamethoxyflavoneOMeHOMeOMeOMeOMeHHOMeH
635,6,7,8,3′,4′-Hexamethoxyflavone
(Nobiletin)
HOMeOMeOMeOMeHOMeOMeHH
645,6,7,3′,4′,5′-HexamethoxyflavoneHOMeOMeOMeHHOMeOMeOMeH
655,7,8,3′,4′,5′-Hexamethoxyflavone
(Bannamurpanisin)
HOMeHOMeOMeHOMeOMeOMeH
663-Hydroxy-5,6,7,8,3′,4′-hexamethoxyflavone
(Natsudaidain)
OHOMeOMeOMeOMeHOMeOMeHH
675-Hydroxy-3,6,7,8,3′,4′-hexamethoxyflavoneOMeOHOMeOMeOMeHOMeOMeHH
685-Hydroxy-6,7,8,3′,4′,5′-hexamethoxyflavone
(Gardenin A)
HOHOMeOMeOMeHOMeOMeOMeH
697-Hydroxy-3,5,6,8,3′,4′-hexamethoxyflavoneOMeOMeOMeOHOMeHOMeOMeHH
708-Hydroxy-3,5,6,7,3′,4′-hexamethoxyflavoneOMeOMeOMeOMeOHHOMeOMeHH
713,5,6,7,8,3′,4′-HeptamethoxyflavoneOMeOMeOMeOMeOMeHOMeOMeHH
Table 10. 13C-NMR of polymethoxyflavones’ (171) skeleton isolated in Citrus genus.
Table 10. 13C-NMR of polymethoxyflavones’ (171) skeleton isolated in Citrus genus.
No.C2C3C4C5C6C7C8C9C10C1′C2′C3′C4′C5′C6′SolventReferences
1161.7-177.6125.4124.5134.3118.3155.9123.0111.2147.1153.0111.2120.5116.0DMSO-d6[26]
2160.6110.2177.1124.7125.2134.1118.4155.8123.1112.3159.599.1163.2106.3130.4DMSO-d6[160]
3162.7105.7177.0124.7125.3134.0118.5155.6123.3123.3109.4149.0151.9111.7119.9DMSO-d6[160]
4160.1106.9176.4159.0107.2134.1109.9157.5113.7122.9127.8114.5161.9114.5127.8DMSO-d6[161]
5163.9109.0177.4160.592.8160.896.1159.8109.3131.5125.9128.8131.1128.8125.9CDCl3[162]
6162.3105.2176.3126.1114.5163.7100.9157.4117.1123.3128.0114.5162.0114.5128.0DMSO-d6[161]
7163.1102.9182.0146.1129.9154.290.9149.5105.0122.8127.9114.3162.0114.3127.9DMSO-d6[163]
8146.4135.9176.0144.9129.2154.690.7148.8104.3123.3129.2113.9160.4113.9129.2DMSO-d6[164]
9163.7102.7182.1152.7131.3157.394.3152.4104.0121.5110.1150.7148.0115.7120.3DMSO-d6[165]
10165.0103.6183.7149.8132.4151.3117.0146.7104.6123.4129.2116.5162.0116.5129.2Acetone-d6[166]
11160.9110.0176.6124.4124.0133.8118.0155.9122.8108.0152.0141.9155.6108.0118.0DMSO-d6[26]
12152.0141.4173.6160.595.5163.592.1158.4109.1130.5127.7128.0129.0128.0127.7CDCl3[167]
13162.0106.3176.6124.9124.3133.5118.0155.2123.0126.0104.1152.9140.8152.9104.1DMSO-d6[26]
14160.4106.8176.0160.093.3163.796.2159.3108.5123.2127.6114.4161.9114.4127.6DMSO-d6[168]
15161.0106.4178.3159.7111.1133.6108.6158.2114.5123.9110.1149.2151.8108.0119.7CDCl3[169]
16161.3110.4176.6104.6156.6123.3120.1150.6123.8126.0147.1153.0115.8124.5120.6DMSO-d6[161]
17163.6103.3182.3152.0131.9158.791.6152.7105.1122.7128.3114.6162.4114.6128.3DMSO-d6[34]
18163.9103.8182.7157.595.7158.5128.9149.4104.8123.6128.1114.6162.7114.6128.1CDCl3[164]
19159.6105.9176.2151.894.0152.0127.9146.8108.0123.2127.8114.3161.7114.3127.8DMSO-d6[164]
20164.1102.6182.5145.5135.6152.4132.6145.1106.1121.0128.4116.1161.4116.1128.4DMSO-d6[170]
21146.3136.1176.0144.9129.2154.690.8148.7104.3123.4110.7121.4150.3148.3111.4DMSO-d6[171]
22163.1102.7182.0148.1131.3150.5127.7145.1115.7121.4109.7119.9150.5115.7147.8DMSO-d6[164]
23146.5135.6176.2147.2130.9150.5127.6144.3102.3122.1111.2148.8147.3115.6121.6DMSO-d6[164]
24160.3106.1175.697.3139.8161.8157.4153.9114.5123.0127.8114.5151.6114.5127.8DMSO-d6[31]
25160.3106.1175.6151.6139.7157.497.3153.9112.0123.0127.8114.5161.9114.5127.8DMSO-d6[161]
26159.4106.2175.8155.693.5156.2129.8150.9107.8123.1127.5114.5161.7114.5127.5DMSO-d6[164]
27160.4107.5177.4160.696.0164.892.8160.0108.9123.7108.3149.0151.6-119.3CDCl3[172]
28161.2107.1177.296.2157.6140.3152.6154.5112.9123.9127.7114.4162.1114.4127.7CDCl3[32]
29162.9107.3177.7127.1114.4164.2100.5157.9117.8127.0103.7153.5141.4153.5103.7CDCl3[173]
30142.6137.6171.0151.0139.3157.596.9153.0110.0123.5128.8114.0160.1114.0128.8DMSO-d6[34]
31155.4137.9178.2151.6131.6158.691.4151.8105.6122.0130.0114.2161.4114.2130.0DMSO-d6[164]
32155.4137.8178.3156.395.7158.2128.3147.8104.5122.2129.8114.3161.4114.3129.8DMSO-d6[164]
33155.5138.3178.1156.497.9165.292.5156.4105.3122.1111.6148.5151.4111.3122.1DMSO-d6[34]
34163.9103.4182.7148.7136.0152.6132.8145.4106.3122.8128.4114.9162.7114.9128.4DMSO-d6[34]
35163.0104.5182.5153.8133.0159.091.0155.0105.5120.1112.2152.4149.5110.0123.9CDCl3[174]
36160.1105.7175.9141.2140.7146.3137.8143.6114.1123.2127.6114.5161.8114.5127.6DMSO-d6[164]
37159.8105.8175.7147.3139.4149.1132.5147.5110.6123.2127.6114.6161.8114.6127.6DMSO-d6[164]
38160.7105.3175.6147.4143.4150.8137.7147.0114.2121.3127.8115.9160.7115.9127.8DMSO-d6[164]
39163.7104.0182.9148.4130.9148.9127.3145.8104.6123.7108.8149.4152.4111.3120.1CDCl3[175]
40155.3139.7178.8162.098.0165.692.2156.8106.1126.0108.6149.2137.8152.0105.1CDCl3[176]
41164.0102.9182.5148.4135.8152.4132.5145.1106.1121.3110.0148.0151.0115.9120.3DMSO-d6[170]
42152.5139.4172.9138.4148.8103.4139.9145.0118.9122.6129.6114.2161.0114.2129.6DMSO-d6[164]
43152.6141.2174.0158.895.8163.992.5161.1109.5123.4111.3148.7150.9110.8121.6CDCl3[177]
44160.4106.1175.8147.6143.6162.1137.8147.6114.7123.1127.8114.3151.0114.3127.8DMSO-d6[31]
45159.6106.7176.1151.7131.8156.594.2155.8111.1123.4108.9149.1151.2112.3119.3DMSO-d6[31]
46160.5107.2177.9152.092.6156.3130.7156.3109.1124.1108.7149.3151.8111.0119.7CDCl3[178]
47160.3111.5175.7158.196.1163.793.2159.8108.2117.8152.1142.2155.8108.2123.9DMSO-d6[179]
48160.3106.4175.797.3139.7157.4151.5153.9111.7123.2109.2149.0151.7112.0119.4DMSO-d6[31]
49142.9137.8171.2147.0143.1150.7137.4146.2112.3123.5128.8114.2160.3114.2128.8DMSO-d6[34]
50155.7137.8178.6148.1135.4152.3132.5144.4106.7122.1129.9114.4161.5114.4129.9DMSO-d6[164]
51155.2138.0178.3156.495.8158.2128.8147.8104.5122.2111.7148.4151.3110.9121.9DMSO-d6[34]
52163.5103.5182.5148.5135.8152.4132.5145.2106.2122.6111.7148.9152.3109.0119.9DMSO-d6[34]
53163.8105.4182.6153.2132.9158.990.7153.1106.3126.5103.9153.7141.6153.7103.9CDCl3[180]
54151.8140.0171.4137.4142.8158.296.1153.7109.8123.8110.6148.9150.5111.0120.8CDCl3[44]
55161.4106.7177.3145.1140.1140.3138.1145.6114.2124.2108.7149.4152.0111.3119.7CDCl3[181]
56143.2137.9171.8147.5143.5151.5137.2146.9111.7123.6129.0144.1160.9114.1129.0CDCl3[45]
57142.8137.7171.2146.9143.0150.7137.4146.1112.2123.5128.7114.1160.2114.1128.7DMSO-d6[164]
58161.2106.6177.4148.9106.6151.4138.0146.9114.8123.5108.2147.7148.4115.0120.2CDCl3[182]
59152.5139.8172.2147.2143.2150.8137.4146.2114.4122.5129.5114.3161.0114.3129.5DMSO-d6[164]
60152.3140.9174.392.3156.3156.4152.5151.0109.4123.6150.9111.0121.8148.7152.3CDCl3[32]
61150.8140.8174.2152.292.4156.4130.4156.3109.4123.6110.9148.7150.9111.0121.8CDCl3[178]
62153.3140.8173.796.0157.7140.2152.4153.6113.1123.4151.0110.9121.7148.7111.4CDCl3[32]
63160.2106.3175.8143.5137.6150.9147.5147.1114.3123.1108.9149.0151.8111.9119.3DMSO-d6[31]
64161.0108.3177.2154.5140.4157.896.3152.6112.9126.9103.4153.6140.9153.6103.4CDCl3[183]
65160.2107.8177.8151.992.3156.6130.5156.3108.7126.6103.1153.4140.7153.4103.1CDCl3[183]
66146.9143.1172.0143.6137.9150.6137.5147.7111.8123.9111.2149.0151.7110.3121.1CDCl3[181]
67155.6138.1178.7148.5135.5152.5132.4144.4106.8122.1110.9148.1151.4111.8122.0DMSO-d6[34]
68163.6104.8183.0149.5136.6153.1132.9145.8107.0126.3103.6153.6141.5153.6103.6CDCl3[184]
69150.7142.9171.8143.6147.7146.9137.4146.9111.7123.9110.6149.0151.7111.3121.1CDCl3[44]
70142.8137.8171.8147.6143.5151.6137.4146.8111.7123.8110.3148.9150.5111.1121.0CDCl3[45]
71150.9140.0172.3143.1137.4151.9138.9147.9114.4122.5110.0148.5154.1111.7121.5DMSO-d6[31]
Table 11. 13C-NMR of polymethoxyflavones’ (171) methoxy groups occurring in Citrus genus.
Table 11. 13C-NMR of polymethoxyflavones’ (171) methoxy groups occurring in Citrus genus.
No.OMe-3OMe-5OMe-6OMe-7OMe-8OMe-2′OMe-3′OMe-4′OMe-5′OMe-6′SolventReferences
1-----60.255.9---DMSO-d6[26]
2-----56.1-55.7--DMSO-d6[160]
3------55.855.7--DMSO-d6[160]
4-56.1-----55.5--DMSO-d6[161]
5-56.5-55.8------CDCl3[162]
6---55.5---56.0--DMSO-d6[161]
7---56.1---55.4--DMSO-d6[163]
8---56.2---55.3--DMSO-d6[164]
9--59.9---55.9---DMSO-d6[165]
10--60.8-61.8-----Acetone-d6[166]
11-----60.960.255.8--DMSO-d6[26]
1259.755.4-56.0------CDCl3[167]
13------56.159.4-56.1DMSO-d6[26]
14-56.1-55.9---55.4--DMSO-d6[168]
15-56.5----56.156.1--CDCl3[169]
16--55.7--60.656.0---DMSO-d6[161]
17--60.656.5---55.6--DMSO-d6[34]
18---56.361.7--55.5--CDCl3[164]
19-56.3-56.2---55.4- DMSO-d6[164]
20--60.561.861.6-----DMSO-d6[170]
21---56.5--55.655.6--DMSO-d6[171]
22--61.0-59.8--56.0--DMSO-d6[164]
23------55.5---DMSO-d6[164]
2461.0-55.561.8---56.4--DMSO-d6[31]
25-61.861.056.4---55.5--DMSO-d6[161]
26-56.3-56.160.9--55.4--DMSO-d6[164]
27-56.0-55.7--55.756.2--CDCl3[172]
28--56.361.662.2--55.5--CDCl3[32]
29---55.8--56.361.056.3-CDCl3[173]
30-61.961.156.5---55.3--DMSO-d6[34]
3159.7-60.056.4---55.4--DMSO-d6[164]
3259.7--56.561.0--55.4--DMSO-d6[164]
3359.8--56.2--55.755.7--DMSO-d6[34]
34--60.762.061.6--55.7--DMSO-d6[34]
35--61.058.5--56.256.1--CDCl3[174]
36-61.7-61.460.8--55.4--DMSO-d6[164]
37-61.761.1-61.1--55.4--DMSO-d6[164]
38-61.361.461.761.8-----DMSO-d6[164]
39--61.0-61.8-56.156.0--CDCl3[175]
4060.3--56.1---61.155.8-CDCl3[176]
41--60.561.861.4-55.8---DMSO-d6[170]
4259.256.956.7-61.0--55.3--DMSO-d6[164]
4359.955.8-55.9--56.156.4--CDCl3[177]
44-62.055.661.961.5--61.6--DMSO-d6[31]
45-61.155.856.5--56.455.9--DMSO-d6[31]
46-56.5-56.361.5-56.156.0--CDCl3[178]
47-56.0-56.1-61.060.556.1--DMSO-d6[179]
48--55.761.855.9-61.056.4--DMSO-d6[31]
49-61.861.461.961.5--55.3--DMSO-d6[34]
5059.7-60.561.861.4--55.4--DMSO-d6[164]
5159.7--56.561.0-55.355.6--DMSO-d6[34]
52--60.561.761.4-55.755.6--DMSO-d6[34]
53--60.856.4--56.560.856.5-CDCl3[180]
5462.361.656.4---56.156.0--CDCl3[44]
55-62.862.1-61.5-56.556.0--CDCl3[181]
5661.962.361.861.6---55.3--CDCl3[45]
57-61.761.361.961.5--55.3--DMSO-d6[164]
58-61.861.961.662.2-56.0---CDCl3[182]
5959.261.761.361.861.5--55.3--DMSO-d6[164]
6059.956.4-56.661.556.0--55.9-CDCl3[32]
6161.456.5-56.459.9-56.055.9--CDCl3[178]
6260.0-56.361.662.256.1--56.0-CDCl3[32]
63-61.955.661.855.7-61.561.4--DMSO-d6[31]
64-61.161.662.2--56.456.456.4-CDCl3[183]
65-61.0-56.561.4-55.956.155.9-CDCl3[183]
66-62.462.061.961.8-56.155.8--CDCl3[181]
6759.7-60.661.861.5-55.455.7--DMSO-d6[34]
68--62.061.161.7-56.361.156.3-CDCl3[184]
6961.662.361.8-61.7-56.055.9--CDCl3[44]
7061.962.361.861.6--55.955.8--CDCl3[45]
7161.563.759.361.661.4-56.355.4--DMSO-d6[31]
Table 12. Chemical structure of polymethoxyflavanones (72–77) isolated in genus Citrus.
Table 12. Chemical structure of polymethoxyflavanones (72–77) isolated in genus Citrus.
No.NameR1R2R3R4R5R6References
725,6,7,4′-TetramethoxyflavanoneOMeOMeOMeHHOMe[34]
735,6,7,8,4′-PentamethoxyflavanoneOMeOMeOMeOMeHOMe[185]
745,6,7,3′,4′-PentamethoxyflavanoneOMeOMeOMeHOMeOMe[100]
755,7,8,3′,4′-PentamethoxyflavanoneOMeHOMeOMeOMeOMe[186]
765-Hydroxy-6,7,8,3′,4′-pentamethoxyflavanone
(5-Demethylcitromitine)
OHOMeOMeOMeOMeOMe[34]
775,6,7,8,3′,4′-Hexamethoxyflavanone
(Citromitin)
OMeOMeOMeOMeOMeOMe[186]
Table 13. 1H-NMR and 13C-NMR of polymethoxyflavanones’ (7277) skeleton isolated in Citrus genus.
Table 13. 1H-NMR and 13C-NMR of polymethoxyflavanones’ (7277) skeleton isolated in Citrus genus.
Compound727374757677
SolventDMSO-d6CDCl3CDCl3CDCl3DMSO-d6CDCl3CDCl3CDCl3
PositionδCδH
(J in Hz)
δCδH
(J in Hz)
δCδH
(J in Hz)
δCδH
(J in Hz)
δCδH
(J in Hz)
δCδH
(J in Hz)
279.05.34
(dd, 13.5, 3.0)
-5.40
(dd, 13.2, 2.8)
79.45.34
(dd, 13.4, 2.7)
79.05.35
(dd, 13.0, 3.0)
78.75.40
(dd, 13.0, 3.0)
78.05.40
(dd, 13.0, 3.0)
345.33.02
(dd, 16.5, 13.5)
2.75
(dd, 16.5, 3.0)
-3.03
(dd, 16.8, 13.2)
2.88
(dd, 16.8, 2.8)
45.53.03
(dd, 16.7, 13.4)
2.76
(dd, 16.7, 2.7)
45.63.02
(dd, 17.5, 13.0)
2.75
(dd, 17.5, 3.0)
42.13.09
(dd, 17.0, 13.0)
2.91
(dd, 17.0, 3.0)
45.63.05
(dd, 17.5, 13.0)
2.84
(dd, 17.5, 3.0)
4189.3---189.4-189.2-198.2-190.2-
5156.7---154.2-156.2-150.8-151.2-
6137.5---137.5-89.5-133.4-139.5-
7160.1---159.4-157.7-154.9-154.7-
897.56.31
(s)
--96.4-132.0-132.8-141.6-
9160.0---159.7-156.8-148.7-150.5-
10109.4---109.1-107.5-103.9-112.0-
1′131.5---131.1-131.6-130.9-131.4-
2′1128.87.39
(d, 9.0)
-7.01
(m)
109.46.99
(d, 1.8)
114.77.02
(d, 2.5)
110.57.00
(d, 2.5)
114.37.02
(d, 2.5)
3′114.56.95
(d, 9.0)
--149.2-147.0-149.1-146.7-
4′154.1---149.4-148.2-150.4-148.1-
5′114.56.95
(d, 9.0)
-6.90
(d, 8.8)
111.26.90
(d, 8.4)
112.16.90
(d, 9.0)
111.66.90
(d, 9.0)
112.16.90
(d, 9.0)
6′128.87.39
(d, 9.0)
-7.01
(m)
118.87.00
(dd, 8.4, 1.8)
118.07.00
(dd, 9.0, 2.5)
119.16.99
(dd, 9.0, 2.5)
118.07.00
(dd, 9.0, 2.5)
OMe-561.93.94
(s)
-3.90
(s)
61.33.91
(s)
56.53.92
(s)
--62.03.90
OMe-661.53.83
(s)
-4.06
(s)
56.13.91
(s)
--60.64.09
(s)
61.74.05
(s)
OMe-756.93.87
(s)
-3.90
(s)
61.63.82
(s)
61.43.82
(s)
61.23.80
(s)
61.83.82
(s)
OMe-8---3.86
(s)
--56.53.95
(s)
61.03.86
(s)
56.13.88
(s)
OMe-3′----55.93.91
(s)
56.13.87
(s)
55.63.91
(s)
60.83.89
(s)
OMe-4′55.93.83
(s)
-3.86
(s)
55.93.91
(s)
56.03.90
(s)
55.63.91
(s)
56.03.85
(s)
Table 14. Chemical structures of polymethoxychalcones (7879) occurring in Citrus genus.
Table 14. Chemical structures of polymethoxychalcones (7879) occurring in Citrus genus.
No.NameR1R2R3R4R5R6R7
782′-Hydroxy-3,4,4′,5′,6′-
pentamethoxychalcone
OHHOMeOMeOMeOMeOMe
792′-Hydroxy-3,4,3′,4′,5′,6′-hexamethoxychalconeOHOMeOMeOMeOMeOMeOMe
Table 15. 1H-NMR and 13C-NMR data of isolated polymethoxychalcones (78-79) in Citrus genus.
Table 15. 1H-NMR and 13C-NMR data of isolated polymethoxychalcones (78-79) in Citrus genus.
Compounds1H-NMR (CDCl3)13C-NMR (DMSO-d6)References
783.71 (s, 3H), 3.82 (s, 3H), 3.83 (s, 3H), 3.85 (s, 6H), 6.38 (s, 1H, H-3′), 7.02 (d, J = 8 Hz, 1H), 7.29 (dd, J = 8 Hz, 2 Hz, 1H), 7.32 (d, J = 2 Hz, 1H), 7.45 (d, J = 16 Hz, 1H), 7.57 (d, J = 16 Hz, 1H), 12.38 (s, 1H, 2′-OH)55.5, 55.6, 56.0, 60.6, 61.5, 96.4, 110.6, 110.8, 111.7, 122.8, 125.0, 127.4, 134.6, 143.9, 149.0, 151.2, 153.0, 157.5, 157.8, 192.6[34]
793.74 (s, 3H), 3.75 (s, 3H), 3.77 (s, 3H), 3.80 (s, 3H), 3.81 (s, 3H), 3.93 (s, 3H), 7.00 (d, J = 9 Hz, 1H,), 7.16 (d, J = 16 Hz, 1H), 7.27 (dd, J = 9 Hz, 3 Hz, 1H), 7.34 (d, J = 3 Hz, 1H), 7.40 (d, J = 16 Hz, 1H), 10.04 (s, 1H, 2′-OH)55.6, 60.8, 60.9, 61.0, 61.5, 110.7, 111.6, 116.7, 123.2, 125.9, 127.2, 137.2, 138.6, 144.8, 145.7, 147.1, 148.9, 149.0, 151.2, 192.9[34]
Table 16. Glycosylated flavones (8083) extracted from Citrus genus.
Table 16. Glycosylated flavones (8083) extracted from Citrus genus.
No.80818283
Ref.[181][181][187][187]
SolventDMSO-d6DMSO-d6CD3OD-d4DMSO-d6
PositionδCδH
(J in Hz)
δCδH
(J in Hz)
δCδH
(J in Hz)
δCδH
(J in Hz)
2151.0-151.0-157.6-156.3-
3146.2-146.2-134.4-133.2-
4172.2-172.2-178.5-177.7-
5143.5-143.5-157.1-156.3-
6137.4-137.4-99.36.29
(brs)
95.76.57
(s)
7151.0-151.0-157.9-158.0-
8135.5-135.5-128.2-128.3-
9148.1-148.1-149.4-148.7-
10114.2-114.2-104.7-104.2-
1′122.7-122.7-122.1-121.2-
2′111.67.82
(d, 2.0)
111.57.83
(d, 2.0)
113.37.97
(brs)
115.27.63
(brs)
3′147.3-147.3-147.5-144.9-
4′153.6-153.6-150.1-147.7-
5′112.57.13
(d, 9.0)
112.57.12
(d, 9.0)
115.26.94
(d, 8.0)
116.06.86
(d, 8.0)
6′121.87.71
(dd, 9.0, 2.0)
121.77.70
(dd, 9.0, 2.0)
123.17.73
(brd, 8.0)
121.77.65
(dd, 2.0, 8.0)
OMe61.9, 61.8, 61.5,
61.4,
55.7,
55.6
4.03 (s)
3.95 (s)
2× 3.86 (s)
3.84 (s)
3.82 (s)
61.9, 61.7, 61.5,
61.4,
55.7,
55.6
4.02 (s)
3.95 (s)
2× 3.85 (s)
3.84 (s)
3.82 (s)
61.2
(8-OMe)

55.8
(3′-OMe)
3.94
(s, 3H, 8-OMe)

3.98
(s, 3H, 3′-OMe)
61.1
(8-OMe)

56.5
(3′-OMe)
3.81
(s, 3H, 8-OMe)

3.90
(s, 3H, 3′-OMe)
1″101.0-101.0-103.05.36 (d, 7.5)100.95. 46 (d, 7.0)
2″74.3-74.6-74.93.53 (m)74.03.10-3.40 (m)
3″74.3-74.3-77.03.39 (m)76.43.10-3.40 (m)
4″70.1-70.2-70.83.33 (m)69.93.10-3.40 (m)
5″76.4-76.4-75.03.50 (m)77.53.08 (m)
6″62.8-63.3-63.64.14
(dd, 1H, 3.0, 11.0)
4.25
(dd, 1H, 1.0, 11.0)
60.93.40
(m, 1H)
3.57
(brd, 1H, 11.5)
1170.3-170.7-171.3---
247.02.14
(d, 1H, 13.0 Hz)
2.09
(d, 1H, 13.0 Hz)
45.32.39 (2H, m)45.22.52 (2H, m)--
368.5-68.7-69.7---
446.71.86
(d, 1H, 15.0 Hz)
2.06
(d, 1H, 15.0 Hz)
45.02.28 (2H, m)45.12.52 (2H, m)--
5175.7-170.0-171.5---
627.90.87 (s, 3H)27.11.02 (s, 3H)26.91.23 (s, 3H)--
5‴−OMe--50.83.48 (s)----
Table 17. Glycosylated flavones (8487) extracted from Citrus genus.
Table 17. Glycosylated flavones (8487) extracted from Citrus genus.
No.84858687
Ref.[188][188][188][189]
SolventCD3OD-d4CD3OD-d4DMSO-d6Pyridine-d5
PositionδH (J in Hz)δH (J in Hz)δCδH (J in Hz)δCδH (J in Hz)δCδH (J in Hz)
2--156.8-163.8-165.0-
3--135.5-104.76.99 (s)103.97.02 (s)
4--170.1-183.8-183.5-
5--147.1-150.0-149.2-
6--133.0-133.2-137.4-
7--150.1-152.8-150.0-
8--131.3-129.4-134.1-
9--144.3-146.6-146.3-
10--106.5-104.4-108.1-
1′--120.6-126.1-122.6-
2′7.87
(d, 2.0)
7.98
(d, 2.0)
113.07.84
(d, 2.0)
111.17.61 (d, 2.2)110.67.67 (d, 2.2)
3′--152.4-150.6-149.6-
4′--148.2-151.1-152.8-
5′6.96
(d, 9.0)
6.93
(d, 8.0)
115.56.97
(d, 8.0)
117.07.68 (d, 8.6)117.37.30 (d, 8.2)
6′7.85
(dd, 9.0, 2.0)
7.70
(dd, 8.0, 2.0)
122.57.64
(dd, 8.0, 2.0)
120.77.87
(dd, 8.6, 2.2)
121.67.77
(dd, 8.2, 2.2)
−OMe4.08 (s)
3.97 (s)
2x 3.92 (s)
3.90 (s)
3.89 (s)
4.04 (s)
3.91 (s)
3.87 (s)
3.83 (s)
61.7
61.4
60.5
55.6
4.02 (s)
3.91 (s)
3.87 (s)
3.83 (s)
61.6
60.6
56.3
4.09 (s)
4.00 (s)
62.4
61.2
56.2
4.24 (s)
4.21 (s)
1″4.88
(d, 7.0)
5.46
(d, 7.0)
100.95.45
(d, 8.0)
102.35.73
(d, 7.3)
104.56.15
(d, 7.3)
2″--74.22.18-2.33
(m, 2H)
74.7-75.7-
3″--76.2 78.3-78.2-
4″--70.02.18-2.33
(m, 2H)
71.5-71.4-
5″--74.2 75.7-75.9-
6″--63.01.01
(s, 2H)
64.75.02
(dd, 12.0, 2.2)
4.74
(dd, 12.0, 6.5)
64.44.99
(dd, 12.0, 1.0)
4.81
(dd, 12.0, 5.2)
1--173.7-
2--45.3-
3‴--68.6-
4--45.6-
5--177.6-
6--27.0-
COO-----171.6-171.6-
----δCδH (J in Hz)
−CH2−----46.32.88 (d, 14.6)
2.99 (d, 14.6)
−CH2−----46.52.97
(d, 14.5)
3.04
(d, 14.5)
−COH----70.1-
Me----28.01.57 (s)
Table 18. Glycosylated flavones (8891) extracted from Citrus genus.
Table 18. Glycosylated flavones (8891) extracted from Citrus genus.
No.88899091
Ref.[189][189][189][189]
SolventPyridine-d5Pyridine-d5Pyridine-d5Pyridine-d5
PositionδCδH
(J in Hz)
δCδH
(J in Hz)
δCδH
(J in Hz)
δCδH
(J in Hz)
2163.8-163.8-165.3-164.8-
3104.66.98 (s)104.86.99 (s)104.17.01 (s)103.96.99 (s)
4183.3-183.3-183.7-183.4-
5150.0-150.0-150.1-149.2-
6133.2-133.2-137.6-137.36.57 (s)
7152.8-152.8-150.1-150.0-
8129.4-129.4-134.2-133.9-
9146.6-146.6-146.4-146.2-
10104.5-104.4-108.2-107.8-
1′125.8-126.1-122.8-122.6-
2′111.27.64 (d, 1.4)111.27.61 (d, 2.0)110.97.69 (d, 2.0)110.67.67 (d, 2.2)
3′150.5-150.7-150.2-150.1-
4′151.3-151.2-152.9-152.9-
5′116.77.68 (d, 8.5)117.17.74 (d, 8.4)117.47.31 (d, 8.4)117.37.32 (d, 8.6)
6′120.67.71
(dd, 8.5, 1.4)
120.77.91
(dd, 8.4, 2.0)
121.77.77
(dd, 8.4, 2.0)
121.57.78
(dd, 2.2, 8.5)
OMe61.5
60.5
56.4
4.03 (s)
4.00 (s)
61.6
60.6
56.4
4.10 (s)
4.00 (s)
62.4
61.3
56.4
4.28 (s)
4.23 (s)
62.2
61.1
56.2
4.21 (s)
4.19 (s)
1″102.25.83 (d, 7.1)102.85.74 (d, 6.8)104.66.15 (d, 7.2)104.56.28 (d, 7.0)
2″74.9-74.8-75.7-75.9-
3″78.6-78.5-78.4-78.2-
4″71.5-71.6-71.5-71.7-
5″79.1-75.8-76.0-79.2-
6″62.64.59
(dd, 12.1, 2.1)
64.85.14
(dd, 12.1, 1.6)
4.77
(dd, 12.1, 6.8)
64.54.99
(d, 11.6)
4.80
(dd, 11.6, 5.6)
62.74.54
(dd, 11.8, 2.6)
COO---171.7-171.8---
COOH--174.7-174.8---
−CH2−--46.43.13 (d, 14.1)
3.32 (d, 14.1)
46.52,99 (d, 14.5)
3.08 (d, 14.5)
--
−CH2−--46.73.14 (d, 15.3)
3.20 (d, 15.3)
46.53.01 (d, 15.0)
3.07 (d, 15.0)
--
−COH--70.1-70.1---
Me--28.31.74 (s)28.11.63 (s)--
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fontana, G.; Bruno, M.; Sottile, F.; Badalamenti, N. The Chemistry and the Anti-Inflammatory Activity of Polymethoxyflavonoids from Citrus Genus. Antioxidants 2023, 12, 23. https://doi.org/10.3390/antiox12010023

AMA Style

Fontana G, Bruno M, Sottile F, Badalamenti N. The Chemistry and the Anti-Inflammatory Activity of Polymethoxyflavonoids from Citrus Genus. Antioxidants. 2023; 12(1):23. https://doi.org/10.3390/antiox12010023

Chicago/Turabian Style

Fontana, Gianfranco, Maurizio Bruno, Francesco Sottile, and Natale Badalamenti. 2023. "The Chemistry and the Anti-Inflammatory Activity of Polymethoxyflavonoids from Citrus Genus" Antioxidants 12, no. 1: 23. https://doi.org/10.3390/antiox12010023

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

Article Metrics

Back to TopTop