Next Article in Journal
Magnetic Core-Shell Nanoparticles Using Molecularly Imprinted Polymers for Zearalenone Determination
Next Article in Special Issue
Partial Characterization of Lectins Purified from the Surco and Vara (Furrow and Rod) Varieties of Black Phaseolus vulgaris
Previous Article in Journal
N-[1,3-Dialkyl(aryl)-2-oxoimidazolidin-4-ylidene]-aryl(alkyl)sulphonamides as Novel Selective Human Cannabinoid Type 2 Receptor (hCB2R) Ligands; Insights into the Mechanism of Receptor Activation/Deactivation
Previous Article in Special Issue
Triptonoterpene, a Natural Product from Celastrus orbiculatus Thunb, Has Biological Activity against the Metastasis of Gastric Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Resin Glycosides from Convolvulaceae Family: An Update

1
Department of Chemistry, Faculty of Mathematics and Natural Sciences, Universitas Padjadjaran, Jatinangor, Sumedang 45363, Indonesia
2
Central Laboratory, Universitas Padjadjaran, Jatinangor, Sumedang 45363, Indonesia
3
Centre of Natural Products Chemistry and Synthesis Studies, Faculty of Mathematics and Natural Sciences, Universitas Padjadjaran, Jatinangor, Sumedang 45363, Indonesia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(23), 8161; https://doi.org/10.3390/molecules27238161
Submission received: 13 October 2022 / Revised: 13 November 2022 / Accepted: 15 November 2022 / Published: 23 November 2022

Abstract

:
Resin glycoside is a type of secondary metabolite isolated commonly from the Convolvulaceae family. It consists of oligosaccharides conjugated to organic acids with a larger percentage having a macrocyclic structure. The resin glycosides reported in this review is classified mostly based on the number of sugar units constructing the structure, which is correlated to the biological properties of the compounds. According to preliminary reviews, the protocols to isolate the compounds are not straightforward and require a special technique. Additionally, the structural determination of the isolated compounds needs to minimize the structure for the elucidation to become easier. Even though resin glycosides have a complicated structural skeleton, several total syntheses of the compounds have been reported in articles published from 2010 to date. This review is an update on the prior studies of the resin glycosides reported in 2010 and 2017. The review includes the classification, isolation techniques, structural determination, biological properties, and total synthesis of the resin glycosides.

Graphical Abstract

1. Introduction

Resin glycosides are a class of natural products comprising oligosaccharides and organic acids, which are mostly obtained from the plants of the Convolvulaceae family [1,2]. The structures of these compounds are intramolecular macrocyclic esters formed through an intramolecular cyclization of the oligoglycosidic acid containing hydroxy fatty acids. However, several resin glycosides have acyclic aglycone [3,4,5] instead of forming macrocyclic structures, which are mostly acylated at the sugar moiety with varied organic acids [6,7,8,9,10]. The classification of this compound is based on the various number of sugar units inherent in the structure, which also form an ester-type dimer that makes the categorization process even larger. The unique structure of resin glycosides is also related to the biological properties of the compounds. Resin glycosides have broad biological properties, such as cytotoxic activities, MDR modulating properties, α-glucosidase inhibitory potential, antimicrobial activity, anti-metastatic activity, antiproliferative activity, neuroprotective and anticonvulsant activities [10,11,12,13,14,15,16]. Others include a sedative effect and GABA release, anti-inflammatory, antidepressant effect, anti-diarrhea, anti-bacteria, antimalaria, and downregulation of aquaporin 3 [15,17,18,19,20,21,22]. The isolation and characterization of this compound are somehow not straightforward. Therefore, special protocols are required to make successful isolation [23,24]. Mostly, the simplification of the resin glycoside structure through hydrolysis is required prior to the structural elucidation. Even though the structure of resin glycoside itself seems complicated, several total syntheses with numerous synthetic strategies have been reported and described in the present review.
A review of resin glycosides from the Convolvulaceae family in this review is an update to the prior review published in 2010 and 2017 [1,2]. The review in 2010 described many aspects of resin glycosides from the Convolvulaceae, including ethnobotanical aspects, structural diversity, isolation techniques, strategies for synthesis, and biological significance. The discussion in the 2017 review is not as large as the review in 2010. The review discussed two sections only, which are isolation and structural elucidation, and limited the discussion to four species of plants only. Recently, in 2022, a similar review regarding the update of resin glycosides from Convolvulaceae was reported [25]. The review discussed two sections, which are the structural diversity and biological properties of the resin glycosides. The present review described a comprehensive explanation of the classification, isolation, biological properties, structural determination, and total synthesis of resin glycosides of Convolvulaceae obtained from articles published from 2010 until 2022.

2. Classification

This section classified resin glycosides into eight groups, namely, monosaccharides, trisaccharides, tetrasaccharides, pentasaccharides, hexasaccharides, heptasaccharides, bidesmoside, and an ester-type dimer. The classification was mainly based on the number of sugars present in the structure, except for the last two-groups. The resin glycosides in the heptasaccharides and bidesmoside group were not reported in the 2010 review. Table 1 is a list of the classified resin glycosides.

2.1. Monosaccharides

The monosaccharide group, which consists of an aglycone as a decanoic acid (C-10) and one sugar group of either quinovose or fucose attached to C-7 of the decanoic acid, has the simplest structure. Fucose sugar is found in the compound of maltifidinic acid E (1) (Figure 1) isolated from the Quamoclit × multifida plant [26]. Meanwhile, the quinovose sugar is located in the compound of quamoclinic acid B (2) and operculinic acid K (3) (Figure 1) from the Quamoclit pennata [27] and Operculina macrocarpa plants [28]. Both quamoclinic acid B (2) and operculinic acid K (3) are diastereoisomers with methine carbon at C-7. The difference was found by measuring their optical rotation, where the compound (2) and compound (3) were [α ] D 22 −20.0 and [α ] D 22 −39.4, respectively. Furthermore, the acid hydrolysis of compound (3) produces a levo aglycone (a levorotatory aglycone) with an optical rotation of [α ] D 22 −3.3, supporting the presence of the R configuration at C-7. Based on these data, compound (3) was characterized as 7R-hydroxydecanoic acid 7-O-β-D-quinovopyranoside. This compound is rarely found as an aglycone or acyl side chain in the glycosidic resins of Convolvulaceae [28].

2.2. Trisaccharides

Cuses-5–7 (46) (Figure 2) were trisaccharide-type resin glycosides isolated from the Cuscuta chinensis plant, where the three compounds are the result of a reductive amination reaction of a trisaccharide glycoside resin with p-anisidine [4]. The reaction was carried out because the glucose in the first sugar unit group on the anomeric carbon was not protected, easily underwent tautomerization, and was difficult to purify. For phytochemical studies, the ethanol fraction containing the resin glycoside was reduced with p-anisidine in order to protect the reducing sugar, thereby leading to an aminoalditol derivative grouped into disaccharide-type resin glycoside.
Cuses-3 (7) and cuses-4 (8) (Figure 3) are trisaccharide-type resin glycosides isolated from Cuscuta chinensis [4]. The glucose, as the first sugar unit, was exposed and unprotected, which led to tautomerism, resulting in α and β-anomers. Similarly, dichondrin C (9) (Figure 3), a trisaccharide-based resin glycoside isolated from the Dichondra repens, was unprotected glucose in the first sugar unit. In its structure, the aglycone was attached to the second sugar unit, rhamnoside [3]. Poranic acid A-B (1011) and A (12), isolated from Porana duclouxii [5], and stansoic acid A (13) (Figure 3), obtained from Ipomoea stans [16], showed a different structure. The trisaccharide has an aglycone attached to the first sugar unit group, hence, there is no unprotected sugar unit in the structure, and it resists tautomerism. The last four compounds have a sugar skeleton of 11-O-α-Rha-(1 2)-O-β-Glc-(1 2)-O-β-Qui-(1 11)-aglycone.

2.3. Tetrasaccharides

Four macrocyclic aglycones of resin glycoside, albinoside VI–IX (1417), and four of their glycosidic acids, albinosinic acid D g (1821) were isolated from Ipomoea alba (Figure 4) [7]. All were categorized as tetrasaccharide-type resin glycosides, which are unique with varying structural skeletons. These include 11S-hydroxytetradecanoic acid 11-O-α-L-Rha-(1 2)-O-[β-D-Qui-(1 6)]-O-β-D-Glc-(1 2)-O-β-D-Qui of compound (14) with its glycosidic acid (18), then 11S-hydroxytetradecanoic acid-11S-hydroxytetradecanoic acid 11-O-α-L-Rha-(1 2)-O-[β-D-Qui-(1 3)]-O-β-D-Qui-(1 2)-O-β-D-Qui of compound (15) with its glycosidic acid (19), 11S-hydroxytetradecanoic acid 11-O-β-D-Fuc-(1 4)-O-α-L-Rha-(1 2)-O-β-D-Glc-(1 2)-O-β-D-Qui of compound (16) with its glycosidic acid (20), and lastly, 11S-hydroxyhexadecanoic acid 11-O-β-D-Fuc-(1 4)-O-α-L-Rha-(1 2)-O-β-D-Glc-(1 2)-O-β-D-Qui of compound (17) with its glycosidic acid (21).
Aquaterin XI (22) (Figure 4)is a tetrasaccharide-type resin glycoside with a linear chain consisting of three rhamnosyl sugar units and one fucosyl sugar unit [29]. It also has a structural skeleton of 11S-hydroxyhexadecanoic acid 11-O-α-rhamnopyranosyl-(1 4)-O-α-rhamnopyranosyl-(1 4)-O-α-rhamnopyranosyl-(1 2)-O-β-fucopyranoside-(1,2″-lactone). This compound was successfully isolated from Ipomoea aquatica.
The oligosaccharide core of calonyctin E (23) (Figure 4) was determined as rhamnosyl-(1 → 2)-[glucosyl-(1 → 3)]-quinosyl-(1 → 2)-quinosyl with 11S-jalapinolic acid (1,3″-lactone) as aglycone [23]. Calonyctin F-H (2426) (Figure 4) almost have the same aglycone and oligosaccharide core, except the sugar unit D-glucose, which was changed to D-quinovose for the skeleton to become 11-O-α-L-rhamnosyl-(1 2)-O-[β-D-quinosyl-(1 3)]-O-β-D-quinosyl-(1 2)-O-β-D-quinovopyranoside. Calonyctin I (27) (Figure 4) has the same oligosaccharide core and aglycone as compound 23 and similarities with compounds 2426 in which the aglycone is attached to a nilic acid (Nla) group. Calonyctin J (28) (Figure 4) is a glycosidic acid with the same oligosaccharide core and aglycone as compounds (2426). Meanwhile, compounds (2328) were isolated from Ipomoea muricata.
Calysolin I (29) and calysolin X (30) (Figure 4) are tetrasaccharide-based resin glycosides with an oligosaccharide core of 11-O-α-L-rhamnosyl-(1 2)-O-[β-D-glucosyl-(1 3)]-O-β-D-glucosyl-(1 2)-O-β-D-quinovopyranosyl and aglycone of 11S-jalapinolic acid (1,2″‴) [30,31]. Other series of calysolin compounds are widely distributed as various types of pentasaccharides, hexasaccharides, and heptasaccharides, isolated from Calystegia soldanella.
Dichondrins A-B (3132) (Figure 4), isolated from Dichondrin repens, have the same sugar sequence of 11-O-α-L-rhamnosyl-(1 4)-O-β-D-glucosyl-(1 2)-[O-α-L-rhamnosyl-(1 6)]-O-β-D-glucopyranoside with 11S-jalapinolic acid as an aglycone [3]. Dichondrin B (32) contains an additional OH group at C-3 of the aglycone (3,11-jalapinolic acid). Mosher method was used to determine the absolute configuration (Δδ = δS – δR, Δδ16H = −0.011, Δδ2Ha = +0.02), and it revealed that the aglycone has an S configuration. Based on the data, the aglycone was determined as 3S,11S-jalapinolic acid.
The aglycones present in evolvulic acid B-C (3334) (Figure 4) have similarities due to the presence of trihydroxy jalapinolic acid at positions C-3, C-11, and C-14 [32]. The absolute configuration of 3S,11R,14R was determined using the Mosher ester method. The aglycone was determined as 3S,11R,14R-trihydroxyhexadecanoic acid (3S,11R, and 14R-jalapinolic acid). Evolvulic acid B (33) has an oligosaccharide core of 11-O-β-D-galactosyl-(1 3)-O-α-L-rhamnosyl-(1 2)-O-β-D-glucosyl-(1 2)-O-β-D-fucopyranoside. Meanwhile, evolvulic acid C (34) has a very different oligosaccharide skeleton core, which is 11-O-α-L-rhamnosyl-(1 3)-O-β-D-glucosyl-(1 2)-[O-β-D-fucosyl-(1 2)]-O-β-D-fucopyranoside. Both compounds were isolated from Evolvulus alsinoides.
The resin glycosides (3538) (Figure 4) isolated from aerial parts of Ipomoea batatas [18] have structures similar to the oligosaccharide core, which is 11-O-α-L-rhamnopyranosyl-(1 → 4)-O-α-L-rhamnopyranosyl-(1 → 4)-O-α-L-rhamnopyranosyl (1 → 4)-O-β-D-fucopyranoside. However, there are some differences, particularly in the position of the macrocyclic aglycone ester bonds. Ipomotaoside A-B (3536) has a jalapinolic acid aglycone at the position of the 1,2″-ester. Meanwhile, ipomotaoside C-D (3738) has the ester bond at 1,3″.
The structure of muricatin D (39) (Figure 4) isolated from Ipomoea muricata [33] was determined as 11S-jalapinolic acid 11-O-β-D-fucopyranosyl-(1 4)-O-α-L-rhamnopyranosyl-(1 2)-O-β-D-glucopyranosyl-(1 2)-β-D-quinovopyranoside. It was reported to be similar to compound (21) and inconsistent with 1H- and 13C-NMR chemical shifts of 39. Muricatin IX–XI (4042) (Figure 4) [34] was also isolated from the same plant species. All resin glycosides have the same oligosaccharide core as muricatin I–VI [72] with three sugar units, namely quinovose (qui′), quinovose (qui″) and rhamnose (rha‴). The fourth unit can be either quinovose or fucose (qui″‴ or fuc″‴) with the presence of an aglycone known as jalapinolic acid. In muricatin IX (40), the fourth sugar unit, quinovose (qui″‴), was attached to the second, quinovose (qui″). Similarly, the jalapinolic acid ester of 40 was attached to nilic acid (Nla) at C-3. However, this is in contrast to muricatin X–XI (4142) in which the fourth sugar unit quinovose or fucose (qui″‴ or fuc″‴) is attached to rhamnose (rha‴) with the aglycone, which is attached to the qui″ at C-3 (ester-1.3″).
There are a least two tetrasaccharide-type resin glycosides obtained from Ipomoea stans, namely, stansin 6 (43) [10,35] and stansinic acid I (44) (Figure 4) [16]. Both compounds have a similar oligosaccharide core, 11-O-β-D-quinovopyranosyl-(1 4)-O-α-L-rhamnopyranosyl-(1 2)-O-β-D-glucopyranosyl-(1 2)-O-β-D-quinopyranoside, with a varying aglycone moiety. Stansin 6 (43) has a macrocylic aglycone (1,3‴-ester), while stansinic acid I (44) contain an acyclic aglycone.
The other series of tricolorins was also found from Ipomoea tricolor, which are tricolorin K-M (4547) (Figure 4) [12]. These resin glycosides share the same oligosaccharide framework as tricolorins A-D [73], which is 11-O-α-L-rhamnosyl-(1 3)-O-α-L-rhamnosyl-(1 2)-O-β-D-glucosyl-(1 2)-O-β-D-fucopyranoside, with a macrocyclic aglycone that is bound to the second sugar unit (glu″) at C-3 (ester-1,3″). The structural difference is only found in the placement of nilic acid (Nla) and methyl butyric acid (Mba) at C-2 and C-3 in the third sugar unit (rha‴).
The tetrasaccharide-based resin glycoside, stansinic acid I (44), has a structure of 11-O-β-D-Qui-(1 → 4)-O-α-L-Rha-(1 → 2)-O-β-D-Glc-(1 → 2)-O-β-D-Qui with acyclic jalapinolic acid, which acts as the aglycone [17]. The structure is similar to that of tyrianthinic acid VI (48) (Figure 4). This compound was obtained from Ipomoea tyrianthina, where the difference was only in the absence of the methyl butyric acid (Mba) group at C-2 in the second sugar unit (Rha‴).
Wolcottinoside I (49) (Figure 4) isolated from Ipomoea wolcottiana [8] has a similar structure to ipomotaoside C-D compound (3738) obtained from Ipomoea batatas [18]. The difference between both compounds is the absence of a dodecanoyl side group (Dodeca) at C-2 in the third sugar unit Rha‴ and an octanoyl group (Octa) at C-4 in the fourth sugar unit Rha.

2.4. Pentasaccharides

Acutacosides A-B (5051) (Figure 5) are resin glycosides containing a pentasaccharide core in their structure [8]. These compounds were isolated from Argyreia acuta and have the same structural framework that is characterized by the presence of similar carbon signals in the 13C-NMR, particularly of the chemical shift from 177 to 41 ppm. The oligosaccharide core of these compounds is 11-O-α-L-rhamnopyranosyl-(1 3)-O-[α-L-rhamnopyranosyl-(1 4)]-O-α-L-rhamnopyranosyl-(1 4)-O-α-L-rhamnopyranosyl-(1 2)-O-β-fucopyranoside with an intramolecular aglycone 1,2″-ester. The only difference is that acutacoside A (50) and acutacoside B (51) have decanoic and dodecanoic groups, respectively. Acutacosides C-E (5254) [36] were isolated, where the structures had the same framework as acutacosides A-B (5051). However, the first sugar unit of acutacosides C-E (5254) was glucose, while for acutacosides A-B (5051) it was fucose. The difference between the structure of acutacosides C-E (5254) was in the position of the cinnamoyl acid (Cna) and the dodecanoyl (Dodeca) groups at C-2, C-3, and C-4 of the fourth sugar unit (rha″‴). Acutacosides F-I (5558) [37] have the same oligosaccharide core as acutacosides A-B (5051); the difference is in the presence of organic acid groups, namely, methyl butyric acid (Mba), cinnamic acid (Cna), dodecanoyl (Dodeca), and butyl (Bu) at the third (Rha‴) and fourth (Rha″‴) sugar units.
In 2012, albinosides I–III (5961) (Figure 5) were isolated from Ipomoea alba. Based on GC-MS analysis, these resin glycosides revealed the presence of sugar units of rhamnose (Rha), quinovose (Qui), and glucose (Glc) in a ratio of 2:2:1 [38]. HMBC analysis was used to determine the location of the aglycone in the structure. Based on the HMBC analysis, macrolactone was located at C-2, while the aglycone of albinoside I (59) was found in the terminal sugar unit of rhamnose (Rha), which was shown through a correlation of H-2 (5.60) with C-1 (173.5) of convolvulinic acid. The macrolatone of albinoside II (60) is located at C-3 on the terminal sugar unit rhamnose (Rha) with a correlation of H-3 (5.66), as well as C-1 (171.1) of convolvulinic acid. For albinoside III (61), the macrolactone is located at C-3 of the second sugar unit, and glucose (Glc) is indicated by a correlation between H-3 (5.78) and C-1 (173.0) jalapinolic acid. Based on its oligosaccharide core, albinoside IV (62) [7] has the same sugar framework as albinoside I (59). However, there is a tigloyl (Tig) group in albinoside IV (62) (Figure 5) that is attached to the third sugar unit, known as rhamnose (Rha′). Albinoside V (63) (Figure 5) has the same framework as the albinoside II (60). The difference between albinoside V (63) and albinoside II (60) with albinoside I (59) and albinoside III (61) is in the second sugar unit, whereby glucose is replaced with rhamnosyl sugar; hence, the oligosaccharide core becomes 11-O-α-L-rhamnosyl-(1 → 2)-O-β-D-quinosyl-(1 → 2)-O-[α-L rhamnosyl-(1 → 3)]-O-α-L-rhamnosyl-(1 → 4)-O-β-D quinosyl [7].
Resin glycoside from Ipomoea aquatica produced eleven aquaterins I–XI, consisting of ten pentasaccharide-type resin glycoside, aquaterins I–X (6473) (Figure 5), and one tetrasaccharide group, aquaterin XI (22) [29]. Structurally, there are three types of pentasaccharide aquaterins compounds (6473). The first is the oligosaccharide core rhamnosyl-(1 → 4)-O-[rhamnosyl-(1 → 3)]-O-rhamnosyl-(1 → 4)-O-rhamnosyl-(1 → 4)-O-fucopyranosyl, with a macrocyclic aglycone of jalapinolic acid attached to the C-2 rha sugar” (ester-1,2″) for the compound aquaterins I–II (6465), V–VI (6869), and VIII (71). The second contains a similar oligosaccharide core as the first, but the jalapinolic acid is attached to C-3, the rha sugar unit (ester-1,3), for aquaterins III–IV (6667), VII (70), and IX (72). Finally, the third framework has a difference in the sugar units, where the fourth is replaced with a glucose sugar unit, thereby producing an oligosaccharide core of rhamnosyl-(1 → 4)-O-[glucosyl-(1 → 3)]-O-rhamnosyl-(1 → 4)-O-rhamnosyl-(1 → 4)-O-fucopyranosyl with jalapinolic acid aglycone (ester-1,3″) for aquaterin X (73). In the following year, eight resin glycosides were isolated, namely, aquaterins XII–XIX (7481) (Figure 5) [39]. These compounds have the same oligosaccharide core structure as aquaterins I–IX compounds (6472). The difference is only in the placement of the macrocyclic aglycones of jalapinolic acid. The ester-1,2″ type is for aquaterins XII–XIV (7476), XVI (78), XVIII–XIX (8081), while the ester-1,3″ type is for aquaterins XV (77) and XVII (79).
Arvensis K-L (8283) (Figure 5) were found in the plants of Convolvulus arvensis [40]. Both compounds have the same oligosaccharide core, which is glucosyl-(1 → 4)-O-glucosyl-(1 → 4)-O-[rhamnosyl-(1 → 2)]-O-glucosyl-(1 → 4)-O-fucopyranosyl. The acyclic aglycone contains different structures, where arvensis L (83) has a common aglycone, and 11S-hydroxyhexadecanoic acid (or 11S-jalapinolic acid), while arvensis K (82) has a new aglycone with an additional one carbon, which is 11S-hydroxyheptadecanoic acid.
The plant of Ipomoea batatas produced batatinoside VII–IX (8486) (Figure 5) [41], a resin glycoside with the same oligosaccharide framework as aquaterins [29]. Batatinoside VII (84) shares similar oligosaccharide cores with aquaterins I–II (6465) and macrocyclic aglycones at C-2 rha units (ester-1,2″). Furthermore, batatinoside VIII (85) is similar to aquaterin X (73), in which one rhamnose sugar is replaced by glucose. Meanwhile, batatinoside IX (86) has the same characteristics as aquaterin III (66) with a similar ester-1,3″ type. Although they have the same oligosaccharide core, the difference between batatinosides VII–IX (8486) and aquterins is in the presence of a dodecanoyl (dodeca) group at C-2 of the third sugar unit and a rhamnosyl sugar [41].
In 2013, six resin glycosides were isolated from Ipomoea cairica, namely, cairicosides A-F (8792) (Figure 5) [42]. At first glance, the oligosaccharide core is somewhat similar to aquaterins [29] or batatinosides VII–IX (8486)[41], but the first sugar unit of the cairicosides is glucose; hence, the structure becomes rhamnosyl-(1 → 4)-O-[rhamnosyl-(1 → 3)]-O-rhamnosyl-(1 → 4)-O-rhamnosyl-(1 → 2)-O-glucopyranosyl. Cairicoside A-E (8791) has a macrocyclic aglycone attached to the C-3 of the second sugar unit, rha′ (ester-1,3″). Meanwhile, the aglycone of cairicoside F (92) is attached to its C-2 (ester-1,2″ type). The plant I. cairica also produced four resin glycosides, namely cairicosides I–IV (9396) (Figure 5) [43]. These compounds have the same framework as cairicoide F (92), with a 1,2″ ester type. The difference among the four structures is only in the organic acid groups attached to the C-2 of the third sugar unit, which can be either butanoyl (Bu), methyl butyric acid (Mba), octanoyl (Octa), or decanoyl (Deca).
Calonyctins found in Ipomoea muricata are of the tetrasaccharide and pentasaccharide type [23]. The pentasaccharide-typed resin glycosides of calonyctins B-C (9798) (Figure 5) have the structural framework as the oligosaccharide core of rhamnosyl-(1 → 3)-O-[quinosyl-(1 → 4)]-O-rhamnosyl-(1 → 2)-O-quinosyl-(1 → 2)-O-fucosyl. Meanwhile, calonyctin D (99) (Figure 5) is only different in its fourth sugar, in which quinovose is replaced with fucose. Calonyctins B-D (9799) have a macrocyclic aglycone of jalapinolic acid and are attached to C-3 in the second sugar unit, qui″ (ester-1,3″ type).
All plant parts of Calystegia hederacea were found to produce four glycosidic acids, namely, calyhedic acids A-D [44]. However, of these four resin glycosides, only calyhedic acid A (100) (Figure 5) belongs to the type of pentasaccharide group with a composition of three glucoses, one quinovose, and one rhamnose sugar unit, accompanied by an acyclic aglycone 12-hydroxyhexadecanoic acid.
Calysolic acids A and B (101102) (Figure 5) were isolated from the fresh leaves, stems, and roots of Calystegia soldanella [45]. Structurally, both acids differ in the order of the oligosaccharide core, where calysolic acid A (101) has a core of rhamnosyl-(1 → 2)-[O-glucosyl-(1 → 6)-O-glucosyl-(1 → 3)]-O-glucosyl-(1 → 2)-O-quinopyranosyl, while calysolic acid B (102) is glucosyl-(1 → 3)-O-rhamnosyl-(1 → 2)-O-[glucosyl-(1 → 3)]-O-glucosyl-(1 → 2)-O-quinopyranoic acid. Both compounds contain acyclic aglycone 11S-hydroxydecanoic acid. C. soldanella was also found to produce calysolins I–IX [30,46]. Calysolin II–III (103104) (Figure 5) has an oligosaccharide core skeleton, which is α-L-rhamnopyranosyl-(1 2)-[O-β-D-glucopyranosyl-(1 6)-O-β-D-glucopyranosyl-(1 3)]-O-β-D-glucopyranosyl (1 2)-O-β-D-quinopyranoside, and jalapinolic acid aglycones (with ester-1,2″ type). Both compounds have differences in the location of the nilic acid (Nla) group, which is attached to the sugar rhamnose unit at its C-2 or C-3 [30]. Meanwhile, calysolin V–VI (105106) (Figure 5) has a β-D-glucopyranosyl-(1 3)-O-α-L-rhamnopyranosyl (1 2)-[O-β-D-glucopyranosyl-(1 3)]-O-β-D-glucopyranosyl-(1 2)-O-β-quinopyranoside. The difference between the structures is in the calysolin V (105), which has a tiglic acid group (tig) on C-3 of glucose (glc″), as opposed to calysolin VI (106). Another difference is that there is a methyl butyric acid (Mba) group on the C-4 glucose unit (glc′) in calysolin VI (106), which was absent in calysolin V (105) [46].
Evolvulic acid A (107) and evolvulin I–III (108110) (Figure 5) have the same oligosaccharide core as β-D-galactopyranosyl-(1 3)-O-α-L-rhamnopyranosyl (1 2)-O-β-D-glucopyranosyl-(1 2)-[O-β-D-fucopyranosyl-(1 4)]-O-β-D-fucopyranoside [47,48]. These compounds are found in Evolvulus alsinoides plants. The most interesting fact about these compounds is the presence of a novel aglycone, which is a 3,11,14-trihydroxyhexadecanoid acid unit, a new fatty acid derivative obtained from nature. The absolute configuration of the aglycone was determined by the Mosher method using an MTPA reagent. Meanwhile, the aglycone was derived as 3S,11R, and 14R-trihydroxyhexadecanoid acid [47].
The two new glycoside resins obtained by the plant Ipomoea pes-caprae, ipomeolides A and B (111112) (Figure 5), had an unusual structure with the aglycone of 11R-jalapinolic acid [49]. The configuration of macrolactones with 11R has not been reported, where the majority of resin glycosides have an 11S configuration. Both compounds have an oligosaccharide core rhamnosyl-(1 → 4)-O-rhamnosyl-(1 → 4)-[O-rhamnosyl-(1 → 3)]-O-rhamnosyl-(1 → 2)-O-fucopyranosyl-(1,2″-lactone).
Merremins A-G (113119) (Figure 5), isolated from Merremia hederacae, have an oligosaccharide core of glucosyl-(1 → 3)-O-[rhamnosyl-(1 → 4)]-O-rhamnosyl-(1 → 4)-O-rhamnosyl-(1 → 2)-O-fucosyl [50]. The difference between these merremin compounds is found in the attachment position of the lactone aglycone. Macrocyclic aglycones of merremins A-D (113116) are attached in the C-2 second sugar (1,2″-ester type). Meanwhile, merremin E (117) has its aglycone attached to the C-3 second sugar (1,3-ester type), with the presence of an acylic aglycone in the merremins F-G (118119) structure.
Seven resin glycosides have succeeded in isolation from Quamoclit × multifida plants [51]. These consist of four resin glycosides with a pentasaccharide core, multifidins III–VI (120123) (Figure 5), two resin glycosides with a hexasaccharide core, multifidin VII–VIII (184185) (Figure 6), and one type of the bidesmode group, multifidine IX (218) (Figure 7). Multifidins III–VI (120123) compounds have two differences; firstly, the aglycone can be either macrocyclic lactone (ester-1,2″ type) for multifidin III–IV (120121) or acyclic aglycone for multifidin V–VI (122122). Secondly, there are two distinct oligosaccharide cores, namely, a glucosyl-(1 → 3)-O-[rhamnosyl-(1 → 4)]-O-rhamnosyl-(1 → 4)-O-rhamnosyl-(1 → 2)-O-fucopyranosyl for multifidin III (120) and multifidin V (122). This is in addition to a similar oligosaccharide core that differs in the first sugar unit, which is quinovose for multifidin IV (121) and multifidin VI (123).
The isolation of resin glycoside from Ipomoea batatas resulted in four murasakimasarins I–IV (124127) (Figure 5) [52]. The difference in the structures of these compounds is in the first sugar unit, which built the oligosaccharide core. The first sugar unit in Murasakimasarin I (124) and murasakimasarin II–IV (125127) are glucose and fucose, respectively. For example, murasakimasarin II (125) has an oligosaccharide core of glucosyl-(1 → 4)-rhamnosyl-(1 → 4)-rhamnosyl-(1 → 4)-rhamnosyl-(1 → 2)-fucosyl. The macrocyclic aglycone of murasakimasarins I–III (124126) is attached to the C-2 of the second sugar unit (ester-1,2″ type). In contrast to murasakimasarin IV (127), its aglycone is attached to the C-3 of the second sugar unit (ester-1,3″ type).
A new glycosidic acid, named murucinic acid II (128) (Figure 5), with an oligosaccharide resin core of glucosyl-(1→2)-O-rhamnosyl-(1→4)-O-rhamnosyl-(1→4)-O-rhamnosyl-(1→2)-O-fucopyranoside was isolated from Ipomoea murucoides [16].
Ten resin glycosides, pescapreins XXI–XXX (129138) (Figure 5), were found in Ipomoea pes-caprae [53]. The resins have frameworks of 11-O-α-L-rhamnopyranosyl-(1 3)-O-[α-L-rhamnopyranosyl-(1 4)]-O-α-L-rhamnopyranosyl-(1 4)-O-α-L-rhamnopyranosyl-(1 2)-O-β-D-fucopyranoside. The aglycone macrocyclic location, pescapreins XXI–XXVIII (129136), is at the C-3 of the second sugar unit with an ester-1,3″ type. Meanwhile, pescapreins XXIX–XXX (137–138) has the aglycone attached to the C-2 of the second sugar unit (ester-1,2″ type).
The discovered resin glycosides in Pharbitis nil seeds have a unique acyclic aglycone, namely methyl 3S,11S-ipurotare in PM-6–PM-7 (139–140) (Figure 5) [54]. PM-6–PM-7 (139140), which has a skeleton of quinosyl-(1 → 4)-rhamnosyl-(1 → 2)-O-glucosyl-(1 → 2)-O-[rhamnosyl-(1) oligosaccharides → 6)]-O-glucopyranosyl. The oligosaccharide sequence indicates that the aglycone is attached to the second sugar unit of glucose.
The pentasaccharide found in the Ipomoea purga plant, known as purginoside I–IV (141144) (Figure 5) [6,9], have an oligosachharide core framework of glucosyl-(1 → 3)-O-[rhamnosyl-(1 → 4)]-O-rhamnosyl-(1 → 4)-O-rhamnosyl-(1 → 2)-O-fucopyranoside, which were confirmed to be related to the pescapreins and murucoidins series [74,75]. The macrocyclic aglycones of purginoside I–IV (141144) are all attached to the C-2 of the second sugar, rhamnose (ester-1,2″).
The quamoclin V–VII (145147) (Figure 5) isolated from Quamoclit pennata have a framework similar to purginoside, which is β-D-glc-(1 → 3)-O-[α-L-rha-(1 → 4)]-O-α-L-rha-(1 → 4)-O-α-L-rha-(1 → 2)-O-β-D-fuc [55]. The aglycones of the compounds differed from the structures of quamoclin V–VII (145147). Similarly, Quamoclin V (145), quamoclin VI (146), and quamoclin VII (147) contain a 1,3″ ester-typed macrocyclic aglycone, 1,2″ ester-typed macrocyclic aglycon, and an acyclic aglycone.
The Operculina turpethum plant produced glycosidic acids, consisting of turpethic acids A-C (148150) (Figure 5), glycoside resin with lactone-1,2″ type, and turpethosides A-B (151152) (Figure 5) [56]. The oligosaccharide core of these compounds is β-D-glucosyl-(1 → 3)-O-[α-L-rhamnosyl-(1 → 4)]-O-α-L-rhamnosyl-(1 → 4)-O-α-L-rhamnosyl-(1 → 2)-O-β-D-glucopyranoside. Meanwhile, the aglycones are unique because there are three variations of the fatty acids, namely, C15, C16, and C17. C15 fatty acids, such as 12S-hydroxypentadecanoic acid found in turpethic acid A (148) and turpethoside A (151). C16 fatty acids are found in turpethic acid B (149) and turpethoside (152) as a common aglycone of the resin glycosides, jalapinolic acid, and 12S-hydroxyhexadecanoic acid. Finally, C17 fatty acid and 12S-hydroxyheptadecanoic acid are found in turpethic acid C (150).
Four linear glycosidic acids, namely, tyrianthinoic acid (153) and tyrianthinic acids III–V (154156) (Figure 5), have an oligosaccharide core of β-D-quinosyl-(1 → 4)-O-β-D-quinosyl-(1 → 4)-O-α-L-rhamnosyl-(1 → 2)-O-β-D-glucosyl-(1 → 2)-O-β-D-quinovopyranoside [17]. These compounds were isolated from Ipomoea tyrianthina.
The pentasaccharide core of wolcottinosides III–IV (157159) (Figure 5) is glucosyl-(1 → 3)-O-[rhamnosyl-(1 → 4)]-O-rhamnosyl-(1 → 4)-O-rhamnosyl-(1 → 2)-O-glucopyranoside, which is almost similar to arboresins 1–6 with a difference in the last sugar unit (glc′) [24]. According to the report, the last sugar unit does not have a branch at C-3 of Rha″ but forms a linear pentasaccharide. Furthermore, wolcottinoside V (158) has a different oligosaccharide core of glucosyl-(1 → 3)-O-[rhamnosyl-(1 → 4)]-O-rhamnosyl-(1 → 4)-O-rhamnosyl-(1 → 2)-O-fucopyranoside, which is similar to intrapilosine VII (112). The macrolactonization of wolcottinosides III–V (157159) took place at the C-2 of the second sugar unit. These three compounds were found in the plant Ipomoea wolcottiana.
The new resin glycoside compound (160) (Figure 5), isolated from the Ipomoea maxima plant, was named compound 1 by the first author [57]. Moreover, compound 1 (160) was reported to have an oligosaccharide core of rhamnosyl-(1 → 3)-O-[rhamnosyl-(1 → 4)]-O-rhamnosyl-(1 → 4)-O-rhamnosyl-(1 → 2)-O-fucopyranoside (ester-1.2″) similar to the aquaterins series [29].

2.5. Hexasaccharides

The plant of Calystegia hederacea produced resin glycosides containing hexasaccharide core, calyhedic acids, and calyhedins series [44,58,59,60]. Calyhedic acid B (161) (Figure 6) comprises β-D-glucosyl-(1 3)-O-α-L-rhamnosyl-(1 2)-[O-β-D-glucosyl-(1 6)-O-β-D-glucosyl-(1 3)]-O-β-D-glucosyl-(1 2)-O-β-D-quinovopyranoside with acyclic aglycone 12S-hydroxyhexadecanoic acid [44]. This acid has similar sugar units with calyhedic acid E-F (162163) (Figure 6), namely four glucose, one rhamnose, and one quinovose. However, in calyhedic acid E-F (162163)[58], one glucose sugar unit changed position and was not bound to C-3 units of rhamnose sugar but was bound to C-6 glucose units. Hence, the hexasaccharide core became α-L-rhamnosyl-(1 2)-[O-β-D-glucosyl-(1 6)-O-β-D-glucosyl-(1 6)-O-β-D-glucosyl-(1 3)]-O-β-D-glucosyl-(1 2)-O-β-D-quinovopyranoside. The difference in the position of the hydroxyl group in the aglycone is between these two compounds. The acyclic aglycone of 11S-hydroxyhexadecanoic and 12S-hydroxyhexadecanoic acids are found in calyhedic acid E (162) and calyhedic acid F (163), respectively. The macrocyclic form of calyhedic acids is the calyhedin series. Calyhedins I–III (164166) (Figure 6) have the same hexasaccharide core as calyhedic acid B (161). The macrolactonization sites of these three compounds are located at the C-2 position of the glc sugar unit. Calyhedin I (164) and calyhedins II–III (165166) have aglycone of 11S-hydroxyhexadecanoic acid and 12S-hydroxyhexadecanoic acid [59]. Calyhedin X (167) (Figure 6) [60] has the same hexasaccharide core as calyhedic acid F (163). The difference is that calyhedin X (166) is a macrolactone aglycone bound to the C-2 sugar unit of Glc.
Calysepins I–VII (168174) (Figure 6) has a structural skeleton of β-D-glucopyranosyl-(1 3)-O-α-L-rhamnopyranosyl-(1 2)-[O-β-D-glucopyranosyl-(1 6)-O-β-D-glucopyranosyl-(1 3)]-O-β-D-glucopyranosyl-(1 2)-α-D-quinovopyranoside [61]. The aglycone of 11S-jalpinolic acid forms a lactone that is attached to the C-2 of the glucose (glc‴). The difference between the calysepins I–VII (168174) is in the variety of organic acids, namely methyl butyric acid (Mba) and Nilic acid (Nla).
In the plant of Calystegia soldanella, a series of hexasaccharide-typed resin glycosides, calysolic acids, and calysolins were isolated. Calysolic acids C-D (175176) (Figure 6) are isolated as glycosidic acid with the aglycone of 11S-hydroxyhexadecanoic acid. Calysolic acid C (175) has the same oligosaccharide core as calyhedic acid B (161). In contrast to calysolic acid D (176), the sugar unit attached to C-6 Glc′ moves to C-4 Glc‴, hence the skeleton becomes β-D-glucosyl-(1 4)-O-β-D-glucosyl-(1 3)-α-L-rhamnosyl-(1 2)-[O-β-D-glucosyl-(1 3)]-O-β-D-glucosyl-(1 2)-O-β-D-quinovopyranoside [45]. Calysolin IV (177) (Figure 6) has the same hexasaccharide framework as calysolic acid D (176). In contrast, the aglycone of calysolin IV (177) forms a cyclic structure at C-2 of the Glc′ sugar unit [30]. Calysolins series with a hexasaccharide core, namely calysolins VII–IX (178180) (Figure 6), were isolated [46]. Calysolin VII (178) has an oligosaccharide core with a similar aglycone location as calysolin IV (177). Meanwhile, calysolin VIII–IX (179180) has the same hexasaccharide core as calysolic acid C (175) and calyhedic acid B (161) with a varying location. Macrolactonization took place at C-2 sugar units Glc′ for calysolin VIII (179) and at C-2 sugar units Glc″ for calysolin IX (180). Calysolins XI–XIII (181183) (Figure 6) [31] are an advanced series of calysolins, with the same hexasaccharide core as calysolin IV (177) and VII (178). The difference between these compounds is in the variety of organic acids. In the following year, calysolins XIV–XVII (184187) (Figure 6) were isolated to obtain a structural skeleton of glucosyl-(1 → 3)-O-rhamnosyl-(1 → 2)-[O-glucosyl-(1 → 6)-O-glucosyl-(1 → 6)-O-glucosyl-(1 → 6)-O-glucosyl-(1 → 6)-(1 → 3)]-O-glucosyl-(1 → 2)-O-quinovopyranoside [62]. This skeleton is similar to the hexasaccharide core of calysolin VIII–IX (179180). Calysolin XIV (184) has the same macrolactone aglycone of C-2 sugar units Glc″‴ as calysolin IX (180). Meanwhile, calysolins XV–XVII (185187) are resin glycosides with acyclic aglycones [62]. The last series of calysolin are calysolin XIII (188) (Figure 6) isolated from the leaves, stems, and roots of C. soldanella [14]. There was no difference in the hexasaccharide core compared to the previous calysolins VIII–IX (179180) and XIV–XVII (184187). The macrolactone was built at the C-2 sugar unit Glc″‴.
Resin glycoside of macrocarposidic acids A-C (189191) (Figure 6) has a unique hexasaccharide core of β-D-glucosyl-(1 3)-O-α-L-rhamnosyl-(1 2)-[O-β-D-glucosyl-(1 3)]-O-β-D-glucosyl-(1 2)-[O-α-L-rhamnosyl-(1 6)]-O-β-D-glucopyranoside [28]. These resin glycosides have similarities in their acyclic aglycone, 3S,12S-dihydroxyhexadecanoic acid, and the presence of exogonoyl organic acids. The difference between the structures is in the presence of isovaleroyl (Iva) and tigloyl (Tig) organic acids at C-2 and C-3 of Glc‴ sugar units. Macrocarposidic acids A-C (189191) are isolated from the plant of Operculina macrocarpa.
Quamoclit × multifida produced hexasaccharide-based resin, maltifidinic acids C-D (192193) (Figure 6) [26]. Both compounds have an oligosaccharide core of β-D-quinovopyranosyl-(1 2)-O-β-D-glucopyranosyl-(1 3)-[O-β-D-fucopyranosyl-(1 4)]-O-α-L-rhamnopyranosyl-(1 2)-O-β-D-glucopyranosyl-(1 2)-O-β-D-quinovopyranoside and categorized to glycosidic acid, with two hydroxyl positions at C-3 and C-11 on the aglycone. Maltifidinic acids C-D (192193) are homologous compounds, and based on HR-Positive Ion-FAB-MS, maltifidinic acid D (193) has a mass unit of 28 (C2H4) larger than maltifidinic acid C (192). Therefore, the difference lies in their aglycone. 1H- and 13C-NMR analysis and the determination of absolute configuration using Mosher method with MTPA reagent revealed that maltifidinic acid C (192) is 3S,11S-dihydroxytetradecanoic acid, 3S,11S-convolvulinic acid or 3S,11S-ipurolic acid, while maltifidinic acid D (193) is 3S,11S-dihydroxyhexadecanoic acid. Another hexasaccharide glycosidic acids, multifidins VII–VIII (194195) (Figure 6), were found in Q. multifida [51]. Multifidin VII (194) has a similar oligosaccharide core as maltifidinic acid C-D (192193), as opposed to multifidin VIII (195), which has a different first sugar unit in the form of fucose. The hexasaccharide core of multifidin VIII (195) becomes quinosyl-(1 → 2)-O-glucosyl-(1 → 3)-[O-fucosyl-(1 → 4)]-O-rhamnosyl-(1 → 2)-O-glucosyl-(1 → 2)-O-fucopyranoside. The aglycones of these two compounds are 3S and 11S-ipurolic acid.
Operculinic acids H-J (196198) (Figure 6) are glycosidic acids obtained from the roots of O. macrocarpa [28]. Structurally, the hexasaccharide core of these two compounds is similar to macrocarposidic acids A-C (189191). The hydroxy group in the aglycone of operculinic acid H-I (196197) is 12S-hydroxyhexadecanoic acid, and 11S-hydroxyhexadecanoic acid found in operculinic acid J (198).
The acylated hydroxy of fatty acids was found in the seeds of Pharbitis nil, PM1–PM5 (199203) (Figure 6) [54]. The framework for these compounds is α-L-rhamnosyl-(1 3)-O-β-D-quinosyl-(1 4)-α-L-rhamnosyl-(1 2)-O-β-D-glucosyl-(1 2)-[O-α-L-rhamnosyl-(1 6)]-O-β-D-glucopyranoside, which is a linear hexasaccharide that differs from the other series of hexasaccharide-typed resin glycosides. The aglycone of PM3 (201) is methyl 3S,11S-dihydroxyhexadecanoate. Meanwhile, PM1–PM2 (199200) and PM4–PM5 (202203) have an aglycone of methyl 3S,11S-dihydroxytetradecanoate.
The new of methyl pyranoside was discovered in the Quamoclit pennata plant, namely, QM4 (204) (Figure 6) [63]. In this compound, the common aglycone of 11S-jalapinolic acid and 11S-convolvulic acid in the resin glycoside is replaced by methyl groups. The hexasaccharide framework of this compound is β-D-quinosyl-(1 2)-O-β-D-glucosyl-(1 3)-O-[β-D-quinosyl-(1 4)]-O-α-L-rhamnosyl-(1 2)-O-β-D-glucosyl-(1 2)-α-D-fucopyranoside. Another resin glycoside found in Q. pennata is quamoclinic acids C-D (205206) (Figure 6), which contain dihydroxy fatty acids, 3S,11S-ipurolic acid [27]. The oligosaccharide core of quamoclinic acid C (205) is similar to QM4 (204), while in quamoclinic acid D (206), the quinovose sugar unit attached to C-4 of the Rha sugar group is replaced with a fucose sugar unit. The hexasaccharide core of quamoclinic acid D (206) becomes β-D-quinosyl-(1 2)-O-β-D-glucosyl-(1 3)-O-[β-D-fucosyl-(1 4)]-O-α-L-rhamnosyl-(1 2)-O-β-D-glucosyl-(1 2)-O-β-D-fucopyranoside.

2.6. Heptasaccharides

Four heptasaccharide-typed resin glycosides were found in the plant of Convolvulus arvensis, namely, arvensis A-D (207210) (Figure 7) [64]. The oligosaccharide core of these compounds is β-D-fucopyranosyl-(1 4)-O-β-D-glucopyranosyl-(1 3)-O-α-L-rhamnopyranosyl-(1 2)-[O-β-D-glucopyranosyl-(1 3)]-O-β-D-glucopyranosyl-(1 2)-[O-α-L-rhampyranosyl-(1 6)]-O-β-D-glucopyranoside, which is similar to the hexasaccharide core of PM1–PM5 (199203). However, a fucose unit is added to the last sugar unit of arvensis A-D (207210). There are two variations of acyclic aglycones in these four compounds, firstly, 12S-pentadecanoic acid for arvensis A (207) and C (209), and secondly, 12S-hexadecanoic acid for arvensis B (208) and D (210).
The plant Calystegia hederacae produced a series of heptasaccharide-type resin glycosides [44,59,60]. Calyhedic acids C-D (211212) (Figure 7) have a similar heptasaccharide core, namely, 11-O-D-glucosyl-(1 → 3)-O-α-L-rhamnosyl-(1 → 2)-[O-β-D-glucosyl-(1 → 6)-O-β-D-glucosyl-(1 → 6)-O-β-D-glucosyl-(1 → 3)]-O-β-D-glucosyl-(1 → 2)-O-β-D-quinovopyranoside. Acyclic aglycones distinguish 11S-hydroxyhexadecanoic and 12S-hydroxyhexadecanoic acids for calyhedic acid C (211) and calyhedic acid D (212), respectively [44]. Three heptasaccharide-type resin glycoside compounds, namely, calyhedins IV–VI (213215) (Figure 7) [59]. The heptasaccharide core of these three compounds is similar to calyhedic acids C-D (211212). The aglycones of calyhedic acids IV–V (213214) form an intramolecular ester bond with glc‴ (1,2‴-ester type), while calyhedic acid VI (215) forms lactone bonds with glc′ (1,2′-ester type). The most recent findings of the heptasaccharides-based calyhedins series from C. hederacea are calyhedins VII–IX (216218) (Figure 7) [60]. Structurally, there is no significant difference between the heptasaccharide cores of calyhedins VII–IX (216218), calyhedins IV–VI (213215), and calyhedic acids C-D (211212), but the difference is in the location of the intramolecular ester. Calyhedins VII–IX (216218) have two types of rings, a 27-membered ring for calyhedins VII (216) and a 22-membered ring for calyhedins VIII–IX (217218).
Heptasaccharides-type resin glycosides with a series of QMs and quamoclinic acid were isolated from Quamoclit pennata [27,63,65]. QM1 (219) and QM7 (220) (Figure 7) have a heptasaccharide core that is almost similar to the heptasaccharide of maltifidinic acid C-D (192193). However, there is an addition of one more sugar unit, rha″, at the end of the oligosaccharide, giving a framework of α-L-rhamnopyranosyl-(1 3)-O-β-D-quinovopyranosyl-(1 2)-O-β-D-glucopyranosyl-(1 3)-[O-β-D-quinovopyranosyl-(1 4)]-O-α-L-rhamnopyranosyl-(1 2)-O-β-D-glucopyranosyl-(1 2)-O-β-D-fucopyranoside. The aglycone of QM1 (219) is determined as methyl 3S,11S-dihydroxytetradecanoate [63], while QM7 (220) is only methyl [65]. Another series of the heptasaccharide resin, namely quamoclinic acid E-F (221222) (Figure 7), has the same aglycone moiety with 3S,11S-ipurolic acid, and 3S,11S-dihydroxytetradecanoic acid [27]. However, the heptasaccharide core of both resin glycosides is different. Quamoclinic acid F (222) has a similar oligosaccharide core to QM1 (219), as opposed to quamoclinic E (221), in which the two sugar units, rhamnose, and quinovose, are replaced by fucose resulting in the core of β-D-fucopyranosyl-(1 3)-O-β-D-quinovopyranosyl-(1 2)-O-β-D-glucopyranosyl-(1 3)-[O-β-D-fucopyranosyl-(1 4)]-O-α-L-rhamnopyranosyl-(1 2)-O-β-D-glucopyranosyl-(1 2)-O-β-D-fucopyranoside.

2.7. Bidesmosides

Jalapinoside B (223) (Figure 8) was isolated from the methanolic extract of the roots of Ipomoea purga [16]. Structurally, this compound is categorized as resin glycoside in the macrocylic bidesmoside class and is characterized by the presence of additional sugar units attached to the aglycone outside the main framework of the oligosaccharide core. This core is similar to the hexasaccharide of multifidin VII (194). Jalapinoside B (223) contains two aglycones and one macrocyclic aglycone 11S-convolvulinic acid (1,3″-lactone) attached to the skeleton of the main oligosaccharide. The second aglycone is an aglycone placed outside the oligosaccharide framework, which has an additional sugar unit of quinovose at the decanoic acid (C-7), which is structurally similar to the monosaccharide-type resin glycoside, quamoclinic acid B (2). Therefore, the framework to the jalapinoside B (223) is 11S-hydroxytetradecanoic acid 11-O-β-D-quinovopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 3)-[O-β-D-fucopyranosyl-(1 → 4)-(2-O-7S-hydroxydecanoic acid-O-β-D-quinovopyranosyl-(1 → 2)]-O-α-L-rhamnopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 2)-O-β-D-quinovopyranoside (1,3″-lactone). Jalapinosides I–II (224225) (Figure 8) were reportedly isolated [66,67]. These compounds have a shared similar oligosaccharide framework with jalapinoside B (223). However, an additional quinovose unit is not attached to the decanoic acid, unlike the 3S,11S-ipurolic acid (C-3). Jalapinoside I–II (224225) has a structural framework of 3S,11S-hydroxytetradecanoic acid 11-O-β-D-quinovopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 3)-[O-β-D-fucopyranosyl-(1 → 4)]-O-α-L-rhamnopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 2)-O-β-D-quinovopyranosyloxy-3-[O-β-D-quinovopyranosyloxy]-1,3″-lactone.
A resin glycoside with a macrocylic bidesmoside framework, named multifidin IX (226) (Figure 8), was isolated from Quamoclit × multifida [51]. This compound has a similar oligosaccharide core as multifidin VIII (195) with additional sugar units in the form of quamoclinic acid B (2) and the acyclic aglycone of methyl 3S,11S-ipurolate. Its structural skeleton is methyl 3S,11S-ipurolate 11-O-β-D-quinovopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 3)-[O-β-D-fucopyranosyl-(1 → 4)]-(2-O-7S-hydroxydecanoic acid 7-O-β-D-quinovopyranosyl-(1 → 2)]-O-α-L-rhamnopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 2)-O-β-D-fucopyranoside. Other macrocylic bidesmoside-type resin glycosides that were isolated from this plant included multifidinic acids F-G (227228) (Figure 8) [68]. Multifidinic acid F (227) has an oligosaccharide core, which is similar to the hexasaccharide core of multifidin VII (194), and the additional glucose unit attached to the 3S,11S-ipurolic acid (C-3), meaning that the structural skeleton of 3S,11S-hydroxytetradecanoic acid can be given as 11-O-β-D-quinovopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 3)-[O-β-D-fucopyranosyl-(1 → 4)]-O-α-L-rhamnopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 2)-O-β-D-quinovopyranosyl-3-O-β-D-glucopyranoside. Multifidinic acid G (228) shares a similar oligosaccharide core as multifidin VIII (195), with an additional glucose unit attached to the 3S,11S-ipurolic acid (C-3), revealing the structural skeleton to be 3S,11S-hydroxytetradecanoic acid 11-O-β-D-quinovopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 3)-[O-β-D-fucopyranosyl-(1 → 4)]-O-α-L-rhamnopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 2)-O-β-D-fucopyranosyl-3-O-β-D-glucopyranoside.
Purgic acids C-D (229230) (Figure 8) are a series of bidesmosidic-type resin glycosides found in Ipomoea purga [67]. The main oligosaccharide core of purgic acid C (229) is similar to the hexasaccharide core of multifidin VII (194) with an acyclic aglycone, 3S,11S-hydroxytetradecanoic acid. Structurally, the compound is similar to that of multifidinic acid F (227), but a bidemosidic moiety is a quinovose unit attached to the dihydroxy aglycone (C3). This leads to the structural framework of purgic acid C (229), which is 11-O-β-D-quinovopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 3)-[O-β-D-fucopyranosyl-(1 → 4)]-O-α-L-rhamnopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 2)-O-β-D-quinovopyranosyl-3S,11S-hydroxytetradecanoic acid 3-O-β-D-quinovopyranoside. Purgic acid D (230) has a similar oligosaccharide core and bidesmosidic sugar units as purgic acid C (229). However, the aglycone of this compound is 3S,11S-hydroxyhexadecanoic acid, hence the skeleton is 11-O-β-D-quinovopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 3)-[O-β-D-fucopyranosyl-(1 → 4)]-O-α-L-rhamnopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 2)-O-β-D-quinovopyranoside-3S,11S-hydroxyhexadecanoic acid 3-O-β-D-quinovopyranoside.
Most macrocylic bidemoside-types of resin glycosides are found in the QMs series of Quamoclit pennata plants [63,65,69]. These QMs are classified as bidemosides because the structures have one additional sugar unit, quamoclinic acid B (2), which is commonly attached to the C-2 unit of the rhamnose sugar. QM2 (231) (Figure 8) has a heptasaccharide core and aglycones similar to QM1 (219). However, QM2 (231) has a bidemosidic part that is absent in QM1 (219), while QM2 (231) has a fully structural skeleton which is given as methyl 3S,11S-dihydroxytetradecanoate 11-O-α-L-rhamnopyranosyl-(1 → 3)-O-β-D-quinovopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 3)-[O-β-D-quinovopyranoside-(1 → 4)]-O-(2-O-7S-hydroxydecanoic acid 7-O-β-D-quinovopyranoside))-α-L-rhamnopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 2)-O-β-D-fucopyranoside. QM3 (232) (Figure 8) has a similar heptasaccharide core as quamoclinic acid E (221) with a bidesmoside part of quamoclinic acid B (2) and an aglycone of methyl 3S,11S-ipurolate in QM3 (232), giving a structural framework of 3S,11S-dihydroxytetradecanoate 11-O-β-D-fucopyranosyl-(1 → 3)-O-β-D-quinovopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 3)-[O-β-D-fucopyranosyl-(1 → 4)]-O-(2-O-7S-hydroxydecanoic acid 7-O-β-D-quinovopyranoside))-α-L-rhamnopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 2)-O-β-D-fucopyranoside. QM5 (233) (Figure 8) has a similar hexasaccharide core to QM4 (204), which comprises a resin without an aglycone but a methyl group QM5 (233). It has a quamoclinic acid B (2) as a bidemosidic moiety to obtain methyl-D-quinovopyranosyl-(1 → 3)-O-β-D-quinovopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 3)-[O-β-D-quinovopyranosyl-(1 → 4)]-O-(2-O-7S-hydroxydecanoic acid 7-O-β-D-quinovopyranosyl-(1 → 2))-α-L-rhamnopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 2)-O-β-D-fucopyranoside as its skeleton [63]. QM8 (235) (Figure 8) [65] has a similar structure to QM5 (233), but QM8 (235) contains α oriented methyl group, while QM5 (233) is in the β-oriented methyl group. Methylated glycoside is present in the QM6 (234) (Figure 8), which has the structural skeleton of a heptasaccharide core, similar to the α-oriented methyl group of QM7 (220). In QM6 (234), there is a bidemosidic quamoclinic acid B (2); hence, the framework becomes methyl-L-rhamnosyl-(1 → 3)-O-β-D-quinosyl-(1 → 2)-O-β-D-glucosyl-(1 → 3)-[O-β-D-quinosyl-(1 → 4)]-O-(2-O-7S-hydroxydecanoic acid 7-O-β-D-quinosyl-(1 → 2)-α-L-rhamnosyl-(1 → 2)-O-β-D-glucosyl-(1 → 2)-O-β-D-fucopyranoside [65]. The heptasaccharide core, bidemosidic moiety, and aglycone of QM9 (236) (Figure 8), are similar to QM3 (232), which is methyl 3S,11S-dihydroxytetradecanoate 11-O-β-D-Fuc-(1 → 3)-O-β-D-Qui-(1 → 2)-O-β-D-Glc-(1 → 3)-[O-β-D-Fuc-(1 → 4)]-O-(2-O-7S-hydroxydecanoic acid 7-O-β-D-Qui)-α-L-Rha-(1 → 2)-O-β-D-Glc-(1 → 2)-O-β-D-Fuc. The difference between QM9 (236) and QM3 (232) is in the organic acid group attached to the C-4 Fuc′ sugar unit, in which nilic acid (Nla) and tiglic acid (Tig) are found in QM3 (232) and QM9 (236), respectively. QM10 (237) (Figure 8) has a similar hexasaccharide core and bidesmosidic moiety as QM5 (233). However, its methyl ipurolate aglycone provides a fully structural skeleton as methyl 3S,11S-dihyroxytetradecanoate 11-O-β-D-Qui-(1 → 3)-O-β-D-Qui-(1 → 2)-O-β-D-Glc-(1 → 3)-[O-β-D-Qui-(1 → 4)]-O-(2-O-7S-hydroxydecanoic acid 7-O-β-D-Qui-(1 → 2))-α-L-Rha-(1 → 2)-O-β-D-Glc-(1 → 2)-O-β-D-Fuc [65]. Another series of QMs resin glycosides are QM11–QM13 (238240) (Figure 8) [69], which have a similar structural framework to QM10 (237). The difference between QM11 and QM13 (238240) lies in the variation of the position of the organic nilic acid (Nla). QM14 (241) (Figure 8) was the last series of QM compounds found in Q. pennata. Structurally, the oligosaccharide core is similar to QM13 (240). Meanwhile, the structural characteristic of QM14 (241) is the presence of two bidesmosidic parts, quamoclinic acid B (2), which is common in the QMs series, and an additional glucose unit attached to C-3 methyl ipurolate as a second bidemosidic. Therefore, the skeleton of this compound is identified as 11-O-β-D-quinovopyranosyl-(1 → 3)-O-β-D-quinovopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 3)-[O-β-D-quinovopyranosyl-(1 → 4)]-O-(2-O-7S-hydroxydecanoic acid 7-O-β-D-quinovopyranosyl-(1 → 2))-α-L-rhamnopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 2)-O-β-D-fucopyranosyl-methyl 3S,11S-dihyroxytetradecanoate 3-O-β-D-glucopyranoside [69].
Quamoclit pennata also produced quamoclinic acid G-H (242243) (Figure 8) [70]. On the other hand, the bidesmosidic quamoclinic acid G-H (242243) has a glucose unit attached to methyl ipurolate (C-3 aglycone). Quamoclinic acid G (242) has a hexasaccharide core similar to that of QM10–QM13 (237240), although the bidesmosidic moiety is unattached to the rhamnose unit compared to the aglycone. Therefore, the compound’s skeleton becomes 11-O-β-D-quinosyl-(1 → 3)-O-β-D-quinosyl-(1 → 2)-O-β-D-glucosyl-(1 → 3)-[O-β-D-quinosyl-(1 → 4)]-O-α-L-rhamnosyl-(1 → 2)-O-β-D-glucosyl-(1 → 2)-O-β-D-fucosyl-methyl 3S,11S-dihyroxytetradecanoate 3-O-β-D-glucopyranoside. Quamoclinic acid H (243) has a heptasaccharide core that is similar to QM2 (231) in the absence of the quamoclinic acid B (2) group in the rhamnosyl sugar. The additional glucose unit is attached to the C-3 aglycone, methyl ipurolate. The compound has a structural skeleton of 11-O-α-L-rhamnosyl-(1 → 3)-O-β-D-quinosyl-(1 → 2)-O-β-D-glucosyl-(1 → 3)-[O-β-D-quinosyl-(1 → 4)]-O-α-L-rhamnosyl-(1 → 2)-O-β-D-glucosyl-(1 → 2)-O-β-D-fucosyl-methyl 3S,11S-dihydroxytetradecanoate 3-O-β-D-glucopyranoside.

2.8. Oligosaccharide Ester-Type Dimer

Batatins III–VI (244247) (Figure 9) were obtained from Ipomoea batatas and grouped into an oligosaccharides ester-type dimer [71]. Unlike the common resin glycoside, an ester-type dimer contains two oligosaccharide units esterified by two hexadecanoic (palmitic acids) [76]. These compounds were identified to have a similar oligosaccharide core as the operculinic acid C [77], with a tetrasaccharide skeleton α-L-rhamnosyl-(1 → 4)-O-α-L-rhamnosyl-(1 → 4)-O-α-L-rhamnosyl-(1 → 2)-O-β-D-fucopyranoside for both units A and B. In unit A, the aglycone of jalapinolic acid is in the form of macrocylic lactone-1,3″, while B has an acyclic aglycone in which the C-1 ester of jalapinolic acid (unit B) is bonded to the C-3 sugar Rha′ (unit A). The difference in the structure of batatins III–VI (244247) is due to the variety of organic acids [71]. Furthermore, other oligosaccharide ester-type dimers from I. batatas is batatin VII (248) (Figure 9) [41], which comprises units A and B with a pentasaccharide skeleton of 11-O-α-L-rhamnosyl-(1 → 3)-O-[α-L-rhamnosyl-(1 → 4)]-O-α-L-rhamnosyl-(1 → 4)-O-α-L-rhamnosyl-(1 → 2)-O-β-D-fucopyranoside, similar to those of the aquaterins [39] and pescapreins series [53]. The 11S-jalapinolic acid aglycone in unit A forms a macrocylic lactone with a 1,3″ ester type, which is an acyclic aglycone in unit B. Batatin VII (248) has a connecting point of the C-1 jalapinolic acid (unit B), bound to the C-3 of Rha″ (unit A).
Five dimeric resins were found in Cuscuta chinensis, namely cuses 8–12 (249253) (Figure 9) [4]. Unlike other dimeric ester-type compounds, where the term ‘dimer’ refers to the similarity of the oligosaccharide framework between the A and B units, cuses 8–12 (249253) are different. The A unit of the first sugar group on the anomeric carbon is poorly protected, and tautormerism occurs. Therefore, the anomeric carbon of the first sugar in unit A of cuses 8–12 (249253) reacts with p-anisidine to obtain an aminoalditol derivative. Unit A of cuses 8–12 (249253) is a trisaccharide that is similar to cuses 5–7 (46), while unit B is a tetrasaccharide with variations in the first sugar and the aglycone, which distinguishes the structure from cuses 8–12 (249253). Unit B of cuses 8 (249) is the tetrasaccharide of 11-O-α-L-Rha-(1 → 3)-O-α-L-Rha-(1 → 2)-O-β-D-Glc-(1 → 2)-O-β-D-Glc, which is attached to the C-11 aglycone, 11S-convolvulic acid. Furthermore, cuses 9 (250) has a similar tetrasaccharide core to cuses 8 (249), but cuses 9 (250) is attached to the aglycone 11S-jalapinolic acid. The tetrasaccharide core of unit B of cuses 10 (251) is first replaced with quinovose sugar and attached to the aglycone of convolvulic acid. This turns the structural framework into 11S-convolvulic acid 11-O-α-L-Rha-(1 → 3)-O-α-L-Rha-(1 → 2)-O-β-D-Glc-(1 → 2)-O-β-D-Qui. Furthermore, cuses 11 (252) and 12 (253) have a xylose unit as the first sugar unit in the tetrasaccharide to give the structural skeleton of 11-O-α-L-Rha-(1 → 3)-O-α-L-Rha-(1 → 2)-O-β-D-Glc-(1 → 2)-O-β-D-Xyl. The difference between cuses 11 (252) and 12 (253) is the aglycone, in which 11S-convolvulic and 11S-jalapinolic acids, respectively.
Purgin I–III (254256) (Figure 9) was isolated from Ipomoea purga, which has a structural framework of the dimeric ester of resin glycoside [6,9]. Meanwhile, Purgin I (254) has a similar oligosaccharide core to the pentasaccharide series of merremins [50], murasakimasarins [52], purginosides [9], and quamoclins [55], with a skeleton of 11-O-β-D-glucosyl-(1 → 3)-O-[α-L-rhamnosyl-(1 → 4)]-O-α-L-rhamnosyl-(1 → 4)-O-α-L-rhamnosyl-(1 → 2)-O-β-D-fucopyranoside. The aglycone 11S-hydroxyhexadecanoic acid in unit A is a 1,2″-macrocylic lactone, while the pentasaccharide in B is attached to the C-11 acyclic aglycone, with the C-1 aglycone, located at C-4 of the glucose unit [9]. Purgin II–III (255256)[6] has a different oligosaccharide core from purgin I (254), which is 11-O-β-D-glucosyl-(1 → 3)-O-[α-L-rhamnosyl-(1 → 4)]-O-α-L-rhamnosyl-(1 → 4)-O-α-L-rhamnosyl-(1 → 2)-O-β-D-quinovopyranoside. The structure is similar to the pentasaccharide-type multifidin IV (121) [51] and operculinic acid B [77]. Furthermore, the connecting site of units A and B and their aglycones is similar to purgin I (254).
Three dimeric esters of resin glycosides were isolated tyrianthins C-E (257259) (Figure 9) from Ipomoea tyrianthina [17]. The tetrasaccharides core in units A and B is similar to the skeleton of tyrianthinic acid VI (48), which is 11-O-β-D-quinovopyranosyl-(1 → 4)-O-α-L-rhamnopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 2)-O-β-D-quinovopyranoside. In unit A, the aglycone, 11S-hydroxyhexadecanoic acid forms a macrocylic lactone-1,3‴, while in unit B, the structure has an acyclic aglycone, with the connecting site between the C-1 aglycone and C-4 qui′ (unit A). The difference between these three compounds is the variety of organic acids.
The Ipomoea stans plant produced a resin glycoside, stansin A (260) (Figure 9), which was classfied into the dimeric ester type because it has two trisaccharide units, which are esterified with hexadecanoic acid [16]. The trisaccharide core of this compound is 11-O-α-L-rhamnopyranosyl-(1 → 2)-O-β-D-glucopyranosyl-(1 → 2)-O-β-D-quinovopyranoside, which is similar to the core stansoic acid A (13). In unit A, the 11S-hydroxyhexadecanoic acid aglycone forms a macrocylic lactone-1,3‴. Meanwhile, there is an acyclic aglycone in unit B, with the linking site of the C-1 aglycone (unit B) attached to the C-4 of rhamnose (unit A).
Wolcottine I (261) (Figure 9) is a dimeric ester-type resin glycoside obtained from Ipomoea wolcottina [24]. The oligosaccharide core of this compound is similar to operculinic acid C (168), a tetrasaccharide with the structural skeleton of 11-O-α-L-rhamnopyranosyl-(1 → 4)-O-α-L-rhamnopyranosyl-(1 → 4)-O-α-L-rhamnopyranosyl-(1 → 4)-O-β-D-fucopyranoside. The 11S-jalapinolic acid aglycone in unit A forms a macrocylic lactone 1,3″, while the oligosaccharides in unit B are attached to the C-11 of the acyclic aglycone, with the connecting site between the C-1 aglycone and the C-3 sugar Rha″ unit A.
This classification comprises several pieces of information with a particular species of plants responsible for producing a specific group of resin glycosides. For example, the Operculina and Jalapae species only produced hexasaccharides-typed resin glycosides, while Porana and Dichondra had trisaccharides and tetrasaccharides groups. Meanwhile, the Ipomoea species is the only species that produced a wide range of resin glycosides, from trisaccharides to octasaccharides, down to the dimer class. The distribution of resin glycosides in several species is shown in Table 2.

3. Biological Activity

Resin glycosides are not only unique in their structures but also have a wide range of biological properties. Several studies on glycosides revealed cytotoxic activities [4,11,20,29,39,42,47,51,59,75,78,79], multidrug-resistance (MDR) modulating properties [3,6,7,12,23,24,49,50,53,57,66,67,80], α-glucosidase inhibitory potential [13,36], antiviral [14], anti-metastatic [15], anti-proliferative [16,29], neuroprotective and anticonvulsant activities [10], sedative effect and gamma-aminobutyric acid (GABA) release [17,81], anti-inflammatory [18,20], antimycobacterial activity [82], the antidepressant effect [19], anticonvulsant effect/anti-seizure for epilepsy [19], the antimalarial effect [22], anti-diarrhoea [20], downregulation aquaporin 3 [15,20], and antibacterial activities [21] (Table 3).

4. Isolation Techniques

The previous reviews described the use of various processes to obtain resin glycosides from plants. However, this review analysed the steps in the isolation of resin glycosides from Convolvulaceae plants.
Most isolation techniques of resin glycosides were initiated by standard extractive procedures. The aerial part, whole part, stem, root, or leaves of the plant was macerated either by methanol [4,5,14,22,30,34,45,55,57,58,66,67,83,84], ethanol [8,29,36,37,39,40,42,43,47,48,50,53,56,64,85,86], hexane [12,74,87], a mixture of acetone:water [23], subsequent hexane/chloroform/dichloromethane [7,17,30], chloroform [6,9,24,41,88], ethyl acetate [18], subsequent hexane/dichloromethane/methanol [82], or dichloromethane [19]. The process of percolation using successive hexane, dichloromethane, and methanol [28] or chloroform/acetone [89], and also the process of soxhlet extraction using hexane/chloroform, were reported [49].
Mostly, methanolic and ethanolic extracts were suspended in water before the extraction. The suspension was extracted either by successive petroleum ether/dichloromethane/ethyl acetate/butanol [29,32,39,40,47,48,64], petroleum ether/ethyl acetate/butanol [56], n-hexane/butanol [55], 70% methanol/hexane [58], ethyl acetate/butanol [30,34,45], petroleum ether/chloroform [36], chloroform/water [8,85] petroleum ether/ethyl acetate [50], ethyl acetate/water [5,89], butanol/water [83], or hexane/chloroform/methanol [84]. Several studies took advantage of the butanol fraction to be chromatographed further to obtain resin glycosides [48,64]. Others used petroleum ether [50], ethyl acetate [34], or dichloromethane [29] fractions to obtain the resin glycosides. According to another report, the first methanolic extract obtained from maceration was directly purified by chromatographic techniques without prior partition [57]. Chromatographic techniques that were reportedly used are based on silica gel, NH2 gel, sephadex LH-20 [89], microporous resin D101 [4,53], Diaion HP-20, ODS, and preparative HPLC. Purifications were not straightforward, with peak-shaving and heart-cutting techniques used to recycle the preparative HPLC to obtain the resin glycosides [12]. Numerous reports describe the conversion of fractions into their acylated [64] or reductive forms [4] to improve the poor resolution of glycosidic acid. The chloroform extract was directly chromatographed [6,41] through the preparative-scale recycling of HPLC [66] to obtain resin glycosides.
Table 4 shows the representative details of isolation techniques of resin glycosides, starting from different solvents in the maceration process.

5. Structural Determination

The structural determination of resin glycosides has been described in preliminary reviews thoroughly [1,2]. In this review, the sequence of the general structural determination of resin glycosides will still be explained based on the reports, even though the majority are mostly similar to the ones reported in the prior reviews.
The structural determination of resin glycosides, spectroscopic, and analytical methods involved infrared, ultraviolet, nuclear magnetic resonance, mass spectrometry, specific rotation, HPLC, and GC-MS [3,4,5,6,7,12,16,17,23,24,29,32,34,36,37,40,41,42,43,47,48,49,50,55,56,58,63,64,65,66,67,68,69,90]. Most structural determination sequences involve alkaline and acidic hydrolysis, which breaks the structure into a simpler and more straightforward structures. The first alkaline hydrolysis was attempted to decay the structure into water-soluble resin glycoside and organic fatty acids, which were then analysed through the GC-MS experiment and compared with authentic samples. Furthermore, their absolute configurations were determined by optical rotation measurements, which were compared with the authentic samples. The water-soluble glycoside was then hydrolysed in acidic conditions to break the resin glycoside into glyosidic acid fractions. These sugar fractions are analysed via several methods, including GC-MS, HPLC, and specific rotations. Several studies stated that the water-soluble glycoside was partitioned with n-BuOH per acetylated, recycled through HPLC, and then characterized by NMR [24,67].
The absolute configuration of the sugars is determined by comparing their specific rotations and GC-MS data with the given standards. The relative configuration of the β-configuration is determined via the large coupling constants of the anomeric protons (J = 7.5, 8.0, 8.5 Hz). The α-configuration was determined via the chemical shifts of C-5 in the carbon NMR, while the Mosher method was used for the 11S-configuration [42].
An infrared experiment was used to determine the main functionalities present in the resin glycoside structure. The three functional groups contained in the infrared spectrum are hydroxyl, alkyl, and carbonyl groups [8,36,37,57]. Ultraviolet experiments showed that resin glycoside absorbs at a wavelength of 280 nm [36].
Mass spectroscopy, including ESI-MS, FAB-MS, HRAPCI-MS [57], APCI-MS [57], can be used to analyse fragments obtained during the experiments to erase overall structural determination [29,42,65]. Furthermore, the NMR experiment is one of the most powerful methods used in the structural elucidation of resin glycoside. All protons and carbons of the structure are characterized by ID and 2D-NMR. The presence of sugars can be determined by analysing the anomeric signals in the proton and carbon NMR. Meanwhile, carboxyl groups in the structure are evaluated through carbon NMR. All sugar connectivity’s in the structure, including the ester location, are determined using 1H-NMR, 13C-NMR, HSQC, COSY, and HMBC. TOCSY was also used in the structural determination [41]. However, it is not mandatory but can be added to the manuscript if the discussion is unusually long or complex. The steps in the structural elucidation of resin glycosides are described in Table 5, which presents the structural determination steps of eight representatives of the resin glycosides, including monosaccharide-, trisaccharide-, tetrasaccharide-, pentasaccharide-, hexasaccharide-, heptasaccharide-, bidesmoside-, and ester dimer-based resin glycosides.

6. Synthesis

The synthesis of resin glycoside is challenging because it involves stereoselective glycosylation, macrolactonization, and deprotection steps. Some of the glycosylations used in the synthesis process include Schmidt glycosylation which requires TMSOTF as an activator [11,90,91,92,93] and interrupted Pummerer reaction-mediated (IPRm) glycosylation [94,95].
Macrocyclization administered to the resin glycoside was undertaken in several ways, including Keck Macrolactonization [90,92,94], the Corey–Nicolaou macrolactonization method [93], ring closing metathesis [11], and through intramolecular glycosylation [91].
An optimized total synthesis of the 22-membered ring of analogue (273) and ipomoeassin F was reported in 2020 [11]. The dissacharide core of the synthetic target was built from the glucose donor (262) and fucose acceptor (263) mediated by TMSOTf to obtain dissacharide (264). The deprotection of two levulinoyl groups using NH2NH2.H2O, AcOH in Pyridine led to a diol that was ready to be subsequently incorporated by the tiglate group (Mukaiyama method) and was sylilated using TESOTf to produce compound (267) in good yield. This compound was further converted to diol (268) using TFA in chloroform before it was conjugated with 4-oxo-10-undecenoic acid and mediated by EDC. The product (269)/(270) was then cyclized through ring-closing metathesis followed by hydrogenation to produce the cyclic product (271)/(272). The hydroxy group was then esterified using the Mukaiyama method with cinnamic acid, followed by TES and TBS deprotections to obtain the desired analogue (273)/ipomoeassin F (274), as shown in Scheme 1.
The synthetic approach in the total synthesis of tricolorin A (tetrasaccharide) was interrupted using the Pummerer reaction-mediated (IPRm) glycosylation method toward the construction of complex oligosaccharides and glycoconjugates [94]. IPRm was applied to connect two disaccharide fragments (280 and 287) and also to build them individually.
Fragment (280) was constructed from the OPSB glycosyl donor and the SPTB glycosyl acceptor (278), as shown in Scheme 2. The acceptor (278) was synthesized via the subsequent deacetylation of rhamnoside (275) and isopropylidenation leading to (276) with free hydroxyl groups at position four. The obtained compound was esterified with (S)-(+)-2-methylbutyric acid in the presence of EDCI and DMAP to give a high yield (277). The following ketal deprotection and acylation of the hydroxyl group at position two resulted in the SPTB glycosyl acceptor (278) that was ready to be glycosylated with the OPSB donor using the IPRm method. The glycosylation produced (279) that was further selectively oxidized using m-CPBA to produce the high-yield SPSB donor (280).
Fragment (287) was prepared from the SPSB donor (281), as shown in Scheme 3. This donor was glycosylated with methyl 11(S)-jalapinolate (282) in the presence of triflic anhydride and 2,6-di-tert-butyl-4-methylpyridine (DTBMP) to obtain compound (283). The deacetylation, isopropylidenization, IPRm glycosylation, and safonification resulted in the aliphatic chain-bearing disaccharide (286). The macrolactonization of (286) was found to be critical in the overall synthesis. Meanwhile, the use of 2,2′-pyridyl disulfide and triphenylphosphine gave a 19-membered ring, and 74% yield of the macrolactone (287).
After obtaining fragments (280) and (287), the last step was to combine both using IPRm glycosylation, as shown in Scheme 4. However, this step only yielded moderate results and about 74% when it was run under −78 oC conditions. The replacement of SPSB-based (289) donors with OPSB-based (290) increased the yield to 92%. A subsequent desopropylidenization, de-benzylidenization, and debenzylation led to the desired tricolorin A (293).
Total synthesis of the most complex resin glycosides, which were isolated to the date-calysolin IX 27-membered ring, macrolactone was carried out using intramolecular glycosylation during the ring-closing step [91]. This approach aimed to reduce the steps in the synthesis with the retrosynthetic analysis consisting of three fragments, namely, the glycosyl donor (301), glycosyl acceptor (298), and disaccharide unit. Fragment (298) was synthesized from the jalapinolic ester (294) and was previously prepared from undecyne and hexanal, as shown in Scheme 5. The reaction of (294) and 6-deoxyglucal (295) stereoselectively conjugated using oxone, followed by zinc chloride, led to the opening of the oxirane ring to produce β-glycoside (296) with the free OH group at the C-2 position suitable for the next glycosylation step. The glycosylation step was carried out using trichloroacetimidate in the presence of catalytic amounts of TMSOTf to produce disaccharide (298) in 83% yields.
Fragment (301) was synthesized using the same method as the synthesis of fragment (298), which involved glucal (299) and thioglycoside (300), as shown in Scheme 6.
Before it was conjugated with (301), the protecting groups of OH and COOH were deprotected to produce intermediate (302), as shown in Scheme 7. The product (302) was conjugated with (301) by placing the bulky protected group of tert-butyldimethylsilyloxy at a neighbouring C-3 position to prevent intramolecular esterification between the free OH and COOH of (302). By applying this strategy, (304) could be successfully combined with fragment (301) using DIC/DMAP to obtain (303).
The regioselective intramolecular glycosylation of (303) was undertaken to produce compound (306), as shown in Scheme 8. The introduction of the disaccharide unit has not been reported in due course.
Two interrupted Pummerer reactions mediated with (IPRm) glycosylations were developed in the synthesis of resin glycoside [95]. This made it easy for the latent glycosides to oxidize the active glycosyl donors for the subsequent glycosylations with high efficacy. Murucoidins IV and V were synthesised through the application of the IPRm glycosylations and a sequential transient protection−glycosylation−deprotection protocol. Murucoidins IV and V were assembled in a convergent manner via a [3 + 2] IPRm glycosidic coupling and a macrolactonization.
There were efficient syntheses of the three macrocyclic lactone units by the use of a Keck macrolactonization approach [92]. The Keck macrolactonization is a carbodiimide-promoted esterification to form a macrolactone in the presence of a nitrogenous base. This reaction was applied in the preparation of macrolactones. The synthesis of macrolactone (315) started from allyl 3-O-PMB-a-L-rhamnopyranoside (307), as shown in Scheme 9. The overall reaction involved protection, deprotection, and glycosylation. The activation of anomeric hemiacetal was carried out using trichloroacetonitrile (CCl3CN) and 1,8-diazabicyclo[5.4.0] undec-7-ene (DBU) to produce trichloroacetimidate (309). The glycosilation steps (309) and (310) involved the Schmidt condition of a catalytic amount of trimethylsilyl trifluoromethane-ulfonate (TMSOTf) and the furnishing of disaccharide lipid (311) in a 97% yield. Macrolactonization took advantage of the Keck method using DCC/DMAP/PPTS.
Zhu et al. (2013) reported a total synthesis of dimeric batatin VI [90] using the convergent approach through [5 + 3] glycosidic coupling for the synthesis of (329) and (335). Pentasaccharide alcohol (329) was prepared by assembling three building blocks [2 + 2 + 1] (323324327), while trirhamnosyl trichloroacetimidate (335) was easily prepared from (332) and (333). An acyl-type functionality acting as a neighbouring participating group was incorporated at the 2-position of each donor to ensure the formation of 1,2-trans stereochemistry of each glycosylation. Groups of p-methoxybenzyl (PMB), levulinoyl (Lev), and isopropylidene groups were protected. Macrolactonization took advantage of a modified Keck method using DCC, pyridine, PPTS, 1,2-dichloroethane (DCE), and reflux at 2 hour.
Pentasaccharide alcohol (329) was built from fragments of (323), (324), and (327), as shown in Scheme 10. Fragment (324) was prepared from known compound using Schmidt glycosylation (TMSOTf). Meanwhile, (329) was built from preparing fragment (323) through the subsequent PMB deprotection of (316), followed by the esterification of the product alcohol (317) with the jalapinolic acid derivative (318) using DCC/DMAP. TBS deprotection of the resulting (319), TMSOTf activation, coupled with rhamnosyl imidate to give disaccharide (322). The deallylation of (322) resulted in compound (323), which was ready to be coupled with fragment (324) prior to activation using TMSOTf. Schmidt glycosylation was trialled at first with a 45% yield of the product. An inverse glycosylation was trialled to improve the yield. Fragment (324) was activated using TMOSTf and reacted with (323) to give compound (326) with an 82% yield. The conjugation of product (326) and (327), followed by Lev deprotection was carried out to obtain pentasaccharide alcohol (329).
Intermediate (335) was synthesized through subsequent Lev protection and TBS deprotection before it was glycosylated with donor (333), as shown in Scheme 11. Finally, fragments (335) and (329) were combined through inverse O-glycosilation to produce a 96% yield of octasaccharide (336). Global deprotection of the groups in (336) was carried out to obtain batatin VI (337).
Xie et al. (2010) reported a total synthesis of batatoside L (362) [93], which was carried out using a convergent strategy between fragments (358) and (360) through glycosylation. The Corey–Nicolaou macrolactonization method was involved in the formation of macrocyclic subunit (358). Four glycosyls trichloroacetimidate donors (350), (351), (343), and (345), and one L-rhamnosyl acceptor (348) were used in the preparation of fragments (360) and (356). The acyl groups serving as the necessary neighbouring participatory groups were incorporated at the 2-OH position of each donor to ensure that the desired 1,2-trans stereoselectivity was obtained for the glycosylation process. Subsequential protection followed by activation (imidation) was involved in the formation of fragment (343) from (338). Similar steps were used in the formation of (345). In addition, the (348) compounds were prepared through the aminolysis of intermediate (345), followed by respective esterification and desylilation, as shown in Scheme 12.
Macrolactone (358) was synthesized using the Schmidt glycosylation method, as shown in Scheme 13. A chiral hydroxy ester (349) was glycosylated with (350) and (351) using the activating agent TMSOTf. The deacetylation process was problematic in determining the condition of DBU in methanol:dichloromethane (40 °C). This is because it resulted in the desired product that was ready to be glycosylated with donor (343) and the TMSOTf-activated step. The process was followed by the saponification of product (355) into (338) before the application of Corey–Nicolaou’s lactonization protocol using 2,20-pyridyl disulfide ((PyS)2) and triphenylphosphine (Ph3P). This was conducted in the highly dilute toluene (7.5 × 10−4 M) upon heating for 5 days to obtain compound (357). The product was then desylilated to obtain fragment (358), which was coupled to (360).
In preparing the exocyclic dirhamno-pyranose fragment (360), the imidate donor (345) was glycosylated with (348) and catalyzed by TMSOTf to produce intermediate (359). This compound was then concerted to imidate over two steps resulting in (360), which was coupled to fragment (358) through TMSOTf-activated glycosylation, followed by a global deprotection to give batatoside L (362), as shown in Scheme 14.
Synthesis of a disaccharide constituent of tricolorin A (370) and tricolorin F was successfully carried out through the MeOTf-promoted intermolecular glycosylation of dodecyl thioglycosyl donors [96]. Son et al. (2009) studied the intramolecular glycosyation using racemic: a mixture with MeOTf as a catalyst. The result showed that the R isomer went through glycosylation easier than the S isomer. Furthermore, the studies were applied to the synthesis of the disaccharide of tricolorin A and the total synthesis of tricolorin F.
In the synthesis of disaccharide and tricolorin A (370), the D-fucosyl donor (363) was conjugated to enantiomerically pure (S)-methyl 11-jalapionate using MeOTf. The resulting product was then safonified and condensed to the glucosyl donor (368) to obtain a low-yield product (369). This compound was then cyclized through MeOTf-activated glycosylation to produce (370) as a disaccharide constituent of tricolorin A, as shown in Scheme 15.
A total synthesis of tricolorin F (377) was then attempted, as shown in Scheme 16. The synthesis was initiated from the starting material donor (371), which was initially MeOTf-promoted and glycosylated with the fucosyl acceptor (366) to produce intermediate (372). The intermediate was then hydrolysed and conjugated with thioquinovoside (374) to give compound (375) with a 6% yield. However, the use of excess (374) increased the yield by 72%. The intramolecular glycosylation of compound (375) was undertaken using a similar MeOTf-activated glycosylate to give compound (376) at a 72% yield. This product was then deprotected globally to produce the desired tricolorin F (377).
In the synthesis of ipomoeassin A-F and analogues, macrocyclic rings were obtained through the ring-closing metathesis (RCM) method [97]. The C-sylillated strategy was applied successfully to synthesize the synthetic target. One of the precursors was the syllilated cinnamic acid (382), which was synthesized through the hydroalumination of (378), followed by the addition of iodine and subsequent O-silylation under standard conditions to produce product (379), which was subsequently converted to (382) through three reaction steps, as shown in Scheme 17. Lithium for the halogen exchange triggered a retro-Brook rearrangement that installed the C-silyl substituent and released the primary alcohol in (380), which was then oxidized to the corresponding acid (382) in an excellent overall yield.
The new strategy was used to esterify the sylillated cinnamic acid (382) by alcohol (383). Furthermore, the Yamaguchi method was used to analyse product (384), and after the PMB deprotection of the 4-oxo-8-nonenoic acid ester segments, diene (386) was produced. This was followed by a macrocycle using the commercial “second-generation” ruthenium alkylidene complex (387). After the hydrogenation and syllil processes, OPMB deprotection, ipomoeassin A (392), and B (390) were obtained. A similar strategy was also applied to the synthesis of ipomoeassin C-F, as shown in Scheme 18.

7. Summary

This updated review shows that resin glycosides are a class of natural products with unique structures, comprising sugar units and the aglycone. The number of sugar unit and the variety of the aglycone, including the organic acids, make a large variation in the structure of resin glycoside. The review on the structural diversity of resin glycosides revealed that the pentasaccharides-typed resin glycosides are the dominant compounds discovered in the Convolvulaceae family, followed by the hexasaccharide type. The most complex structure of resin glycosides was found in the group of the oligomer ester. Meanwhile, the monosaccharide-typed resin glycoside made it the simplest structure of all. The unique structure of the resin glycosides is associated with the varied biological properties of the compounds, covering cytotoxic activities, MDR modulating properties, α-glucosidase inhibitory potential, antiviral, anti-metastatic, and anti-proliferative activities. It also covers the neuroprotective and anticonvulsant activities, sedative effect, GABA release, anti-inflammatory, antimycobacterial, antidepressant, anticonvulsant, antimalarial, anti-diarrhoea, downregulation aquaporin 3, and antibacterial activities. Among all of the activity, resin glycosides are known to be a good MDR modulator when they are combined with resistant drugs and have the potential to be explored more in the future. The isolation and structural characterization of resin glycosides shown in this review provide useful information. For the chemical synthesis part, the review gave information for the chemical access to the resin glycoside, and it can possibly stimulate new synthetic strategies in the future, including strategies for glycosylation to make an oligosaccharide core, cyclization, including macrolactonization, to make a macrocyclic core and also in the selection, protection, and deprotection of the hydroxyl groups during synthesis.

Author Contributions

R.M. and M.F. wrote the manuscript and U.S. provided critical inputs. U.S. supervised the development of the work and manuscript evaluation. All authors discussed the results and contributed to the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the PTUPT Grant, Ministry of Education and Culture, Indonesia, grant number 5413/UN6.3.1/PT.00/2021 and Academic Leadership Grant with No: 1959/UN6.3.1/PT.00/2022 by Unang Supratman.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declared no conflicts of interest.

References

  1. Pereda-Miranda, R.; Rosas-Ramírez, D.; Castañeda-Gómez, J. Resin Glycosides from the Morning Glory Family. In Fortschritte der Chemie Organischer Naturstoffe/Progress in the Chemistry of Organic Natural Products; Springer: Berlin/Heidelberg, Germany, 2010; Volume 92, pp. 77–153. [Google Scholar]
  2. Ono, M. Resin Glycosides from Convolvulaceae Plants. J. Nat. Med. 2017, 71, 591–604. [Google Scholar] [CrossRef] [Green Version]
  3. Song, W.B.; Wang, W.Q.; Zhang, S.W.; Xuan, L.J. Multidrug Resistance-Reversal Effects of Resin Glycosides from Dichondra repens. Bioorg. Med. Chem. Lett. 2015, 25, 795–798. [Google Scholar] [CrossRef] [Green Version]
  4. Fan, B.Y.; Luo, J.G.; Gu, Y.C.; Kong, L.Y. Unusual Ether-Type Resin Glycoside Dimers from the Seeds of Cuscuta chinensis. Tetrahedron 2014, 70, 2003–2014. [Google Scholar] [CrossRef]
  5. Ding, W.B.; Zhang, D.G.; Liu, C.J.; Li, G.H.; Li, Y.Z. Resin Glycosides from Porana duclouxii. J. Asian Nat. Prod. Res. 2014, 16, 135–140. [Google Scholar] [CrossRef]
  6. Castañeda-Gómez, J.; Figueroa-González, G.; Jacobo, N.; Pereda-Miranda, R. Purgin II, a Resin Glycoside Ester-Type Dimer and Inhibitor of Multidrug Efflux Pumps from Ipomoea purga. J. Nat. Prod. 2013, 76, 64–71. [Google Scholar] [CrossRef]
  7. Cruz-Morales, S.; Castañeda-Gómez, J.; Rosas-Ramírez, D.; Fragoso-Serrano, M.; Figueroa-González, G.; Lorence, A.; Pereda-Miranda, R. Resin Glycosides from Ipomoea alba Seeds as Potential Chemosensitizers in Breast Carcinoma Cells. J. Nat. Prod. 2016, 79, 3093–3104. [Google Scholar] [CrossRef]
  8. Yin, Y.Q.; Pan, J.T.; Yu, B.W.; Cui, H.H.; Yan, Y.S.; Chen, Y.F. Two Pentasaccharide Resin Glycosides from Argyreia acuta. Nat. Prod. Res. 2016, 30, 20–24. [Google Scholar] [CrossRef]
  9. Castañeda-Gómez, J.; Pereda-Miranda, R. Resin Glycosides from the Herbal Drug Jalap (Ipomoea purga). J. Nat. Prod. 2011, 74, 1148–1153. [Google Scholar] [CrossRef]
  10. León-Rivera, I.; Villeda-Hernández, J.; Campos-Peña, V.; Aguirre-Moreno, A.; Estrada-Soto, S.; Navarrete-Vázquez, G.; Rios, M.Y.; Aguilar-Guadarrama, B.; Castillo-España, P.; Rivera-Leyva, J.C. Evaluation of the Neuroprotective Activity of Stansin 6, a Resin Glycoside from Ipomoea stans. Bioorg. Med. Chem. Lett. 2014, 24, 3541–3545. [Google Scholar] [CrossRef]
  11. Zong, G.; Hu, Z.; Duah, K.B.; Andrews, L.E.; Zhou, J.; O’Keefe, S.; Whisenhunt, L.; Shim, J.S.; Du, Y.; High, S.; et al. Ring Expansion Leads to a More Potent Analogue of Ipomoeassin F. J. Org. Chem. 2020, 85, 16226–16235. [Google Scholar] [CrossRef]
  12. Castañeda-Gómez, J.; Lavias-Hernández, P.; Fragoso-Serrano, M.; Lorence, A.; Pereda-Miranda, R. Acylsugar Diversity in the Resin Glycosides from Ipomoea tricolor Seeds as Chemosensitizers in Breast Cancer Cells. Phytochem. Lett. 2019, 32, 77–82. [Google Scholar] [CrossRef]
  13. Rosas-Ramírez, D.; Escandón-Rivera, S.; Pereda-Miranda, R. Morning Glory Resin Glycosides as α-Glucosidase Inhibitors: In Vitro and in Silico Analysis. Phytochemistry 2018, 148, 39–47. [Google Scholar] [CrossRef]
  14. Ono, M.; Kanemaru, Y.; Yasuda, S.; Okawa, M.; Kinjo, J.; Miyashita, H.; Yokomizo, K.; Yoshimitsu, H.; Nohara, T. A New Resin Glycoside from Calystegia soldanella and Its Antiviral Activity towards Herpes. Nat. Prod. Res. 2017, 31, 2660–2664. [Google Scholar] [CrossRef]
  15. Chen, C.; Ma, T.; Zhang, C.; Zhang, H.; Bai, L.; Kong, L.; Luo, J. Down-Regulation of Aquaporin 5-Mediated Epithelial-Mesenchymal Transition and Anti-Metastatic Effect by Natural Product Cairicoside E in Colorectal Cancer. Mol. Carcinog. 2017, 56, 2692–2705. [Google Scholar] [CrossRef]
  16. León-Rivera, I.; del Río-Portilla, F.; Enríquez, R.G.; Rangel-López, E.; Villeda, J.; Rios, M.Y.; Navarrete-Vázquez, G.; Hurtado-Días, I.; Guzmán-Valdivieso, U.; Núñez-Urquiza, V.; et al. Hepta-, Hexa-, Penta-, Tetra-, and Trisaccharide Resin Glycosides from Three Species of Ipomoea and Their Antiproliferative Activity on Two Glioma Cell Lines. Magn. Reson. Chem. 2017, 55, 214–223. [Google Scholar] [CrossRef]
  17. León-Rivera, I.; Castro, J.M.; Mirón-López, G.; del Río-Portilla, F.; Enríquez, R.G.; Reynolds, W.F.; Estrada-Soto, S.; Rendón-Vallejo, P.; del Carmen Gutiérrez, M.; Herrera-Ruiz, M.; et al. Resin Glycosides from Ipomoea tyrianthina and Their Sedative and Vasorelaxant Effects. J. Nat. Med. 2014, 68, 655–667. [Google Scholar] [CrossRef]
  18. Yoshikawa, K.; Yagi, C.; Hama, H.; Tanaka, M.; Arihara, S.; Hashimoto, T. Ipomotaosides A-D, Resin Glycosides from the Aerial Parts of Ipomoea batatas and Their Inhibitory Activity against COX-1 and COX-2. J. Nat. Prod. 2010, 73, 1763–1766. [Google Scholar] [CrossRef]
  19. Mirón-López, G.; Herrera-Ruiz, M.; Estrada-Soto, S.; Aguirre-Crespo, F.; Vázquez-Navarrete, L.; León-Rivera, I. Resin Glycosides from the Roots of Ipomoea tyrianthina and Their Biological Activity. J. Nat. Prod. 2007, 70, 557–562. [Google Scholar] [CrossRef]
  20. Zhu, D.; Chen, C.; Bai, L.; Kong, L.; Luo, J. Downregulation of Aquaporin 3 Mediated the Laxative Effect in the Rat Colon by a Purified Resin Glycoside Fraction from Pharbitis Semen. Evid.-Based Complement. Altern. Med. 2019, 2019, 9406342. [Google Scholar] [CrossRef] [Green Version]
  21. Nguyen, H.T.; Yu, N.H.; Park, A.R.; Park, H.W.; Kim, I.S.; Kim, J.C. Antibacterial Activity of Pharbitin, Isolated from the Seeds of Pharbitis Nil, against Various Plant Pathogenic Bacteria. J. Microbiol. Biotechnol. 2017, 27, 1763–1772. [Google Scholar] [CrossRef]
  22. Tasdemir, D.; Brun, R.; Franzblau, S.G.; Sezgin, Y.; Çalis, I. Evaluation of Antiprotozoal and Antimycobacterial Activities of the Resin Glycosides and the Other Metabolites of Scrophularia cryptophila. Phytomedicine 2008, 15, 209–215. [Google Scholar] [CrossRef]
  23. Wang, W.Q.; Liu, S.S.; Song, W.B.; Li, J.; Xuan, L.J. Resin Glycosides from the Seeds of Ipomoea muricata and Their Multidrug Resistance Reversal Activities. Phytochemistry 2018, 149, 24–30. [Google Scholar] [CrossRef]
  24. Corona-Castañeda, B.; Rosas-Ramírez, D.; Castañeda-Gómez, J.; Aparicio-Cuevas, M.A.; Fragoso-Serrano, M.; Figueroa-González, G.; Pereda-Miranda, R. Resin Glycosides from Ipomoea wolcottiana as Modulators of the Multidrug Resistance Phenotype in Vitro. Phytochemistry 2016, 123, 48–57. [Google Scholar] [CrossRef]
  25. Fan, B.Y.; Jiang, X.; Li, Y.X.; Wang, W.L.; Yang, M.; Li, J.L.; Wang, A.D.; Chen, G.T. Chemistry and Biological Activity of Resin Glycosides from Convolvulaceae Species. Med. Res. Rev. 2022, 42, 2025–2066. [Google Scholar] [CrossRef]
  26. Ono, M.; Kishida, M.; Ikegami, Y.; Takaki, Y.; Okawa, M.; Kinjo, J.; Yoshimitsu, H.; Nohara, T.; Miyahara, K. Components of Convolvulin from Quamoclit × multifida. J. Nat. Med. 2011, 65, 95–102. [Google Scholar] [CrossRef]
  27. Ono, M.; Takagi-Taki, Y.; Honda-Yamada, F.; Noda, N.; Miyahara, K. Components of Ether-Insoluble Resin Glycoside (Convolvulin) from Seeds of Quamoclit pennata. Regul. Artic. Chem. Pharm. Bull. 2010, 58, 666–672. [Google Scholar] [CrossRef] [Green Version]
  28. Lira-Ricárdez, J.; Pereda-Miranda, R.; Castañeda-Gómez, J.; Fragoso-Serrano, M.; Simas, R.C.; Leitão, S.G. Resin Glycosides from the Roots of Operculina macrocarpa (Brazilian Jalap) with Purgative Activity. J. Nat. Prod. 2019, 82, 1664–1677. [Google Scholar] [CrossRef]
  29. Fan, B.Y.; Gu, Y.C.; He, Y.; Li, Z.R.; Luo, J.G.; Kong, L.Y. Cytotoxic Resin Glycosides from Ipomoea aquatica and Their Effects on Intracellular Ca2+ Concentrations. J. Nat. Prod. 2014, 77, 2264–2272. [Google Scholar] [CrossRef]
  30. Takigawa, A.; Muto, H.; Kabata, K.; Okawa, M.; Kinjo, J.; Yoshimitsu, H.; Nohara, T.; Ono, M. Calysolins I-IV, Resin Glycosides from Calystegia soldanella. J. Nat. Prod. 2011, 74, 2414–2419. [Google Scholar] [CrossRef]
  31. Ono, M.; Kawakami, G.; Takigawa, A.; Kabata, K.; Okawa, M.; Kinjo, J.; Yokomizo, K.; Yoshimitsu, H.; Nohara, T. Calysolins X-XIII, Resin Glycosides from Calystegia soldanella, and Their Antiviral Activity toward Herpes Simplex Virus. Chem. Pharm. Bull. 2014, 62, 839–844. [Google Scholar] [CrossRef]
  32. Xu, J.Y.; He, Y.; Zhang, A.W.; Lu, Y.; Chen, G.T.; Yang, M.; Fan, B.Y. Isolation of Evolvulic Acids B and C, Two New Components of Crude Resin Glycoside Fraction from Evolvulus alsinoides. Nat. Prod. Res. 2021, 35, 3766–3771. [Google Scholar] [CrossRef]
  33. Ono, M.; Taketomi, S.; Kakiki, Y.; Yasuda, S.; Okawa, M.; Kinjo, J.; Miyashita, H.; Yoshimitsu, H.; Nohara, T. A New Glycosidic Acid, Muricatic Acid D, and Resin Glycosides, Muricatins X and XI, from the Crude Resin Glycoside Fraction of the Seeds of Ipomoea muricata. Chem. Pharm. Bull. 2021, 69, 291–297. [Google Scholar] [CrossRef]
  34. Ono, M.; Taketomi, S.; Kakiki, Y.; Yasuda, S.; Okawa, M.; Kinjo, J.; Yoshimitsu, H.; Nohara, T. A New Resin Glycoside, Muricatin IX, from the Seeds of Ipomoea muricata. Chem. Pharm. Bull. 2016, 64, 1408–1410. [Google Scholar] [CrossRef] [Green Version]
  35. Reynolds, W.F.; Yu, M.; Enriquez, R.G.; Gonzalez, H.; Leon, I.; Magos, G. Isolation and Characterization of Cytotoxic and Antibacterial Tetrasaccharide Glycosides from Ipomoea stans. J. Nat. Prod. 1995, 58, 1730–1731. [Google Scholar] [CrossRef]
  36. Wang, L.; Yan, Y.S.; Cui, H.H.; Yin, Y.Q.; Pan, J.T.; Yu, B.W. Three New Resin Glycosides Compounds from Argyreia acuta and Their α-Glucosidase Inhibitory Activity. Nat. Prod. Res. 2017, 31, 537–542. [Google Scholar] [CrossRef]
  37. Yu, B.W.; Sun, J.J.; Pan, J.T.; Wu, X.H.; Yin, Y.Q.; Yan, Y.S.; Hu, J.Y.; McPhee, D.J. Four Pentasaccharide Resin Glycosides from Argyreia acuta. Molecules 2017, 22, 440. [Google Scholar] [CrossRef] [Green Version]
  38. Cruz-Morales, S.; Castañeda-Gómez, J.; Figueroa-González, G.; Mendoza-García, A.D.; Lorence, A.; Pereda-Miranda, R. Mammalian Multidrug Resistance Lipopentasaccharide Inhibitors from Ipomoea alba Seeds. J. Nat. Prod. 2012, 75, 1603–1611. [Google Scholar] [CrossRef]
  39. Fan, B.Y.; Li, Z.R.; Ma, T.; Gu, Y.C.; Zhao, H.J.; Luo, J.G.; Kong, L.Y. Further Screening of the Resin Glycosides in the Edible Water Spinach and Characterisation on Their Mechanism of Anticancer Potential. J. Funct. Foods 2015, 19, 141–154. [Google Scholar] [CrossRef]
  40. Lu, Y.; He, Y.; Yang, M.; Fan, B.Y. Arvensic Acids K and L, Components of Resin Glycoside Fraction from Convolvulus arvensis. Nat. Prod. Res. 2021, 35, 2303–2307. [Google Scholar] [CrossRef]
  41. Rosas-Ramírez, D.; Pereda-Miranda, R. Resin Glycosides from the Yellow-Skinned Variety of Sweet Potato (Ipomoea batatas). J. Agric. Food Chem. 2013, 61, 9488–9494. [Google Scholar] [CrossRef]
  42. Yu, B.; Luo, J.; Wang, J.; Zhang, D.; Yu, S.; Kong, L. Pentasaccharide Resin Glycosides from Ipomoea cairica and Their Cytotoxic Activities. Phytochemistry 2013, 95, 421–427. [Google Scholar] [CrossRef]
  43. Pan, J.T.; Yu, B.W.; Yin, Y.Q.; Li, J.H.; Wang, L.; Guo, L.B.; Shen, Z. Four New Pentasaccharide Resin Glycosides from Ipomoea cairica with Strong α-Glucosidase Inhibitory Activity. Molecules 2015, 20, 6601–6610. [Google Scholar] [CrossRef] [Green Version]
  44. Ono, M.; Ichihara, Y.; Saito, N.; Yamada, M.; Yuuki, K.; Nawata, M.; Tsutsumi, S.; Yasuda, S.; Tsuchihashi, R.; Okawa, M.; et al. Identification and Characterization of Organic and Glycosidic Acids in Crude Resin Glycoside Fraction from Calystegia hederacea. J. Nat. Med. 2020, 74, 200–211. [Google Scholar] [CrossRef]
  45. Takigawa, A.; Setoguchi, H.; Okawa, M.; Kinjo, J.; Miyashita, H.; Yokomizo, K.; Yoshimitsu, H.; Nohara, T.; Ono, M. Identification and Characterization of Component Organic and Glycosidic Acids of Crude Resin Glycoside Fraction from Calystegia soldanella. Chem. Pharm. Bull. 2011, 59, 1163–1168. [Google Scholar] [CrossRef] [Green Version]
  46. Ono, M.; Takigawa, A.; Kanemaru, Y.; Kawakami, G.; Kabata, K.; Okawa, M.; Kinjo, J.; Yokomizo, K.; Yoshimitsu, H.; Nohara, T. Calysolins V-IX, Resin Glycosides from Calystegia soldanella and Their Antiviral Activity toward Herpes. Chem. Pharm. Bull. 2014, 62, 97–105. [Google Scholar] [CrossRef] [Green Version]
  47. Fan, B.Y.; Lu, Y.; Yang, M.; Li, J.L.; Chen, G.T. Evolvulins I and II, Resin Glycosides with a Trihydroxy Aglycone Unit from Evolvulus alsinoides. Org. Lett. 2019, 21, 6548–6551. [Google Scholar] [CrossRef]
  48. Fan, B.Y.; Lu, Y.; Xu, J.Y.; Zhang, A.W.; Yang, M.; Chen, G.T. Evolvulin III, a New Resin Glycoside Isolated from Evolvulus alsinoides. Phytochem. Lett. 2020, 39, 132–134. [Google Scholar] [CrossRef]
  49. Sura, M.B.; Ponnapalli, M.G.; Annam, S.C.V.A.R.; Bobbili, V.V.P. Ipomeolides A and B, Resin Glycosides from Ipomoea pes-caprae and Combination Therapy of Ipomeolide A with Doxorubicin. J. Nat. Prod. 2019, 82, 1292–1300. [Google Scholar] [CrossRef]
  50. Wang, W.Q.; Song, W.B.; Lan, X.J.; Huang, M.; Xuan, L.J. Merremins A-G, Resin Glycosides from Merremia hederacea with Multidrug Resistance Reversal Activity. J. Nat. Prod. 2014, 77, 2234–2240. [Google Scholar] [CrossRef]
  51. Ono, M.; Azuchi, M.; Ichio, M.; Jiyoubi, Y.; Tsutsumi, S.; Yasuda, S.; Tsuchihasi, R.; Okawa, M.; Kinjo, J.; Yoshimitsu, H.; et al. Seven New Resin Glycosides from the Seeds of Quamoclit × multifida. J. Nat. Med. 2019, 73, 11–22. [Google Scholar] [CrossRef]
  52. Ono, M.; Teramoto, S.; Naito, S.; Takahashi, A.; Yoneda, A.; Shinkai, M.; Taga, N.; Yasuda, S.; Tsuchihasi, R.; Okawa, M.; et al. Four New Resin Glycosides, Murasakimasarins I–IV, from the Tuber of Ipomoea batatas. J. Nat. Med. 2018, 72, 784–792. [Google Scholar] [CrossRef]
  53. Yu, B.W.; Luo, J.G.; Wang, J.S.; Zhang, D.M.; Yu, S.S.; Kong, L.Y. Pentasaccharide Resin Glycosides from Ipomoea pes-caprae. J. Nat. Prod. 2011, 74, 620–628. [Google Scholar] [CrossRef]
  54. Ono, M.; Takigawa, A.; Mineno, T.; Yoshimitsu, H.; Nohara, T.; Ikeda, T.; Fukuda-Teramachi, E.; Noda, N.; Miyahara, K. Acylated Glycosides of Hydroxy Fatty Acid Methyl Esters Generated from the Crude Resin Glycoside (Pharbitin) of Seeds of Pharbitis nil by Treatment with Indium(III) Chloride in Methanol. J. Nat. Prod. 2010, 73, 1846–1852. [Google Scholar] [CrossRef]
  55. Ono, M.; Takaki, Y.; Takatsuji, M.; Akiyama, K.; Okawa, M.; Kinjo, J.; Miyashita, H.; Yoshimitsu, H.; Nohara, T. Three New Resin Glycosides and a New Tetrahydropyran Derivative from the Seeds of Quamoclit pennata. Chem. Pharm. Bull. 2012, 60, 1083–1087. [Google Scholar] [CrossRef] [Green Version]
  56. Ding, W.; Jiang, Z.H.; Wu, P.; Xu, L.; Wei, X. Resin Glycosides from the Aerial Parts of Operculina turpethum. Phytochemistry 2012, 81, 165–174. [Google Scholar] [CrossRef]
  57. Koryudzu, K.; Arai, M.A.; Ahmed, F.; Sadhu, S.K.; Ishibashi, M. A New Resin Glycoside from Ipomoea maxima. Nat. Prod. Commun. 2012, 7, 219–220. [Google Scholar] [CrossRef] [Green Version]
  58. Ono, M.; Saito, N.; Minamishima, H.; Yasuda, S.; Tsuchihashi, R.; Okawa, M.; Kinjo, J.; Miyashita, H.; Yoshimitsu, H.; Nohara, T. Two New Glycosidic Acids, Calyhedic Acids E and F, in Crude Resin Glycoside Fraction from Calystegia hederacea. Nat. Prod. Res. 2021, 36, 46–53. [Google Scholar] [CrossRef]
  59. Ono, M.; Yuhara, N.; Shimohara, T.; Matsubara, S.; Yasuda, S.; Tsuchihashi, R.; Okawa, M.; Kinjo, J.; Zhou, J.R.; Yoshimitsu, H.; et al. Calyhedins I–VI: Resin Glycosides from the Rhizomes of Calystegia hederacea. Phytochemistry 2021, 190, 112888. [Google Scholar] [CrossRef]
  60. Ono, M.; Shimohara, T.; Yuhara, N.; Matsubara, S.; Yasuda, S.; Tsuchihashi, R.; Okawa, M.; Kinjo, J.; Yoshimitsu, H.; Nohara, T. Four New Resin Glycosides, Calyhedins VII–X, from the Rhizomes of Calystegia hederacea. Nat. Prod. Res. 2021, 1–10. [Google Scholar] [CrossRef]
  61. Lv, K.Q.; Ji, H.Y.; Du, G.X.; Peng, S.; Guo, P.J.; Wang, G.; Zhu, Y.; Wang, Q.; Wang, W.Q.; Xuan, L.J. Calysepins I-VII, Hexasaccharide Resin Glycosides from Calystegia sepium and Their Cytotoxic Evaluation. J. Nat. Prod. 2022, 85, 1294–1303. [Google Scholar] [CrossRef]
  62. Ono, M.; Takigawa, A.; Muto, H.; Kabata, K.; Okawa, M.; Kinjo, J.; Yokomizo, K.; Yoshimitsu, H.; Nohara, T. Antiviral Activity of Four New Resin Glycosides Calysolins XIV-XVII from Calystegia soldanella against Herpes Simplex Virus. Chem. Pharm. Bull. 2015, 63, 641–648. [Google Scholar] [CrossRef] [Green Version]
  63. Akiyama, K.; Mineno, T.; Okawa, M.; Kinjo, J.; Miyashita, H.; Yoshimitsu, H.; Nohara, T.; Ono, M. Three Acylated Glycosidic Acid Methyl Esters and Two Acylated Methyl Glycosides Generated from the Convolvulin Fraction of Seeds of Quamoclit pennata by Treatment with Indium(III) Chloride in Methanol. Chem. Pharm. Bull. 2013, 61, 952–961. [Google Scholar] [CrossRef] [Green Version]
  64. Fan, B.Y.; Lu, Y.; Yin, H.; He, Y.; Li, J.L.; Chen, G.T. Arvensic Acids A-D, Novel Heptasaccharide Glycosidic Acids as the Alkaline Hydrolysis Products of Crude Resin Glycosides from Convolvulus arvensis. Fitoterapia 2018, 131, 209–214. [Google Scholar] [CrossRef]
  65. Akiyama, K.; Yamamoto, K.; Mineno, T.; Okawa, M.; Kinjo, J.; Yoshimitsu, H.; Nohara, T.; Ono, M. Five New Resin Glycoside Derivatives Isolated from the Convolvulin Fraction of Seeds of Quamoclit pennata after Treatment with Indium(III) Chloride in Methanol. Chem. Pharm. Bull. 2014, 62, 125–133. [Google Scholar] [CrossRef] [Green Version]
  66. Bautista, E.; Fragoso-Serrano, M.; Pereda-Miranda, R. Jalapinoside, a Macrocyclic Bisdesmoside from the Resin Glycosides of Ipomea purga, as a Modulator of Multidrug Resistance in Human Cancer Cells. J. Nat. Prod. 2015, 78, 168–172. [Google Scholar] [CrossRef]
  67. Bautista, E.; Fragoso-Serrano, M.; Pereda-Miranda, R. Jalapinoside II, a Bisdesmoside Resin Glycoside, and Related Glycosidic Acids from the Officinal Jalap Root (Ipomoea purga). Phytochem. Lett. 2016, 17, 85–93. [Google Scholar] [CrossRef]
  68. Ono, M.; Akiyama, K.; Kishida, M.; Okawa, M.; Kinjo, J.; Yoshimitsu, H.; Miyahara, K. Two New Glycosidic Acids, Multifidinic Acids F and G, of the Ether-Insoluble Resin Glycoside (Convolvulin) from the Seeds of Quamoclit × multifida. J. Nat. Med. 2013, 67, 822–826. [Google Scholar] [CrossRef]
  69. Ono, M.; Akiyama, K.; Yamamoto, K.; Mineno, T.; Okawa, M.; Kinjo, J.; Miyashita, H.; Yoshimitsu, H.; Nohara, T. Four New Acylated Glycosidic Acid Methyl Esters Isolated from the Convolvulin Fraction of Seeds of Quamoclit pennata after Treatment with Indium(III) Chloride in Methanol. Chem. Pharm. Bull. 2014, 62, 830–835. [Google Scholar] [CrossRef] [Green Version]
  70. Ono, M.; Imao, M.; Miyahara, K. Two New Glycosidic Acids, Quamoclinic Acids G and H, of the ResinGlycosides (Convolvulin) from the Seeds of Quamoclit pennata. Chem. Pharm. Bull. (Tokyo) 2010, 58, 1232–1235. [Google Scholar] [CrossRef] [Green Version]
  71. Rosas-Ramírez, D.; Escalante-Sánchez, E.; Pereda-Miranda, R. Batatins III-VI, Glycolipid Ester-Type Dimers from Ipomoea batatas. Phytochemistry 2011, 72, 773–780. [Google Scholar] [CrossRef]
  72. Bah, M.; Pereda-Miranda, R. Detailed FAB-Mass Spectrometry and High Resolution NMR Investigations of Tricolorins A-E, Individual Oligosaccharides from the Resins of Ipomoea tricolor (Convolvulaceae) 1. Tetrahedron 1996, 52, 13063–13080. [Google Scholar] [CrossRef]
  73. Noda, N.; Tsuji, K.; Kawasaki, T.; Miyahara, K.; Hanazono, H.; Yang, C.-R. A Novel Resin Glycoside, Merremin (Tuguajalapin X Dimer), from Merremia hungaiensis. Chem. Pharm. Bull. 1995, 43, 1061–1063. [Google Scholar] [CrossRef]
  74. Ono, M.; Kawasaki, T.; Miyahara, K. Resin Glycosides. V. Identification and Characterization of the Component Organic and Glycosidic Acids of the Ether-Soluble Crude Resin Glycosides (“Jalapin”) from Rhizoma Jalapae Braziliensis (Roots of Ipomoea operculata). Chem. Pharm. Bull. 1989, 37, 3209–3213. [Google Scholar] [CrossRef] [Green Version]
  75. Chérigo, L.; Pereda-Miranda, R. Resin Glycosides from the Flowers of Ipomoea murucoides. J. Nat. Prod. 2006, 69, 595–599. [Google Scholar] [CrossRef]
  76. Maria Gaspar, E.M. Soldanelline B The First Acylated Nonlinear Tetrasaccharide Macrolactone from the European Convolvulaceae Calystegia soldanella. Eur. J. Org. Chem. 2001, 2001, 369–371. [Google Scholar] [CrossRef]
  77. Zhu, D.; Chen, C.; Xia, Y.; Kong, L.Y.; Luo, J. A Purified Resin Glycoside Fraction from Pharbitidis Semen Induces Paraptosis by Activating Chloride Intracellular Channel-1 in Human Colon Cancer Cells. Integr. Cancer Ther. 2019, 18, 1534735418822120. [Google Scholar] [CrossRef]
  78. Li, J.H.; Pan, J.T.; Yin, Y.Q. Two Novel Resin Glycosides Isolated from Ipomoea cairica with α-Glucosidase Inhibitory Activity. Chin. J. Nat. Med. 2016, 14, 227–231. [Google Scholar] [CrossRef]
  79. León-Rivera, I.; Herrera-Ruiz, M.; Estrada-Soto, S.; Gutiérrez, M.D.C.; Martínez-Duncker, I.; Navarrete-Vázquez, G.; Rios, M.Y.; Aguilar, B.; Castillo-España, P.; Aguirre-Moreno, A. Sedative, Vasorelaxant, and Cytotoxic Effects of Convolvulin from Ipomoea tyrianthina. J. Ethnopharmacol. 2011, 135, 434–439. [Google Scholar] [CrossRef]
  80. León-Rivera, I.; Mirón-López, G.; Molina-Salinas, G.M.; Herrera-Ruiz, M.; Estrada-Soto, S.; Gutiérrez, M.D.C.; Alonso-Cortes, D.; Navarrete-Vázquez, G.; Ríos, M.Y.; Said-Fernández, S. Tyrianthinic Acids from Ipomoea Tyrianthina and Their Antimycobacterial Activity, Cytotoxicity, and Effects on the Central Nervous System. J. Nat. Prod. 2008, 71, 1686–1691. [Google Scholar] [CrossRef]
  81. Ono, M.; Fukuda, H.; Murata, H.; Miyahara, K. Resin Glycosides from the Leaves and Stems of Ipomoea Digitata. J. Nat. Med. 2009, 63, 176–180. [Google Scholar] [CrossRef]
  82. Çaliş, I.; Sezgin, Y.; Dönmez, A.A.; Rüedi, P.; Tasdemir, D. Crypthophilic Acids A, B, and C: Resin Glycosides from Aerial Parts of Scrophularia crypthophila. J. Nat. Prod. 2007, 70, 43–47. [Google Scholar] [CrossRef] [Green Version]
  83. Yin, Y.-Q.; Huang, X.-F.; Kong, L.-Y.; Niwa, M. Three New Pentasaccharide Resin Glycosides from the Roots of Sweet Potato (Ipomoea batatas). Chem. Pharm. Bull. 2008, 56, 1670–1674. [Google Scholar] [CrossRef] [Green Version]
  84. Tao, H.; Hao, X.; Liu, J.; Ding, J.; Fang, Y.; Gu, Q.; Zhu, W. Resin Glycoside Constituents of Ipomoea pes-caprae (Beach Morning Glory). J. Nat. Prod. 2008, 71, 1998–2003. [Google Scholar] [CrossRef]
  85. Escalante-Sánchez, E.; Rosas-Ramírez, D.; Linares, E.; Bye, R.; Pereda-Miranda, R. Batatinosides II-VI, Acylated Lipooligosaccharides from the Resin Glycosides of Sweet Potato. J. Agric. Food Chem. 2008, 56, 9423–9428. [Google Scholar] [CrossRef]
  86. Escobedo-Martínez, C.; Pereda-Miranda, R. Resin Glycosides from Ipomoea pes-caprae. J. Nat. Prod. 2007, 70, 974–978. [Google Scholar] [CrossRef]
  87. Bah, M.; Chérigo, L.; Cardoso Taketa, A.T.; Fragoso-Serrano, M.; Hammond, G.B.; Pereda-Miranda, R. Intrapilosins I-VII, Pentasaccharides from the Seeds of Ipomoea intrapilosa. J. Nat. Prod. 2007, 70, 1153–1157. [Google Scholar] [CrossRef]
  88. Noda, N.; Horiuchi, Y. The Resin Glycosides from the Sweet Potato (Ipomoea batatas L. LAM.). Chem. Pharm. Bull. 2008, 56, 1607–1610. [Google Scholar] [CrossRef] [Green Version]
  89. Zhu, S.Y.; Huang, J.S.; Zheng, S.S.; Zhu, K.; Yang, J.S. First Total Synthesis of the Proposed Structure of Batatin VI. Org. Lett. 2013, 15, 4154–4157. [Google Scholar] [CrossRef]
  90. Nawój, M.; Grobelny, A.; Mlynarski, J. Macrolide Core Synthesis of Calysolin IX Using an Intramolecular Glycosylation Approach. Eur. J. Org. Chem. 2020, 2020, 47–51. [Google Scholar] [CrossRef] [Green Version]
  91. Huang, H.; Yang, C.S.; Zeng, L.N.; Zhang, L.L. Synthesis of Key Macrolactone Structure of Resin Glycosides Using a Keck Macrolactonization Method. J. Asian Nat. Prod. Res. 2015, 17, 289–298. [Google Scholar] [CrossRef]
  92. Xie, L.; Zhu, S.Y.; Shen, X.Q.; He, L.L.; Yang, J.S. Total Synthesis of Batatoside L. J. Org. Chem. 2010, 75, 5764–5767. [Google Scholar] [CrossRef]
  93. Sun, J.; Fang, J.; Xiao, X.; Cai, L.; Zhao, X.; Zeng, J.; Wan, Q. Total Synthesis of Tricolorin A via interrupted Pummerer Reaction-Mediated Glycosylation and One-Pot Relay Glycosylation. Org. Biomol. Chem. 2020, 18, 3818–3822. [Google Scholar] [CrossRef]
  94. Fang, J.; Zeng, J.; Sun, J.; Zhang, S.; Xiao, X.; Lu, Z.; Meng, L.; Wan, Q. Total Syntheses of Resin Glycosides Murucoidins IV and V. Org. Lett. 2019, 21, 6213–6216. [Google Scholar] [CrossRef]
  95. Son, S.H.; Yanagiya, N.; Furukawa, J.I.; Sakairi, N. Intramolecular Glycosylation Approach toward Constructing the Macrocyclic Structure of Resin Glycosides. Synlett 2009, 2009, 2957–2960. [Google Scholar]
  96. Nagano, T.; Pospíšil, J.; Chollet, G.; Schulthoff, S.; Hickmann, V.; Moulin, E.; Herrmann, J.; Müller, R.; Fürstner, A. Total Synthesis and Biological Evaluation of the Cytotoxic Resin Glycosides Ipomoeassin A-F and Analogues. Chem. - A Eur. J. 2009, 15, 9697–9706. [Google Scholar] [CrossRef] [Green Version]
  97. Noda, N.; Kobayashi, H.; Miyahara, K.; Kawasaki, T. Resin Glycosides. III. Isolation and Structural Study of the Genuine Resin Glycosides, Muricatins I-VI, from the Seeds of Ipomoea Muricata. Chem. Pharm. Bull. 1988, 36, 920–929. [Google Scholar] [CrossRef]
Figure 1. Structures of monosaccharide-based resin glycosides 13.
Figure 1. Structures of monosaccharide-based resin glycosides 13.
Molecules 27 08161 g001
Figure 2. Structures of trisaccharide-based resin glycosides 46.
Figure 2. Structures of trisaccharide-based resin glycosides 46.
Molecules 27 08161 g002
Figure 3. Structures of trisaccharide-based resin glycosides 713.
Figure 3. Structures of trisaccharide-based resin glycosides 713.
Molecules 27 08161 g003
Figure 4. Structures of tetrasaccharide-based resin glycosides 1449.
Figure 4. Structures of tetrasaccharide-based resin glycosides 1449.
Molecules 27 08161 g004aMolecules 27 08161 g004bMolecules 27 08161 g004c
Figure 5. Structures of pentasaccharide-based resin glycosides 50160.
Figure 5. Structures of pentasaccharide-based resin glycosides 50160.
Molecules 27 08161 g005aMolecules 27 08161 g005bMolecules 27 08161 g005cMolecules 27 08161 g005dMolecules 27 08161 g005eMolecules 27 08161 g005fMolecules 27 08161 g005gMolecules 27 08161 g005h
Figure 6. Structures of hexasaccharide-based resin glycosides 161206.
Figure 6. Structures of hexasaccharide-based resin glycosides 161206.
Molecules 27 08161 g006aMolecules 27 08161 g006bMolecules 27 08161 g006cMolecules 27 08161 g006dMolecules 27 08161 g006e
Figure 7. Structures of heptasaccharide-based resin glycosides 207222.
Figure 7. Structures of heptasaccharide-based resin glycosides 207222.
Molecules 27 08161 g007aMolecules 27 08161 g007b
Figure 8. Structures of bidesmoside-based resin glycosides 223243.
Figure 8. Structures of bidesmoside-based resin glycosides 223243.
Molecules 27 08161 g008aMolecules 27 08161 g008bMolecules 27 08161 g008c
Figure 9. Structures of dimer-based resin glycosides 244261.
Figure 9. Structures of dimer-based resin glycosides 244261.
Molecules 27 08161 g009aMolecules 27 08161 g009bMolecules 27 08161 g009cMolecules 27 08161 g009d
Scheme 1. Optimized total syntheses of 22-membered ring analogue two (273) and ipomoeassin F (274).
Scheme 1. Optimized total syntheses of 22-membered ring analogue two (273) and ipomoeassin F (274).
Molecules 27 08161 sch001
Scheme 2. Preparation of SPSB-disaccharide (280).
Scheme 2. Preparation of SPSB-disaccharide (280).
Molecules 27 08161 sch002
Scheme 3. Synthesis of macrolactone disaccharide (287) and (288).
Scheme 3. Synthesis of macrolactone disaccharide (287) and (288).
Molecules 27 08161 sch003
Scheme 4. Synthesis of tricolorin A (293).
Scheme 4. Synthesis of tricolorin A (293).
Molecules 27 08161 sch004
Scheme 5. Synthesis of disaccharide (298).
Scheme 5. Synthesis of disaccharide (298).
Molecules 27 08161 sch005
Scheme 6. Synthesis of disaccharide (301).
Scheme 6. Synthesis of disaccharide (301).
Molecules 27 08161 sch006
Scheme 7. Synthesis of (303).
Scheme 7. Synthesis of (303).
Molecules 27 08161 sch007
Scheme 8. Synthesis of (306).
Scheme 8. Synthesis of (306).
Molecules 27 08161 sch008
Scheme 9. Synthesis of macrolactone (315).
Scheme 9. Synthesis of macrolactone (315).
Molecules 27 08161 sch009
Scheme 10. Synthesis of Pentasaccharide acceptor (329).
Scheme 10. Synthesis of Pentasaccharide acceptor (329).
Molecules 27 08161 sch010
Scheme 11. Completion of total synthesis of Batatin VI (337).
Scheme 11. Completion of total synthesis of Batatin VI (337).
Molecules 27 08161 sch011
Scheme 12. Preparation of monosaccharide building blocks.
Scheme 12. Preparation of monosaccharide building blocks.
Molecules 27 08161 sch012
Scheme 13. Formation of macrolactone framework.
Scheme 13. Formation of macrolactone framework.
Molecules 27 08161 sch013
Scheme 14. Completion of total synthesis of Batatoside L (362).
Scheme 14. Completion of total synthesis of Batatoside L (362).
Molecules 27 08161 sch014
Scheme 15. Synthesis of a disaccharide constituent of tricolorin A (370).
Scheme 15. Synthesis of a disaccharide constituent of tricolorin A (370).
Molecules 27 08161 sch015
Scheme 16. Total synthesis of tricolorin F 377.
Scheme 16. Total synthesis of tricolorin F 377.
Molecules 27 08161 sch016
Scheme 17. Synthesis of 382.
Scheme 17. Synthesis of 382.
Molecules 27 08161 sch017
Scheme 18. Total synthesis of ipomoeassin A (392) and B (390).
Scheme 18. Total synthesis of ipomoeassin A (392) and B (390).
Molecules 27 08161 sch018
Table 1. Classification of resin glycosides from 2010–2021.
Table 1. Classification of resin glycosides from 2010–2021.
NoTypeSourcesRefs.
Monosaccharides
1Maltifidinic acid E (1)Quamoclit × multifida[26]
2Quamoclinic acid B (2)Quamoclit pennata[27]
3Operculinic acid K (3)Operculina macrocarpa[28]
Trisaccharides
1Cuses 5 (4)Cuscuta chinensis[4]
2Cuses 6 (5)Cuscuta chinensis[4]
3Cuses 7 (6)Cuscuta chinensis[4]
4Cuses 3 (7)Cuscuta chinensis[4]
5Cuses 4 (8)Cuscuta chinensis[4]
6Dichondrin C (9)Dichondra repens[3]
7Poranic acid A (10)Porana duclouxii[5]
8Poranic acid B (11)Porana duclouxii[5]
9Poranaside A (12)Porana duclouxii[5]
10Stansoic acid A (13)Ipomoea stans[16]
Tetrasaccharides
1Albinoside VI (14)Ipomoea alba[7]
2Albinoside VII (15)Ipomoea alba[7]
3Albinoside VIII (16)Ipomoea alba[7]
4Albinoside IX (17)Ipomoea alba[7]
5Albinosinic acid D (18)Ipomoea alba[7]
6Albinosinic acid E (19))Ipomoea alba[7]
7Albinosinic acid F (20)Ipomoea alba[7]
8Albinosinic acid G (21)Ipomoea alba[7]
9Aquaterin XI (22)Ipomoea aquatica[29]
10Calonyctin E (23)Ipomoea muricata[23]
11Calonyctin F (24)Ipomoea muricata[23]
12Calonyctin G (25)Ipomoea muricata[23]
13Calonyctin H (26)Ipomoea muricata[23]
14Calonyctin I (27)Ipomoea muricata[23]
15Calonyctin J (28)Ipomoea muricata[23]
16Calysolin I (29)Calystegia soldanella[30]
17Calysolin X (30)Calystegia soldanella[31]
18Dichondrins A (31)Dichondra repens[3]
19Dichondrins B (32)Dichondra repens[3]
20Evolvulic acid B (33)Evolvulus alsinoides[32]
21Evolvulic acid C (34)Evolvulus alsinoides[18]
22Ipomotaoside A (35)Ipomoea batatas[18]
23Ipomotaoside B (36)Ipomoea batatas[18]
24Ipomotaoside C (37)Ipomoea batatas[18]
25Ipomotaoside D (38)Ipomoea batatas[18]
26Muricatic acid D (39)Ipomoea muricata[33]
27Muricatin IX (40)Ipomoea muricata[34]
28Muricatin X (41)Ipomoea muricata[33]
29Muricatin XI (42)Ipomoea muricata[33]
30Stansin 6 (43)Ipomoea stans[10,35]
31Stansinic acid I (44)Ipomoea stans[16]
32Tricolorin K (45)Ipomoea tricolor[12]
33Tricolorin L (46)Ipomoea tricolor[12]
34Tricolorin M (48)Ipomoea tricolor[12]
35Tyrianthinic acids VI (48)Ipomoea tyrianthina[17]
36Wolcottinoside I (49)Ipomoea wolcottiana[24]
Pentasaccharides
1Acutacoside A (50)Argyreia acuta[8]
2Acutacoside B (51)Argyreia acuta[8]
3Acutacoside C (52)Argyreia acuta[36]
4Acutacoside D (53)Argyreia acuta[36]
5Acutacoside E (54)Argyreia acuta[36]
6Acutacoside F (55)Argyreia acuta[37]
7Acutacoside G (56)Argyreia acuta[37]
8Acutacoside H (57)Argyreia acuta[37]
9Acutacoside I (58)Argyreia acuta[37]
10Albinoside I (59)Ipomoea alba[38]
11Albinoside II (60)Ipomoea alba[38]
12Albinoside III (61)Ipomoea alba[38]
13Albinoside IV (62)Ipomoea alba[7]
14Albinoside V (63)Ipomoea alba[7]
15Aquaterin I (64)Ipomoea aquatica[29]
16Aquaterin II (65)Ipomoea aquatica[29]
17Aquaterin III (66)Ipomoea aquatica[29]
18Aquaterin IV (67)Ipomoea aquatica[29]
19Aquaterin V (68)Ipomoea aquatica[29]
20Aquaterin VI (69)Ipomoea aquatica[29]
21Aquaterin VII (70)Ipomoea aquatica[29]
22Aquaterin VIII (71)Ipomoea aquatica[29]
23Aquaterin IX (72)Ipomoea aquatica[29]
24Aquaterin X (73)Ipomoea aquatica[29]
25Aquaterin XII (74)Ipomoea aquatica[39]
26Aquaterin XIII (75)Ipomoea aquatica[39]
27Aquaterin XIV (76)Ipomoea aquatica[39]
28Aquaterin XV (77)Ipomoea aquatica[39]
29Aquaterin XVI (78)Ipomoea aquatica[39]
30Aquaterin XVII (79)Ipomoea aquatica[39]
31Aquaterin XVIII (80)Ipomoea aquatica[39]
32Aquaterin XIX (81)Ipomoea aquatica[39]
33Arvensic acid K (82)Convolvulus arvensis[40]
34Arvensic acid L (83)Convolvulus arvensis[40]
35Batatinoside VII (84)Ipomoea batatas[41]
36Batatinoside VIII (85)Ipomoea batatas[41]
37Batatinoside IX (86)Ipomoea batatas[41]
38Cairicoside A (87)Ipomoea cairica[42]
39Cairicoside B (88)Ipomoea cairica[42]
40Cairicoside C (89)Ipomoea cairica[42]
41Cairicoside D (90)Ipomoea cairica[42]
42Cairicoside E (91)Ipomoea cairica[42]
43Cairicoside F (92)Ipomoea cairica[42]
44Cairicoside I (93)Ipomoea cairica[43]
45Cairicoside II (94)Ipomoea cairica[43]
46Cairicoside III (95)Ipomoea cairica[43]
47Cairicoside IV (96)Ipomoea cairica[43]
48Calonyctin B (97)Ipomoea muricata[23]
49Calonyctin C (98)Ipomoea muricata[23]
50Calonyctin D (99)Ipomoea muricata[23]
51Calyhedic acid A (100)Calystegia hederacea[44]
52Calysolic acid A (101)Calystegia soldanella[45]
53Calysolic acid B (102)Calystegia soldanella[45]
54Calysolin II (103)Calystegia soldanella[30]
55Calysolin III (104)Calystegia soldanella[30]
56Calysolin V (105)Calystegia soldanella[46]
57Calysolin VI (106)Calystegia soldanella[46]
58Evolvulic acid A (107)Evolvulus alsinoides[47]
59Evolvulin I (108)Evolvulus alsinoides[47]
60Evolvulin II (109)Evolvulus alsinoides[47]
61Evolvulin III (110)Evolvulus alsinoides[48]
62Ipomeolide A (111)Ipomoea pes-caprae[49]
63Ipomeolide B (112)Ipomoea pes-caprae[49]
64Merremin A (113)Merremia hederacea[50]
65Merremin B (114)Merremia hederacea[50]
66Merremin C (115)Merremia hederacea[50]
67Merremin D (116)Merremia hederacea[50]
68Merremin E (117)Merremia hederacea[50]
69Merremin F (118)Merremia hederacea[50]
70Merremin G (119)Merremia hederacea[50]
71Multifidin III (120)Quacmoclit × multifida[51]
72Multifidin IV (121)Quacmoclit × multifida[51]
73Multifidin V (122)Quacmoclit × multifida[51]
74Multifidin VI (123)Quacmoclit × multifida[51]
75Murasakimasarin I (124)Ipomoea batatas[52]
76Murasakimasarin II (125)Ipomoea batatas[52]
77Murasakimasarin III (126)Ipomoea batatas[52]
78Murasakimasarin IV (127)Ipomoea batatas[52]
79Murucinic acid II (128)Ipomoea murucoides[16]
80Pescaprein XXI (129)Ipomoea pes-caprae[53]
81Pescaprein XXII (130)Ipomoea pes-caprae[53]
82Pescaprein XXIII (131)Ipomoea pes-caprae[53]
83Pescaprein XXIV (132)Ipomoea pes-caprae[53]
84Pescaprein XXV (133)Ipomoea pes-caprae[53]
85Pescaprein XXVI (134)Ipomoea pes-caprae[53]
86Pescaprein XXVII (135)Ipomoea pes-caprae[53]
87Pescaprein XXVIII (136)Ipomoea pes-caprae[53]
88Pescaprein XXIX (137)Ipomoea pes-caprae[53]
89Pescaprein XXX (138)Ipomoea pes-caprae[53]
90PM6 (139)Pharbitis nil[54]
91PM7 (140)Pharbitis nil[54]
92Purginoside I (141) Ipomoea purga[9]
93Purginoside II (142)Ipomoea purga[9]
94Purginosides III (143)Ipomoea purga[6]
95Purginosides IV (144)Ipomoea purga[6]
96Quamoclin V (145)Quamoclit pennata[55]
97Quamoclin VI (146)Quamoclit pennata[55]
98Quamoclin VII (147)Quamoclit pennata[55]
99Turpethic acids A (148)Operculina turpethum[56]
100Turpethic acids B (149)Operculina turpethum[56]
101Turpethic acids C (150)Operculina turpethum[56]
102Turpethoside A (151)Operculina turpethum[56]
103Turpethoside B (152)Operculina turpethum[56]
104Tyrianthinoic acid (153)Ipomoea tyrianthina[17]
105Tyrianthinic acid III (154)Ipomoea tyrianthina[17]
106Tyrianthinic acid IV (155)Ipomoea tyrianthina[17]
107Tyrianthinic acid V (156)Ipomoea tyrianthina[17]
108Wolcottinoside II (157)Ipomoea wolcottiana[24]
109Wolcottinoside III (158)Ipomoea wolcottiana[24]
110Wolcottinoside IV (159)Ipomoea wolcottiana[24]
111A new resin glycoside (160)Ipomoea maxima[57]
Hexasaccharides
1Calyhedic acid B (161)Calystegia hederacea[44]
2Calyhedic acids E (162)Calystegia hederacea[58]
3Calyhedic acids F (163)Calystegia hederacea[58]
4Calyhedin I (164)Calystegia hederacea[59]
5Calyhedin II (165)Calystegia hederacea[59]
6Calyhedin III (166)Calystegia hederacea[59]
7Calyhedin X (167)Calystegia hederacea[60]
8Calysepin I (168)Calystegia sepium[61]
9Calysepin II (169)Calystegia sepium[61]
10Calysepin III (170)Calystegia sepium[61]
11Calysepin IV (171)Calystegia sepium[61]
12Calysepin V (172)Calystegia sepium[61]
13Calysepin VI (173)Calystegia sepium[61]
14Calysepin VII (174)Calystegia sepium[61]
15Calysolic acid C (175)Calystegia soldanella[45]
16Calysolic acid D (176)Calystegia soldanella[45]
17Calysolin IV (177)Calystegia soldanella[30]
18Calysolin VII (178)Calystegia soldanella[46]
19Calysolin VIII (179)Calystegia soldanella[46]
20Calysolin IX (180)Calystegia soldanella[46]
21Calysolin XI (181)Calystegia soldanella[31]
22Calysolin XII (182)Calystegia soldanella[31]
23Calysolin XIII (183)Calystegia soldanella[31]
24Calysolin XIV (184)Calystegia soldanella[62]
25Calysolin XV (185)Calystegia soldanella[62]
26Calysolin XVI (186)Calystegia soldanella[62]
27Calysolin XVII (187)Calystegia soldanella[62]
28Calysolin XVIII (188)Calystegia soldanella[14]
29Macrocarposidic acids A (189)Operculina macrocarpa[28]
30Macrocarposidic acids B (190)Operculina macrocarpa[28]
31Macrocarposidic acids C (191)Operculina macrocarpa[28]
32Maltifidinic acid C (192)Quamoclit × multifida[26]
33Maltifidinic acid D (193)Quamoclit × multifida[26]
35Multifidin VII (194)Quamoclit × multifida[51]
36Multifidin VIII (195)Quamoclit × multifida[51]
37Operculinic acids H (196)Operculina macrocarpa[28]
38Operculinic acids I (197)Operculina macrocarpa[28]
39Operculinic acids J (198)Operculina macrocarpa[28]
40PM1 (199)Pharbitis nil[54]
41PM2 (200)Pharbitis nil[54]
42PM3 (201)Pharbitis nil[54]
43PM4 (202)Pharbitis nil[54]
44PM5 (203)Pharbitis nil[54]
45QM4 (204)Quamoclit pennata[63]
46Quamoclinic acid C (205)Quamoclit pennata[27]
47Quamoclinic acid D (206)Quamoclit pennata[27]
Heptasaccharides
1Arvensic acid A (207)Convolvulus arvensis[64]
2Arvensic acid B (208)Convolvulus arvensis[64]
3Arvensic acid C (209)Convolvulus arvensis[64]
4Arvensic acid D (210)Convolvulus arvensis[64]
5Calyhedic acid C (211)Calystegia hederacea[44]
6Calyhedic acid D (212)Calystegia hederacea[44]
7Calyhedin IV (213)Calystegia hederacea[59]
8Calyhedin V (214)Calystegia hederacea[59]
9Calyhedin VI (215)Calystegia hederacea[59]
10Calyhedin VII (216)Calystegia hederacea[60]
11Calyhedin VIII (217)Calystegia hederacea[60]
12Calyhedin IX (218)Calystegia hederacea[60]
13QM1 (219)Quamoclit pennata[63]
14QM7 (220)Quamoclit pennata[65]
15Quamoclinic acid E (221)Quamoclit pennata[27]
16Quamoclinic acid F (222)Quamoclit pennata[27]
Macrocyclic Bidesmoside
1Jalapinoside B (223)Ipomoea purga[16]
2Jalapinoside I (224)Ipomoea purga[66]
3Jalapinoside II (225)Ipomoea purga[67]
4Multifidin IX (226)Quamoclit × multifida[51]
5Multifidinic acid F (227)Quamoclit × multifida[68]
6Multifidinic acid G (228)Quamoclit × multifida[68]
7Purgic acid C (229)Ipomoea purga[67]
8Purgic acid D (230)Ipomoea purga[67]
9QM2 (231)Quamoclit pennata[63]
10QM3 (232)Quamoclit pennata[63]
11QM5 (233)Quamoclit pennata[63]
12QM6 (234)Quamoclit pennata[65]
13QM8 (235)Quamoclit pennata[65]
14QM9 (236)Quamoclit pennata[65]
15QM10 (237)Quamoclit pennata[65]
16QM11 (238)Quamoclit pennata[69]
17QM12 (239)Quamoclit pennata[69]
18QM13 (240)Quamoclit pennata[69]
19QM14 (241)Quamoclit pennata[69]
20Quamoclinic Acid G (242)Quamoclit pennata[70]
21Quamoclinic Acid H (243)Quamoclit pennata[70]
Oligomers Ester-type dimer
1Batatin III (244)Ipomoea batatas[71]
2Batatin IV (245)Ipomoea batatas[71]
3Batatin V (246)Ipomoea batatas[71]
4Batatin VI (247)Ipomoea batatas[71]
5Batatin VII (248)Ipomoea batatas[41]
6Cuses 8 (249)Cuscuta chinensis[4]
7Cuses 9 (250)Cuscuta chinensis[4]
8Cuses 10 (251)Cuscuta chinensis[4]
9Cuses 11 (252)Cuscuta chinensis[4]
10Cuses 12 (253)Cuscuta chinensis[4]
11Purgin I (254)Ipomoea purga[9]
12Purgin II (255)Ipomoea purga[6]
13Purgin III (256)Ipomoea purga[6]
14Tyrianthin C (257)Ipomoea tyrianthina[17]
15Tyrianthin D (258)Ipomoea tyrianthina[17]
16Tyrianthin E (259)Ipomoea tyrianthina[17]
17Stansin A (260)Ipomoea stans[16]
18Wolcotinne I (261)Ipomoea wolcottiana[24]
Table 2. Distribution of resin glycosides in species of Convolvulaceae.
Table 2. Distribution of resin glycosides in species of Convolvulaceae.
GenusTriTetraPentaHexaHeptaBidesmosideDimer
Cuscuta
Ipomoea
Porana
Evolvulus
Dichondra
Calystegia
Convolvulus
Argyreia
Merremia
Quamoclit
Operculina
Jalapae
Pharbitis
Table 3. Biological properties of resin glycosides.
Table 3. Biological properties of resin glycosides.
NoName of CompoundActivityReferences
1Ipomoeassin F and analogues (273274)Cytotoxic against MDA-MB 231 and Michigan Cancer Foundation-7 (MCF-7).[11]
2Evolvulin I (108)Cytotoxic against MCF-7 cells (IC50 value of 3.12 μM).[47]
3Tricolorin A (293)The MDR reversal factor of vinblastine cytotoxicity enhanced by 2164-fold in MCF/Vin cells.[12]
4Ipomeolides A (111)Inducing apoptosis cell death in K562 cells when combined with doxorubicyn.[49]
5Calonyctins E (23), J (28), Muricatic acid CThe MDR reversal activity of vincristine cytotoxicity enhanced by 2.5–407.1-fold at 25 μM in KB/VCR cells (Calonyctin E was the most active).[23]
6Cairicosides I–IV (9396)α-Glucosidase inhibitory potential in vitro comparable to acarbose.[43]
7Cairicoside E (91)Anti-metastatic effect through EMT inhibition via down-regulation of AQP5 in CRC cells.[15]
8Calysolin XVIII (188)Anti-HSV-1 with EC50 value 2.3 µM and relatively high cytotoxicity (IC50 8.7 μM).[14]
9Albinosides VII-VIII (1516)The MDR reversal activity of vinblastine enhanced with reversal factors (RFMCF-7/Vin+) of 201 and >2517-fold, respectively, in MCF/Vin cells.[7]
10Jalapinoside B (223), purgic acid D (230), stansoic acid A (13), stansin A (43), murucinic acid II (128), stansinic acid I (44)
  • A discrete effect against the proliferative rate of HCT-15 (ED50 > 24.0 µg/mL) and OVCAR (ED50 > 25.0 µg/mL (jalapinoside B, purgic acid D, stansoic acid A, murucinic acid II, and stansinic acid I).
  • Active against the proliferation of OVCAR (ED50 5.5 µg/mL), and less cytotoxic toward HCT-15 and UISO-SQC-1 (stansin A).
  • Showing cell viability effects on C6 and RG2 cell-lines.
[16]
11Jalapinoside II (225)The MDR reversal activity of vinblastines enhanced with reversal factors (RFMCF-7/Vin+) of >1906 fold in MCF/Vin cells.[67]
12Acutacoside C-E (5254)A weak α-glucosidase inhibitory activity.[3]
13Wolcottinosides I–IV (261), (157159)The potentiation effect of the antibiotic susceptibility up to eightfold when combined with etracycline, kanamycin, or chloramphenicol.[24]
14Aquaterins XII-XIII (7475), XV (77)Cytotoxic against various cancer cell lines (IC50 3.0–8.9 μM).[39]
15Aquaterin II (65)Inducing G0/G1 arrest regulated by related proteins CDK4/6, cyclin D/E and p21. This was caused by mitochondria-mediated apoptosis with a decrease in MMP, ROS accumulation, caspase cascade activation and Bax/Bcl-2 alteration. Additionally, the suppression of the PI3K/Akt signalling pathway was observed, suggesting that it participated in the Aquaterin II–induced cell growth inhibition (antiproliferative mechanism).[39]
16Dichondrins A-C (31-32), (9)The MDR reversal activity of vincristine enhanced with reversal factors (RFMCF-7/Vin+) by 1.03–1.78-fold in KB/VCR cells.[3]
17Jalapinoside I (224)The MDR reversal activity of vincristine enhanced with reversal factors (RFMCF-7/Vin+) by >1906-fold in MCF-7/Vin cells.[66]
18Aquaterins I−IV (6467), Aquaterin VII (70), Aquaterin IX-X (7273)
  • Active against HepG2 cells with an IC50 value of 2.4 μM (aquaterin IV). Through the inhibition of the proliferation of HepG2 cells via G0/G1 arrest and apoptosis induction.
  • Elevating Ca2+ in HepG2 cells (aquaterins I−IV, aquaterin VII, aquaterin IX, aquaterin X), which is involved in the regulation of the cytotoxic activities observed.
[29]
19Merremin A (113), Merremin E-F (117118), murucoidin VThe MDR reversal activity of vincristine enhanced with reversal factors by 2.3−142.5-fold in KB/VCR cells[50]
20Stansin 6 (43)Neuroprotective and anticonvulsant activities.[10]
21Tyrianthins C-E (257259)Sedative and vasorelaxant effect.[17]
22Cuses 5–12 (46), (249253)Cytotoxic toward MCF-7, SMMC-7721, and MG-63 cell lines with IC50 values ranging from 8.72 to 59.35 mg/mL.[4]
23Cairicosides A–E (8791)Moderately cytotoxic against a small panel of human tumor cell lines with IC50 values in a range from 4.28 to 14.31 μM. Cairicosides A-D are more cytotoxic than cairicoside E.[42]
24Purgin III (256), Purginosides I–IV (141144), Purgin I (254)The MDR reversal activity of vinblastine enhanced with reversal factors (RFMCF-7/Vin+) by 1.4 to 6.5-fold in MCF-7/Vin cells.[6]
25ConvolvulinSedative and vasorelaxant effect. Cytotoxic against nasopharyngeal carcinoma cell line KB (ED50 2.3 μg/mL).[81]
26Pescapreins XXI–XXX (129138)Multidrug resistance modulator in the human breast cancer cell line MCF-7/ADR.[53]
27Ipomotaosides A (35)Anti-inflammatory activity against cyclooxygenase (COX)-1 and -2.[18]
29Cairicosides A-B (8788)Strong α-glucosidase inhibitory activity with IC50 25.3 ± 1.6 and 28.5 ± 3.3 μM, compared to acarbose.[80]
30Calysolin I–IX (29), 103104, 177, 105106, 178180) Antiviral activity against HSV-1 with IC50 ranging from 4.3 to 18.5 μM and EC50 from 1.9 to 19.2 μM. [46]
31Calysolin X–XIII (30), 181183) Antiviral activity against HSV-1 with IC50 from 14.6 to 32.6 μM and EC50 from 2.6 to 7.5 μM. [31]
32Calysolin XIV–XVII (184187) Antiviral activity against HSV-1 with IC50 ranging from 7.1 to 50.4 μM and EC50 ranging from 3.5 to 12.8 μM. [62]
Table 4. Representative details of isolation techniques of resin glycosides.
Table 4. Representative details of isolation techniques of resin glycosides.
NoPlantPart of PlantExtraction/SeparationRefs.
1Calystegia
soldanella
Fresh leaves, stems and roots
  • Step 1: Maceration of the plant using methanol at room temperature (1 month).
  • Step 2: Extraction of water suspension from the methanol extract with subsequent ethyl acetate and butanol.
  • Step 3: Separation and purification of ethyl acetate fraction using subsequent silica gel column chromatography, using gradient mixtures of CHCl3–MeOH–H2O and HPLC, and using 95% and 90% MeOH as an eluent producing calysolin IX and compound 1, respectively.
[14]
2Ipomoea tricolorSeeds
  • Step 1: Maceration of the plant using n-hexane at room temperature.
  • Step 2: Treatment of the hexane extract with chloroform and analysis of the crude by TLC and HPLC with a standard resin glycoside, tricolorin A. This step was carried out to identify the lipophilic resin glycoside mixtures.
  • Step 3: Separation and purification of the hexane extract by preparative HPLC using the peak-shaving and heart cutting techniques to obtain ten eluates of resin glycosides.
[12]
3Ipomoea muricataSeeds
  • Step 1: Maceration of the plant using acetone/water (7:3 v/v) at room temperature (3 × 24 h).
  • Step 2: Extraction of water suspension in the acetone/water extract with subsequent petroleum ether and ethyl acetate.
  • Step 3: Separation and purification of the ethyl acetate fraction using silica gel column chromatography, eluting with CHCl3/MeOH mixtures, preparative HPLC (MeCN-H2O), or NH2 column chromatography and eluting with MeOH/H2O to yield resin glycosides. All compounds were further purified by Sephadex LH-20.
[23]
4Ipomoea albaSeeds
  • Step 1: Maceration of the plant using subsequent n-hexane and chloroform at room temperature.
  • Step 2: Precipitation of chloroform-soluble fractions with methanol.
  • Step 3: Analysis of the crude using RP-HPLC on the C-18 column for comparison with reference compounds.
  • Step 4: Separation and purification of the crude using preparative-scale recycling HPLC, a Symmetry C18 column (Waters; 7 μm, 19 × 300 mm), isocratic elution with MeOH−CH3CN−H2O (10:7:3), and a flow rate of 8 mL/min for the first subfraction. Collection was performed by the heart-cutting technique. For the remaining fractions, isocratic elution with MeOH−CH3CN (7:3) and a flow rate of 8.5 mL/min was used. These procedures afforded pure resin glycosides.
[7]
5Ipomoea wolcottianaDried flower
  • Step 1: Maceration of the plant using subsequent chloroform at room temperature.
  • Step 2: Separation of chloroform extract with open silica gel column chromatography with a gradient of MeOH in CHCl3.
  • Step 3: Separation of fraction-containing resin using a reversed-phase C18 (350 g) column chromatography, eluted with MeOH to eliminate pigmented residues.
  • Step 4: Purification of resin glycoside-containing fractions using preparative recycle HPLC on a Symmetry C18 column (Waters; 7 μm, 19 x 300 mm), an isocratic elution with MeOH-CH3CN (9:1) at a flow rate of 8 mL/min, and a sample injection of 500 μL (sample concentration: 180 mg/mL). Eluates were collected through the technique of “heart cutting”. Each peak was independently reinjected (sample injection, 500 μL; concentration, 0.1 mg/μL), by employing the same C18 column and an isocratic elution MeOH-CH3CN (9:1) with a flow rate of 8.5 mL/min. These procedures afforded pure resin glycosides.
[24]
6Ipomoea batatasAerial part
  • Step 1: Maceration of the plant using subsequent ethyl acetate at room temperature (six weeks).
  • Step 2: Separation of ethyl acetate extract using silica gel column chromatography eluted with hexane-EtOAc.
  • Step 3: Separation of fraction five using silica gel column chromatography, with hexane-EtOAc for elution.
  • Step 4: Separation of fraction 5–2 using silica gel column chromatography eluted with hexane-EtOAc.
  • Step 5: Purification using preparative HPLC (ODS, 88% MeOH) to obtain purified resin glycosides.
[18]
7Ipomoea tyrianthinaRoot
  • Step 1: Maceration of the plant using subsequent n-hexane, dichloromethane, and methanol at room temperature.
  • Step 2: Separation of methanolic extract with C-18 silica gel column chromatography over the reversed-phase (C18) with a gradient of CH3OH in H2O for elution.
  • Step 3: Separation of resinous fraction using Sephadex LH-20 column chromatography and preparative TLC, obtaining two chromatographic fractions.
  • Step 4: Purification of a more polar chromatographic fraction using preparative HPLC using an Ultrasil ODS column (10 mm i.d. × 300 mm, 5 μm, Altex), eluting with a mixture of CH3CN-H2O (7:3), at a flow rate of 1 mL/min at 25 oC and detection with UV at 215 nm. Chromatographic peaks were collected and reinjected until pure. This technique afforded pure resin glycosides.
[17]
8Operculina macrocarpaRoot
  • Step 1: Percolation of the plant using subsequent n-hexane, dichloromethane, and methanol at room temperature.
  • Step 2: Suspending the methanolic extract with water and sonication of the suspension.
  • Step 3: Dissolving the water-insoluble fraction with methanol.
  • Step 4: Liquid–liquid partion of the supernatant with butanol and evaporation of the fraction to obtain resin glycosides.
[28]
Table 5. Steps in the structural elucidation of representatives of resin glycosides.
Table 5. Steps in the structural elucidation of representatives of resin glycosides.
NoName of Resin GlycosideStructural ElucidationRefs.
1Quamoclinic acid B (2)
Monosaccharide
Quamoclit pennata
  • Mass spectrometry (FAB-MS): showing the [M–H] ion peak at m/z 333 (C16H29O7) along with a fragment ion peak at m/z 187 [333–146 (6-deoxyhexose unit)] in the negative-ion mode.
  • 13C-NMR: showing 16 carbons comprising one carbonyl carbon, one oxygenated methine carbon, and one terminal quinovopyranosyl unit.
  • 1H-NMR: showing two equivalent methylene protons adjacent to a carbonyl group, and a primary methyl group ascribable to the 7-hydroxydecanoic acid moiety along with one β-D-quinovopyranosyl unit which has the 4C1 conformation.
[27]
2Poranic acid A (10)
Trisaccharides
Porana duclouxii
  • Mass spectrometry (ESI-MS): exhibiting an [M–H]-ion peak at m/z 697 as well as a series of fragment ions at m/z 551 [697–146 (C6H10O4)], 389 [551–162 (C6H10O5)] and 243 [389–146 (C6H10O4)].
  • 1H-NMR: showing three anomeric protons [δH 6.33 (br s); 5.87 (d, J 1⁄4 7.0Hz); 4.87 (d, J 1⁄4 7.5Hz)].
  • 13C-NMR: showing 32 signals, of which 18 carbons were assignable to a trisaccharide moiety and 14 carbons to the aglycone.
  • 1H–1H COSY, HSQC, and HMBC experiments: helping to assign 1H and 13C NMR signals.
  • Specific optical rotation-1: determining the absolute configuration of 11-hydroxytetradecanoic acid as 11S by the comparison of specific rotations of methyl ester with those obtained from Mexican jalap roots.
  • Specific optical rotation-2: determining the absolute configurations of sugars by the direct comparison of specific optical rotation with authentic samples.
  • GC-MS: confirming the absolute configuration through their TMSi-derivatives.
  • HMBC: supporting the connectivities between the hydroxyl fatty acid aglycone, oligosaccharide, and among monosaccharides.
[5]
3Calonyctins E (23)
Tetrasaccharides
Ipomoea muricata
  • Mass spectrometry (HRESIMS): showing a molecular ion peak at m/z 919.4528 [M + Na]+ (calcd. for C42H72NaO20, 919.4509).
  • 1H-1H COSY: showing four spin systems, which attributed to one hexose and three 6-deoxyhexose units.
  • HPLC analysis and optical rotations: revealing that the hexose unit belonged to D-glucose, and the three 6-deoxyhexose units belonged to two D-quinovose and one rhamnose.
  • 1H NMR: suggesting the β-configurations of D-quinovose and D-glucose through the coupling constants of the anomeric protons at δH 4.75 (d, J = 7.8 Hz, 1H), 5.63 (d, J = 7.7 Hz, 1H), and 5.23 (d, J = 8.1 Hz, 1H).
  • 13C NMR spectrum: suggesting the α-configuration of L-rhamnose through the chemical shift in C-5 of rhamnose (δC 66.9).
  • HMBC-1: Determining the oligosaccharide core as rhamnosyl-(1->2)-[glucosyl-(1->3)]-quinosyl-(1->2)-quinosyl through the long-range HMBC correlations: H-1 of β-Qui′ (δH 5.63) to C-2 of b-Qui (δC 79.8), H-1 of α-Rha (δH 5.94) to C-2 of β-Qui′ (δH 78.7), and H-1 of β-Glc (δH 5.23) to C-3 of β-Qui′ (δC 85.1).
  • NMR data and EIMS data: determining the aglycone to be the (11S)-jalapinolic acid (Jal) (after alkaline and acid hydrolyses).
  • HMBC-2: determining the linkage between β-Qui (1/11) Jal through correlations from the H-1 of Qui (δH 4.75) to the C-11 of Jal (δC 81.8).
  • HMBC-3: Determining the lactonization position to be H-2 of β-Glc, which indicated that the compound contained a 22-membered ring, and this was inferred by the correlation from H-2 of β-Glc (δH 5.67) to C-1 of Jal (δC 174). The compound contained an additional acetyl group (Ac) for the correlation between methyl (δH 4.75) and the carbonyl (δC 171.2). The additional HMBC correlations from δH 5.48 (H-4 of Rha) to δC 171.2 (C-1 of Ac) indicated that the OH-4 of Rha was acylated by the acetyl group.
[23]
4Ipomeolides A (111)
Pentasaccharides
Ipomoea pes-caprae
  • Mass spectrometry (HRESIMS/MS): showing a deprotonated molecular ion at m/z 1249.7345 [M − H]− (calcd. 1249.7309) in the negative ion mode, consistent with the molecular formula C63H110O24.
  • MALDITOF: showing a sodium adduct ion at m/z 1273.76. Other significant MS ions at m/z 1068.5688 were [M+ − 182 (C12H22O + H)], 1049.5551 [M+ − 201(C10H16O4 + H)], 903.4960 [M+ − H −183(C12H23O) − 163(C6H11O5)], 599.4520 [M+ − H − 475(C24H43O9) − 174(C8H14O4) − H], 417.5515 [M+ − 834(C41H70O17 + H)], 247.1173 [M+ − 1003 (C52H91O18)], and 205.6764 [M+ − H − 1044(C55H96O18)].
  • Infrared: showing hydroxy (3442 cm−1) and ester (1723, 1064 cm−1) functionalities.
  • 1H-NMR-1: showing six secondary methyls at δ 1.52 (d, J = 6.6 Hz), 1.62 (d, J = 6.0 Hz), 1.66 (d, J = 5.4 Hz), 1.61 (d, J = 5.4 Hz), 1.40 (d, J = 6.6 Hz), and 1.21 (d, J = 6.6 Hz); three terminal methyls at δ 0.94 (t, J = 7.2 Hz), 0.88 (t, J = 6.5 Hz), and 0.89 (t, J = 7.5 Hz); five anomeric protons at δ 4.75 (d, J = 7.8 Hz), 6.17 (br s), 5.49 (br s), 5.61 (br s), 5.94 (br s), and the resonance of a long chain fatty acid.
  • 13C-NMR: showing 63 carbon signals, including the signals of three carbonyls at δ 173.3, 173.4, and 176.6 ppm and five anomeric carbons at δ 104.6, 99.5, 99.1, 104.9, and 103.9 ppm.
  • 1H-NMR-2: assigning the anomeric configuration of the fucosyl moiety as β, based on the large coupling constant (J = 7.8 Hz) of H-1′ while the proton signals arising from the rhamnose residues at δH 6.17 (br s), 5.49 (br s), 5.61 (br s), and 5.94 (br s) were consistent with an α-configuration of the anomeric protons.
  • GC-MS: identifying the fatty acids as the n-dodecanoic acid and (2S)-methylbutanoic acid (a product of alkaline hydrolysis in the resin)
  • Specific rotation: additionally identifying fatty acids through the comparison of their specific rotation value ([α]27D +18 (c 0.2, CHCl3)) with that of an authentic sample.
  • Mosher’s method: assigning (11R) configurations of (11R)-hydroxyhexadecanoate (the product of subsequent acid hydrolysis of glycosidic acid methyl ester).
  • EIMS data: assigning the position of the hydroxy group at C-11 in jalapinolic acid based on crucial fragments observed at m/z 71, 101, and 201 in the EIMS data.
  • GC-MS: identifying the sugars as L-rhamnopyranose and D-fucopyranose through the tetramethylsilane (TMS) derivatives.
  • 1H−1H COSY: showing spin systems H-2−H-3−H-4 corresponding to the 2-methylbutanoyl (Mba) moiety.
  • HMBC: determining the locations of the ester substituents, and lactonization The following key HMBC correlations were observed: H-1′′ of α-Rha′′ (δH 6.17) showed correlations with C-2′ of β-Fuc′ (δC 80.5), H-1′′′ of α-Rha′′′ (δH 5.49) with C-4′′ of α-Rha′′ (δC 80.4), H-1′′′′ of α-Rha′′′′ (δH 5.61) with C-4′′′ of α-Rha′′′ (δC 80.4), and H-1′′′′′ of α-Rha′′′′′ (δH 5.94) with C-3′′ of α-Rha′′ (δC 73.9). In addition, the position of Mba was located at C-4′′′′′ of α-Rha′′′′′, and the n-dodecanoyl moiety at C-2′′′ of α-Rha′′′ was based on the long-range correlations between H-4′′′′′ of α-Rha′′′′′ (δH 5.81) with C-1 of Mba (δC 176.6) and H-2′′′ of α-Rha′′′ (δH 5.97) with the C-1 of the dodecanoyl (δC 173.4) moiety, respectively. The position of lactonization at C-2′′ of the rhamnosyl unit was established by the correlations between H-2′′ (δH 6.02) of Rha′′ and C-1 (δC 173.3) of the jalapinolic acid moiety, respectively.
  • NOESY: providing the diagnostic NOE correlations between the interglycosidic protons, H-1′/H-11, H-2′/H-1′′, H-3′′/H-1′′′′′, H-1′′/H-1′′′, and H-4′′′/H-1′′′′.
[49]
5Calysolin XVIII (188)
Hexasaccharides
Calystegia soldanella
  • Mass spectrometry: finding the molecular formula to be C72H120O35 using high-resolution positive-ion FAB-MS.
  • 1H-NMR: comparing the chemical shifts of signals arising from the sugar moieties in calysolin XVIII 1 and its methyl ester 3a, and revealing downfield shifts (Δδ = δ1–δ3a) due to the acylation of signals assignable to H-2 (Δδ = ca. 1.10) and H-4 (Δδ = 1.46) of the rhamnosyl residue (Rha), H-3 (Δδ = ca. 1.34) and H-4 (Δδ = ca. 1.56) of the second glucosyl residue (Glc′), and H-2 (Δδ = 1.46) of the fourth glucosyl residue (Glc′′′). Signals due to H2-2 of the aglycone (11S-jalapinolic acid) moiety (Jla) of calysolin XVIII were non-equivalent at δ 2.55 and 2.71, whereas those due to H2-2 of Jla and its methyl ester were equivalent at δ 2.32 (2H, t, J = 7.5 Hz) [45].
  • HMBC: observing the key cross peaks between the H-4 of Rha and C-1 of the 3-hydroxy-2-methylenebutyryl residue (Hma), H-3 of Glc′ and C-1 of the second 2-methylbutyryl residue (Mba′), H-4 of Glc′ and C-1 of the third 2-methylbutyryl residue (Mba′′), and H-2 of Glc′′′ and C-1 of Jla in the HMBC spectrum. Although no cross-peak was identified between the H-2 of Rha and C-1 of the first 2-methylbutyryl unit (Mba), the ester linkages of Mba, Mba′, Mba′′, Hma, and Jla were located at OH-2 of Rha, OH-3 of Glc′, OH-4 of Glc′, OH-4 of Rha, and OH-2 of Glc′′′, respectively. These linkages were supported by the fragment ion peaks observed at m/z 1399 [1543 (M−H)−144 (C H O)]−, 1301 684 [1399 − 98 (2-methylenebutyryl residue)]−, 1053 [1543 − 162 (hexosyl residue) − 146 (6-deox-yhexosyl residue) − 98 − 84 (2-methylbutyryl residue)]−, 579 [1053 − 162 − 144 − 84 × 2]−, 417 [579 − 162]−, and 271 [417 − 146]− in the negative-ion FAB-MS of 1 [72].
  • The absolute configuration of the component 2-methylbutyric acid of the crude resin glycoside fraction of this plant was previously determined to be S [45].
[14]
6Arvensic acid A-D (207210)
Heptasaccharides
Convolvulus arvensis
  • Mass spectrometry: determining molecular formula to be C57H100O35, C58H102O35, C57H100O36, and C58H102O36, respectively, according to the HRESIMS data of [M − H]− at m/z 1343.5999, 1357.6155, 1359.5959, 1373.6099, respectively.
  • NMR: The 1H and 13C NMR spectra of compounds 1–4 were similar with each other, especially in the signal region of sugar moieties (δH 3.5–6.5, δC 60–110).
  • 1H NMR: seven anomeric proton signals at δ 4.95 (1H, d, J = 7.5 Hz), 5.80 (1H, d, J = 7.5 Hz), 6.35 (1H, br s), 5.39 (1H, d, J = 7.5 Hz), 4.92 (1H, d, J = 7.5 Hz), 5.46 (1H, br s), and 4.91 (1H, d, J = 7.5 Hz), and three secondary methyl resonances at δ 1.78 (3H, d, J = 6.0 Hz), 1.48 (3H, d, J = 6.5 Hz), and 1.66 (3H, d, J = 6.0 Hz), were assignable to the CH3–6 of 6-deoxyhexoses.
  • 13C NMR: seven anomeric carbon signals in the range δ 102.2–106.5.
  • 2D NMR spectra (HSQC, HMBC, 1He1H COSY and TOCSY): allowing the assignment of the heptasaccharide skeleton, which was comprised of one D-fucose, two L-rhamnose, and four D-glucose units.
  • 1H NMR: determining the β-configurations of the fucose and glucose units by the large coupling constants of the anomeric protons.
  • 13C NMR: determining the α-configurations for L-rhamnose units by the chemical shifts of C-5 of rhamnose.
  • HMBC: showing the connnectivities between the aglycones (Ag) and sugar chains, and among the monosaccharides in 1–4, which were established according to their HMBC spectra in which, also using one as an example, long range correlations were observed from δH 4.95 (H-1, Glc) to δC 81.6 (C-12, Ag), from δH 5.80 (H-1, Glc′) to δC 79.6 (C-2, Glc), from δH 6.35 (H-1, Rha) to δC 78.0 (C-2, Glc′), from δH 5.39 (H-1, Glc″) to δC 84.1 (C-3, Rha), from δH 4.92 (H-1, Fuc) to δC 81.9 (C-4, Glc ″), from δH 5.46 (H-1, Rha′) to δC 68.7 (C-6, Glc), and from δH 4.91 (H-1, Glc‴) to δC 90.3 (C-3, Glc′).
[64]
7Jalapinoside I (224)
Bidesmoside
Ipomoea purga
  • Mass spectrometry (HRESIMS and MALDI-TOFMS): identifying quasi-molecular ions at m/z 1689.8448 [M + H]+ (C80H137O37).
  • LC-ESIMS/MS showing a quasi-molecular ion of low abundance at m/z 1313 [M − H]−, corresponding to the molecular formula C56 H97O34, and an intense ion at m/z 1167 [M − H − 146]−, corresponding to the elimination of one methylpentose unit which, in turn, produced the fragmentation pattern of a hexasaccharide with the same glycosidation sequence as purgic acid A but with a difference of 16 amu based on a negative ion FAB with diagnostic ions at m/z 1021 [1167 − 146]−, 859 [1021 − 162]−, 713 [859 − 146]−, 567 [713 − 146]−, and 405 [567 − 162]−, indicating the presence of ipurolic acid as the aglycone.
  • 1H and 13C NMR: the data were compared with purgic acids A and B. The main differences were in the esterifying residues and macrolactonization, as well as the presence of ipurolic acid as the aglycone, containing an additional quinovose which was glycosidically linked at C-3 to constitute a macrocyclic resin glycoside bisdesmoside.
  • COSY and TOCSY-1: assigning chemical shift values after identifying and differentiating each of the 1H NMR signals.
  • 1H−13C HSQC: sssigning the 13C NMR signals
  • NMR: confirming seven anomeric signals at δH 4.90 (1H, d, J = 7.6 Hz; δC 102.9, Qui-1); 5.86 (1H, d, J = 7.8 Hz; δC 101.6, Glc-1); 6.47 (1H, d, J = 1.3 Hz; δC 96.8, Rha-1); 6.42 (1H, d, J = 7.8 Hz; δC 101.6, Glc′-1); 5.15 (1H, d, J = 7.8 Hz; δC 106.0, Qui′-1); 5.78 (1H, d, J = 7.6 Hz; δC 101.8, Fuc-1); and 4.82 (1H, d, J = 7.8 Hz; δC 103.0, Qui′′-1).
  • COSY and TOCSY-2: distinguishing seven separate spin systems for sugar skeletons.
  • HMBC: establishing the glycosylation sequence and the positions of the esterification, lactonization, and ester-type linkage. The following key correlations were observed: connectivities between H-1 (δH 4.90, Qui-1) and C-11 (δC 81.2, Ipur-11); H-1 (δH 4.82, Qui′′-1) and C-3 (δC 68.6, Ipur-3); H-1 (δH 5.86, Glc-1) and C-2 (δC 80.4, Qui-2); H-2 (δH 4.28−4.26, Glc-2) and C-1 (δC 96.8, Rha-1); H-1 δH 6.42, Glc′-1) and C-3 (δC 74.0, Rha-3); H-1 (δH 5.15, Qui′-1) and C-2 (δC 86.3, Glc′-2); and H-4 (δH 4.61, Rha-4) and C-1 (δC 101.8, Fuc-1). The locations of the acyl residues were established by the following HMBC correlations: H2-6 (δH 5.69 and 5.66, Glc-6) and the carbonyl of an acetate group (δC 168.5); H-2 (δH 5.91, Rha-2) and C-1 of a 2-methylbutyrate residue (δC 176.6); H-3 (δH 5.35, Qui′-3) and C-1 of a second methylbutyrate moiety (δC 176.2); and H-3 (δH 5.99, Fuc-3) with the carbonyl of a dodecanoate group (δC 173.1).
[66]
8Batatin VII (248)
Oligomer ester type dimer
Ipomoea batatas
  • Mass spectrometry (ESI and FAB in the negative ion detection): allowing for the detection of the quasi-molecular ion [M − H], indicating a composistion of C126H219O48. A fragment [M/2 − H]− was observed in both C126H219O48 ionization techniques, which indicated the rupture of the ester type bond and represented the high-mass fragment ions for the two monomeric units (m/z 1249 [A − H] and m/z 1251 [B − H]) as previously reported for the batatin series. This technique provided an easily detectable and intense quasi-molecular ion [M − H]− as well as all of the characteristic ions resulting from the glycosidic cleavage. The FAB/MS of batatinosides VII showed a peak at m/z 1181 [M − H]−, indicating a molecular formula of C58H102O24.
  • MALDI/MS: confirming the molecular formula. The fragmentation pattern was determined with the peaks observed at m/z 271, 417, 545, 837, and 999 that are common to resin glycosides containing operculinic acid A as their glycosidic acid. The fragment at m/z 999 [1181 − C12H22O] wasindicated by the loss of glucose as identified by the fragment at m/z 837 [999 − C6H10O5].
  • NMR: determining the presence of two units of simonic acid B, through a comparison of its spectroscopic constants with those previously reported for batatins I and II. This dimer illustrates the four-step approach for the identification of carbohydrate structural elements in the oligosacharaide core. Estimations of the sugar units were conducted through the anomeric protons around 4.8−6.4 ppm, which were used as “structural reporter groups” for the dimeric structure.
  • COSY and TOCSY: identifying each constitutive monosaccharide through the assignment of chemical shift values, after identifying and differentiating the 1H NMR signals.
  • HSQC: providing the assignments for the 13C NMR data.
  • HMBC: observing the glycosylation sequence and the positions of esterification and lactonization. The carbonyl resonances at δC 175.2 were assigned to the unit A lactone functionality due to its 2JCH coupling with the C-2 diastereotopic methylene protons centered at δH 2.26−2.30. The macrolactonization site at C-3 of the first rhamnose unit was established by the observed 3J correlation between this CH carbonyl carbon and the H-3 of rhamnose (δH 5.63). The ester-type linkage established by the acyclic unit B at the macrocyclic unit A was identified by 3JCH correlation between the carbonyl group for the ester (δC 174.9; unit B) and H-2 of the terminal rhamnose (Rha′′′, δH 5.77) on unit A. The dodecanoyl residue on both units A and B (δC 173.4) was located at C-2 of the third saccharide unit (Rha′, δH 4−99). The methylbutanoyl residue (δC 166.9, unit A; δC 166.4, unit B) was located as the acylating group at positions C-4 (δH 5.86, unit A; and δH 5.93, unit B) of the terminal rhamnose unit (Rha′′).
[41]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Maharani, R.; Fajar, M.; Supratman, U. Resin Glycosides from Convolvulaceae Family: An Update. Molecules 2022, 27, 8161. https://doi.org/10.3390/molecules27238161

AMA Style

Maharani R, Fajar M, Supratman U. Resin Glycosides from Convolvulaceae Family: An Update. Molecules. 2022; 27(23):8161. https://doi.org/10.3390/molecules27238161

Chicago/Turabian Style

Maharani, Rani, Mohamad Fajar, and Unang Supratman. 2022. "Resin Glycosides from Convolvulaceae Family: An Update" Molecules 27, no. 23: 8161. https://doi.org/10.3390/molecules27238161

Article Metrics

Back to TopTop