Next Article in Journal
Pattern Evolution during Double Liquid-Vapor Phase Transitions under Weightlessness
Previous Article in Journal
Subcritical Water Chromatography with Electrochemical Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Isolation of Monoterpene Dihydrochalcones from Piper montealegreanum Yuncker (Piperaceae)

by
Harley Da Silva Alves
1,*,
Wilma Raianny Vieira da Rocha
1,
Raimundo Braz-Filho
2,3 and
Maria Célia de Oliveira Chaves
4
1
Department of Pharmacy/Post-Graduation Program in Pharmaceutical Sciences, State University of Paraíba, 58429-500 Campina Grande-PB, Brazil
2
Chemical Science Laboratory, CCT, State University of Fluminense North Darcy Ribeiro, 28013-600 Campo dos Goytacazes-RJ, Brazil
3
Department of Chemistry, DEQUIM-ICE-Federal University Rural of Rio de Janeiro, CEP 23894-374 Seropédica-RJ, Brazil
4
Research Institute for Drugs and Medicines, Federal University of Paraíba, CEP 58051-900 João Pessoa-PB, Brazil
*
Author to whom correspondence should be addressed.
Molecules 2017, 22(6), 874; https://doi.org/10.3390/molecules22060874
Submission received: 5 April 2017 / Revised: 18 May 2017 / Accepted: 23 May 2017 / Published: 9 June 2017
(This article belongs to the Section Natural Products Chemistry)

Abstract

:
Four new compounds were isolated from the branches of Piper montealegreanum Yuncker, a shrub found in the Amazon rainforest, including two new dihydrochalcones named claricine (1) and maisine (2), a cinnamic acid derivative 3 and a phenylalkanoid 4, along with a porphyrin identified as the known compound phaeophytin a (5). The structures were established using spectroscopic experiments, including 1D and 2D NMR and HRESIMS experiments, performed on the two monoterpene dihydrochalcones and their monoacetyl derivatives. The structural diversity of these substances is very important for the Piper genus chemotaxonomy.

1. Introduction

Piper (Piperaceae) is a diverse pantropical plant genus, consisting of approximately 2000 species, and is a rich source of numerous biologically active products [1,2]. Several species of Piper are used in traditional medicine as analgesics, anti-inflammatory compounds, and for the treatment of snakebites [3]. The phytochemical investigation of the genus Piper, widely distributed in the tropical and subtropical regions of the world, has allowed the identification of many classes of compounds such as aristolactams [4,5], alkamides [6,7], lignans [8], chalcones, and dihydrochalcones [9]. Recently we reported the isolation of some bioactive agents including (S)-8-formyl-3′,5-dihydroxy-7-methoxy-6-methylflavanone (A), 3′-formyl-3,4′,6′-trihydroxy-2′-methoxy-5′-methylchalcone (B) and the phenylpropanoid ethyl-3,4-methylenedioxy-5-methoxy-7,8-dihydrocinnamate (C) from a crude ethanolic extract (CEE) of the branches of Piper montealegreanum [10], a native species of the Amazon region, north of Brazil [11]. We also recently disclosed the analgesic and anti-inflammatory effects of compounds A and B [12].
Piper montealegreanum is a branching shrub upper with elongate internodes and oblong or lance elliptic leaves with an acuminate apex and subequilaterally rounded. It presents spikes of 3 mm thick and up to 9 cm long; drupes are oblong or obovoid, laterally compressed. It was first recorded in 1908 in the city of Monte Alegre, Pará-Brazil [11].
In the present paper, we report the isolation and structure elucidation of two new compounds 1 and 2, which we have given the trivial names claricine and maisine, respectively, ethyl-3,4,5-trimethoxy-7,8-dihydrocinnamate (3), (5-(3E,5E)-undeca-3,5-dienyl)benzo[d][1,3]dioxole (4) and the known compound phaeophytin a (5). The structures of the compounds were determined by interpretation of the spectral data, including 1D and 2D (HMQC, HMBC, COSY and NOESY) NMR and HRESIMS experiments, of the new monoterpene dihydrochalcones and their monoacetyl derivatives in conjunction with data reported in the literature.

2. Results and Discussion

Chromatographic procedures led to the isolation of five compounds from P. montealegreanum (Figure 1). Compounds 1 and 2 were identified as monoterpene dihydrochalcones. One of the most characteristic features of the structural chemistry of chalcones and dihydrochalcones is their ability to act as dienophiles in enzyme-catalyzed Diels-Alder reactions [13]. In the biosynthetic route proposed by Hano et al. [14] it was possible predicted how an initial adduct reacted with the 2′,4′,2,4-tetrahydroxychalcone to form the monoterpene dihydrochalcone (Scheme 1). Few plant families such as the Anonnaceae [15], Moraceae [16] and Piperaceae [17] have a group of adducts. The crinatusins, obtained from the stems of Cyathocalyx crinatus, showed lethal toxicity in a brine shrimp bioassay [15]. Palodesangrens A–E, obtained from the bark of Brosimum rubescens, effectively inhibited the formation of a 5a-dihydrotestosterone androgen receptor complex implicated in androgen-dependent diseases [18]. Cytotoxicity toward KB cells was recorded for both fissistin and isofissistin [19].
The general metabolism of phenylpropanoid 3 can generate an enormous array of secondary metabolites (flavonoids, coumarins, lignins, hydrolyzable tannins) based on the few intermediates of the shikimate pathway as the core unit [20]. Phenylalkanoids like 4 act as chemotaxonomy markers and exhibit antifungal activity [21]. The porphyrinic compounds (such as phaeophytin a) are derived from chlorophyll a and are widely present in the plant kingdom [22].
The various compounds isolated from different Piper species have been classified under twelve categories, viz. alkaloids/amides, propenylphenols, lignans, neolignans, terpenes, steroids, kawapyrones, piperolides, chalcones and dihydrochalcones, flavones, flavanones and miscellaneous compounds [4]. Most of the chalcones and dihydrochalcones isolated from the Piper species are oxygenated at the C-2′, C-4′ and C-6′ positions and the B-ring is unsubstituted in all the chalcones and dihydrochalcones, except asebogenin [23], flavokawains A and C [24] in which C-4 position is also oxygenated. Five new unusual monoterpene-substituted dihydrochalcones, adunctins A–E [25], were isolated from the leaves of P. aduncum. Various propenylphenols have been isolated so far from different Piper species, such as ω-hydroxyisodillapiole [26] and safrole [27]. Besides the above mentioned groups of compounds, several long chain alkanes, alcohols, acids, esters and oxygenated cyclohexanes have been isolated from Piper species (e.g., villiramulin B) [28].
Compound 1 was obtained as yellow gummy solid and its IR spectrum showed absorption bands at νmax 3433 cm−1 (OH), 2850 cm−1 (formyl C–H), 1618 cm−1 (suggesting conjugated carbonyl), and 1458 and 1419 cm−1 (aromatic ring). The molecular formula C28H32O6 was determined by negative-mode HRESIMS measurement of the molecular ion at m/z 463.2150 ([M − H], calcd. for C28H31O6, 463.2126, delta 5.2 ppm), corresponding to 13 double-bond equivalents.
The UV spectrum (in MeOH) showed two absorption bands at λmax 274 and 338 (sh) nm. The use of AlCl3 produced bathochromic shift to λmax 299, 313 and 371 (sh) nm, respectively, suggesting the presence of a flavonoid-sustaining chelated hydroxyl group and/or ortho-dihydroxy system [13,29].
The 13C-NMR spectrum revealed carbon signals (Table 1) corresponding to the presence of four methyl (including methoxyl group at δC 67.23), four sp3 methylene (δC 29.79, 26.38, 36.88 and 37.23), nine methine (two sp3C 45.19 and 51.0) and seven sp2 (including one formyl group at δC 192.36), and eleven sp2 non-hydrogenated (including one conjugated carbonyl at δC 209.55) carbon atoms, compatible with partial expanded molecular formula (C=O)(C)10(CHO)(CH)8(CH2)4(CH3)3(CH3O) = C28H29O3, suggesting the presence of three hydroxyl groups consistent with the 13C chemical shifts observed for oxygenated aromatic carbon atoms at δC 155.46 (C-3), 165.21 (C-4′, para-conjugated with carbonyl) and 167.23 (C-6′, ortho-conjugated with carbonyl), along with one methoxylated C-2′ at δC 166.65 (Table 1).
This data, involving a comparison with 1H- and 13C-NMR data of hostmanin D [30], suggested 1 to be a prenylated dihydrochalcone with a totally substituted ring A. This deduction was supported by differences justified by the presence of formyl and hydroxyl groups at rings A and B of 1, in agreement with HRESITOFMS (negative mode) compatible with the proposed fragmentation summarized in Scheme 2.
The 1H-NMR and HMQC spectra of the compound 1 (Table 1) exhibited signals attributed to dimethylallyl groups CH3-8″ (δH 1.62 (d, J = 0.5 Hz)/δC 17.84, shielded by γ-effect of the methylene group CH2-5″) and CH3-9″ (δH 1.70 (d, J = 1.0 Hz, 3H)/δC 25.69), the aromatic methoxy group MeO-2′ (δH 3.90 (s)/δC 67.23, located in a position steric hindrance), the aryl formyl hydrogen 4-CHO (δH 10.04 (s)/δC 192.36), the aromatic methyl group CH3-7′ (δH 1.89 (s)/δC 6.56), two olefinic signals CH-2″ (δH 5.53 (m)/δC 118.73) and CH-6″ (δH 5.12 (qt, J = 7.0, 1.5 Hz)/δC 124.02), two hydrogen-bonded OH at δH 12.88 (1H, s, OH-6′) and 12.46 (1H, s, OH-4′), and a disubstituted phenyl group. Concerning this phenyl group, the NMR and HMQC spectra displayed four methine carbons: CH-2 at δH 6.58 (dd, J = 2.5, 1.5 Hz)/δC 114.48, CH-4 at δH 6.48 (ddd, J = 7.5, 2.5, 1.5 Hz)/δC 113.30, CH-5 at δH 6.94 (t, J = 7.5 Hz)/δC 129.35, and CH-6 at δH 6.64 (dd, J = 7.5, 1.5 Hz)/δC 119.82. These data suggested a 1,3-disubstituted aromatic ring. The HMBC spectrum showed heteronuclear correlations of the signals corresponding to non-hydrogenated carbon C-5′ at δC 108.85 with hydrogens 3H-7′ at δH 1.89 (2JCH), HO-4′ at δH 12.46 (3JCH) and HO-6′ at δH 12.88 (3JCH). Thus, these data were used to locate the methyl group (δH 1.89) at C-5′ and the hydroxyl groups (δH 12.88 and 12.46) in C-6′ and C-4′, respectively. The correlations between the signal at δC 107.97 (C-3′) with the chelated hydroxyl group at δH 12.46 (HO-4′, 3JCH) and the formyl hydrogen at δH 10.04 (CHO-8′, 3JCH) were used to assign the aldeidic group to C-3′ and to confirm the chelated hydroxyl group at δH 12.46 to C-4′. Correlations observed in the HMBC spectrum between δH 3.90/δc 166.65 permitted us to assign the methoxy group at δH 3.90 in C-2′. The presence of signals corresponding to two methine carbons α (δC 51.06, CH-8) and β (δC 45.19, CH-7) of the dihydrochalcone skeleton suggested that a diprenylated moiety was located in these positions. This deduction is confirmed by the correlations revealed by the HMBC spectrum: δC 209.55 (C-9), 45.19 (CH-7), 29.79 (CH2-1″) and 36.88 (CH2-10″) with δH 4.13 (H-8) (Figure 2). Additional heteronuclear long-range couplings are summarized in Table 1.
The relative stereochemistry at CH-7 and CH-8 was deduced from the J-value between H-7 and H-8 (J = 11.0 Hz) characteristic of a trans axial-axial coupling [30]. The HMBC correlations observed between H-4″ (δH 2.02 (m)), H-10″ (δH 2.23 (m) and 2.20 (m)) and H-5″ (δH 2.14 (m)) with C-3″ (δC 137.41) confirmed the attachment of the prenyl unit at C-3″ (Table 1). However, there was two structural alternatives for the diprenylated moiety. The COSY spectrum visualized correlations between δHH 2H-10″/H-7 and δHH 2H-1″/H-8 and established the first structural alternative [15,30] (Figure 3).
In the 13C-NMR spectrum of the acetylated derivative (1a) of 1, a magnetic deshield was observed at CH-2 and CH-4 (ortho positions) and at CH-6 (para position) of δC 5.7, 6.1 and 5.2 ppm, respectively, in accordance with the acetylation of OH-3 located in the B aromatic ring and confirming that this type of monosubstituition is relatively rare in the B ring of flavonoids. In addition, the presence of the acetate group was confirmed from absorption of methyl carbon at δC 20.86 and of the ester carbonyl at δC 169.05.
Compound 2 was isolated as yellow gum with molecular formula C28H32O7 determined by ESI-Q-TOF-HRMS at m/z 479.2107 ([M − H], calcd. 480.2149) (Scheme 3).
The IR spectrum showed a broad band at νmax 3429 cm−1 due to hydrogen bonded OH, conjugated formyl carbonyl (1716 cm−1), conjugated cetonic carbonyl (νmax 1620 cm−1) and aromatic ring (νmax 1454 and 1419 cm−1). In its UV spectrum in MeOH, the presence of an absorption band was observed at λmax 274 nm plus a shoulder at 338 nm. The use of a shift reagent (AlCl3) promoted a bathochromic shifts λmax 312 nm and 368 nm, suggesting the presence of one dihydrochalcone [12,29]. In the 1H-NMR spectrum of 2, characteristic signals of a 2′-methoxy-3′-formyl-4′,6′-dihydroxy-5′-(methyl-phenyl)-7-phenyl-(3-hydroxy)-cyclohex-2″-enyldihydrochalcone derivative were observed (Table 1). These spectroscopic data are quite similar to those of 1. The 13C-NMR spectrum contained six additional resonances with one quaternary carbon (δC 147.33, C-7″), one methyl at δC 17.63 (CH3-9″), three methylenes at δC 33.20 (CH2-4″), 32.91 (CH2-5″) and 111.19 (CH2-8″), and one methine at δC 75.73 (CH-6″) bearing an O-atom. The presence of an olefinic methylene signal (δC 111.19) and only a methyl signal (δC 17.63), associated with molecular formula (C28H32O7), suggested that 2 has a hydroxyl group at C-6″ and a double bond at C-7″, in contrast to 1 that has one quaternary carbon (δC 131.70) and two methyl (δC 17.84 and 25.69), two methylenes (δC 37.23 and 26.38) and one methine (δC 124.02) signals. The spectroscopic data of 6″-hydroxy-7″-methylpentenyl moiety was established with combined results of HMQC and HMBC experiments and was supported by HMBC correlations between CH-6″ (δC 75.73) and 2H-8″ (δH 4.93 (m), δH 4.84 (m) and Me-9″ (δH 1.73 (m)).
Spectral analyses of 2a showed a magnetic deshielding effect at CH-2, CH-4 and CH-6 for δC 120.30, 119.35 and 124.97, respectively. The absorptions at δC 170.30 (C-10) and 21.08 (CH3-11) confirmed the presence of the acetate group. A marked difference was also observed in the chemical shift at CH-6″ for δC 71.17, suggesting the presence of the acetate group at C-6′ and a shield of C-7″ and CH2-5″ by γ-effect of the carbonyl carbon.
The phenylpropanoid was obtained of a mixture with two substances: the majority being 1 and other encoded as 3 (Figure 3). In the 13C-NMR spectrum, the presence of 39 signals was observed (APT experiment), 11 assigned to 3, including four for quaternary carbons (δC: 136.00 (C-1), 153.21 (C-3/5), 136.50 (C-4) and 173.80 (C-9)), one for methine carbon (δC 105.36 (CH-2/CH-6)), three for methylenes carbons (δC: 31.37 (CH2-7), 36.09 (CH2-8) and 60.47 (CH2-10)) and three methyl carbons (δC 60.82 (MeO-4), 56.09 (MeO-3/5), and 14.22 (CH3-11)). In the 1H-NMR spectrum, three singlets were seen at δH: 6.40 (2H, brs, H-2/H-6), 3.82 (6H, s, MeO-3/MeO-5) and 3.80 (3H, s, MeO-4), the last two being characteristic of aromatic methoxyls, suggesting the presence of a tetrasubstituted aromatic ring in the structure. Other signals observed were δH: 2.87 (t, J = 8.0 Hz, 2H-7), 2.59 (t, J = 8.0 Hz, 2H-8), 4.10 (m, 2H-10) and 1.22 (t, J = 7.0 Hz, 3H-11). The NMR assignments and comparison with literature data [31] allowed us to deduce 3 was a new cinnamic acid derivative, 3,4,5-trimethoxy-hydrocinnamic acid ethyl ester.
Compound 4 was isolated as a yellow oil. The IR spectrum showed bands at νmax 2924 and 2852 cm−1 characteristic of aliphatic C–H stretching; at νmax 1489 and 810 cm−1 consistent with C=C trisubstituted aromatic ring; and at νmax 1442 and 1246 cm−1 suggestive of C–H and C–O stretch, respectively. Broad bands at νmax 1718, 1653 and 939 cm−1 suggested the presence of bond-conjugated C=C with trans stereochemistry [32].
The 1H-NMR (δ, CDCl3, 500 MHz) showed three signals at δH: 6.70 (d, J = 8.0 Hz, H-5), 6.65 (brd, J = 1.5 Hz, H-2) and 6.60 (dd, J = 8.0, 1.5 Hz, H-6) and a singlet at δH 5.89 (2H, OCH2O-3-4), characteristic of methylenedioxy group. These data suggested the presence of trisubstituted aromatic ring. The signals at δH: 5.99 (brt, J = 14.5, 1.0 Hz, H-4′/H-5′), 5.58–5.52 (m, H-4′/H-6′), 2.59 (t, J = 7.5 Hz, 2H-1′), 2.30 (td, J = 7.5, 8.0 Hz, 2H-2′), 2.02 (td, J = 7.5, 7.0 Hz, 2H-7′), 1.35–1.31 (m, 2H-8′), 1.29–1.26 (m, 2H-9′, 2H-10′) and 0.86 (t, J = 7.0 Hz, 3H-11′) suggested the presence of an unsaturated acyclic chain with a terminal aliphatic methyl. In the 13C-NMR APT experiment, 18 signals were observed , three of which were for quaternary carbons at δC 147.47 (C-3), 145.54 (C-4) and 135.80 (C-1); seven for methine carbons at δC 133.10 (CH-6′), 130.99 (CH-4′), 130.84 (CH-3′), 130.08 (CH-5′), 121.08 (CH-6), 108.85 (CH-2), 108.05 (CH-5); seven for methylene carbons at δc 100.69 (OCH2O), 35.62 (CH2-1′), 34.70 (CH2-2′), 32.59 (CH2-7′), 31.89 (CH2-9′), 29.40 (CH2-8′) and 22.67 (CH2-10′); and one methyl carbon at δc 14.12 (Me-11′) [22]. The structure of 4 was supported by 2D NMR experiments (Table 2) and literature data [28] and elucidated as (5-(3E,5E)-undeca-3,5-dienyl)benzo[d][1,3]dioxole (Figure 3). Using similar methods as above along with literature data [22,33], compound 5 was identified as phaeophytin a (Figure 3), isolated for the first time from P. montealegreanum.

3. Materials and Methods

3.1. General Procedures

UV spectra were recorded on a Vankel-50 UV-Vis spectrophotometer (Varian, Palo Alto, CA, USA) and AlCl3 was used as shift reagent. IR spectra were obtained on an FT-IR-1750 spectrophotometer (Perkin-Elmer, Waltham, MA, USA) using KBr disks. The spectra mass were obtained on a LCMS-IT-TOF (225-07100-34) (Shimadzu, Tokay, Japan) equipped with a Z-spray ESI (electrospray) source and operated in negative mode. 1H- and 13C-NMR (1D and 2D) spectra were recorded on a Mercury 500 and 200 MHz spectrometer (Varian, Palo Alto, CA, USA) in CDCl3 using TMS as internal standard. Sephadex LH-20 and silica gel 60 (PF 254 art. 7749 and art. 7731) were purchased from Merck (Kenilworth, NJ, USA). Hexane, chloroform, ethyl acetate and methanol were of analytical grade and were purchased from Merck. The acetylated flavonoids were obtained by dissolving the samples in pyridine and acetic anhydride (1:1—v/v), and the mixture was maintained under stirring for 48 h. The reaction progress was monitored by TLC. After this period, the mixture was treated with ice-cold distilled water until the formation of a precipitate. The precipitate was filtered and washed several times with distilled water, dissolved in chloroform, and dried with anhydrous sodium sulfate. The remaining solvent was evaporated in a rotating evaporator [34].

3.2. Plant Material

Branches from P. montealegreanum Yuncker were collected in Belém (Pará State, Brazil; Latitude 14°10′00″ S, longitude 53°05′00″ W). The botanical material was identified by Dra. Elsie F. Guimarães, at the Botanical Department of UFRJ, and a specimen voucher was deposited at Emilio Goeldi Museum, Belém, under serial number MSP-010.

3.3. Extraction and Isolation

A powered material of P. montealegreanum (1.3 kg) was treated with EtOH (4 × 2.0 L) at room temperature (30 °C) and for three consecutive days, then the solvent was removed under reduced pressure to furnish the crude ethanolic extract (CEE, 115 g), which was partitioned in a series of solvents to yield hexane (44 g), chloroform (37 g), ethyl acetate (8 g) and hydroalcoholic (4.5 g) fractions. A portion (9.5 g) of the hexane fraction was chromatographed over a silica gel column using a gradient of solvents with increasing polarity (hexane, chloroform and methanol) yielding 109 fractions. The fractions 38–42 (1.169 mg) were rechromatographed using Sephadex LH-20 and eluted with methanol. Of the 19 fractions obtained, fractions 12–13 and 14 were purified over silica gel column and preparative TLC to yield 1 (9 mg) and 2 (6 mg), respectively. Fraction 10–11 was chromatographed on Sephadex LH-20 and purified by preparative TLC yielding 4 (12 mg). The supernatant of 12–13 was chromatographed by preparative TLC to obtain a mixture of 1 and 3 (14 mg). 11.0 g of chloroform phase, eluted over Sephadex LH-20 column with a binary mixture of methanol−chloroform (1:1—v/v), provided compound 5 (25 mg).
(2′-Methoxy-3′-formyl-4′,6′-dihydroxy-5′-methylphenyl)-[3″-(7″-dimethylbut-6″-enyl)-7-phenyl-(3-hydroxy)-cyclohex-2″-enyl]-methyl-9-one (claricine, 1). Yellow gum; UV (MeOH) λmax 274, 338 (sh) nm; (MeOH + AlCl3) λmax 299, 313, 371 (sh) nm; IR (KBr disc) νmax 3.433, 2.922, 1.618, 1.458, 1.419 cm−1; for 1H-NMR and 13C-NMR spectroscopic data, see Table 1; HRESIMS m/z 463, 2150 ([M − H], (calcd. for C28H32O6, 464,2200).
Acetylated monoterpene dihydrochalcone 1a. For 1H-NMR 1H-NMR and 13C-NMR spectroscopic data, see Table 1.
(2′-Methoxy-3′-formyl-4′,6′-dihydroxy-5′-methylphenyl)-[3″-(6′-hydroxy-7″-methylpentenyl)-7-phenyl-(3-hydroxy)-cyclohex-2″-enyl]-methyl-9-one (maisine, 2). Yellow gum; UV (MeOH) λmax 274, 338 (sh) nm; (MeOH + AlCl3) λmax 300, 312, 368 (sh) nm; IR (KBr disc) νmax 3429, 2924, 1716, 1620, 1454, 1419 cm−1; for 1H-NMR and 13C-NMR spectroscopic data, see Table 1; HRESIMS m/z 479,2107 [M − H] (calcd. for C28H32O7, 480,2149).
Acetylated monoterpene dihydrochalcone 2a. For 1H-NMR and 13C-NMR spectroscopic data, see Table 1.
Ethyl-3,4,5-trimethoxy-7,8-dihydrocinnamate (3). 1H-NMR (CDCl3, 500 MHz) δH 6.40 (2H, s), 4.10 (2H, m), 3.82 (6H, s, OCH3-3 and 5), 3.80 (3H, s, OCH3-4), 2.87 (2H, t, J = 8.0 Hz, H-7), 2.59 (2H, J = 8.0 Hz, H-8), 1.22 (3H, t, J = 7.0 Hz); 13C-NMR (CDCl3) δC 14.22, 31.37, 36.09, 56.09 (2C), 60.47, 60.82, 105.36, 136.00, 136.50, 153.21, 173.80.
(5-(3E,5E)-Undeca-3,5-dienyl)benzo[d][1,3]dioxole (4). Yellow oil; IR (KBr disc) νmax 2924, 2852, 1718, 1653, 1489, 1442, 1246, 939, 810 cm−1; for 1H-NMR and 13C-NMR spectroscopic data, see Table 2.
Phaeophytin a (5). Green powder; IR (KBr disc) νmax 3433, 2924, 2850, 1731, 1701, 1620, and 1377 cm−1; 1H-NMR and 13C-NMR data were consistent with literature [22,33].

4. Conclusions

The present study, based on the phytochemical investigations of P. montealegreanum, using chromatographic and spectroscopic techniques, led to the identification of five substances, including two flavonoids with rare skeletons, called monoterpene dihydrochalcones. The acetylated derivatives of the dihydrochalcones claricine (1) and maisine (2) were prepared to identify their structures. The cinnamic acid derivative, ethyl-3,4,5-trimethoxy-7,8-dihydrocinnamate (3) and the phenylalkanoid, (5-(3E,5E)-undeca-3,5-dienyl)benzo[d][1,3]dioxole (4) were isolated for the first time as natural products, along with the known compound phaeophytin a (5). Monoterpene dihydrochalcones are unusual compounds restricted to a few plant genus, such as Piper. Besides that, phenylpropanoids are precursors of some classes of secondary metabolites, such as flavonoids. Thus, these structures are important for the chemotaxonomy of the Piper genus. These data corroborate with previous studies on the species evaluating the biological properties of the isolated products.

Supplementary Materials

Supplementary File 1

Acknowledgments

The authors thank: Conselho Nacional de Desenvolvimento Científico e Tecnológico for financial support; Maria das Graças Bichara Zoghbi for plant material; Laboratório de Espectrometria de Massas do Nordeste for MS spectra; and Ipefarm/Federal University of Paraíba for IR, UV and NMR spectra.

Author Contributions

Harley da Silva Alves prepared the extracts, isolated the compounds, worked on structural elucidation and participated in manuscript writing. Maria Célia de Oliveira Chaves worked on structural elucidation and wrote the manuscript. Raimundo Braz-Filho participated in structural elucidation and editing of manuscript. Wilma Raianny Vieira da Rocha revised the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Asri, S.F.M.; Ramli, E.S.M.; Soelaiman, I.N.; Noh, M.A.M.; Rashid, A.H.A.; Suhaimi, F. Piper sarmentosum effects on 11-hydroxysteroid dehydrogenase type 1 enzyme in serum and bone in rat model of glucocorticoid-induced osteoporosis. Molecules 2016, 21. [Google Scholar] [CrossRef]
  2. Jeffrey, C.S.; Leonard, M.D.; Glassmire, A.E.; Dodson, C.D.; Richards, L.A.; Kato, M.J.; Lee, A.D. Antiherbivore prenylated benzoic acid derivatives from Piper kelleyi. J. Nat. Prod. 2014, 77, 148–153. [Google Scholar] [CrossRef] [PubMed]
  3. Di Stasi, L.C.; Oliveira, G.P.; Carvalhaes, M.A.; Queiroz, M., Jr.; Tien, O.S.; Kakinami, S.H.; Reis, M.S. Medicinal plants popularly used in the Brazilian tropical atlantic forest. Fitoterapia 2002, 73, 69–91. [Google Scholar] [CrossRef]
  4. Parmar, V.S.; Jain, S.C.; Bisht, K.S.; Jain, R.; Tanela, P.; Jha, A.; Tyagi, O.D.; Prasad, A.D.; Wengel, J.; Olsen, C.E.; et al. Phytochemistry of the Genus Piper. Phytochemistry 1997, 49, 597–673. [Google Scholar] [CrossRef]
  5. Cardozo-Júnior, E.L.; Chaves, M.C.O. Caldensin, A new natural N-methylaristolactam from Piper caldense. Pharm. Biol. 2003, 41, 216–218. [Google Scholar] [CrossRef]
  6. Chaves, M.C.O.; Cunha, E.V.L. Two amides from Piper tuberculatum fruits. Fitoterapia 2001, 72, 197–199. [Google Scholar]
  7. Rukachaisirikul, T.; Siriwattanakit, P.; Sukcharoenphol, K.; Wongvein, C.; Ruttanaweang, P.; Wongwattanavuch, P.; Suksamrarn, A. Chemical constituents and bioactivity of Piper sarmentosum. J. Ethnopharmacol. 2004, 93, 173–176. [Google Scholar] [CrossRef] [PubMed]
  8. Ma, Y.; Han, G.Q.; Li, C.L.; Cheng, J.R.; Arison, B.H.; Hwang, S.B. Neolignans from Piper polysyphorum C. DC. Acta Pharm. Sin. 1991, 26, 345–350. [Google Scholar]
  9. Jenett-Siems, K.; Mockenhaupt, F.P.; Bienzle, U.; Gupta, M.P.; Eich, E. In vitro antiplasmodial activity of Central American medicinal plants. Trop. Med. Int. Health 1999, 4, 611–615. [Google Scholar] [CrossRef] [PubMed]
  10. Alves, H.S.; Souza, M.F.V.; Chaves, M.C.O. Three New Compounds from Piper montealegreanum Yuncker (Piperaceae). J. Braz. Chem. Soc. 2011, 22, 1610–1615. [Google Scholar] [CrossRef]
  11. Yuncker, T.G. New species of Piperaceae from Brazil. Bull. Inst. Bot. 1966, 3, 1–370. [Google Scholar]
  12. Queiroz, A.C.; Alves, H.S.; Silva, L.H.A.C.; Dias, T.L.M.F.; Santos, M.S.; Melo, G.M.A.; Campesatto, E.A.; Chaves, M.C.O.; Moreira, M.S.A. Antinociceptive and anti-inflammatory effects of flavonoids PMT1 and PMT2 isolated from Piper montealegreanum Yuncker (Piperaceae) in mice. Nat. Prod. Res. 2014, 28, 403–406. [Google Scholar] [CrossRef] [PubMed]
  13. Øyvind, M.A.; Kenneth, R.M. Flavonoids: Chemistry, Biochemistry, and Applications, 1st ed.; CRC Press: New York, NY, USA, 2006. [Google Scholar]
  14. Hano, Y.; Aida, M.; Nomura, T.; Ueda, S. A novel way of determining the structure of artonin I, an optically active Diels–Alder type adduct, with the aid of an enzyme system of Morus alba cell cultures. J. Chem. Soc. Chem. Commun. 1992, 17, 1177–1178. [Google Scholar] [CrossRef]
  15. Shibata, K.; Tatsukawa, A.; Umeoka, K.I.; Lee, H.S.; Ochi, M. Crinatusins, bioactive Diels-Alder adducts from Cyathocalyx crinatus. Tetrahedron 2000, 56, 8821–8824. [Google Scholar] [CrossRef]
  16. Yang, D.S.; Li, Z.L.; Yang, Y.P.; Xiao, W.L.; Li, X.L. A New Geranylated 2-Arylbenzofuran from Morus alba. Chin. Herb. Med. 2015, 7, 191–194. [Google Scholar] [CrossRef]
  17. Joshi, A.S.; Li, X.C.; Ninrod, A.C.; Elsohly, H.N.; Walker, L.A.; Clark, A.M. Dihydrochalcones from Piper longicaudatum. Plant Med. 2001, 67, 186–188. [Google Scholar] [CrossRef] [PubMed]
  18. Shirota, O.; Takizawa, K.; Sekita, S.; Satake, M.; Hirayama, Y.; Hakamata, Y.; Hayashi, T.; Yanagawa, T. Antiandrogenic natural Diels-Alder-type adducts from Brosimum rubescens. J. Nat. Prod. 1997, 60, 997–1002. [Google Scholar] [CrossRef] [PubMed]
  19. Alias, Y.; Awang, K.; Hadi, A.H.; Thoison, O.; Sévenet, T.; Païs, M. An antimitotic and cytotoxic chalcone from Fissistigma lanuginosum. J. Nat. Prod. 1995, 58, 1160–1166. [Google Scholar] [CrossRef] [PubMed]
  20. Vogt, T. Phenylpropanoid Biosynthesis. Mol. Plant 2010, 3, 2–20. [Google Scholar] [CrossRef] [PubMed]
  21. Brú, J.; Guzman, J.D. Folk medicine, phytochemistry and pharmacological application of Piper marginatum. Rev. Bras. Farmacogn. 2016, 26, 767–779. [Google Scholar] [CrossRef]
  22. Maciel, J.K.S.; Chaves, O.S.; Filho, S.G.B.; Teles, Y.C.F.; Fernandes, M.G.; Assis, T.S.; Fernandes, P.D.; Andrade, A.P.; Felix, L.P.; Silva, T.M.S.; et al. New alcamide and anti-oxidant activity of Pilosocereus gounellei A. Weber ex K. Schum. Bly. ex Rowl. (Cactaceae). Molecules 2016, 21. [Google Scholar] [CrossRef] [PubMed]
  23. Orjala, J.; Wright, A.D.; Behrends, H.; Folkers, G.; Sticher, O. Cytotoxic and antibacterial dihydrochalcones from Piper aduncum. J. Nat. Prod. 1994, 57, 18–26. [Google Scholar] [CrossRef] [PubMed]
  24. Haensel, R.; Wiess, D.; Schmidt, B. Fungistatische wirkung der kawadroge und ihrer inhaltsstoffe. Plant Med. 1966, 14, 1–9. [Google Scholar] [CrossRef]
  25. Orjala, J.; Wright, A.D.; Erdelmeier, C.A.J.; Sticher, O.; Rali, T. New Monoterpene-Substituted Dihydrochalcones from Piper aduncum. Helv. Chim. Acta 1993, 76, 1481–1488. [Google Scholar] [CrossRef]
  26. Falkiner, M.; Loder, J.W.; Russell, G.B.; Shelton, M. ω-Hydroxyisodillapiole, a new cinnamyl alcohol from Piper novae-hollandiae. Aust. J. Chem. 1972, 25, 2417–2420. [Google Scholar] [CrossRef]
  27. Leão da Silva, M.; Maia, J.G.S.; Mourao, J.C.; Pedreira, G.; Marx, M.C.; Gottlieb, O.R.; Magalhães, M.T. Essential oils from Amazonica VI. Acta Amaz. 1973, 3, 41–42. [Google Scholar]
  28. Galinis, D.L.; Wiemer, D.F.J. Villiramulins A and B: New phenol derivatives from Piper villiramulum. Org. Chem. 1993, 58, 7804–7807. [Google Scholar] [CrossRef]
  29. Bohm, B.A. Introduction to Flavonoids, 1st ed.; Harwood Academic Publishers: Vancouver, UK, 1998. [Google Scholar]
  30. Portet, B.; Fabre, N.; Roumy, V.; Gornitzka, H.; Bourdy, G.; Chevalley, S.; Sauvin, M.; Valentin, A.; Moulis, S. Activity-guided isolation of antiplasmodial dihydrochalcones and flavanones from Piper hostmannianum var. berbicense. Phytochemistry 2007, 68, 1312–1320. [Google Scholar] [CrossRef] [PubMed]
  31. Chaves, O.S.; Teles, Y.C.F.; Monteiro, M.M.O.; Junior, L.G.M.; Agra, M.F.; Braga, V.A.; Silva, T.M.S.; Souza, M.F.V. Alkaloids and Phenolic Compounds from Sida rhombifolia L. (Malvaceae) and Vasorelaxant Activity of Two Indoquinoline Alkaloids. Molecules 2017, 22. [Google Scholar] [CrossRef] [PubMed]
  32. Pavia, D.L.; Lampman, G.M.; Kriz, G.S. Introdução à Espectroscopia, 4th ed.; Cengage Learning: Boston, MA, USA, 2010. [Google Scholar]
  33. Jerz, G.; Arrey, T.N.; Wray, V.; Du, Q. Structural characterization of 132-hydroxy-(132-S)-phaeophytin a from leaves and stems of Amaranthus tricolor isolated by high-speed countercurrent chromatography. Innov. Food Sci. Emerg. Technol. 2007, 8, 413–418. [Google Scholar] [CrossRef]
  34. Silva, T.M.S.; Carvalho, M.G.; Braz-Filho, R. Estudo espectroscópico em elucidação estrutural de flavonoides de Solanum jabrense Agra & Nee e S. paludosum Moric. Quim. Nova 2009, 32, 1119–1128. [Google Scholar]
Sample Availability: Compound samples are no longer available.
Scheme 1. Probable biosynthetic route for monoterpene dihydrochalcones.
Scheme 1. Probable biosynthetic route for monoterpene dihydrochalcones.
Molecules 22 00874 sch001
Figure 1. Compounds isolated from P. montealegreanum.
Figure 1. Compounds isolated from P. montealegreanum.
Molecules 22 00874 g001
Scheme 2. Proposed fragmentation mechanisms (only peaks classified as principals) in the HRESITOFMS of 1.
Scheme 2. Proposed fragmentation mechanisms (only peaks classified as principals) in the HRESITOFMS of 1.
Molecules 22 00874 sch002
Figure 2. The 2D NMR correlations of compound 1: 1H-1H correlation spectroscopy (1H-1H COSY) (bold) and selected heteronuclear multiple bond correlations (HMBC) (arrows).
Figure 2. The 2D NMR correlations of compound 1: 1H-1H correlation spectroscopy (1H-1H COSY) (bold) and selected heteronuclear multiple bond correlations (HMBC) (arrows).
Molecules 22 00874 g002
Figure 3. Structural alternatives preliminarily proposed for compound 1.
Figure 3. Structural alternatives preliminarily proposed for compound 1.
Molecules 22 00874 g003
Scheme 3. Fragments proposed to justify the main peaks observed in the HRESITOFMS of 2.
Scheme 3. Fragments proposed to justify the main peaks observed in the HRESITOFMS of 2.
Molecules 22 00874 sch003
Table 1. 1H (500 MHz) and 13C (125 MHz) NMR for 1, 1a, 2 and 2a, including results obtained by heteronuclear 2D shift-correlated HMQC (1JCH) and HMBC (nJCH, n = 2 and 3), using CDCl3 as solvent. Chemical shifts (δ, ppm) and coupling constants (J, Hz) are shown in parentheses *.
Table 1. 1H (500 MHz) and 13C (125 MHz) NMR for 1, 1a, 2 and 2a, including results obtained by heteronuclear 2D shift-correlated HMQC (1JCH) and HMBC (nJCH, n = 2 and 3), using CDCl3 as solvent. Chemical shifts (δ, ppm) and coupling constants (J, Hz) are shown in parentheses *.
11a22a
C TypeδCδHHMBCδCδCδHHMBCδC
C-1145.56 5, 6, 7144.27145.39 5145.14
CH-2114.486.58, dd (2.5, 1.5)4, 6120.24114.406.56, m6, 7120.30
C-3155.46 2, 5150.63155.52 5150.66
CH-4113.306.48, ddd. (7.5, 2.5, 1.5)2, 6119.43113.356.47, md2, 6119.35
CH-5129.356.94, t (7.5) 128.98129.276.92, dt (7.5, 2.0) 127.82
CH-6119.826.64 dd (7.5, 1.5)2, 4125.08119.796.62, md2, 4, 7124.97
CH-745.193.01, dt (11.0, 5.5)2, 8, 2H-1″44.8445.083.00, dt (11.0, 5.5)2, 644.79
CH-851.064.13, dt (11.0, 5.5)7, 2H-10″50.9351.014.09, m750.87
C-9209.55 8209.21209.40 209.03
C-10 169.05 170.30
C-1′109.34 HO-6′108.95109.28 HO-6′109.00
C-2′166.65 CH3O-2′166.60166.60 CH3O-2′166.65
C-3′107.97 8′, HO-4′108.22107.98 8′, HO-4′108.06
C-4′165.21 8′, HO-4′, 3H-7′165.34165.26 HO-4′, 8′, 3H-7′165.36
C-5′108.85 3H-7′, HO-4′, HO-6′108.88108.89 3H-7′, HO-4′, HO-6′109.00 119.35
C-6′167.23 HO-6′, 3H-7′167.24167.22 HO-6′, 3H-7′167.28
CH2-1″29.792.40, 2.38H-7, H-829.7829.712.42, m 2.37, m 29.79
CH-2″118.735.53, m 118.69119.155.55, m 119.53
C-3″137.41 2H-4″, 2H-10″, 2H-5″137.28137.17 136.36
CH2-4″37.232.02 m2H-5″37.1733.202.23-2.18, m 30.60
CH2-5″26.382.14 m2H-4″26.3332.911.65, m 30.37
CH-6″124.025.12 qt (7.0, 1.5)2H-5″, 3H-8″, 3H-9″123.9375.734.08, m2H-8″, 3H-9″71.17
C-7″131.70 H-6″, 3H-8″, 3H-9″131.63147.33 3H-9″145.14
CH3-8″17.841.62 d (0.5)3H-9″18.00111.194.93, m 4.84, m6″, 3H-9″112.98
CH3-9″25.691.70 d (1.0)3H-8″25.6617.631.73, s2H-8″18.07
CH2-10″36.882.23, m, 2.20 mH-7, H-836.7836.672.23, m 2.18, m 36.76
OCH3-2′67.233.90, s 67.2767.243.90, s 67.33
CH3-7′6.561.89, s 6.526.571.87, s 6.56
CHO-8′192.3610.04, s 192.37192.3510.03, s 192.41
HO-4′-12.46, s 12.44, s
HO-6′-12.88, s 12.89, s
CH3-11 20.86 21.08
* Number of hydrogens bound to carbon atoms deduced by comparative analysis of {1H} and APT-13C-NMR spectra. Chemical shifts and coupling constants (J) obtained from 1D 1H-NMR spectrum. Superimposed 1H signals are described without multiplicity, and chemical shifts deduced by HMQC, HMBC and 1H-1H-COSY spectra. All 1H and 13C chemical shift assignments were also based on homonuclear 1H-1H-COSY and heteronuclear 2D shift-correlated HMQC (1JCH) and HMBC (nJCH, n = 2 and 3) NMR.
Table 2. 1H (500 MHz) and 13C (125 MHz) NMR for 4, including results obtained by heteronuclear 2D shift-correlated HMQC (1JCH), HMBC (nJCH, n = 2 and 3), COSY and NOESY, using CDCl3 as solvent. Chemical shifts (δ, ppm) and coupling constants (J, Hz) are given in parentheses.
Table 2. 1H (500 MHz) and 13C (125 MHz) NMR for 4, including results obtained by heteronuclear 2D shift-correlated HMQC (1JCH), HMBC (nJCH, n = 2 and 3), COSY and NOESY, using CDCl3 as solvent. Chemical shifts (δ, ppm) and coupling constants (J, Hz) are given in parentheses.
HMQCHMBCCOSYNOESY
δCδH2JCH3JCH1H-1H1H-1H
C
1135.80 H-1′H-5, H-2′
3147.47 H-5, H-7
4145.54 H-6, H-2, H-7
CH
2108.856.65 brd (1.5) H-1′
5108.056.70 d (8.0)
6121.086.60 dd (8.0, 1.5) H-2, H-1′
3′130.845.58–5.52 (m) H-1′H-2′H-5′
4′130.995.99 tt (14.5, 1.0) H-2′, H-3′, H-6′
5′130.085.99 tt (14.5, 1.0) H-7′, H-3′ H-3′, H-7′
6′133.105.58–5.52 mH-7′ H-7′
CH2
7100.695.89 s
1′35.622.59 t (7.5)H-2′H-2, H-6, H-3′H-2′
2′34.702.30 dt (7.5, 8.0)H-1′, H-3′H-4′H-1′, H-3′
7′32.592.02 dt (7.0, 7.5)H-6′H-5′, H-8′H-6′, H-8′H-9′
8′29.401.35–1.31 mH-7′ H-7′
9′31.921.29–1.25 mH-10′H-11′ H-7′, H-11′
10′22.681.29–1.25 mH-9′, H-11′ H-11′
CH3
11′14.120.86 t (7.0) H-10′H-9′

Share and Cite

MDPI and ACS Style

Alves, H.D.S.; Rocha, W.R.V.d.; Braz-Filho, R.; Chaves, M.C.d.O. Isolation of Monoterpene Dihydrochalcones from Piper montealegreanum Yuncker (Piperaceae). Molecules 2017, 22, 874. https://doi.org/10.3390/molecules22060874

AMA Style

Alves HDS, Rocha WRVd, Braz-Filho R, Chaves MCdO. Isolation of Monoterpene Dihydrochalcones from Piper montealegreanum Yuncker (Piperaceae). Molecules. 2017; 22(6):874. https://doi.org/10.3390/molecules22060874

Chicago/Turabian Style

Alves, Harley Da Silva, Wilma Raianny Vieira da Rocha, Raimundo Braz-Filho, and Maria Célia de Oliveira Chaves. 2017. "Isolation of Monoterpene Dihydrochalcones from Piper montealegreanum Yuncker (Piperaceae)" Molecules 22, no. 6: 874. https://doi.org/10.3390/molecules22060874

Article Metrics

Back to TopTop