Next Article in Journal
Biological Evaluation of Natural and Synthesized Homovanillic Acid Esters as Inhibitors of Intestinal Fatty Acid Uptake in Differentiated Caco-2 Cells
Next Article in Special Issue
Quantitative Analysis of Terpenic Compounds in Microsamples of Resins by Capillary Liquid Chromatography
Previous Article in Journal
Identification of Cannabis sativa L. (hemp) Retailers by Means of Multivariate Analysis of Cannabinoids
Previous Article in Special Issue
Recent Achievements in Medicinal and Supramolecular Chemistry of Betulinic Acid and Its Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Throughput 1H-Nuclear Magnetic Resonance-Based Screening for the Identification and Quantification of Heartwood Diterpenic Acids in Four Black Pine (Pinus nigra Arn.) Marginal Provenances in Greece

by
Kostas Ioannidis
1,2,*,
Eleni Melliou
2 and
Prokopios Magiatis
2
1
Laboratory of Forest Genetics and Biotechnology, Institute of Mediterranean and Forest Ecosystems, Hellenic Agricultural Organization “Demeter”, Ilissia, 11528 Athens, Greece
2
Department of Pharmacognosy and Natural Products Chemistry, Faculty of Pharmacy, University of Athens, Panepistimiopolis Zografou, 15771 Athens, Greece
*
Author to whom correspondence should be addressed.
Molecules 2019, 24(19), 3603; https://doi.org/10.3390/molecules24193603
Submission received: 30 August 2019 / Revised: 26 September 2019 / Accepted: 1 October 2019 / Published: 7 October 2019
(This article belongs to the Special Issue Terpenes and Terpene Derivatives)

Abstract

:
A high-throughput quantitative Nuclear Magnetic Resonance 1H-NMR method was developed and applied to screen the quantity of the diterpenic resin acids in the heartwood of black pine, due to the renewed scientific interest in their medicinal properties and use in various diseases treatment. The 260 samples were taken from Pinus nigra clones, selected from four provenances of the Peloponnese (Greece), participating in a 35-year-old clonal seed orchard. Total resin acids per dry heartwood weight (dhw) varied greatly, ranging from 30.05 to 424.70 mg/gdhw (average 219.98 mg/gdhw). Abietic was the predominant acid (76.77 mg/gdhw), followed by palustric acid (47.94 mg/gdhw), neoabietic acid (39.34 mg/gdhw), and pimaric acid (22.54 mg/gdhw). Dehydroabietic acid was at moderate levels (11.69 mg/gdhw), while levopimaric, isopimaric, and sandaracopimaric acids were in lower concentrations. The resin acid fraction accounted for 72.33% of the total acetone extractives. Stilbenes were presented in significant quantities (19.70%). The resin acid content was composed mainly of the abietane type resin acids (83.56%). Peloponnesian Pinus nigra heartwood was found to be the richest source of resin acids identified to date and is considered the best natural source for the production of such bioactive extracts. The results indicate a high potential for effective selection and advanced breeding of pharmaceutical and high economic value bioactive substances from Pinus nigra clones.

1. Introduction

In Greece, as well as in other European countries, black pine constitutes extensive natural forests. Due to its advantages, it is considered one of the most important silvicultural coniferous species and it is extensively used in reforestation programs throughout the country [1]. Its significance lead to its planting outside its natural range, e.g., North and South America, Australia, and New Zealand. Furthermore, its importance is verified by its potential use for the production of high added value products, i.e., bioactive compounds produced from wood and wood waste materials [2,3]. Such bioactive substances are, inter alia, the resin acids occurring in pine’s oleoresin.
Oleoresin consists of a non-volatile fraction, also referred as colophony or rosin, dissolved in a volatile monoterpenic part, the turpentine. Oleoresin, in coniferous species, is synthesized by the epithelial cells surrounding the resin canals, from which it is exuded as a response to mechanical wounds or biotic attacks, thus constituting a defense mechanism of living trees [4]. Resin acids, among other substances, are part of the non-volatile fraction, composed of a mixture of diterpenic resinous compounds. Diterpenic resin acids are the most common members of oleoresin [4] and the most abundant compounds in heartwood and knots [5,6,7,8]. Their fractions can vary among individual trees and species [9,10,11]. Many softwoods, and particularly Pinus species, contain relatively high oleoresin amounts, moving through the wood’s extensive vertical and radial networks of resin canals.
Resin acids are mostly tricyclic compounds, arranged in two types depending on the presence or absence of double bonds in their aromatic rings. The first, called abietane-type, includes acids such as abietic acid (1), dehydroabietic acid (2), neoabietic acid (3), palustric acid (4) and levopimaric acid (5), and the second, named pimarane-type, includes acids such as pimaric acid (6), sandaracopimaric acid (7), and isopimaric acid (8) (Figure 1). The cyclic diterpene acids originate from a common acyclic biosynthetic precursor, geranylgeranyl diphosphate (GGPP) [12]. Diterpene synthase enzymes act on the 20-carbon GGPP substrate to form diterpenes, which are subsequently hydroxylated and then oxidized, in two independent transformations, by other enzymes to the diterpene acids [13].
The diterpene synthase genes involved in resin acid formation prove their highly genetically-correlated heartwood properties [14,15]. Selection and breeding in Pinus species could result in doubling the oleoresin yields [16] and further genetic gains in extractive yield could be achieved [17]. Genetic control of extractive production has considerable economic potential in some species [18].
Pine resin has been highly applied in industry. Resin acids and their derivatives, apart from being used in intermediate chemical products, such as polymer additives, tackifiers, emulsifiers in synthetic rubber, paper sizing to control water absorptivity, and aroma industry, are also deployed in adhesives, surface coatings, printing inks, chewing gums, and in the Greek wine industry [19,20]. Furthermore, several medicinal uses concerning the treatment of abscesses, boils, cancers, toothache, and skin diseases, such as psoriasis and ringworm [19], have been reported even from antiquity. According to Dioscorides, during the first century A.D., a wine called retsina, i.e., wine that had been mixed with resin from Pinus halepensis, was used for lung and stomach disorders and as an antidote for cough [19].
In recent years, many studies have shown that several substances in rosin have antimicrobial properties [21], cardiovascular effects [22,23], antiallergic properties [24], anti-inflammatory properties [25,26], anticonvulsant effects [27], antiulcer-gastroprotective properties [28,29], and cytotoxic activity toward human fibroblasts [29]. Their medicinal properties, due to the bioactivity of natural abietane and pimarane diterpene acids, which have been reviewed by San Feliciano et al. [30] and Reveglia et al. [31] respectively, have caused a renewed scientific interest in pine extracts.
The aim of the present work was to analyze the resin acids composition and variability from the heartwood extractives of four Pinus nigra L. subsp. pallasiana marginal provenances from the Peloponnese, southern Greece, growing in a clonal seed orchard. The heartwood of these trees produced an exceptionally high extractive content, approximately 30% w/w [3], making it particularly interesting for the present analysis. The target compounds were the abietane-type resin acids: abietic, dehydroabietic, neoabietic, palustric, and levopimaric acids, and the pimarane-type: pimaric, isopimaric, and sandaracopimaric acids. The 1H-NMR spectroscopy was used to analyze, qualitatively and quantitatively, the highly complex diterpenic acid mixtures of heartwood rosin.
Many studies have been carried out on pine oleoresin composition, some of which have demonstrated the potential of NMR spectroscopy, a modern analytical methodology with applications in complex mixtures [32,33]. The applied technique highlights its feasibility and wide practical applicability in characterizing various complex matrixes, from leaves [34] and walnuts [35] for profiling “green” extracts and different genotypes, respectively, to foodstuffs [36,37] for quality/safety or discrimination purposes, and biological [38] for lipid profiling. 13C-NMR spectroscopy has already been used for the quantitation of resin acids in wood extracts from Pinus radiata [39] and Pinus nigra ssp. laricio [20,40]. Skakovskii et al. [41], through the use of high-field NMR spectrometers introduced the alternative 1H-NMR analysis of resin acids, and this technique is here preferred as quicker and straightforward. Moreover, the possible simultaneous record of 1H-NMR and 13C-NMR experiments can pave the way to the robust analysis of resin acid by combined data as performed for other systems [42].

2. Results

1H-NMR spectrum of black pine (P. nigra Arn.) heartwood extracts, showing the characteristic peaks of the studied resin acids and internal standard, are presented in Figure 2. The whole spectrum is available as supplementary Figure S1. Their chemical shifts are presented in Table 1. Recovery and intraday precision data are presented in Table 2 and Table 3, respectively.
In the frame of this investigation 260 samples were analyzed. The heartwood of the studied trees was exceptionally rich in acetone extractive content (TAE), averaging 304.15 mg/g of dry heartwood (dhw), which contained high resin acid quantities. The mean resin acid fraction accounted for 72.33% of the total extractive content, while stilbenes were present in significant quantities, comprising 19.70%. The remaining 7.97% is referred to as other substances present in the TAE, i.e., minor unidentified resin acids, fatty acids, unsaponifiables, triglycerides, other phenols, waxes, sterols, etc.
The mean concentration of all studied diterpenic resin acids as their min and max values are presented in Table 4. The total resin acid (TRA) content showed an extent variation ranging from 30.05 to 424.70 mg/gdhw, with an average of 219.98 mg/gdhw (±96.20). Abietic acid was the predominant acid (76.77 ± 37.39 mg/gdhw), followed by palustric acid (47.94 ± 23.31 mg/gdhw) and neoabietic acid (39.34 ± 21.21 mg/gdhw), all belonging to the abietane-type resin acids. The next in abundance acid was the pimaric acid, with a mean concentration of 22.54 ± 11.28 mg/gdhw. Dehydroabietic acid was found at moderate levels (11.69 ± 5.73 mg/gdhw), while the rest, levopimaric acid, isopimaric acid, and sandaracopimaric acid, were observed in lower concentrations, i.e., 8.07 ± 211.38 mg/gdhw, 10.91 ± 6.53 mg/gdhw, and 2.72 ± 1.49 mg/gdhw, respectively.
In Table 5, the mean resin acid concentrations, the results from the analysis of variance (ANOVA), as well as Duncan’s Multiple Range Test (MRT) results, of the four Pinus nigra L. provenances are presented. The ANOVA and MRT showed that there were statistically significant (p < 0.001) differences among clones for all examined acids (data not shown) and also among provenances in concentrations of both abietic acid (p < 0.01) and pimaric acid (p < 0.01).
In Figure 3, the mean percentage of each individual resin acid to total resins, as well as their types, are presented. Abietic, palustric, and neoabietic acids, which were the most abundant resin acids, accounted for 34.90%, 21.79%, and 17.88%, respectively. Dehydroabietic and levopimaric acids comprised 5.32% and 3.67% of the TRA, respectively. Pimaric acid was in a fairly large quantity, corresponding to 10.25% of the total studied resin acids, while sandaracopimaric and isopimaric acids corresponded to 1.24% and 4.96%, respectively, of the total studied resin acids. The resin acid content was composed mainly of the abietane resin acid type (≈84%), while the pimarane-type accounted only for the rest, ≈16%, so the abietane- to pimarane-type resin acids ratio was circa 5.2. In Figure 4, the mean percentage content of the constituents of black pine’s heartwood acetone extraction from the Peloponnese is presented.
In Figure 5, the frequency distribution of the total resin acids (TRA) of black pine from the Peloponnese are presented.

3. Discussion

The current study aimed to demonstrate that 1H-NMR spectroscopy can be effectively used for the analysis of Pinus nigra heartwood resin acids although their spectra are complicated and partially overlapping. Despite the structural similarity of the terpenic acids included in the oleoresin, we were able to select non overlapping peaks, belonging to specific protons, that could lead to the identification and quantitation of the target resin acids. Using 1H-NMR spectroscopy, we achieved to minimize time-consuming procedures of separation of every analyte before quantitation.
As it was previously shown in the cases of olive oil [43,44], beer [45], wine [33], and stilbenes from Pinus nigra heartwood [3], qNMR is the most appropriate method for quantitation. The developed methodology for the analysis of heartwood resin acids requires less than a minute to record the spectrum and perform the analytes quantitation, whereas using the previously reported 13C-NMR methodology requires a considerably longer time, i.e., approximately two hours [40] or even more [46]. The advantage of the current methodology allowed the analysis of a very large amount of samples in a short period and led to reliable quantitative data.
The results indicated obvious individual tree and clone variation in heartwood resin acids content due to genetic and environmental factors affecting rosin production and secretion [4,47]. Venäläinen et al. [48] also reported high variation among trees in the resin acid concentrations of Pinus sylvestris. Heartwood resin acid contents present provenance variation as well (Table 5) because of the heterogeneity among populations [3]. In general, the southeastern provenance of Parnonas, grown at the most xerothermic environment [49] and subjected to the strongest winds [50], showed higher values of total resin acids, i.e., for neoabietic acid, palustric acid, and sandaracopimaric acid, while the southwestern of Taigetos had the lowest concentrations of all except for dehydroabietic and levopimaric acids. For the latter, Taigetos provenance showed the highest content of all provenances. Although Feneos provenance (northeastern origin) predominated in dehydroabietic acid and pimaric acid concentrations, only the latter acid statistically differed from Taigetos origin. The origin of Zarouhla presented the highest value for abietic and isopimaric acids, having a statistically significant difference only for the abietic acid found in Taigetos. Willför et al. [51] found similar results, mentioning that there were no significant differences between trees from the southern and northern regions, thus confirming Song’s [52] view that the oleoresin of trees within several Pinus species show similar basic chemical characteristics, regardless of their geographic location.
The present study stated that acetone extractives of Pinus nigra heartwood mainly consisted of resin acids. These results were in accordance with those of Martínez-Inigo et al. [53], Willför et al. [51], and Belt et al. [54] for Pinus sylvestris. Hafizoğlu [55], in a review article, mentioned that several researchers have found that most of the oleoresin in Pinus nigra and P. sylvestris were comprised of resin acids. In the same direction were the results of Uprichard and Lloyd [56] in P. radiata, reporting that resin acids predominate in heartwood total extractives. Yildirim and Holmbom [57] have found a remotely lower percentage in stemwood extractives of Pinus brutia.
Lower resin acid concentrations compared to the present study were detected in Pinus sylvestris by Martínez-Inigo et al. [53], Hovelstad et al. [58], and Belt et al. [54]. Concerning Pinus sylvestris heartwood, Harju et al. [59], Venäläinen et al. [48], Willför et al. [51], and Arshadi et al. [60] estimated lower resin acid content, too. Moreover, in Pinus nigra, Yildirim and Holmbom [61] found similar results with the previous researchers. Correspondingly, medium resin acid concentrations were found in P. radiata by Lloyd [7]. On the contrary, Willför et al. [62] found in both Pinus nigra and P. brutia heartwood even lower amounts of resin acids than the present study, without referring to predominant acids or the fatty/resin and abietane/pimarane ratios. Low levels of resin acid were noted in Pinus contorta by Lewinsohn et al. [63] and Arshadi et al. [60]. Benouadah et al. [64] surprisingly found even lower concentrations in dry Pinus halepensis heartwood.
The fatty acid/resin acid ratio in the present study was estimated as very low due to the high presence of resin acids, as well as stilbenes, in total acetone extractives and could not exceed 0.18. Lower fatty/resin acid ratio in Pinus sylvestris was reported by Martínez-Inigo et al. [53] and Belt et al. [54], depending on the heartwood sample position, and in P. radiata by Uprichard and Lloyd [56]. Arshadi et al. [60] reported that in Pinus sylvestris and Pinus contorta, in general, the concentration of resin acids was higher than that of fatty acids. In Pinus contorta and Pinus attenuata, the results of Anderson et al. [65] confirmed the above findings as well as those of the present study. In contrast, Hafizoğlu’s [55] review article reported that fatty/resin acid ratio was a multiple of the unit concerning P. nigra and P. sylvestris, and that only in P. brutia the resin acids were double the fatty ones. The results of Uçar and Fengel [66], as well as those of Uçar and Balaban [67], for Pinus nigra were in agreement with Hafizoğlu’s. In both studies, the fatty/resin ratio depended on the Pinus nigra variety or provenance.
In the present study, the abietane-type dominated over the pimarane-type resin acids and the estimated (abietane/pimarane) ratio was calculated at 5.1. This is consistent with the results of Harju et al. [68] and Willför et al. [51] in Pinus sylvestris, who estimated a little higher ratio values, as well as with Uçar and Fengel [66], who estimated a little lower ratio values, depending on variety. The common element in all studies was that the resin acids of abietane-type were multiples compared to the pimarane ones.
Abietic acid was the most abundant resin acid, which is in accordance with the findings of Martínez-Inigo et al. [53], Willför et al. [51,62], and Yildirim and Holmbom [61] in Pinus sylvestris, of Yildirim and Holmbom [57] in Pinus brutia, and of Anderson et al. [65] in Pinus attenuata. Hafizoğlu’s review article [55] reported that, in most cases, abietic, mainly in Pinus nigra, and levopimaric, mainly in Pinus sylvestris, acids were the most abundant, followed by palustric acid though dependent on species and origins. Abietic acid was found to be the most abundant resin acid in Scots pine according to Harju et al. [68], but followed by the mixture of palustric/levopimaric acids and neoabietic or pimaric acids. Abietic acid was the most abundant resin acid in P. radiata too [56]. Different results were those of Rezzi et al. [20] who distinguished two clusters of Pinus nigra ssp. laricio. In the first one, levopimaric acid was the most abundant resin, whereas in the second, dehydroabietic acid and levopimaric acid were approximately in equal concentrations. Cannac et al. [46] identified Pinus nigra ssp. laricio, from Corsica too, as belonging to the first cluster.
In the examined Pinus nigra provenances, abietic acid was the most abundant one, followed by palustric acid, neoabietic acid, and pimaric acid. Similar results, but with a few differences in the classification of resin acid concentrations, were also those of Ekeberg et al. [69] in Scots pine and Benouadah et al. [64] in Aleppo pine. In both studies, abietic acid was the major acid though dehydroabietic acid was ranked second or third. Papajannopoulos et al. [70] reported that the basic composition of the oleoresin of three Greek pine species, Pinus halepensis, P. brutia, and P. pinea, were found to be similar concerning the abundance of several resin acids. The mixture of palustric/levopimaric acids were the most abundant followed by the abietic, in similar concentrations, and neoabietic or isopimaric acids. The above were in accordance with the results of Lange and Stefanovic Janezic [71] for Pinus sylvestris and Lloyd [7] for Pinus radiata.
Different were the results of Uçar and Fengel [66] who found palustric acid to be the richest resin acid in the stemwood of Pinus nigra varieties, followed by isopimaric acid or dehydroabietic acid (depending on the variety), while abietic acid was ranked third. Palustric acid was found to be the most abundant Pinus nigra resin acid by Uçar and Balaban [67] followed by abietic acid, neoabietic acid, or levopimaric acid, depending on the individual tree.
The results of Lewinsohn et al. [63] obtained in the case of Pinus contorta are in contrast with our results. They found that the total resin acids consisted mainly of levopimaric acid, followed by palustric, isopimaric, abietic, and neoabietic acids at moderate levels. The above composition resembled those of mature Pinus contorta rosin stated by Anderson et al. [65]. In Pinus attenuata, Anderson et al. [65] found that dehydroabietic acid was the major resin acid followed by the mixture of levopimaric/palustric acids, and finally by neoabietic acid.
The present study sets a framework for future research on a number of issues, which could lead to new scientific results. Exploiting the individual clone and provenance variation of the genetic material of Pinus nigra from the Peloponnese, its heartwood has been found to be the richest source in resin acids, as well as in stilbenes [3], identified to date. Effective selection and advanced breeding require further research towards the estimation of several genetic parameters (i.e., heritability, breeding values, and genetic gains for all the above mentioned traits) of the studied provenances and clones.

4. Materials and Methods

4.1. Plant Material

The plant material was sampled from a 10 hectares Pinus nigra Arnold clonal seed orchard (CSO), established in the western part of the Peloponnese in 1978. The CSO comprises 52 clones and a total number of 2700 grafts, derived from intensively selected plus trees, originating from four marginal provenances (Zarouhla, Feneos, Parnonas, and Taigetos) of the natural black pine forest of the Peloponnese (Figure 6). Clones (one ramet/clone) were randomly assigned at 6 m × 6 m spacing within replications (single tree plot design), without blocking, with the only restriction that no grafts of the same clone were planted closer than 30 m [72].

4.2. Sampling

Sampling, coring, heartwood discrimination and orientation, and extraction protocol are extensively described in Ioannidis et al. [3]. In brief, increment cores, approximately 30 cm above ground and in a north-south orientation, of 12 mm in diameter were extracted from a total of 260 healthy individuals that were sampled during October and November 2013 and covered all 52 clones participating in the seed orchard (five ramets per clone), and stored in darkness at −76 °C. Heartwood was separated from the rest of the core, using the benzidine discrimination method, and milled to produce ≤0.75 mm particles, which were freeze-dried for 48 h at −52 °C and 0.03 mbar pressure, to ensure almost complete removal of moisture and volatile compounds.

4.3. Extraction Protocol

In brief, resin acids were extracted with 12 ml acetone from 200 mg (±0.1 mg) of freeze-dried, ground heartwood. The mixture was first placed in an orbital shaker at 350 rpm (Edmund Bühler GmbH, Bodelshausen, Germany) in darkness at room temperature for 24 h, followed by its transfer to an ultrasonic bath (Semat, UK) for 1 h to complete the extraction. The liquid phase was separated by centrifuging (3075× g for 15 min) (Eppendorf 5810R, Germany) and the solvent was evaporated in a heated vacuum rotary evaporator (Buchi, Switzerland) at 40 °C to determine the weights. Recovery of the first extraction was calculated by employing successive extraction procedures and quantitation in each one.

4.3.1. H-NMR Spectral Analysis/Quantitation Methodology

The dried extractives were submitted to chemical analysis by 1H-NMR using syringaldehyde as internal standard and deuterated chloroform (CDCl3) as solvent. The extract of each sample obtained from the extraction procedure was dissolved in 600 μL deuterated chloroform (CDCl3) (Euriso-Top) and the solution was transferred to a 5 mm NMR tube. 1H-NMR spectra were recorded at 400 MHz (Bruker DRX400). Typically, 16 scans were collected into 32K data points over a spectral width of 0–16 ppm with a relaxation delay of 10 s. Prior to Fourier transformation (FT), an exponential weighting factor corresponding to a line broadening of 0.3 Hz was applied. The spectra were phase corrected automatically using TopSpin software (Bruker, Billerica, MA, USA). For the peaks of interest, accurate integration was performed manually. Quantitative determination of the resin acids (RA), i.e., the concentration CRA, was obtained by comparing the area of the selected signals with that of the internal standard (IS), i.e., the relative integration (IRA). The following formula was used, which include IRA as well as the weight (msample = 200 mg), the molecular weight (MWRA), and the average recovery of each resin acid (RRA, Table 2) of the analyzed dry heartwood powder:
CRA = [(0.002745 × IRA × MWRA)/msample]/RRA
The recovery was calculated as the ratio of successive extractions of heartwood samples for the studied resin acids. The samples contained different levels of extractives in order to obtain unbiased results. The intraday precision (i.e., the experimental error) was determined by analyzing three replicates of three random samples in the same day. Concentrations were based on freeze-dried heartwood (dhw) and expressed in mg/gdhw. Syringaldehyde (Acros Organics) internal standard (IS) solution was prepared in acetonitrile at a concentration of 0.5 mg/mL. Prior to use, the IS solution reached room temperature and the solutions of all the extracted samples were mixed, before evaporation, with 1 mL of a syringaldehyde solution, i.e., 0.00274 moles of syringaldehyde.
The identity of all compounds was defined by the literature data [41,73,74,75]. Due to rather complicated resin acids’ 1H-NMR spectra, nonoverlapping, undoubtedly defined peaks of the olefinic or aromatic protons were selected for quantitation. Concerning the abietane-type resin acids, abietic acid (1), neoabietic acid (2), dehydroabietic acid (3), palustric acid (4), and levopimaric acid (5) were distinguished by their single C(H)-14 proton at 5.77 ppm, 6.20 ppm, 6.88 ppm, 5.39 ppm, and 5.53 ppm, respectively. For the pimarane-type resin acids, pimaric acid (6) and isopimaric acid (8) were distinguished by the C(H)-15 proton at 5.67 ppm and 5.84 ppm, respectively, and sandaracopimaric acid (7) was distinguished by the C(H)-14 proton at 5.22 ppm. It should be noted that for pimaric and isopimaric acids the selected signal for quantitation was only one of the four equivalent peaks of the corresponding doublet to avoid overlapping with other peaks. In these two cases, the integration value was multiplied by a factor of four before comparing the area with that of the internal standard in order to estimate the concentration.
To exclude the possibility of overlapping peaks we have performed 2D HMQC (2 Dimensional Heteronuclear Multiple Quantum Correlation) experiment presented in Supplementary Material Figures S2 and S3.

4.3.2. Statistical Analysis

Descriptive statistics, analysis of variance (ANOVA), as well as Duncan’s Multiple Range Test (MRT), based on the 0.05 level of significance, were performed in order to check the hypothesis that there were statistically significant differences among the mean concentrations in the samples of all clones and among the four provenances, using SPSS v.20 software for Windows (IBM SPSS Statistics 2011, IBM Corp. New York, NY, USA).

5. Conclusions

The 1H-NMR spectroscopy proved to be a fast, sensitive, accurate, and comfortable methodology for high-throughput identification and quantitation of resin acids. According to this study, the genetic material of black pine heartwood, originating from the Peloponnese, was proved to be the richest natural source of resin acids compared to any other European pine species/population referred to in the literature. The high resin acid concentrations, as well as differences among origins for some of them, can be attributed to their adaptation over time to individual environments that differ in bioclimatic and soil parameters. Such marginal populations growing in the edge of their natural range, under the pressure and impact of climate change, will be populations for further investigations and adaptability testing. Furthermore, the presence of great variation in resin acid concentrations among trees and origins from the Peloponnese, as well as the consensus of genetic control of resin acid production-segregation, highlight the great potential for effective selection and advanced breeding of the studied genetic material for pharmaceutical and high economic value bioactive substances derived from Pinus nigra L. heartwood.

Supplementary Materials

The supplementary materials are available online.

Author Contributions

Conceptualization, P.M., E.M., and K.I.; methodology, P.M., E.M., and K.I.; formal analysis, K.I.; investigation, K.I.; resources, P.M., E.M., and K.I.; data curation, K.I.; writing—original draft preparation, K.I.; writing—review and editing, K.I., P.M., and E.M..; supervision, P.M.; funding acquisition, P.M.

Funding

This research received no external funding.

Acknowledgments

We would like to express our appreciation to the Laboratory of Forest Genetics and Tree Improvement, Faculty of Forestry and Natural Environment, School of Agriculture, Forestry and Natural Environment, Aristotle University of Thessaloniki, Greece, for providing the increment borer in order to extract increment cores. The authors want to express their sincere thankfulness to the librarian of the Forest Research Institute of Athens, D. Panayiotopoulou (MSc Library and Information Science), for information seeking and retrieving processes as well as her additional proof reading service.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Matziris, D. Variation in growth and branching characters in black pine (Pinus nigra Arnold) of Peloponnesos. Silvae Genet. 1989, 38, 77–81. [Google Scholar]
  2. Pietarinen, S.P.; Willför, S.M.; Ahotupa, M.O.; Hemming, J.E.; Holmbom, B.R. Knotwood and bark extracts: Strong antioxidants from waste materials. J. Wood Sci. 2006, 52, 436–444. [Google Scholar] [CrossRef]
  3. Ioannidis, K.; Melliou, E.; Alizoti, P.; Magiatis, P. Identification of black pine (Pinus nigra Arn.) heartwood as a rich source of bioactive stilbenes by qNMR. J. Sci. Food Agric. 2017, 97, 1708–1716. [Google Scholar] [CrossRef] [PubMed]
  4. Hillis, W.E. Heartwood and Tree Exudates; Springer Verlag: Berlin, Germany, 1987; p. 268. [Google Scholar]
  5. Porter, L.J. The resin and fatty acid content of living Pinus radiata wood. N. Z. J. For. Sci. 1969, 12, 687–693. [Google Scholar]
  6. Hemingway, R.W.; Hillis, W. Changes in fats and resins of Pinus radiata associated with heartwood formation. Appita 1971, 24, 439–443. [Google Scholar]
  7. Lloyd, J.A. Distribution of extractives in Pinus radiata earlywood and latewood. N. Z. J. For. Sci. 1978, 8, 288–294. [Google Scholar]
  8. Nisula, L. Wood Extractives in Conifers: A Study of Stemwood and Knots of Industrially Important Species; Åbo Akademi University Press: Åbo, Finland, 2018; p. 372. [Google Scholar]
  9. Joye, N.M.; Lawrence, R.V. Resin acid composition of pine oleoresins. J. Chem. Eng. Data 1967, 12, 279–282. [Google Scholar] [CrossRef]
  10. Conner, A.H.; Diehl, M.A.; Rowe, J.W. Tall oil precursors in three western pines: Ponderosa, Lodgepole, and Limber pine. Wood Sci. 1980, 12, 183–191. [Google Scholar]
  11. Conner, A.; Diehl, M.; Rowe, J. Tall oil precursors and turpentine in jack and eastern white pine. Wood Sci. 1980, 12, 194–200. [Google Scholar]
  12. Lichtenthaler, H.K. The 1-deoxy-d-xylulose-5-phosphate pathway of isoprenoid biosynthesis in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1999, 50, 47–65. [Google Scholar] [CrossRef]
  13. Keeling, C.I.; Bohlmann, J. Diterpene resin acids in conifers. Phytochemistry 2006, 67, 2415–2423. [Google Scholar] [CrossRef] [PubMed]
  14. Fries, A.; Ericsson, T.; Gref, R. High heritability of wood extractives in Pinus sylvestris progeny tests. Can. J. For. Res. 2000, 30, 1707–1713. [Google Scholar] [CrossRef]
  15. Ericsson, T.; Fries, A.; Gref, R. Genetic correlations of heartwood extractives in Pinus sylvestris progeny tests. For. Genet. 2001, 8, 73–79. [Google Scholar]
  16. Squillace, A.; Harrington, T. Olustee’s high-yielder produces 487 bbls pine gum per crop for four straight years. Nav. Stores Rev. 1968, 77, 4–5. [Google Scholar]
  17. Franklin, E.; Taras, M.; Volkman, D. Genetic gains in yields of oleoresin, wood extractives and tall oil. Tappi 1970, 53, 2302–2304. [Google Scholar]
  18. Taylor, A.M.; Gartner, B.L.; Morre, J.I. Heartwood formation and natural durability—A review. Wood Fiber Sci. 2002, 34, 587–611. [Google Scholar]
  19. Langenheim, J.H. Plant Resins: Chemistry, Evolution, Ecology, and Ethnobotany; Timber Press: Portland, OR, USA, 2003; p. 612. ISBN 0-88192-574-8. [Google Scholar]
  20. Rezzi, S.; Bighelli, A.; Castola, V.; Casanova, J. Composition and chemical variability of the oleoresin of Pinus nigra ssp. laricio from Corsica. Ind. Crops Prod. 2005, 21, 71–79. [Google Scholar] [CrossRef]
  21. Spessard, G.O.; Matthews, D.R.; Nelson, M.D.; Rajtora, T.C.; Fossum, M.J.; Giannini, J.L. Phytoalexin-like activity of abietic acid and its derivatives. J. Agric. Food Chem. 1995, 43, 1690–1694. [Google Scholar] [CrossRef]
  22. De Oliveira, A.M.; Tirapelli, C.R.; Ambrosio, S.R.; da Costa, F.B. Diterpenes: A therapeutic promise for cardiovascular diseases. Recent Patents Cardiovasc. Drug Discov. 2008, 3, 1–8. [Google Scholar] [CrossRef]
  23. Gonzalez, M.A.; Correa-Royero, J.; Agudelo, L.; Mesa, A.; Betancur-Galvis, L. Synthesis and biological evaluation of abietic acid derivatives. Eur. J. Med. Chem. 2009, 44, 2468–2472. [Google Scholar] [CrossRef]
  24. Ulusu, N.N.; Ercil, D.; Sakar, M.K.; Tezcan, E.F. Abietic acid inhibits lipoxygenase activity. Phytother. Res. 2002, 16, 88–90. [Google Scholar] [CrossRef] [PubMed]
  25. Fernandez, M.; Tornos, M.; Garcia, M.; De las Heras, B.; Villar, A.; Saenz, M. Anti-inflammatory activity of abietic acid, a diterpene isolated from Pimenta racemosa var. grissea. J. Pharm. Pharmacol. 2001, 53, 867–872. [Google Scholar] [CrossRef] [PubMed]
  26. Takahashi, N.; Kawada, T.; Goto, T.; Kim, C.-S.; Taimatsu, A.; Egawa, K.; Yamamoto, T.; Jisaka, M.; Nishimura, K.; Yokota, K. Abietic acid activates peroxisome proliferator-activated receptor-γ (PPARγ) in RAW264. 7 macrophages and 3T3-L1 adipocytes to regulate gene expression involved in inflammation and lipid metabolism. FEBS Lett. 2003, 550, 190–194. [Google Scholar] [CrossRef]
  27. Talevi, A.; Cravero, M.S.; Castro, E.A.; Bruno-Blanch, L.E. Discovery of anticonvulsant activity of abietic acid through application of linear discriminant analysis. Bioorg. Med. Chem. Lett. 2007, 17, 1684–1690. [Google Scholar] [CrossRef] [PubMed]
  28. Schmeda-Hirschmann, G.; Astudillo, L.; Rodríguez, J.; Theoduloz, C.; Yáñez, T. Gastroprotective effect of the Mapuche crude drug Araucaria araucana resin and its main constituents. J. Ethnopharmacol. 2005, 101, 271–276. [Google Scholar] [CrossRef]
  29. Schmeda-Hirschmanna, G.; Astudillo, L.; Sepúlveda, B.; Rodríguez, J.A.; Theoduloz, C.; Yáñez, T.; Palenzuela, J.A. Gastroprotective effect and cytotoxicity of natural and semisynthetic labdane diterpenes from Araucaria araucana resin. Z. Naturforschung C 2005, 60, 511–522. [Google Scholar] [CrossRef]
  30. San Feliciano, A.; Gordaliza, M.; Salinero, M.A.; del Corral, J.M.M. Abietane acids: Sources, biological activities, and therapeutic uses. Planta Med. 1993, 59, 485–490. [Google Scholar] [CrossRef] [PubMed]
  31. Reveglia, P.; Cimmino, A.; Masi, M.; Nocera, P.; Berova, N.; Ellestad, G.; Evidente, A. Pimarane diterpenes: Natural source, stereochemical configuration, and biological activity. Chirality 2018, 30, 1115–1134. [Google Scholar] [CrossRef]
  32. Simmler, C.; Napolitano, J.G.; McAlpine, J.B.; Chen, S.-N.; Pauli, G.F. Universal quantitative NMR analysis of complex natural samples. Curr. Opin. Biotechnol. 2014, 25, 51–59. [Google Scholar] [CrossRef]
  33. Nikolantonaki, M.; Magiatis, P.; Waterhouse, A.L. Direct analysis of free and sulfite-bound carbonyl compounds in wine by two-dimensional quantitative proton and carbon nuclear magnetic resonance spectroscopy. Anal. Chem. 2015, 87, 10799–10806. [Google Scholar] [CrossRef]
  34. Cerulli, A.; Masullo, M.; Montoro, P.; Hošek, J.; Pizza, C.; Piacente, S. Metabolite profiling of “green” extracts of Corylus avellana leaves by 1H NMR spectroscopy and multivariate statistical analysis. J. Pharm. Biomed. Anal. 2018, 160, 168–178. [Google Scholar] [CrossRef] [PubMed]
  35. Popescu, R.; Ionete, R.E.; Botoran, O.R.; Costinel, D.; Bucura, F.; Geana, E.I.; Alabedallat, Y.F.J.; Botu, M. 1H-NMR profiling and carbon isotope discrimination as tools for the comparative assessment of walnut (Juglans regia L.) cultivars with various geographical and genetic origins—A preliminary study. Molecules 2019, 24, 1378. [Google Scholar] [CrossRef] [PubMed]
  36. Merkx, D.W.H.; Hong, G.T.S.; Ermacora, A.; van Duynhoven, J.P.M. Rapid quantitative profiling of lipid oxidation products in a food emulsion by 1H NMR. Anal. Chem. 2018, 90, 4863–4870. [Google Scholar] [CrossRef] [PubMed]
  37. Fan, S.; Zhong, Q.; Fauhl-Hassek, C.; Pfister, M.K.-H.; Horn, B.; Huang, Z. Classification of Chinese wine varieties using 1H NMR spectroscopy combined with multivariate statistical analysis. Food Control 2018, 88, 113–122. [Google Scholar] [CrossRef]
  38. Barrilero, R.; Gil, M.; Amigó, N.; Dias, C.B.; Wood, L.G.; Garg, M.L.; Ribalta, J.; Heras, M.; Vinaixa, M.; Correig, X. LipSpin: A new bioinformatics tool for quantitative 1H NMR lipid profiling. Anal. Chem. 2018, 90, 2031–2040. [Google Scholar] [CrossRef] [PubMed]
  39. Suckling, I.D.; Ede, R. A quantitative 13C nuclear magnetic resonance method for the analysis of wood extractives and pitch samples. Appita J. 1990, 43, 77–80. [Google Scholar]
  40. Rezzi, S.; Bighelli, A.; Castola, V.; Casanova, J. Direct identification and quantitative determination of acidic and neutral diterpenes using 13C-NMR spectroscopy. application to the analysis of oleoresin of Pinus nigra. Appl. Spectrosc. 2002, 56, 312–317. [Google Scholar] [CrossRef]
  41. Skakovskii, E.D.; Tychinskaya, L.Y.; Gaidukevich, O.A.; Kozlov, N.G.; Klyuev, A.Y.; Lamotkin, S.A.; Shpak, S.I.; Rykov, S.V. NMR determination of the composition of balsams from Scotch pine resin. J. Appl. Spectrosc. 2008, 75, 439–443. [Google Scholar] [CrossRef]
  42. Rotondo, A.; Mannina, L.; Salvo, A. Multiple assignment recovered analysis (MARA) NMR for a direct food labeling: The case study of olive oils. Food Anal. Methods 2019, 12, 1238–1245. [Google Scholar] [CrossRef]
  43. Karkoula, E.; Skantzari, A.; Melliou, E.; Magiatis, P. Direct measurement of oleocanthal and oleacein levels in olive oil by quantitative 1H NMR. Establishment of a new index for the characterization of extra virgin olive oils. J. Agric. Food Chem. 2012, 60, 11696–11703. [Google Scholar] [CrossRef]
  44. Karkoula, E.; Skantzari, A.; Melliou, E.; Magiatis, P. Quantitative measurement of major secoiridoid derivatives in olive oil using qNMR. Proof of the artificial formation of aldehydic oleuropein and ligstroside aglycon isomers. J. Agric. Food Chem. 2014, 62, 600–607. [Google Scholar] [CrossRef] [PubMed]
  45. Manoukian, P.; Melliou, E.; Liouni, M.; Magiatis, P. Identification and quantitation of benzoxazinoids in wheat malt beer by qNMR and GC–MS. LWT Food Sci. Technol. 2016, 65, 1133–1137. [Google Scholar] [CrossRef]
  46. Cannac, M.; Barboni, T.; Ferrat, L.; Bighelli, A.; Castola, V.; Costa, J.; Trecul, D.; Morandini, F.; Pasqualini, V. Oleoresin flow and chemical composition of Corsican pine (Pinus nigra subsp. laricio) in response to prescribed burnings. For. Ecol. Manag. 2009, 257, 1247–1254. [Google Scholar] [CrossRef]
  47. Zobel, B.; Jett, J. Genetics of Wood Production; Springer-Verlag: Berlin, Germany, 1995; 337p. [Google Scholar]
  48. Venäläinen, M.; Harju, A.M.; Kainulainen, P.; Viitanen, H.; Nikulainen, H. Variation in the decay resistance and its relationship with other wood characteristics in old Scots pines. Ann. For. Sci. 2003, 60, 409–417. [Google Scholar] [CrossRef]
  49. Matziris, D. Genetic variation in morphological and anatomical needle characteristics in the Black pine of Peloponnesos. Silvae Genet. 1984, 33, 164–169. [Google Scholar]
  50. Ioannidis, K. Genetic Improvement of Multiple Traits in Black Pine (Pinus nigra Arn.). Ph.D. Thesis, Aristotle University of Thessaloniki, Thessaloniki, Greece, 2018. [Google Scholar]
  51. Willför, S.; Hemming, J.; Reunanen, M.; Holmbom, B. Phenolic and lipophilic extractives in Scots pine knots and stemwood. Holzforschung 2003, 57, 359–372. [Google Scholar] [CrossRef]
  52. Song, Z. Characteristics of Oleoresin and Classification of Pinus in China; China Forestry Publishing House: Beijing, China, 1998; 206p. [Google Scholar]
  53. Martínez-Inigo, M.J.; Immerzeel, P.; Gutierrez, A.; del Río, J.C.; Sierra-Alvarez, R. Biodegradability of extractives in sapwood and heartwood from Scots pine by sapstain and white-rot fungi. Holzforschung 1999, 53, 247–252. [Google Scholar] [CrossRef]
  54. Belt, T.; Keplinger, T.; Hänninen, T.; Rautkari, L. Cellular level distributions of Scots pine heartwood and knot heartwood extractives revealed by Raman spectroscopy imaging. Ind. Crops Prod. 2017, 108, 327–335. [Google Scholar] [CrossRef]
  55. Hafizoglu, H. Wood extractives of Pinus sylvestris L., Pinus nigra Arn. and Pinus brutia Ten. with special reference to nonpolar components. Holzforschung 1983, 37, 321–326. [Google Scholar]
  56. Uprichard, J.M.; Lloyd, J.A. Influence of tree age on the chemical composition of Radiata pine. N. Z. J. For. Sci. 1980, 10, 551–557. [Google Scholar]
  57. Yildirim, H.; Holmbom, B. Investigations on the wood extractives of pine species from Turkey: III Non-volatile, nonpolar components in Pinus brutia. Acta Acad. Abo. 1977, B37, 1–9. [Google Scholar]
  58. Hovelstad, H.; Leirset, I.; Oyaas, K.; Fiksdahl, A. Screening analyses of pinosylvin stilbenes, resin acids and lignans in Norwegian conifers. Molecules 2006, 11, 103–114. [Google Scholar] [CrossRef] [PubMed]
  59. Harju, A.M.; Venäläinen, M.; Anttonen, S.; Viitanen, H.; Kainulainen, P.; Saranpää, P.; Vapaavuori, E. Chemical factors affecting the brown-rot decay resistance of Scots pine heartwood. Trees 2003, 17, 263–268. [Google Scholar]
  60. Arshadi, M.; Backlund, I.; Geladi, P.; Bergsten, U. Comparison of fatty and resin acid composition in boreal Lodgepole pine and Scots pine for biorefinery applications. Ind. Crops Prod. 2013, 49, 535–541. [Google Scholar] [CrossRef]
  61. Yildirim, H.; Holmbom, B. Investigations on the wood extractives of pine species from Turkey: II Composition of fatty and resin acids in Pinus sylvestris and Pinus nigra. Acta Acad. Abo. 1977, B37, 1–6. [Google Scholar]
  62. Willför, S.; Hafizo-Glu, H.; Tümen, I.; Yazici, H.; Arfan, M.; Ali, M.; Holmbom, B. Extractives of Turkish and Pakistani tree species. Holz Roh. Werkst. 2007, 65, 215–221. [Google Scholar] [CrossRef]
  63. Lewinsohn, E.; Savage, T.J.; Gijzen, M.; Croteau, R. Simultaneous analysis of monoterpenes and diterpenoids of conifer oleoresin. Phytochem. Anal. 1993, 4, 220–225. [Google Scholar] [CrossRef]
  64. Benouadah, N.; Pranovich, A.; Aliouche, D.; Hemming, J.; Smeds, A.; Willför, S. Analysis of extractives from Pinus halepensis and Eucalyptus camaldulensis as predominant trees in Algeria. Holzforschung 2018, 72, 97–104. [Google Scholar] [CrossRef]
  65. Anderson, A.B.; Riffer, R.; Wong, A. Monoterpenes, fatty and resin acids of Pinus contorta and Pinus attenuata. Phytochemistry 1969, 8, 2401–2403. [Google Scholar] [CrossRef]
  66. Uçar, G.; Fengel, D. Variation in composition of extractives from wood of Pinus nigra varieties. Phytochemistry 1995, 38, 877–880. [Google Scholar] [CrossRef]
  67. Uçar, G.; Balaban, M. Cyclohexane extracts of black pine wood naturally grown in eastern Thrace. Holz Roh. Werkst. 2002, 60, 34–40. [Google Scholar] [CrossRef]
  68. Harju, A.M.; Kainulainen, P.; Venäläinen, M.; Tiitta, M.; Viitanen, H. Differences in resin acid concentration between brown-rot resistant and susceptible Scots pine heartwood. Holzforschung 2002, 56, 479–486. [Google Scholar] [CrossRef]
  69. Ekeberg, D.; Flæte, P.-O.; Eikenes, M.; Fongen, M.; Naess-Andresen, C.F. Qualitative and quantitative determination of extractives in heartwood of Scots pine (Pinus sylvestris L.) by gas chromatography. J. Chromatogr. A 2006, 1109, 267–272. [Google Scholar] [CrossRef] [PubMed]
  70. Papajannopoulos, A.D.; Song, Z.Q.; Liang, Z.Q.; Spanos, J.A. GC-MS analysis of oleoresin of three Greek pine species. Holz Roh. Werkst. 2001, 59, 443–446. [Google Scholar] [CrossRef]
  71. Lange, W.; Stevanović Janežić, T. Chemical composition of some Pinus sylvestris L. oleoresins from southern Serbia, Bosnia and Makedonia: Two entities of Scotch Pine present on the Balkan peninsula. Holzforschung 1993, 47, 207–212. [Google Scholar] [CrossRef]
  72. Matziris, D. Variation in cone production in a clonal seed orchard of black pine. Silvae Genet. 1993, 42, 136–141. [Google Scholar]
  73. Landucci, L.L.; Zinkel, D.F. The 1H and 13C-NMR spectra of the abietadienoic resin acids. Holzforschung 1991, 45, 341–346. [Google Scholar] [CrossRef]
  74. Muto, N.; Tomokuni, T.; Haramoto, M.; Tatemoto, H.; Nakanishi, T.; Inatomi, Y.; Murata, H.; Inada, A. Isolation of apoptosis- and differentiation-inducing substances toward human promyelocytic leukemia HL-60 cells from leaves of Juniperus taxifolia. Biosci. Biotechnol. Biochem. 2008, 72, 477–484. [Google Scholar] [CrossRef] [PubMed]
  75. Olate, V.R.; Usandizaga, O.G.; Schmeda-Hirschmann, G. Resin diterpenes from Austrocedrus chilensis. Molecules 2011, 16, 10653–10667. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Samples of the extracts are available from the authors.
Figure 1. The structures and numbering of the diterpenic resin acids of the Pinus nigra heartwood extracts.
Figure 1. The structures and numbering of the diterpenic resin acids of the Pinus nigra heartwood extracts.
Molecules 24 03603 g001
Figure 2. 1H-NMR spectrum (400 MHz, CDCl3, δ-values in ppm) of black pine (P. nigra Arn.) heartwood extracts, showing the characteristic peaks of studied resin acids and internal standard. The numbers indicate the corresponding resin acids’ peaks used for quantitation: abietic acid (1), neoabietic acid (2), dehydroabietic acid (3), palustric acid (4), levopimaric acid (5), pimaric acid (6), sandaracopimaric acid (7), and isopimaric acid (8). Stilbenes’ peaks are also indicated (P = pinosylvin, PMME and PDME = monomethylether and dimethylether of pinosylvin, respectively).
Figure 2. 1H-NMR spectrum (400 MHz, CDCl3, δ-values in ppm) of black pine (P. nigra Arn.) heartwood extracts, showing the characteristic peaks of studied resin acids and internal standard. The numbers indicate the corresponding resin acids’ peaks used for quantitation: abietic acid (1), neoabietic acid (2), dehydroabietic acid (3), palustric acid (4), levopimaric acid (5), pimaric acid (6), sandaracopimaric acid (7), and isopimaric acid (8). Stilbenes’ peaks are also indicated (P = pinosylvin, PMME and PDME = monomethylether and dimethylether of pinosylvin, respectively).
Molecules 24 03603 g002
Figure 3. The mean percentage of abietane and pimarane types of resin acids.
Figure 3. The mean percentage of abietane and pimarane types of resin acids.
Molecules 24 03603 g003
Figure 4. The mean percentage content of the constituents in the black pine’s heartwood acetone extraction from the Peloponnese (P = pinosylvin, PMME and PDME = monomethylether and dimethylether of pinosylvin, respectively).
Figure 4. The mean percentage content of the constituents in the black pine’s heartwood acetone extraction from the Peloponnese (P = pinosylvin, PMME and PDME = monomethylether and dimethylether of pinosylvin, respectively).
Molecules 24 03603 g004
Figure 5. Total resin acids frequency distribution of black pine (P. nigra Arn.).
Figure 5. Total resin acids frequency distribution of black pine (P. nigra Arn.).
Molecules 24 03603 g005
Figure 6. The positions of the provenances (circles) and the clonal seed orchard (CSO).
Figure 6. The positions of the provenances (circles) and the clonal seed orchard (CSO).
Molecules 24 03603 g006
Table 1. 1H-NMR proton’s chemical shifts of the studied resin acids selected for quantitation (400 MHz, CDCl3, δ-values in ppm).
Table 1. 1H-NMR proton’s chemical shifts of the studied resin acids selected for quantitation (400 MHz, CDCl3, δ-values in ppm).
Resin AcidAbietane TypePimarane Type
ProtonAbietic Acid (1)Neoabietic Acid (2)Dehydroabietic Acid (3)Palustric Acid (4)Levopimaric Acid (5)Pimaric Acid (6)Sandaracopimaric Acid (7)Isopimaric Acid (8)
C(H)-15-----5.67-5.84
C(H)-145.776.206.885.395.53-5.22-
Table 2. Recovery data and their coefficient of variance for the studied resin acids.
Table 2. Recovery data and their coefficient of variance for the studied resin acids.
Abietic Acid (1)Neoabietic Acid (2)Dehydroabietic Acid (3)Palustric Acid (4)Levopimaric Acid (5)Pimaric Acid (6)Sandaracopimaric Acid (7)Isopimaric Acid (8)
Recovery (%)90.2693.5288.7890.5090.7392.7592.0787.37
Coefficient of Variance (%)1.061.820.852.398.401.933.821.48
Table 3. Intraday precision data for the studied resin acids, expressed as relative standard deviations (%).
Table 3. Intraday precision data for the studied resin acids, expressed as relative standard deviations (%).
IDAbietic Acid (1)Neoabietic Acid (2)Dehydroabietic Acid (3)Palustric Acid (4)Levopimaric Acid (5)Pimaric Acid (6)Sandaracopimaric Acid (7)Isopimaric Acid (8)
Sample 15.556.298.842.179.075.099.829.07
Sample 26.517.383.495.184.857.552.883.31
Sample 33.685.402.489.928.817.662.987.57
Average5.256.364.935.767.586.775.226.65
Table 4. The mean resin acid concentrations of Pinus nigra L. heartwood samples (n = 260) as determined by quantitative 1H-NMR.
Table 4. The mean resin acid concentrations of Pinus nigra L. heartwood samples (n = 260) as determined by quantitative 1H-NMR.
Abietic Acid (1)Neoabietic Acid (2)Dehydroabietic Acid (3)Palustric Acid (4)Levopimaric Acid (5)Pimaric Acid (6)Sandaracopimaric Acid (7)Isopimaric Acid (8)Total Resin Acids
Average mg/gdhw76.7739.3411.6947.948.0722.542.7210.91219.98
Min. mg/gdhw7.002.912.569.760.082.200.160.5030.05
Max. mg/gdhw181.75101.8238.59105.2264.9159.426.6734.09424.70
Table 5. The mean resin acid concentrations (mg/gdhw) of the four Pinus nigra L. provenances’ heartwood after applying ANOVA and Duncan’s MRT at p = 0.05, Zarouhla: 65 samples, Feneos: 85 samples, Parnonas: 45 samples, and Taigetos: 65 samples, total 260 samples.
Table 5. The mean resin acid concentrations (mg/gdhw) of the four Pinus nigra L. provenances’ heartwood after applying ANOVA and Duncan’s MRT at p = 0.05, Zarouhla: 65 samples, Feneos: 85 samples, Parnonas: 45 samples, and Taigetos: 65 samples, total 260 samples.
ProvenanceAbietic Acid (1)Neoabietic Acid (2)Dehydroabietic Acid (3)Palustric Acid (4)Levopimaric Acid (5)Pimaric Acid (6)Sandaracopimaric Acid (7)Isopimaric Acid (8)Total Resin Acids
Zarouhla81.99 a40.06 a11.10 a48.83 a7.41 a21.60 ab2.61 a11.45 a224.46 a
Feneos78.64 a37.84 a12.29 a46.43 a8.91 a25.09 a2.74 a11.03 a222.768 a
Parnonas79.99 ab41.87 a11.74 a49.86 a6.88 a23.57 ab2.92 a11.20 a227.83 a
Taigetos66.05 b36.36 a11.82 a45.61 a9.69 a20.35 b2.54 a9.89 a201.25 a
The means followed by the same letter (a, b as superscript) are not statistically different.

Share and Cite

MDPI and ACS Style

Ioannidis, K.; Melliou, E.; Magiatis, P. High-Throughput 1H-Nuclear Magnetic Resonance-Based Screening for the Identification and Quantification of Heartwood Diterpenic Acids in Four Black Pine (Pinus nigra Arn.) Marginal Provenances in Greece. Molecules 2019, 24, 3603. https://doi.org/10.3390/molecules24193603

AMA Style

Ioannidis K, Melliou E, Magiatis P. High-Throughput 1H-Nuclear Magnetic Resonance-Based Screening for the Identification and Quantification of Heartwood Diterpenic Acids in Four Black Pine (Pinus nigra Arn.) Marginal Provenances in Greece. Molecules. 2019; 24(19):3603. https://doi.org/10.3390/molecules24193603

Chicago/Turabian Style

Ioannidis, Kostas, Eleni Melliou, and Prokopios Magiatis. 2019. "High-Throughput 1H-Nuclear Magnetic Resonance-Based Screening for the Identification and Quantification of Heartwood Diterpenic Acids in Four Black Pine (Pinus nigra Arn.) Marginal Provenances in Greece" Molecules 24, no. 19: 3603. https://doi.org/10.3390/molecules24193603

Article Metrics

Back to TopTop