Skip to content
Publicly Available Published by De Gruyter June 29, 2017

A succinct access to ω-hydroxylated jasmonates via olefin metathesis

  • Guillermo H. Jimenez-Aleman EMAIL logo , Selina Seçinti and Wilhelm Boland EMAIL logo

Abstract

In higher plants, jasmonates are lipid-derived signaling molecules that control many physiological processes, including responses to abiotic stress, defenses against insects and pathogens, and development. Among jasmonates, ω-oxidized compounds form an important subfamily. The biological roles of these ω-modified derivatives are not fully understood, largely due to their limited availability. Herein, a brief (two-step), simple and efficient (>80% yield), versatile, gram-scalable, and environmentally friendly synthetic route to ω-oxidized jasmonates is described. The approach utilizes olefin cross-metathesis as the key step employing inexpensive, commercially available substrates and catalysts.

1 Introduction

Jasmonates are signaling molecules that are ubiquitous in higher plants. They are key regulators of plant growth and development, as well as of responses to both biotic and abiotic stresses [1], [2]. Jasmonates are biosynthesized from α-linolenic acid, which, after several enzymatic steps, yields jasmonic acid [JA (1)] as the (3R,7S)-isomer (Scheme 1). However, some degree of epimerization at C7 may occur, and, moreover, the 7R epimer can be found within the cell [3]. Like JA (1), JA-derivatives exist in planta as a mixture of the 7S and 7R epimers. For instance, compound JKL (7) has been described as the 7R epimer [4], [5]. In response to environmental and developmental cues, JA (1) is activated by its conjugation to L-isoleucine (Ile) [6]. This reaction produces the most active jasmonate (3R,7S)-N-jasmonoyl-L-isoleucine (JA-Ile, 4). The ω-oxidation of active JA-Ile (4) is carried out by members of the CYP94 family. This reaction produces the ω-hydroxylated acid 12-OH-JA-Ile (8) that is further converted into other ω-modified (12-modified) jasmonates (Scheme 1). Many 12-modified jasmonates (e.g. compounds 510, Scheme 1) exist in relatively high abundance (compared to other jasmonates) in various plant species [7]; however, little is known about their physiological roles. These chemical species may be stored in cell compartments such as vacuoles and may contribute to maintaining the homeostasis of jasmonates within plant tissues, which is of crucial importance for plant fitness and defense [8]. On the one hand, naturally occurring 12-modified jasmonates are believed to be partially or completely inactive with regard to the canonical jasmonate pathway [9], [10]. However, 12-OGlc-JA (5) was found to activate leaf-closing movement in Samaneasaman [11]. The action mechanism of 12-OGlc-JA (5) is independent of the COI1-JAZ complex, the canonical jasmonate receptor. This suggests that endogenous 12-modified jasmonates may have biological roles beyond the classical jasmonate functions known to date. On the other hand, some synthetic 12-modified jasmonate derivatives have been described as inducers of plant defenses similarly to bioactive endogenous jasmonates [12]. Furthermore, these derivatives were recently employed to uncouple growth and defense in Nicotiana attenuata plants, opening a whole new field of research with potential application in agriculture [13]. Nonetheless, the production costs of 12-modified jasmonates (and jasmonates in general) must be sufficiently low if these compounds are to be employed in agricultural systems. The above-mentioned observations highlight the need to develop new synthetic strategies to prepare 12-modified jasmonates from a commercial product, such as methyljasmonate, in order to evaluate their mode of action and eventually to employ these compounds in agriculture.

Scheme 1: Simplified scheme of jasmonate biosynthesis.LOX, Lipoxygenase; AOS, allene oxide synthase; AOC, allene oxide cyclase; OPR3, 12-oxophytodienoate reductase 3; JMT, JA carboxyl methyltransferase; MJE, methyljasmonate esterase; JAR1, JA-amino synthetase; ILL6/IAR3, JA-Ile amidohydrolases; CYP94B1/B3/C1, cytochromes P450; LFE, lactone forming enzyme. Red squares depict ω modifications. A dashed arrow represents a proposed transformation. Nomenclature and numbering commonly used for jasmonates is employed for clarity. Epimerization at C7 may occur, lactone JKL (7) exists in nature. For comprehensive reviews on jasmonate biosynthesis, signaling and activity, please see reference [1].
Scheme 1:

Simplified scheme of jasmonate biosynthesis.

LOX, Lipoxygenase; AOS, allene oxide synthase; AOC, allene oxide cyclase; OPR3, 12-oxophytodienoate reductase 3; JMT, JA carboxyl methyltransferase; MJE, methyljasmonate esterase; JAR1, JA-amino synthetase; ILL6/IAR3, JA-Ile amidohydrolases; CYP94B1/B3/C1, cytochromes P450; LFE, lactone forming enzyme. Red squares depict ω modifications. A dashed arrow represents a proposed transformation. Nomenclature and numbering commonly used for jasmonates is employed for clarity. Epimerization at C7 may occur, lactone JKL (7) exists in nature. For comprehensive reviews on jasmonate biosynthesis, signaling and activity, please see reference [1].

The jasmonates’ biological relevance made them the target of several synthetic approaches aiming to obtain not only naturally occurring compounds but also synthetic analogs; all of these contributed greatly to our understanding of how these phytohormones act [14], [15], [16]. A few studies have reported on the synthesis of 12-modified jasmonates [17], [18]. However, because the procedures for synthesis are lengthy and provide low overall yield, access to these compounds for both bioassays and agricultural applications may be limited [19]. To circumvent these limitations, a synthetic route to ω-modified jasmonates based on the cross-metathesis (CM) of alkenes was explored (Scheme 2). The preferential formation of the more stable E over the Z isomer and the generation of complex mixtures disadvantaged CM for years. Nowadays, at least the stereoselectivity in favor of the Z isomer is no longer an issue, thanks to the recent development of molybdenum- [20], [21], tungsten- [20], and ruthenium-based (commercially available, inexpensive, and versatile) Z-selective catalysts [22], [23].

Scheme 2: Alternative routes to 12-modified jasmonates via CM metathesis.(ia, iib) CM of 2 with different terminal alkenes; (ib, iia) saponification/conjugation to selected amino acids (aa).
Scheme 2:

Alternative routes to 12-modified jasmonates via CM metathesis.

(ia, iib) CM of 2 with different terminal alkenes; (ib, iia) saponification/conjugation to selected amino acids (aa).

Jasmonates (commercially available or synthetic) are commonly employed in bioassays as mixtures of all possible diastereoisomers – in particular, with respect to the stereocenters at the cyclopentanone ring [16], [24], [25]. This is not at all disadvantageous, as cross-inhibition by the non-natural isomers or by the natural but less active (3R,7R) form does not occur. Therefore, inexpensive, commercially available MeJA (2) [a mixture of the (3R,7R), (3S,7S), (3S,7R), and (3R,7S) isomers in a ratio of approximately 45:45:5:5, respectively] was employed throughout this study in order to develop and optimize the CM method. And hence, all ω-modified jasmonates were prepared as a diastereomeric mixture unless otherwise stated.

2 Results and discussion

2.1 Optimization of the reaction conditions for CM

First, the CM reaction of MeJA (2) with 14a was investigated employing the Grubbs second-generation catalyst 11 (Scheme 3). Despite the broad functional group tolerance of 11 [26], the alcohol 14a barely reacted with 2, which meant that 2 could be recovered almost completely from the reaction mixture. This result may be explained by the reported low reactivity of homoallylic free alcohols in metathesis reactions [27]. Therefore, the protected alcohol 14b was evaluated in the CM with 2 (Scheme 3). Catalyst 11 was active at loadings as low as 0.5 mol% (relative to 2) in the CM. Product 15b was obtained in 61% yield after 1 h using 1 mol% of 11 (Table 1, entry 2). To further optimize the reaction, solvents commonly employed in metathesis were screened (Table 1, entries 2, 4, and 5). All tested solvents performed similarly, but THF gave higher yields of 15b (Table 1, entry 2). Conveniently, the reaction slowly proceeded under solvent-free conditions, so the process was environmentally friendly (Table 1, entry 6). The performance of catalysts 12 (Hoveyda–Grubbs second generation) and 13 (Grubbs Z-selective) was also evaluated. Reactions were run in THF at 27°C with substrates at a concentration of approximately 0.2 M and 1 mol% catalyst loading in order to better assess differences between catalysts [23]. Catalysts 11 and 12 produced similar yield towards 15b after 1 h of reaction (Table 1, entries 2 and 7). However, the Z-selective catalyst 13 needed a longer induction time (compared to 11 and 12), and for up to 1 h, no CM product was detected. This was unexpected, as 13 has been reported active in the homodimerization of monosubstituted alkenes at catalyst loadings as low as 0.1 mol% [23]. The low reactivity of 13 may be explained by the nature of both substrate 2 and catalyst 13. MeJA (2) is a 1,2-disubstituted olefin with a bulky cyclopentanone ring that could favor the homodimerization of 14b over the CM product due to steric hindrance with the bulky adamantyl group of 13. Following this reasoning, both catalyst loadings of 13 and different ratios of the starting substrates 2 and 14b were screened (Table 2). It must be noted that under solvent-free conditions and at room temperature, 13 was active from a catalyst loading of 5 mol%, giving an excellent 95% of Z-selectivity (Table 2, entry 2). On the other hand, a six-molar excess of the terminal alkene 14b improved the yield toward 15b significantly, while the amount of 16 (homodimerization product of 2) did not exceed 5% (Table 2, entry 6). While increasing the temperature may speed up a CM reaction, it may also negatively influence the yield and/or the Z-selectivity of the desired CM product [23]. Strikingly, increasing the temperature in the CM of 2 with 14b shortened reaction times from 24 h to 1 h and slightly improved the Z-selectivity for 15b (Table 2, entries 8–10). It is worth mentioning that at short reaction times, the self-metathesis product 16 was not observed.

Scheme 3: CM reactions explored in this study and catalysts used.(i) Screening of catalysts 11, 12 and 13; catalyst loadings; solvents (THF, CH2Cl2 and C6H6); ratio of reactants; temperatures (27, 40, 60 and 80°C).
Scheme 3:

CM reactions explored in this study and catalysts used.

(i) Screening of catalysts 11, 12 and 13; catalyst loadings; solvents (THF, CH2Cl2 and C6H6); ratio of reactants; temperatures (27, 40, 60 and 80°C).

Table 1:

CM of 2 and 14b: screening of catalysts, catalyst loading and solvents.

EntryCatalystCatalyst loading (mol%)Solvent15b (%)a
1110.5THF21
2111THF61
3112THF60
4111Benzene44
5111CH2Cl224
6111No solvent4
7121THF50
8131THFn.d.b
  1. aYield determined by GC-MS. bn.d., not detected.

Table 2:

CM of 2 and 14b using catalyst 13: screening of catalyst loading, ratio of substrates and temperature.

EntryCatalyst loading (mol%)14b:2 (mol/mol)t (h)T (°C)15b (%)aZ-15b (%)b16 (%)a,c
111:12427~1
251:1242725953
3101:1242734957
451:2242725945
552:1242724943
656:1242764935
756:1127n.d.d
856:11401597n.d.
956:11601497n.d.
1056:11802196n.d.
  1. aYield determined by GC-MS. bPercentage of the Z-isomer of 15b. cResponse factor assumed as for15b. dn.d., not detected.

2.2 Influence of the -R group of 14

The reaction of 2 with substrates 14b–d was studied with complexes 11 and 12 in order to generalize the method for these unselective but widely employed catalysts. As expected, 11 and 12 preferentially afforded the E-15b–d alkene (Table 3). Interestingly, the Z/E ratio of the CM product 15 was influenced by the nature of the protecting group. The amount of Z-15 increased about two-fold with 14d as the substrate (THP as the protecting group; Table 3, entry 3) compared to 14b (Table 3, entry 1). Therefore, the E/Z ratio of the CM-product 15 can be fine-tuned by choosing the best protective group in substrate 14. This observation may be useful in other CM reactions when the protection of alcohol groups is needed. Envisioning the possibility of directly installing more complex substituents at C12 (e.g. a sugar moiety as in jasmonate 5, Scheme 1), the CM of 2 and the more challenging substrate 14e (sterically hindered and more likely to chelate Ru) was explored. Under the employed conditions [13 (20 mol%), not optimized], the CM product 15e was obtained in 40% yield after purification and 60% of Z-selectivity. This result suggests that it is possible to install complex moieties at C12, such as a glycoside, by employing CM. Nonetheless, the preparation of a glycoside such as 12-OGlc-JA (5) might be made more efficient by the glycosylation of Z-3, which can be readily obtained from Z-15b (more than 90% Z-selectivity). Furthermore, it is reasonable to speculate that the CM of 2 and terminal alkenes bearing a protected sulfonic acid, aldehyde, or carboxylic acid group may offer an efficient way to prepare compounds 6, 9, or 10, respectively.

Table 3:

CM of 2 and 14b–d using catalyst 11 or 12.

EntrySubstrateCatalystZ-15 (%)a,b
114b1117
214c1120
314d1129
414b1218
514c1219
614d1224
  1. aPercentage of Z-CM product 15 determined by GC-MS. bAverage of two separate runs.

2.3 Gram scale preparation of Z-15b

Previous reports have shown that CM reactions perform better under a static vacuum, which helps to remove undesired volatile olefins from the reaction mixture [28]. Following this idea, the CM of 2 and 14b was carried out under a dynamic vacuum in order to remove both 1-butene and ethylene formed during the CM (Scheme 4). The versatile jasmonate 15b (1.05 g, >80% after column purification) was obtained with more than 90% Z-selectivity. It is worth mentioning that 17b (a valuable 1,2-disubstituted, structurally symmetric alkene, as no ethylene is produced in self-metathesis [29]) was recovered and may be employed in further reactions. This CM allows for the gram-scale preparation of the valuable scaffold 15b, a scaffold which could be used as the substrate to prepare many other 12-modified jasmonates.

Scheme 4: Gram-scale preparation of 15b employing catalyst 13.
Scheme 4:

Gram-scale preparation of 15b employing catalyst 13.

2.4 Preparation of 12-modified derivatives of JA-Ile (4)

Several 12-modified jasmonates exist in nature as conjugates with amino acids [30], [31], for example, compound 8 (Scheme 1). Therefore, it was briefly explored whether the preparation of such derivatives would be more efficient with the CM reaction preceding the conjugation to the amino acid (Scheme 2, ia–iia) or vice versa (Scheme 2, ib–iib). The Z-selective catalyst 13 (catalyst loading up to 7 mol%) failed to promote the CM reaction of JA-Ile (4) [mixture of diastereoisomers] and 14b. This result is consistent with the low activity reported for 13 catalyzing the CM of substrates bearing aldehyde or carboxylic acid groups [29]. However, catalyst 11 promoted CM at a catalyst loading of 3 mol% (Figure S1). These results indicate that the CM reaction should precede the conjugation to amino acids when aiming the synthesis of Z-enriched 12-modified JA-amino acid conjugates.

3 Conclusions

The CM reaction described here can be conducted on the lab bench employing inexpensive, commercially available catalysts 11, 12, or 13, and the substrates MeJA (2) and 14b. The Z-selective catalyst 13 should be used to prepare the (Z)-jasmonates resembling the natural signals. In addition, (E)-enriched 12-modified derivatives can be prepared by employing catalyst 11 or 12. This feature may be of interest when conducting structure activity relationship (SAR) studies. The MeJA (2) employed in this study was a mixture of all possible diastereoisomers containing the (3R,7R)- and (3S,7S)-diastereoisomers (both trans- relative to the cyclopentanone ring) in a roughly 1:1 ratio. A similar 1:1 ratio of (3R,7R)- and (3S,7S)-diastereoisomers was found for two diastereomeric macrolactones prepared by using the CM method presented here as a key step [13]. This result indicates that the CM reaction of substrates 2 and 14b is not stereocontrolled by the preexisting stereocenters on MeJA (2). Epimerization at C7 may occur both in vivo and when handling solutions of synthetic jasmonates [3]; therefore, our synthetic approach starting with inexpensive MeJA (2) is extremely useful.

In conclusion, an efficient synthetic approach which allows for the preparation of many 12-modified jasmonates has been developed based on olefin CM. Due to its efficiency and versatility, this method will likely be widely employed to prepare both naturally occurring and rationally designed 12-modified jasmonates of biological interest. The availability of such compounds (grams if needed) makes it possible to study their biological properties and, eventually, to apply them in agriculture. Additionally, the described method allows for the preparation of molecular probes in order to better understand a major signaling pathway in plants.

4 Experimental section

4.1 General experimental methods and materials

Reactions were carried out in oven-dried glassware. Flash-column chromatography was performed on silica gel 60 (230–400 mesh). Plates of silica gel 60 F254 were used for thin-layer chromatography. Compounds were visualized by soaking the plates in a potassium permanganate staining agent. All solvents, catalysts, and substrates 2, 14a, and 14b were commercially available and were used as received. Compounds 1 [25], 4 [25], 14c [32], 14d [33], and 14e [34] were prepared according to procedures described in the literature. Spectroscopic data from 15c and 15d matched those previously described [17], [35]. Nuclear magnetic resonance (NMR) spectra were acquired at 400 MHz for 1H NMR and 100 MHz for 13C NMR. Chemical shifts were assigned relative to the residual solvent peak, which served as an internal standard. Gas chromatography-mass spectrometry (GC-MS) spectra were measured using electron impact (70 eV) as the ionization method in positive mode (EI+). High-resolution mass spectrometry (HRMS) electrospray ionization in positive mode (ESI+) was performed on high-resolution time-of-flight (TOF) equipment.

4.2 General procedure for CM reactions

CM reactions were carried out following the methodology described in [36]. Briefly, the desired amount of catalyst was added to a 4 mL vial. The vial was sealed with a screw cap to which a rubber septum was attached and purged with argon for 5 min. A syringe was used to add first the solvent, then a mixture of the substrates. The reaction was allowed to stir in a heating block at the desired temperature. Aliquots of the reaction mixture were taken at the specified time periods. Samples for GC-MS analysis were obtained by adding 20 μL of the reaction to 0.5 mL of a solution of ethyl vinyl ether (3 mol L−1 in CH2Cl2); reactions under solvent-free conditions were stopped in a similar manner, but 1 μL aliquots were taken unless otherwise stated. The sample was shaken for 5 min and then analyzed via GC-MS. All reactions were performed in duplicate. Conditions for each particular experiment were as follows:

  • Screening catalyst loadings of 11, Table 1: 11 (0.9, 1.8, 3.7, or 9.0 mg; 1.1, 2.2, 4.4, or 11 μmol, respectively); THF (1 mL); 2 (50.6 mg; 0.22 mmol); 14b (24.9 mg; 0.22 mmol); 27°C.

  • Screening solvents, Table 1: 11 (1.8 mg; 2.2 μmol, 1 mol%); benzene, THF, toluene, or dichloromethane (1 mL); 2 (50.6 mg; 0.22 mmol); 14b (24.9 mg; 0.22 mmol); 27°C. For solvent free; 2 (101.2 mg; 0.44 mmol); 14b (49.8 mg; 0.44 mmol).

  • Screening catalysts 11, 12, and 13 at 1 mol%, Table 1: 11, 12, or 13 (2.2 μmol; 1 mol%; 1.8, 1.4, or 1.4 mg, respectively); THF (1 mL); 2 (50.6 mg; 0.22 mmol); 14b (24.9 mg; 0.22 mmol); 27°C.

  • Screening catalyst loadings of 13, Table 2: 13 (2.8, 13.9, or 27.8 mg; 4.4, 22, or 44 μmol, respectively); 2 (101.2 mg; 0.44 mmol); 14b (49.8 mg; 0.44 mmol); Neat; 27°C.

  • Screening substrate ratios, Table 2: 13 (6.0 mg, 9.5 μmol, 5 mol%); 14b/2 (0.1/0.2, 0.2/0.2, 0.4/0.2, or 1.2/0.2, mmol/mmol); Neat; 27°C. Experiments were conducted in 1.5 mL vials.

  • Screening temperatures, Table 2: 13 (6 mg, 9.5 μmol, 5 mol%); 2 (46.0 mg; 0.2 mmol); 14b (135.6 mg; 1.2 mmol); Neat; 27, 40, 60, or 80°C. Experiments were conducted in 1.5 mL vials.

  • Screening protecting groups, Table 3: 11 (2.0 mg, 2.4 μmol, 1 mol%); 2 (55.2 mg; 0.24 mmol); 14b, 14c, or 14d (27.1, 44.0, or 37.5 mg, respectively; 0.24 mmol). 12 (2.0 mg, 3.2 μmol); 2 (73.6 mg; 0.32 mmol); 14b, 14c, or 14d (36.2, 59.6, or 50.5 mg, respectively; 0.32 mmol). Reactions were carried out under solvent-free conditions in 1.5 mL vials. Aliquots of 2 μL were taken after 1 h for GC-MS analysis.

  • CM of JA-Ile (4) and 14b, Figure S1: 11 (6.0 mg, 7 μmol, 3 mol% relative to 4) or 13 (17 mg, 27 μmol, 7 mol% relative to 4); 4 (74.4 mg, 0.23 mmol); 14b (259.0 mg, 2.3 mmol); Neat; 60°C. The reaction was followed by thin-layer chromatography. Samples for GC-MS analysis were obtained by adding 2 μL of the reaction mixture to 500 μL of ethyl vinyl ether (3 mol L−1 in CH2Cl2), shaking for 5 min and derivatizing with trimethylsilyldiazomethane.

4.3 Synthetic procedures

4.3.1 Methyl (Z)-2-(2-(5-acetoxypent-2-en-1-yl)-3-oxocyclopentyl)acetate (Z-15b); gram-scale

A 10 mL Schlenk flask was charged with 2 (1.02 g, 4.5 mmol), 14b (3.05 g, 27 mmol, 6.0 equiv.) and catalyst 13 (0.14 g, 0.23 mmol, 5 mol%). The flask was connected to a vacuum line (400 mbar, dynamic vacuum), and gentle warming (45°C) was applied while the contents were being stirred. After 3 h, the reaction was quenched by the addition of 5 mL of a solution of ethyl vinyl ether (3 mol L−1 in CH2Cl2). Stirring continued for 30 min, at which time Pb(AcO)4 (1.5 equivalents relative to the catalyst) was added and the mixture stirred overnight. The reaction mixture was then filtered through a short pad of silica (approximately 10 g/0.05 mmol of catalyst) and eluted with four times 55 mL of CH2Cl2:AcOEt (10:1, v/v). The combined organic fractions were concentrated at reduced pressure and purified by flash chromatography on SiO2 (150 g) eluted with AcOEt:n-pentane (3:7, v/v). Product 15b (1.05 g; 83%, Z:E=93:7) was obtained as a pale yellow oil. GC-MS (EI+): m/z (%): 43 (100), 67 (43), 83 (52), 130 (57), 149 (83), 193 (20), 282 (3) [M]+; HRMS (ESI+-TOF): m/z=305.1354 [M + Na]+ (calc. for C15H22NaO5=305.1359); 1H NMR (CDCl3, 400 MHz): δ 5.39 (m; 2H), 4.00 (td; J=6.9, 6.9, 1.3 Hz; 2H), 3.63 (s; 3H), 2.63 (m; 1H), 2.40–2.12 (m; 8H), 2.12–1.9 (m; 1H), 1.98 (s; 3H), 1.86 (dt; J=10.3, 5.4, 5.4 Hz; 1H), 1.45 ppm (m; 1H); 13C NMR (CDCl3, 100 MHz): δ 218.7, 172.5, 171.1, 128.6, 127.0, 63.7, 53.9, 51.7, 38.7, 38.0, 37.7, 27.2, 26.9, 25.5, 20.9 ppm.

4.3.2 (2R,3R,4S,5R)-2-(acetoxymethyl)-6-(((Z)-5-(2-(2-methoxy-2-oxoethyl)-5-oxocyclopentyl)pent-3-en-1-yl)oxy)tetrahydro-2H-pyran-3,4,5-triyl triacetate (15e)

A two-neck flask fitted with a condenser was charged with 2 (59 mg, 0.26 mmol, 2 equiv.), 14e (53 mg, 0.13 mmol, α:β=1:1), THF (1 mL) and catalyst 13 (16.6 mg, 0.03 mmol, 20 mol% relative to 14e). A gentle stream of argon was passed through the solution, and the mixture was warmed to 45°C. The reaction was stirred overnight (approximately 18 h) and worked up as described for 15b [flash chromatography using AcOEt:n-hexane (1:1, (v/v)]. Product 15e (30 mg; 40%, Z:E=60:40) was obtained as a viscous oil. HRMS (ESI+-TOF): m/z=593.2181 [M + Na]+ (calc. for C27H38NaO13=593.2205); 1H NMR (CDCl3, 400 MHz): δ 5.44 (m; 2H), 5.19 (m; 1H), 5.03 (m; 2H),4.50 (d; J=2.9 Hz; 1H), 4.26 (m; 1H), 4.05 (m; 2H), 3.87 (m; 1H), 3.70 (m; 4H), 3.46 (m; 1H), 2.66 (m; 1H), 2.32 (m; 7H), 2.07 (m; 13H), 1.91 (m; 1H), 1.66 ppm (m; 1H); 13C NMR (CDCl3, 100 MHz): δ 218.9, 172.6, 170.8, 170.4, 169.6, 169.5, 128.4, 127.5, 100.9, 73.0, 71.9, 71.4, 69.6, 68.6, 62.1, 54.1, 51.8, 38.6, 38.1, 37.9, 27.4, 25.7, 25.7, 20.9, 20.8, 20.8, 20.8 ppm (β-anomer).


Dedication: This article is dedicated to the memory of Prof. Dr. Lothar Jaenicke.


Acknowledgments

All experimental work of this study was supported by the Max Planck Society and the DFG (ChemBioSys, CRC 1127). Special thanks go to Kerstin Ploss for HRMS measurements and to Emily Wheeler for editorial assistance.

References

1. Wasternack C, Hause B. Jasmonates: biosynthesis, perception, signal transduction and action in plant stress response, growth and development. An update to the 2007 review in Annals of Botany. Ann Bot 2013;111:1021–58.10.1093/aob/mct067Search in Google Scholar PubMed PubMed Central

2. Machado RAR, McClure M, Hervé MR, Baldwin IT, Erb M. Benefits of jasmonate-dependent defenses against vertebrate herbivores in nature. Elife 2016;5:e13720.10.7554/eLife.13720Search in Google Scholar PubMed PubMed Central

3. Fonseca S, Chini A, Hamberg M, Adie B, Porzel A, Kramell R, et al. (+)-7-iso-jasmonoyl-l-isoleucine is the endogenous bioactive jasmonate. Nat Chem Biol 2009;5:344–50.10.1038/nchembio.161Search in Google Scholar PubMed

4. Inoue M, Nishi T, Kitahara T. Enantioselective synthesis of both the enantiomers of jasmine ketolactone and its epimer. Biosci Biotechnol Biochem 1999;63:1122–4.10.1271/bbb.63.1122Search in Google Scholar PubMed

5. Nishi T, Kitahara T. Synthesis of natural (+)-methyl tuberonate, (−)-jasmine ketolactone and its epimer. Proc Japan Acad 1995;71B:20–3.10.2183/pjab.71.20Search in Google Scholar

6. Staswick PE, Tiryaki I. The oxylipin signal jasmonic acid is activated by an enzyme that conjugates it to isoleucine in Arabidopsis. Plant Cell 2004;16:2117–27.10.1105/tpc.104.023549Search in Google Scholar PubMed PubMed Central

7. Miersch O, Neumerkel J, Dippe M, Stenzel I, Wasternack C. Hydroxylated jasmonates are commonly occurring metabolites of jasmonic acid and contribute to a partial switch-off in jasmonate signaling. New Phytol 2008;177:114–27.10.1111/j.1469-8137.2007.02252.xSearch in Google Scholar PubMed

8. Koo AJ, Howe GA. The wound hormone jasmonate. Phytochemistry 2009;70:1571–80.10.1016/j.phytochem.2009.07.018Search in Google Scholar PubMed PubMed Central

9. Wasternack C, Song S. Jasmonates: biosynthesis, metabolism, and signaling by proteins activating and repressing transciption. J Exp Bot 2017;68:1303–21.10.1093/jxb/erw443Search in Google Scholar PubMed

10. Koo AJ, Thireault C, Zemelis S, Poudel AN, Zhang T, Kitaoka N, et al. Endoplasmic reticulum-associated inactivation of the hormone jasmonoyl-L-isoleucine by multiple members of the cytochrome P450 94 family in Arabidopsis. J Biol Chem 2014;289:29728–38.10.1074/jbc.M114.603084Search in Google Scholar PubMed PubMed Central

11. Nakamura Y, Mithöfer A, Kombrink E, Boland W, Hamamoto S, Uozumi N, et al. 12-hydroxyjasmonic acid glucoside is a COI1-JAZ-independent activator of leaf-closing movement in Samanea saman. Plant Physiol 2011;155:1226–36.10.1104/pp.110.168617Search in Google Scholar PubMed PubMed Central

12. Jimenez-Aleman GH, Machado RA, Görls H, Baldwin IT, Boland W. Synthesis, structural characterization and biological activity of two diastereomeric JA-Ile macrolactones. Org Biomol Chem 2015;13:5885–93.10.1039/C5OB00362HSearch in Google Scholar PubMed

13. Jimenez-Aleman GH, Machado RAR, Baldwin IT, Boland W. JA-Ile-macrolactones uncouple growth and defense in wild tobacco. Org Biomol Chem 2017;15:3391–5.10.1039/C7OB00249ASearch in Google Scholar

14. Kiyota H, Oritani T, Kuwahara S. Synthesis and biological evaluation of abscisic acid, jasmonic acid, and its analogs. In: Clark J, Ohkawa H, editors. Discoveries in agrochemicals. ACS symposium series, Vol. 892. Washington, DC: American Chemical Society, 2004:246–54.Search in Google Scholar

15. Lauchli R, Boland W. Indanoyl amino acid conjugates: tunable elicitors of plant secondary metabolism. Chem Rec 2003;3:12–21.10.1002/tcr.10043Search in Google Scholar PubMed

16. Blechert S, Bockelmann C, Brümmer O, Füßlein M, Gundlach H, Haider G, et al. Structural separation of biological activities of jasmonates and related compounds. J Chem Soc Perkin Trans 1 1997:3549–59.10.1039/a702494kSearch in Google Scholar

17. Heinrich CF, Widemann E, Sanz J, Lugan R, Heitz T, Pinot F, et al. A route for the total synthesis of enantiomerically enriched jasmonates 12-COOH-JA and 12-COOH-JA-Ile. Eur J Org Chem 2015;2015:1130–6.10.1002/ejoc.201403347Search in Google Scholar

18. Nakamura Y, Miyatake R, Matsubara A, Kiyota H, Ueda M. Enantio-differential approach to identify the target cell for glucosyl jasmonate-type leaf-closing factor, by using fluorescence-labeled probe compounds. Tetrahedron 2006;62:8805–13.10.1016/j.tet.2006.06.092Search in Google Scholar

19. Wasternack C. Action of jasmonates in plant stress responses and development–applied aspects. Biotechnol Adv 2014;32:31–9.10.1016/j.biotechadv.2013.09.009Search in Google Scholar PubMed

20. Yu M, Schrock RR, Hoveyda AH. Catalyst-controlled stereoselective olefin metathesis as a principal strategy in multistep synthesis design: a concise route to (+)-neopeltolide. Angew Chem Int Ed 2015;54:215–20.10.1002/anie.201409120Search in Google Scholar PubMed PubMed Central

21. Meek SJ, O’Brien RV, Llaveria J, Schrock RR, Hoveyda AH. Catalytic Z-selective olefin cross-metathesis for natural product synthesis. Nature 2011;471:461–6.10.1038/nature09957Search in Google Scholar PubMed PubMed Central

22. Herbert MB, Grubbs RH. Z-selective cross metathesis with ruthenium catalysts: synthetic applications and mechanistic implications. Angew Chem Int Ed 2015;54:5018–24.10.1002/anie.201411588Search in Google Scholar PubMed

23. Keitz BK, Endo K, Patel PR, Herbert MB, Grubbs RH. Improved ruthenium catalysts for Z-selective olefin metathesis. J Am Chem Soc 2012;134:693–9.10.1021/ja210225eSearch in Google Scholar PubMed PubMed Central

24. Vadassery J, Reichelt M, Hause B, Gershenzon J, Boland W, Mithöfer A. CML42-mediated calcium signaling coordinates responses to Spodoptera herbivory and abiotic stresses in Arabidopsis. Plant Physiol 2012;159:1159–75.10.1104/pp.112.198150Search in Google Scholar PubMed PubMed Central

25. Vadassery J, Reichelt M, Jimenez-Aleman GH, Boland W, Mithöfer A. Neomycin inhibition of (+)-7-iso-jasmonoyl-L-isoleucine accumulation and signaling. J Chem Ecol 2014;40:676–86.10.1007/s10886-014-0448-7Search in Google Scholar PubMed

26. Connon SJ, Blechert S. Recent developments in olefin cross-metathesis. Angew Chem Int Ed Engl 2003;42:1900–23.10.1002/anie.200200556Search in Google Scholar PubMed

27. Caggiano L, Castoldi D, Beumer R, Bayón P, Telser J, Gennari C. Effects of allylic and homoallylic substituents on the ring closing metathesis reaction used to synthesise simplified eleuthesides. Tetrahedron Lett 2003;44:7913–9.10.1016/j.tetlet.2003.09.010Search in Google Scholar

28. Keitz BK, Endo K, Herbert MB, Grubbs RH. Z-selective homodimerization of terminal olefins with a ruthenium metathesis catalyst. J Am Chem Soc 2011;133:9686–8.10.1021/ja203488eSearch in Google Scholar PubMed PubMed Central

29. Koh MJ, Khan RK, Torker S, Yu M, Mikus MS, Hoveyda AH. High-value alcohols and higher-oxidation-state compounds by catalytic Z-selective cross-metathesis. Nature 2015;517:181–6.10.1038/nature14061Search in Google Scholar PubMed

30. Widemann E, Grausem B, Renault H, Pineau E, Heinrich C, Lugan R, et al. Sequential oxidation of jasmonoyl-phenylalanine and jasmonoyl-isoleucine by multiple cytochrome P450 of the CYP94 family through newly identified aldehyde intermediates. Phytochemistry 2015;117:388–99.10.1016/j.phytochem.2015.06.027Search in Google Scholar PubMed

31. Kitaoka N, Kawaide H, Amano N, Matsubara T, Nabeta K, Takahashi K, et al. CYP94B3 activity against jasmonic acid amino acid conjugates and the elucidation of 12-O-beta-glucopyranosyl-jasmonoyl-L-isoleucine as an additional metabolite. Phytochemistry 2014;99:6–13.10.1016/j.phytochem.2013.12.019Search in Google Scholar PubMed

32. Hartmann O, Kalesse M. The total synthesis of (−)-Aurafuron A. Org Lett 2012;14:3064–7.10.1021/ol3011387Search in Google Scholar PubMed

33. Hernández D, Nielsen L, Lindsay KB, López García MA, Bjerglund K, Skrydstrup T. Stereocontrolled synthesis of 2-substituted-1,3-azasilaheterocycles. Org Lett 2010;12:3528–31.10.1021/ol101379tSearch in Google Scholar PubMed

34. Konstantinovic S, Predojevic J, Gojkovic S, Pavlovic V. Synthesis of (ω)-unsaturated O-protected glucosides as precursors for unsaturated neutral bolaforms. Indian J Chem 2003;42B:666–9.10.1002/chin.200324176Search in Google Scholar

35. Nakamura Y, Miyatake R, Inomata S, Ueda M. Synthesis and Bioactivity of Potassium β-D-Glucopyranosyl 12-Hydroxy Jasmonate and Related Compounds. Biosci Biotechnol Biochem 2008;72:2867–76.10.1271/bbb.80338Search in Google Scholar PubMed

36. Ritter T, Hejl A, Wenzel AG, Funk TW, Grubbs RH. A standard system of characterization for olefin metathesis catalysts. Organometallics 2006;25:5740–5.10.1021/om060520oSearch in Google Scholar


Supplemental Material:

The online version of this article (DOI: https://doi.org/10.1515/znc-2017-0104) offers supplementary material, available to authorized users.


Received: 2017-6-7
Revised: 2017-6-7
Accepted: 2017-6-8
Published Online: 2017-6-29
Published in Print: 2017-7-14

©2017 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 27.4.2024 from https://www.degruyter.com/document/doi/10.1515/znc-2017-0104/html
Scroll to top button