Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2024 Jul 18;89(15):10451–10461. doi: 10.1021/acs.joc.4c00501

Pd-Catalyzed C(sp2)–H/C(sp2)–H Coupling of Limonene

Marco Di Matteo , Anna Gagliardi , Alexandre Pradal †,*, Luis F Veiros , Fabrice Gallou §, Giovanni Poli †,*
PMCID: PMC11301683  PMID: 39025478

Abstract

graphic file with name jo4c00501_0014.jpg

Limonene undergoes a regioselective Pd(II)-catalyzed C(sp2)–H/C(sp2)–H coupling with acrylic acid esters and amides, α,β-unsaturated ketones, styrenes, and allyl acetate, affording novel 1,3-dienes. DFT computations gave results in accord with the experimental results and allowed for the formulation of a plausible mechanism. The postfunctionalization of one of the coupled products was achieved via a large-scale Sonogashira reaction conducted under micellar catalysis.

Introduction

It is widely agreed that the achievement of sustainability with the stop of global warming in our society must pass through the steady decrease of the use of fossil feedstock.1 On this basis, the gradual replacement of nonrenewable feedstock with biomass appears to be one main measure to take toward this goal.2 While this transition is already in full swing for the industrial production of alternative fuels, it is not systematically applied in fine chemistry.3 Accordingly, a major role of today’s synthetic chemist consists in valorizing molecules made in bulk amounts by nature, or easily available from waste material, and establishing robust protocols to transform them into drop-in4 or dedicated5 biobased molecules6 of interest in fields such as material and crop sciences, flavor, fragrance, or medicinal chemistry. On the other hand, the catalytic functionalization of an intrinsically unactivated C–H bond in a molecule is a key topic in green chemistry, as it complies with atom- as well as step-economy, adding the C–H bond to the catalogue of the ordinary functional groups, such as halides, alcohols, or carbonyls.7

Terpenes are natural hydrocarbons featuring vast structural diversity. Although this class of molecules is considerably less abundant than other primary feedstocks such as lignin, starch, cellulose, proteins, chitin, or triglycerides, they possess a multitude of chemically different C–H bonds.8 Therefore, their use as starting material for catalytic C–H activation/functionalization processes appears to be an ideal, yet challenging, combination toward sustainability in chemical synthesis. As part of a new, long-term project dedicated to the catalytic C–H/C–H functionalization of terpenes, we decided to first focus our attention on the palladium-catalyzed C–H/C–H cross-coupling of (+)-limonene (Scheme 1).

Scheme 1. General Project: Discovery of Catalytic C–H/C–H Cross-Coupling Reactions Starting from Monoterpenes.

Scheme 1

Available directly as a byproduct of the citrus juice industry or produced biotechnologically by yeasts, limonene is the cheapest monoterpene. With two unsaturations, five allylic C–H bond types (11 C–H bonds in total), and two vinylic C–H bond types (3 C–H bonds in total), it appears a perfect substrate for the development of new C–H activation/functionalization protocols. In this article, we show our results on the Pd-catalyzed regioselective direct C(sp2)–H/C(sp2)–H (dehydrogenative) cross-coupling913 between (+)-limonene and alkene partners. To the best of our knowledge, no such cross-coupling has been reported so far, and only two previous C–H functionalizations involving limonene have been described to date. Watson’s group14 reported the Pd-catalyzed C–H borylation of limonene to give the corresponding vinylborane, which could be engaged in situ in a Suzuki-Miyaura reaction (Scheme 2, eq 1). A second precedent deals with the Rh-catalyzed hydroformylation of limonene, pioneered by Kollár15 in 1990, and improved first by Gusevskaya,16 then by Rieger17 in 2020 (Scheme 2, eq 2).18 In a different context, some years ago the group of Loh pioneered nondirected cross-dehydrogenative couplings (CDCs) of intrinsically unactivated 1,1-disubstituted alkenes. In these highly challenging oxidative Pd(II)-catalyzed couplings, alkene reaction partners that are not intrinsically biased for a directed intramolecular metalation come together through C–H bond breaking (Scheme 2, eqs 3 and 4). Thus, building on Loh’s results, we undertook the study of undirected Pd-catalyzed C–H/C–H couplings of terpenes, starting with limonene.

Scheme 2. (Top) Precedents of Limonene C–H Functionalization and (Bottom) Precedents of Nondirected Alkene/Alkene CDCs.

Scheme 2

Results and Discussion

Choice of the Model Reaction and Optimization

We started our study by reacting (R)-(+)-limonene 1a and ethyl acrylate 2a(19) in a 1-to-2 ratio, applying the same reaction conditions as those reported by Loh in 2013. Gratifyingly, under these reaction conditions, a regioselective C9–H/C3–H (limonene numbering) dehydrogenative coupling between the exocyclic alkene of limonene and the β-position of the acrylate took place, generating the corresponding sorbate derivative 3aa in a promising 28% spectroscopic (1H NMR) yield (Table 1, entry 1). The reaction was not stereoselective, generating an almost inseparable E/Z mixture in a 69:31 ratio. Optimization of the reaction conditions was then undertaken. Among the ethereal solvents (1,4-dioxane, THF, diglyme, and 2-methyltetrahydrofuran), THF turned out to be the best compromise in terms of conversion, yield, and diastereomeric ratio (Table 1, entries 2–5).

Table 1. Optimization of the Model C–H-to-C–C Coupling of Limonenea.

graphic file with name jo4c00501_0013.jpg

entry X n solvent yieldb E/Z
1 OAc 2.0 DCE 28 69:31
2 OAc 2.0 dioxane 37 67:33
3 OAc 2.0 THF 36 68:32
4 OAc 2.0 diglyme c  
5 OAc 2.0 2-Me-THF 34 66:34
6 OAc 2.0 DMSO <5 50:50
7 OAc 2.0 AcOH c  
8 OAc 2.0 HFIP <5 82:18
9 OAc 2.0 DMF 12 51:49
10 OAc 2.0 NMP 9 66:34
11 OAc 2.0 DMSO/PivOH 1:1 5 57:43
12 OCOCF3 2.0 THF 6 64:36
13 PdCl2(ACN)2 2.0 THF 33 65:35
14 PEPPSI-IPrd 2.0 THF 19 62:38
15 OPiv 2.0 THF 41 62:38
16 OPiv 4.0 THF 85 75:25
a

Reaction conditions: (+)-limonene (0.5 mmol), ethyl acrylate (2.0 equiv), PdX2 (15 mol %), solvent (0.2 M), and 110 °C in a sealed vial.

b

Measured by quantitative 1H NMR, using 1,4-dinitrobenzene as internal standard.

c

No conversion was observed.

d

[1,3-Bis(2,6-Diisopropylphenyl)imidazol-2-ylidene](3-chloropyridyl)palladium(II) dichloride.

Switch to other dipolar aprotic or protic solvents such as DMSO, AcOH, HFIP, DMF, NMP, or a 1:1 v/v mixture of DMSO/PivOH resulted in either low yields or no conversion (Table 1, entries 6–11). The use of Pd(OCOCF3)2, PdCl2(CH3CN)2, PEPPSI-IPr, and Pd(OPiv)2 as palladium sources (Table 1, entries 12–15) in THF gave the final product in 6, 33, 19, and 41% yield, respectively. In summary, Pd(OAc)2 and Pd(OPiv)2 (Table 1, entries 3 and 15) in THF turned out to be the best and almost equally performing catalysts. However, due to the variable purity of commercial Pd(OAc)2,20 we decided to stick to Pd(OPiv)2 for the completion of the study. Finally, keeping the reaction conditions of entry 15 and increasing the amount of AgOAc21 to 4.0 equiv generated the conjugated dienoate 3aa in 85% yield and 75:25 4E/4Z ratio (Table 1, entry 16). Unfortunately, it was not possible to reduce the catalytic loading of the palladium catalyst22 nor to switch to greener solvents.23

Study of the Selectivity

The coupling generates only two of the four possible geometrical isomers: A (2E,4E), B (2Z,4E), C (2E,4Z), and D (2Z,4Z) (Figure 1, top). Since the vicinal coupling constant J(Hα/Hβ) in the 1H NMR spectrum of the two observed isomers is 11.5 Hz, we can rule out isomers B and D. Furthermore, a NOESY 2D NMR spectrum of an analytically pure sample of the major isomer showed a correlation between the methyl group on C8 (limonene numbering) and Hβ (Figure 1, bottom). This allowed to assign the configuration 2E,4E to the major isomer A and the 2E,4Z to the minor isomer C.

Figure 1.

Figure 1

Top: the four possible geometrical isomers of the coupling reaction between 1 and 2. Bottom: selected region of the NOESY 2D NMR spectrum of major isomer A.

Scope of the C–H/C–H Coupling

After the model reaction was optimized, the scope of this dehydrogenative coupling was evaluated by testing other alkenes as reaction partners for limonene (Scheme 3). Reacting limonene with ethyl, methyl, t-butyl, benzyl, and n-butyl acrylates under the previously optimized reaction conditions gave the expected corresponding dienoates 3aa–3ae in isolated yields ranging from 61 to 85%.24 We then considered the influence of substitution of the partner alkene. Unfortunately, no conversion was observed with benzyl methacrylate 2z1, while only degradation products were generated with benzyl crotonate 2z2. Thus, the coupling appears to be incompatible with substitutions at the α- and β-positions of the electron-poor alkenes. More complex acrylate esters such as menthyl and geranyl acrylates gave the corresponding coupling products 3af and 3ag, although in low yield (Scheme 3, part b).25

Scheme 3. Substrate Scope.

Scheme 3

All reactions were performed on 0.5 mmol scale. Unless otherwise stated, percentage represent isolated yields. Reaction conditions: (+)-limonene (0.5 mmol), electron-poor alkene (2.0 equiv), Pd(OPiv)2 (15 mol %), AgOAc (4.0 equiv), THF (0.2 M), and 110 °C (reflux).

Yield measured by quantitative 1H NMR using 1,4-dinitrobenzene as internal standard.

Tertiary α,β-unsaturated amides were also good coupling partners, giving the expected coupling products 3ah, 3ai, and 3aj in 53–61% isolated yield. These coupling partners appear to be somewhat less efficient, yet, slightly more stereoselective than acrylates.26 Methyl vinyl ketone and ethyl vinyl ketone reacted too, giving the corresponding conjugated dienes in 21%27 and 54% isolated yields, respectively. Cyclohex-2-enone gave no conversion, confirming the incompatibility of this protocol with the β-substitution of the α,β-unsaturated alkene (Scheme 3, part c).28 A couple of non-electron-poor alkenes were also tested as secondary coupling partners. Accordingly, styrene reacted regioselectively at the less substituted alkene position, affording a 71:29 4E/4Z mixture of the coupled product 3am in 53% yield, while allyl acetate gave a 73:27 4E/4Z mixture of the dienyl allylic acetate 3an in 46% yield (Scheme 3, part d). Keeping ethyl acrylate as the electron-poor coupling partner, the coupling was also tested with other terpenes or terpenoids carrying an isopropenyl moiety besides an endocyclic alkene functional group. Accordingly, carvone, perillyl acetate, valencene, β-pinene, and isopulegol acetate all reacted exclusively at the exocyclic alkene, providing the corresponding coupling products in 67, 60, 56, 55, and 29% yield (Scheme 3, part e).

Check for Potential Racemization/Epimerization

To confirm that the reaction conditions are not affecting the stereogenic allylic center of limonene, we tested and compared the couplings of (R)- and (S)-limonene with that of (S)-N-methyl-N-(1-phenylethyl)acrylamide 2o. We reasoned that in the absence of racemization each coupling should afford a couple of enantiopure EE/EZ products 3ao and 3(ent)ao that are different (diastereoisomeric relation) from each other. Contrarily, the four diastereoisomers (as E/Z couples) should be produced from each experiment, should the racemization of limonene (or the epimerization of the coupled product) take place. The two couplings were successful, affording products 3ao and 3(ent)ao in 59:41 and 57:43 E/Z ratio, and 38 and 43% yield, respectively (Scheme 4). A careful comparison of the 13C spectra of the products of the two coupling reactions (each as an E/Z mixture) showed that they are not fully superimposable. Hence, we confirmed that this coupling does not affect the stereochemical integrity of the allylic stereocenter of limonene.

Scheme 4. Top: Coupling between (R)- or (S)-Limonene, and (S)-N-Methyl-N-(1-phenylethyl)acrylamide.

Scheme 4

Conditions: Pd(OPiv)2 (15 mol %), AgOAc (4.0 equiv), THF, reflux, 24 h. Bottom: Comparison of the 13C-NMR spectra of the products from the coupling between 1 and 2o, and between ent-1 and 2o.

Study of the Reaction Mechanism

The coupling mechanism between limonene and methyl acrylate to give the sorbate derivative 3ab was studied via DFT calculations.29 As four geometrical isomers can in principle be generated, two of which are experimentally observed, we computationally studied all four pathways (the complete profiles are shown in the Supporting Information). The monomeric complex [Pd(II)(bis-κ2(OAc)2] I was considered as the starting active catalyst (ΔG = 0 kcal/mol as a reference) (Scheme 5, part A). Formation of the aggregate between [Pd(II)(bis-κ2(OAc)2] and the exocyclic alkene of limonene is the first step of the catalytic cycle.30 This endergonic step already biases the geometry of the trisubstituted alkene in the final product. Indeed, the two located complexes II4E (3.7 kcal/mol, black path) and II4Z (5.4 kcal/mol, red path) are already en route toward the E and Z trisubstituted alkenes, respectively. As to the first path, following full π coordination of the exocyclic alkene of limonene to palladium generates intermediate III4EG = −4.1 kcal/mol), passing an energetic barrier of 8.8 kcal/mol. In this step, one acetate ligand passes from κ2 to κ1-coordination, to allow the π coordination by the alkene. The next step is concerted metalation deprotonation (CMD) wherein the κ1-bound acetate ligand intramolecularly deprotonates the terminal H atom of exocyclic alkene trans to the C6 ring. The step is endergonic, passing an energetic barrier of 16.0 kcal/mol and generating the σ-vinylpalladium complex IV4E (2.8 kcal/mol) with the newly generated acetic acid molecule still coordinated. Worthy of note is the metal assistance (dPd–H = 2.30 Å) in the CMD transition state TSIII4E→IV4E (see Figure S1 in the Supporting Information). The step is stereoretentive, in that the palladium atom replaces the position of the deprotonated vinylic H atom. Follows a two-step dissociative, globally exergonic ligand exchange between the exiting acetic acid and the entering acrylate, which generates complex VII4E (−4.8 kcal/mol, not shown) through a barrier of 15.5 kcal/mol. The subsequent carbopalladation is very exergonic, generating complex VIII4E (−26.5 kcal/mol), the resting state of the catalytic cycle, via a barrier of 8.6 kcal/mol.

Scheme 5. Part A: First Part of the Free Energy Profile of the Pd(OAc)2-Catalyzed/AgOAc-Mediated Coupling between Limonene and Methyl Acrylate; From the Starting Complex to the Carbopalladation Step.

Scheme 5

Part B: second part of the energy profile of the Pd(OAc)2 catalyzed/AgOAc-mediated coupling between limonene and methyl acrylate; from the carbopalladation complex VIII4E to product 2E,4E-3ab.

The higher energy aggregate II4Z follows an analogous [π-complex formation/CMD/AcOH-acrylate ligand exchange/carbopalladation] sequence of steps as the one starting from II4E, leading to intermediate VIII4Z (−27.9 kcal/mol, red path). The following step for the two paths (black and red) is a dehydropalladation (β-H elimination) (Scheme 5, part B). This globally endergonic two-step transformation lies at the second bifurcation of the mechanism and determines which of the two diastereotopic allylic H atoms at C3 is involved in the agostic interaction, which is in turn associated with the geometry of the 1,2-disubstituted alkene in the product. In one of these two paths (brown path), intermediate VIII4E generates the hydride complex X2E4E (−14.1 kcal/mol) stepping over a barrier of 24.8 kcal/mol and passing through the “agostic” intermediate IX2E4E (−11.9 kcal/mol). The subsequent reductive elimination is exergonic, affording the Pd(0) complex π-coordinated with the final 2E,4E product and σ-coordinated with acetic acid XI2E4E (−24.2 kcal/mol), passing a barrier of 5.2 kcal/mol. Then, the final products 2E,4E-3ab and acetic acid are released from the complex through the entry into the coordination sphere of two solvent molecules that stabilize the Pd atom. This step is endergonic, reaching +1.2 kcal/mol. The alternative path from intermediate VIII4E (orange path) follows an analogous [dehydropalladation/reductive elimination/product release] sequence, affording the hydride complex X2Z4E (−11.8 kcal/mol), passing through the “agostic” intermediate IX2Z4E (−9.8 kcal/mol), while the subsequent reductive elimination gives XI2Z4E (−21.0 kcal/mol), and the product release step generates 2Z,4E-3ab, Pd(THF)2, and AcOH, rising the energy to +3.5 kcal/mol. Inspection of this second part of the mechanism reveals that the brown path lies consistently at much lower energies than the orange path.

It is important to notice that the barrier for going backward from complexes XI2E4E and XI2Z4E to common resting state VIII4E is lower than the one for going forward toward product release (Scheme 5, part B), and this is also true for the corresponding sequence of steps starting from minor resting state VIII4Z (see the Supporting Information).

This scenario implies that equilibration between XI2E4E and XI2Z4E takes place before product release. Accordingly, the 2E/2Z ratios in the products are expected to depend on the energy differences between XI2E4E and XI2Z4E for the major 4E path, and on that between XI2E4Z and XI2Z4Z for the minor 4Z path. The computed ΔGs highly favor the 2E stereochemistry in both cases (ΔG4E = 3.2 kcal/mol; ΔG4Z = 5.0 kcal/mol), which agrees with the fact that only the 2E geometrical isomers are experimentally observed (Figure 2).

Figure 2.

Figure 2

Free energy values of the Pd(0) complexes π-coordinated with the final products and σ-coordinated with acetic acid. Left side: the 4E couple (ΔG4E = 3.2 kcal/mol); right side: the 4Z couple (ΔG4Z = 5.0 kcal/mol).

According to the above results, the catalytic cycle is experimentally viable only for the two paths that at the second branching lead to the 2E configurated products. To further support the above-mentioned equilibration between the intermediates XI2E and XI2Z, a classical Mizoroki-Heck coupling between ethyl acrylate and phenyl iodide was carried out [Pd(OAc)2 (10 mol %), dppe (20 mol %), nBu4NOAc (2.0 equiv), DMF, 120 °C, 2 h] in the presence of Z-methyl cinnamate (Scheme 6). Besides the formation of the expected E-ethyl cinnamate, this corollary experiment restituted an 82:18 Z/E mixture of methyl cinnamate. Visibly, the Pd(0)/AcOH system was responsible for the isomerization of Z-cinnamate (see Scheme S3 in the Supporting Information for the mechanistic detail), which is in accordance with the reversibility of the dehydropalladation/reductive elimination sequence.31

Scheme 6. Corollary Experiment to Prove the Reversibility of the Dehydropalladation/Reductive Elimination Sequence.

Scheme 6

On the Palladium Reoxidation Step

Coming back to the computational study, the last step of the catalytic cycle is the silver-mediated oxidation of the released Pd(THF)2 (Scheme 7).32,33 This state is generated after an endergonic product release that reaches 1.2 and 3.2 kcal/mol through a 25.4 and 27.4 kcal/mol rise from the isoenergetic intermediates XI2E4E and XI2E4Z, respectively. In this redox step, Ag(I) is reduced to Ag(0), whose actual state is undetermined. Consideration of the simplest aggregate Ag2,34 according to the equation [Pd(0)THF2 + 2 AgOAc-THF → Pd(OAc)2 + Ag2 + 4 THF] turns out to be endergonic (ΔrG° = +2.5 kcal/mol). However, the thermodynamics becomes slightly favorable (ΔrG° = −11.1 kcal/mol) when the Ag4 (D2h) aggregate is considered as the silver reduced product, according to the equation [2 Pd(0)-THF2 + 4 AgOAc-THF → 2 Pd(OAc)2 + Ag4 + 8 THF], and clearly favorable (ΔrG° = −84.6 kcal/mol) when the Ag8 (D2h, dodecahedron)34d,35 aggregate is considered as the silver reduced product, according to the equation [4 Pd(0)-THF2 + 8 AgOAc-THF → 4 Pd(OAc)2 + Ag8 + 16 THF]. Thus, using the latter stoichiometry, the reaction free energy changes (ΔrG°) of the catalytic cycles relative to the formation of 2E,4E-3ab as well as 2E,4Z-3ab become −20.0 and −18.0 kcal/mol, respectively. Therefore, assuming that the states relative to Pd(THF)2 are very close in energy and geometry to the corresponding transition states leading to and departing from it, we can predict energetic spans36 close to 27.7 kcal/mol [+1.2 – (−26.5)] for the 2E,4E isomer and to 31.1 kcal/mol [(+3.2 – (−27.9)] for the 2E,4Z isomer. Here again, the computed energy values are qualitatively in accord with the observed 2E,4E (major)/2E,4Z (minor) isomer ratio of the coupled products and also with the experimental conditions (24 h of THF reflux).

Scheme 7. Final Part of the Free Energy Profile of the Pd(OAc)2-Catalyzed/AgOAc-Mediated Coupling between Limonene and Methyl Acrylate: From Intermediates XI2E4E and XI2E4Z to Pd(OAc)2 Regeneration.

Scheme 7

However, the aggregation state of metallic silver is not the only factor affecting the thermodynamics of the palladium reoxidation step. Indeed, the fact that four equivalents of AcOAg are necessary to obtain optimal yields of couplings suggests that on a macroscopic ground, the excess of oxidant37 has the role of pushing the equilibrium toward the product side, thus securing the regeneration of adequate amounts of Pd(OAc)2. Therefore, while the E/Z isomeric ratio at the trisubstituted C4=C5 double bond is kinetic in origin, resulting from a different energy span, that at C2=C3 results (for each isomer at C4=C5) from a local thermodynamic control due to product equilibration before its release from the corresponding Pd(0) complex precursor. According to the above DFT computations, the mechanism of the full catalytic cycle relative to the coupling between limonene and methyl acrylate to give the major isomer 2E,4E-3ab is proposed in Scheme 8.

Scheme 8. Proposed Mechanism for the Pd-Catalyzed Coupling between Limonene and Methyl Acrylate to Give the Major Product 2E,4E-3ab.

Scheme 8

Coupling with Allyl Acetate

The mechanism of the coupling between limonene and allyl acetate to give 3an deserves particular comment.38 Indeed, the intermediate after the carbopalladation step undergoes dehydropalladation, although deacetoxypalladation could have been an alternative plausible path.39 We speculate that the allylic nature of the hydrogen atom involved in the dehydropalladation, which in turn generates a conjugated 1,3-diene, is probably the reason for this behavior (Scheme 9).

Scheme 9. Observed (Top Right) and Nonobserved (Bottom Right) Pathways after the Carbopalladation Step in the Coupling between Limonene and Allyl Acetate.

Scheme 9

In view of the nature of our mixed academic/industrial collaborative project, and to further demonstrate the usefulness of this C–H-activation-based, catalytic decoration of limonene, we decided to test a postfunctionalization protocol based on micellar catalysis.40 This greener type of catalysis may become possible thanks to the presence of surfactants, polymersomes, dendrimers, or nanogels that self-assemble in supramolecular architectures that behave as nanoreactors. In this context, we have recently studied Pd(0) nanoparticles embedded in core–shell nanogels as recoverable catalysts for the Mizoroki–Heck reaction,41 and we and others have reported a series of articles demonstrating the power of surfactant-promoted, transition-metal-catalyzed chemistry in water at room temperature with possible catalyst recovery.42 Accordingly, we decided to perform the coupling between limonene and p-bromostyrene, and, if successful, to test the resulting product in a Sonogashira coupling43 using micellar conditions. The planned C(sp2)–H/C(sp2)–H coupling, using the optimized conditions in THF as the solvent, worked as expected, giving bromodiene 3ap in 34% yield. The Sonogashira coupling was next tackled. After some experimentation (see Supporting Information for the optimization), we found that treatment of 3ao (1.0 mmol scale) with ethynylbenzene in the presence of the catalytic system [CataCXium-A-Pd-G3 (0.20 mol %), NEt3 (3.0 equiv), TPGS-750-M (2 wt)/H2O, glucose (5 mol %)], and THF (15%) as the cosolvent of choice for 24 h at 45 °C gave the desired coupling product 4 in 76% isolated yield (Scheme 10). Notably, these conditions are compared favorably with the classical Sonogashira protocol.44 The demonstrated sequence allows for a rapid further elaboration of a valorized limonene platform.

Scheme 10. Sonogashira Coupling under Micellar Catalysis from a C–H/C–H Coupling Product, as Postfunctionalization Example.

Scheme 10

Conclusions

In conclusion, we have developed the first vinyl/vinyl coupling of limonene with several unsaturated partners such as acrylic esters and amides, α,β-unsaturated ketones, styrenes, and allyl acetate. The reaction turned out to be totally regioselective for both reaction partners, involving exclusively the exocyclic unsaturation of limonene without affecting the integrity of the absolute stereochemistry of the C4 stereogenic center of limonene. As to the stereoselectivity, the reaction constantly produced a mixture of the two geometrical isomers 2E,4E and 2E,4Z, out of the four possible, with the former always prevailing. On the other hand, the 2Z isomers were never detected. In addition, terpenes other than limonene, possessing an isopropenyl function in addition to an endocyclic alkene, such as carvone, perillyl acetate, and valencene, could also be satisfactorily coupled. DFT computation gave results in qualitative accord with the observed selectivity, allowing the proposal of a plausible mechanism for the coupling.

To further valorize the method developed, a large-scale Sonogashira reaction using a C–H/C–H coupled product as a starting substrate was successfully performed under micellar catalysis conditions, allowing its further selective functionalization. Future work will focus on the development of this C(sp2)–H/C(sp2)–H coupling in water, on other selective postfunctionalizations, and on new selective catalytic modifications of terpenes.

Acknowledgments

This work is dedicated to Professor Marta Catellani for her outstanding contribution to palladium chemistry. We warmly thank Dr. Irène Arrata, Project Manager of the European ITN consortium CHAIR https://chair-itn.eu/ for her endless support and encouragement. This project has received funding from the European Union’s Horizon 2020 research and innovation program under the Marie Skłodowska-Curie Grant Agreement No 860762. Centro de Química Estrutural (CQE) and Institute of Molecular Sciences (IMS) acknowledge the financial support of Fundação para a Ciência e Tecnologia (Projects UIDB/00100/2020, UIDP/00100/2020, and LA/P/0056/2020, respectively).

Data Availability Statement

The data underlying this study are available in the published article, in its Supporting Information, and openly available in Zenodo (10.5281/zenodo.10462156).

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.joc.4c00501.

  • Detailed optimization data, experimental procedures, characterization data, copies of NMR spectra of all new compounds, and DFT computations with coordinates (PDF)

Author Present Address

Institut Lavoisier de Versailles (ILV), Université de Versailles Saint-Quentin-en-Yvelines, Université Paris-Saclay, CNRS, UMR 8180, 45 avenue des États-Unis, 78035 Versailles, France

Author Contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. For the specific author contributions, see the Supporting Information.

An earlier version of this manuscript was deposited on a preprint server (10.26434/chemrxiv-2024-nzkq1).45

The authors declare no competing financial interest.

Supplementary Material

jo4c00501_si_001.pdf (5.1MB, pdf)

References

  1. SEI, IISD, ODI, E3G, and UNEP, The Production Gap Report 2021, http://productiongap.org/2021report.
  2. Poluzzi A.; Guandalini G.; D’Amore F.; Romano M. C. The Potential of Power and Biomass-to-X Systems in the Decarbonization Challenge: a Critical Review. Curr. Sustain. Renew. Energy Rep. 2021, 8, 242–252. 10.1007/s40518-021-00191-7. [DOI] [Google Scholar]
  3. a Zhou C. H.; Xia X.; Lin C.-X.; Tong D.-S.; Beltramini J. Catalytic Conversion of Lignocellulosic Biomass to Fine Chemicals and Fuels. Chem. Soc. Rev. 2011, 40, 5588–5617. 10.1039/c1cs15124j. [DOI] [PubMed] [Google Scholar]; b Chen G.; Weselake R. J.; Singer S. D. (Eds.), Plant Bioproducts; Springer New York: New York, NY, 2018. [Google Scholar]
  4. Dutta S. Sustainable Synthesis of Drop-In Chemicals from Biomass via Chemical Catalysis: Scopes, Challenges, and the Way Forward. Energy Fuels 2023, 37, 2648–2666. 10.1021/acs.energyfuels.2c03891. [DOI] [Google Scholar]
  5. Carus M.; Dammer L.; Puente Á.; Raschka A.; Arendt O.. https://renewable-carbon.eu/publications/product/bio-based-drop-in-smart-drop-in-and-dedicated-chemicals-%E2%88%92-full-version/.
  6. Bio-Based Chemicals, A 2020 Update, IEA Bioenergy: Task 42: 2020: 01, https://www.ieabioenergy.com/wp-content/uploads/2020/02/Bio-based-chemicals-a-2020-update-final-200213.pdf.
  7. For review, see:; a Dalton T.; Faber T.; Glorius F. C–H Activation: Toward Sustainability and Applications. ACS Cent. Sci. 2021, 7, 245–261. 10.1021/acscentsci.0c01413. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Rogge T.; Kaplaneris N.; Chatani N.; Kim J.; Chang S.; Punji B.; Schafer L. L.; Musaev D. G.; Wencel-Delord J.; Roberts C. A.; Sarpong R.; Wilson Z. E.; Brimble M. A.; Johansson M. J.; Ackermann L. C–H activation. Nat. Rev. Methods Primer 2021, 43, 1–31. [Google Scholar]; c Roudesly F.; Oble J.; Poli G. Metal-catalyzed C–H Activation/functionalization: The Fundamentals. J. Mol. Catal. Chem. A 2017, 426, 275–296. 10.1016/j.molcata.2016.06.020. [DOI] [Google Scholar]; d de Jesus R.; Hiesinger K.; van Gemmeren M. Preparative Scale Applications of C–H Activation in Medicinal Chemistry. Angew. Chem., Int. Ed. 2023, 62, e202306659 10.1002/anie.202306659. [DOI] [PubMed] [Google Scholar]
  8. a Breitmaier E.. Terpenes: Flavors, Fragrances, Pharmaca, Pheromones, Ed.: Breitmaier E.; Wiley-VCH: Weinheim, 2006. [Google Scholar]; b Brill Z. G.; Condakes M. L.; Ting C. P.; Maimone T. J. Navigating the Chiral Pool in the Total Synthesis of Complex Terpene Natural Products. Chem. Rev. 2017, 117, 11753–11795. 10.1021/acs.chemrev.6b00834. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Winnacker M. Polyamides Derived from Terpenes: Advances in Their Synthesis, Characterization and Applications. Eur. J. Lipid Sci. Technol. 2023, 125, 2300014 10.1002/ejlt.202300014. [DOI] [Google Scholar]
  9. For reviews on C(sp2)–H/C(sp2)–H coupling involving plain alkenes, see:; a Shang X.; Liu Z.-Q. Transition Metal-catalyzed Cvinyl–Cvinyl Bond Formation via Double Cvinyl–H Bond Activation. Chem. Soc. Rev. 2013, 42, 3253–3260. 10.1039/c2cs35445d. [DOI] [PubMed] [Google Scholar]; b Maikhuri V. K.; Maity J.; Srivastavac S.; Prasad A. K. Transition metal-catalyzed double Cvinyl–H bond activation: synthesis of conjugated dienes. Org. Biomol. Chem. 2022, 20, 9522–9588. 10.1039/D2OB01646J. [DOI] [PubMed] [Google Scholar]
  10. For precedents of Pd-catalyzed nondirected C(sp2)–H/C(sp2)–H couplings involving plain alkenes, see:; a da Silva M. J.; Gusevskaya E. V. Palladium-catalyzed oxidation of monoterpenes: novel tandem oxidative coupling–oxidation of camphene by dioxygen. J. Mol. Catal. A: Chem. 2001, 176, 23–27. 10.1016/S1381-1169(01)00240-0. [DOI] [Google Scholar]; b da Silva M. J.; Gonçalves J. A.; Alves R. B.; Howarth O. W.; Gusevskaya E. V. Palladium Catalyzed Transformations of Monoterpenes: Stereoselective Deuteriation and Oxidative Dimerization of Camphene. J. Organomet. Chem. 2004, 689, 302–308. 10.1016/j.jorganchem.2003.10.016. [DOI] [Google Scholar]; c Hatamoto Y.; Sakaguchi S.; Ishii Y. Oxidative Cross-Coupling of Acrylates with Vinyl Carboxylates Catalyzed by a Pd(OAc)2/HPMoV/O2 System. Org. Lett. 2004, 6, 4623–4625. 10.1021/ol047971u. [DOI] [PubMed] [Google Scholar]; d Xu Y. H.; Lu J.; Loh T.-P. Direct Cross-Coupling Reaction of Simple Alkenes with Acrylates Catalyzed by Palladium Catalyst. J. Am. Chem. Soc. 2009, 131, 1372–1373. 10.1021/ja8084548. [DOI] [PubMed] [Google Scholar]; e Xu Y.-H.; Wang W.-J.; Wen Z.-K.; Hartley J. J.; Loh T.-P. Palladium-Catalyzed Direct Cross-coupling Reaction between Indenes and Electron-deficient Alkenes. Tetrahedron Lett. 2010, 51, 3504–3507. 10.1016/j.tetlet.2010.04.097. [DOI] [Google Scholar]; f Yu H.; Jin W.; Sun C.; Chen J.; Du W.; He S.; Yu Z. Palladium-Catalyzed Cross-Coupling of Internal Alkenes with Terminal Alkenes to Functionalized 1,3-Butadienes Using C–H Bond Activation: Efficient Synthesis of Bicyclic Pyridones. Angew. Chem., Int. Ed. 2010, 49, 5792–5797. 10.1002/anie.201002737. [DOI] [PubMed] [Google Scholar]; g Wen Z.-K.; Xu Y.-H.; Loh T.-P. Palladium(II)-catalyzed Cross-coupling of Simple Alkenes with Acrylates: a Direct Approach to 1,3-Dienes through C–H Activation. Chem. Sci. 2013, 4, 4520–4524. 10.1039/c3sc52275j. [DOI] [Google Scholar]; h Wen Z. K.; Xu Y. H.; Loh T.-P. Palladium-Catalyzed Cross-Coupling of Unactivated Alkenes with Acrylates: Application to the Synthesis of the C13–C21 Fragment of Palmerolide A. Chem.—Eur. J. 2012, 18, 13284–13287. 10.1002/chem.201201806. [DOI] [PubMed] [Google Scholar]; i Zhang X.; Wang M.; Zhang M.-X.; Xu Y.-H.; Loh T.-P. Synthesis of Dienyl Ketones via Palladium(II)-Catalyzed Direct Cross-Coupling Reactions between Simple Alkenes and Vinyl Ketones: Application to the Synthesis of Vitamin A1 and Bornelone. Org. Lett. 2013, 15, 5531–5533. 10.1021/ol402692t. [DOI] [PubMed] [Google Scholar]
  11. For precedents of Pd-catalyzed nondirected C(sp2)–H/C(sp2)–H couplings involving heteroatom-enriched alkenes, see:; a Gigant N.; Bäckvall J.-E. Synthesis of Conjugated Dienes via a Biomimetic Aerobic Oxidative Coupling of Two Cvinyl–H Bonds. Chem.—Eur. J. 2013, 19, 10799–10803. 10.1002/chem.201301771. [DOI] [PubMed] [Google Scholar]; b Gigant N.; Gillaizeau I. Palladium(II)-Catalyzed Direct Alkenylation of Nonaromatic Enamides. Org. Lett. 2012, 14, 3304–3307. 10.1021/ol301249n. [DOI] [PubMed] [Google Scholar]; c Chen Y.; Wang F.; Jia A.; Li X. Palladium-Catalyzed Selective Oxidative Olefination and Arylation of 2-Pyridones. Chem. Sci. 2012, 3, 3231–3236. 10.1039/c2sc20869e. [DOI] [Google Scholar]; d Yu Y.-Y.; Georg G. I. Dehydrogenative Alkenylation of Uracils via Palladium-catalyzed Regioselective C–H Activation. Chem. Commun. 2013, 49, 3694–3696. 10.1039/c3cc41130c. [DOI] [PMC free article] [PubMed] [Google Scholar]; e Yu Y.-Y.; Niphakis M. J.; Georg G. I. Palladium(II)-Catalyzed Dehydrogenative Alkenylation of Cyclic Enaminones via the Fujiwara–Moritani Reaction. Org. Lett. 2011, 13, 5932–5935. 10.1021/ol202677g. [DOI] [PMC free article] [PubMed] [Google Scholar]; f Ge H.; Niphakis M. J.; Georg G. I. Palladium(II)-Catalyzed Direct Arylation of Enaminones Using Organotrifluoroborates. J. Am. Chem. Soc. 2008, 130, 3708–3709. 10.1021/ja710221c. [DOI] [PubMed] [Google Scholar]; g Li M.; Li L.; Ge H. Direct C-3-Alkenylation of Quinolones via Palladium-Catalyzed C–H Functionalization. Adv. Synth. Catal. 2010, 352, 2445–2449. 10.1002/adsc.201000364. [DOI] [Google Scholar]; h Kim D.; Hong S. Palladium(II)-Catalyzed Direct Intermolecular Alkenylation of Chromones. Org. Lett. 2011, 13, 4466–4469. 10.1021/ol2018278. [DOI] [PubMed] [Google Scholar]; i Min M.; Kim Y.; Hong S. Regioselective Palladium-catalyzed Olefination of Coumarins via Aerobic Oxidative Heck Reactions. Chem. Commun. 2013, 49, 196–198. 10.1039/C2CC37676H. [DOI] [PubMed] [Google Scholar]; j Moon Y.; Kwon D.; Hong S. Palladium-Catalyzed Dehydrogenation/Oxidative Cross-Coupling Sequence of β-Heteroatom-Substituted Ketones. Angew. Chem., Int. Ed. 2012, 51, 11333–11336. 10.1002/anie.201206610. [DOI] [PubMed] [Google Scholar]
  12. For precedents of Pd-catalyzed directed C(sp2)–H/C(sp2)–H couplings, see:; a Xu Y.-H.; Chok Y.-K.; Loh T.-P. Synthesis and Characterization of a Cyclic Vinylpalladium(II) Complex: Vinylpalladium Species as the Possible Intermediate in the Catalytic Direct Olefination Reaction of Enamide. Chem. Sci. 2011, 2, 1822–1825. 10.1039/c1sc00262g. [DOI] [Google Scholar]; b Zhao Q.; Tognetti V.; Joubert L.; Besset T.; Pannecoucke X.; Bouillon J.-P.; Poisson T. Palladium-Catalyzed Synthesis of 3-Trifluoromethyl-Substituted 1,3-Butadienes by Means of Directed C–H Bond Functionalization. Org. Lett. 2017, 19, 2106–2109. 10.1021/acs.orglett.7b00704. [DOI] [PubMed] [Google Scholar]; c Liu M.; Yang P.; Karunananda M. K.; Wang Y.; Liu P.; Engle K. M. C(alkenyl)–H Activation via Six-Membered Palladacycles: Catalytic 1,3-Diene Synthesis. J. Am. Chem. Soc. 2018, 140, 5805–5813. 10.1021/jacs.8b02124. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Schreib B. S.; Son M.; Aouane F. A.; Baik M.-H.; Carreira E. M. Allene C(sp2)–H Activation and Alkenylation Catalyzed by Palladium. J. Am. Chem. Soc. 2021, 143, 21705–21712. 10.1021/jacs.1c11528. [DOI] [PubMed] [Google Scholar]; e Liang Q.-J.; Yang C.; Meng F.-F.; Jiang B.; Xu Y.-H.; Loh T.-P. Chelation versus Non-Chelation Control in the Stereoselective Alkenyl sp2 C–H Bond Functionalization Reaction. Angew. Chem., Int. Ed. 2017, 56, 5091–5095. 10.1002/anie.201700559. [DOI] [PubMed] [Google Scholar]; f Meng K.; Li T.; Yu C.; Shen C.; Zhang J.; Zhong G. Geminal Group-Directed Olefinic C–H Functionalization via Four- to Eight-membered Exo-metallocycles. Nat. Commun. 2019, 10, 5109. 10.1038/s41467-019-13098-1. [DOI] [PMC free article] [PubMed] [Google Scholar]; g Liu M.; Sun J.; Erbay T. G.; Ni H.-Q.; Martín-Montero R.; Liu P.; Engle K. M. Pd(II)-Catalyzed C(alkenyl)–H Activation Facilitated by a Transient Directing Group. Angew. Chem., Int. Ed. 2022, 61, e202203624 10.1002/anie.202203624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. For other Rh- or Ru-catalyzed directed C(sp2)–H/C(sp2)–H couplings, see:; a Besset T.; Kuhl N.; Patureau F. W.; Glorius F. Rh(III)-Catalyzed Oxidative Olefination of Vinylic C–H Bonds: Efficient and Selective Access to Di-unsaturated α-Amino Acid Derivatives and Other Linear 1,3-Butadienes. Chem.—Eur. J. 2011, 17, 7167–7171. 10.1002/chem.201101340. [DOI] [PubMed] [Google Scholar]; b Boultadakis-Arapinis M.; Hopkinson M. N.; Glorius F. Using Rh(III)-Catalyzed C–H Activation as a Tool for the Selective Functionalization of Ketone-Containing Molecules. Org. Lett. 2014, 16, 1630–1633. 10.1021/ol500258q. [DOI] [PubMed] [Google Scholar]; c Hu X.-H.; Zhang J.; Yang X.-F.; Xu Y.-H.; Loh T.-P. Stereo- and Chemoselective Cross-Coupling between Two Electron-Deficient Acrylates: An Efficient Route to (Z,E)-Muconate Derivatives. J. Am. Chem. Soc. 2015, 137, 3169–3172. 10.1021/ja512237d. [DOI] [PubMed] [Google Scholar]; d Feng R.; Yu W.; Wang K.; Liu Z.; Zhang Y. Ester-Directed Selective Olefination of Acrylates by Rhodium Catalysis. Adv. Synth. Catal. 2014, 356, 1501–1508. 10.1002/adsc.201300944. [DOI] [Google Scholar]; e Li F.; Yu C.; Zhang J.; Zhong G. Weinreb Amide Directed Cross-coupling Reaction between Electron-deficient Alkenes Catalyzed by a Rhodium Catalyst. Org. Biomol. Chem. 2017, 15, 1236–1244. 10.1039/C7OB00026J. [DOI] [PubMed] [Google Scholar]; f Logeswaran R.; Jeganmohan M. Rhodium(III)-Catalyzed Aerobic Oxidative C–H Olefination of Unsaturated Acrylamides with Unactivated Olefins. Org. Lett. 2021, 23, 767–771. 10.1021/acs.orglett.0c03981. [DOI] [PubMed] [Google Scholar]; g Zhanga J.; Loh T.-P. Ruthenium- and Rhodium-catalyzed Cross-coupling Reaction of Acrylamides with Alkenes: Efficient Access to (Z,E)-Dienamides. Chem. Commun. 2012, 48, 11232–11234. 10.1039/C2CC36137J. [DOI] [PubMed] [Google Scholar]; h Hu X.-H.; Yang X.-F.; Loh T.-P. Selective Alkenylation and Hydroalkenylation of Enol Phosphates through Direct C–H Functionalization. Angew. Chem., Int. Ed. 2015, 54, 15535–15539. 10.1002/anie.201506437. [DOI] [PubMed] [Google Scholar]
  14. Watson D. A.; Reid W. B. Synthesis of Trisubstituted Alkenyl Boronic Esters from Alkenes Using the Boryl-Heck Reaction. Org. Lett. 2018, 20, 6832–6835. 10.1021/acs.orglett.8b02949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Kollár L.; Bakos J.; Heil B.; Sándor P.; Szalontai G. J. Hydroformylation of Chiral Terpenes with PtCl(SnCl3)-(bis-phosphine) as Catalyst. Organomet. Chem. 1990, 385, 147–152. 10.1016/0022-328X(90)87153-5. [DOI] [Google Scholar]
  16. Vieira C. G.; de Freitas M. C.; dos Santos E. N.; Gusevskaya E. V. Synthesis of Fragrance Ingredients by Tandem Hydroformylation-Cyclization of Limonene Catalyzed by Rhodium Complexes and Pyridinium p-Toluenesulphonate. ChemCatChem. 2012, 4, 795–801. 10.1002/cctc.201200057. [DOI] [Google Scholar]
  17. Causero A.; Troll C.; Rieger B. (+)-Limonene Functionalization: Syntheses, Optimization, and Scale-up Procedures for Sustainable Polymer Building Blocks. Ind. Eng. Chem. Res. 2020, 59, 15464–15477. 10.1021/acs.iecr.0c02998. [DOI] [Google Scholar]
  18. The complex n-BuLi/TMEDA is known to metalate limonene at C-10:; Crawford R. J.; Erman W. F.; Broaddus C. D. Metalation of Limonene. A Novel Method for the Synthesis of Bisabolane Sesquiterpenes. J. Am. Chem. Soc. 1972, 94, 4298–4306. 10.1021/ja00767a044. [DOI] [Google Scholar]
  19. Acrylic acid, and thus acrylates, can also be biobased, resulting from the dehydration of lactic acid or 3-hydroxypropionic acid, or from dehydration/oxidation of glycerol. See:; Brobbey M. S.; Louw J.; Görgens J. F. Biobased Acrylic Acid Production in a Sugarcane Biorefinery: A Techno-economic Assessment Using Lactic Acid, 3-Hydroxypropionic Acid and Glycerol as Intermediates. Chem. Eng. Res. Des. 2023, 193, 367–382. 10.1016/j.cherd.2023.03.034. [DOI] [Google Scholar]
  20. a Carole W. A.; Colacot T. J. Understanding Palladium Acetate from a User Perspective. Chem.—Eur. J. 2016, 22, 7686–7695. 10.1002/chem.201601450. [DOI] [PubMed] [Google Scholar]; b Carole W. A.; Colacot T. J. Unravelling the Mystery of Palladium Acetate. Investigations of high purity palladium acetate and its two common impurities. Chim. Oggi-Chem. Today 2017, 35, 64–65. [Google Scholar]
  21. For other oxidizing agents tested, see the Supporting Information.
  22. Other terminal oxidants were also considered for this coupling (Cu(OAc)2, CuCl2, AgCO3, Ag2O, DMBQ, PIDA, BQ, Oxone®, H2O2, t-BuO2H, AgNO3 and K2S2O8). However, none of these oxidants performed better than AgOAc.
  23. Experiments of the model coupling reaction using 2-MeTHF, dimethyl isosorbide, or Cyrene® as solvent gave unsatisfactory yields (see Supporting Information).
  24. Preparation of compound 3ab was also performed on a 16.2 mmol scale providing the desired product with 47% isolated yield and a d.r. of 69:31.
  25. The low yield of 3af and 3ag was due to difficulties to separate the product from the starting acrylate during purification.
  26. We only considered tertiary acrylamides, as primary and secondary acrylamides are expected to be incompatible with this coupling.
  27. The low yield of this product is mainly due its high volatility.
  28. Other electron-poor alkene partners such as acrylonitrile, phenylvinylsulfoxide, vinyl diethylphosphonate or 5,6-dihydro-2H-pyran-2-one did not provide the desired coupling product. No conversion was observed with acrylonitrile and phenylvinylsulfoxide. Full conversion but very low yields (<5%) were obtained with vinyl diethylphosphonate or 5,6-dihydro-2H-pyran-2-one.
  29. a Parr R. G.; Yang W.. Density Functional Theory of Atoms and Molecules; Oxford University Press: New York, 1989. [Google Scholar]; (b) Calculations performed at the PBE0-D3/(SDD*,6-311++G(d,p)//PBE0/(SDD*,6-31G(d,p)) level using the GAUSSIAN 09 package. Solvent effects (THF) were considered using the PCM/SMD model. A full account of the computational details and a complete list of references are provided as Supporting Information.
  30. Before performing the DFT studies, three mechanistic hypotheses relative to the first palladation event were initially considered (see Scheme S2).
  31. Sommer H.; Juliá-Hernández F.; Martin R.; Marek I. Walking Metals for Remote Functionalization. ACS Cent. Sci. 2018, 4, 153–165. 10.1021/acscentsci.8b00005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Some studies on Pd-catalysis in the presence of silver salts invoke the intervention of bimetallic Pd/Ag species (see next reference). However, a coupling reaction carried out between limonene and ethyl acrylate using a stoichiometric amount of Pd(OAc)2 and in the absence of AgOAc, gave the expected product 3aa in 50% isolated yield (71:29 E/Z). This result suggests that the silver salt in this coupling is only responsible for the Pd(0) reoxidation step.
  33. a de Carvalho R. L.; Diogo E. B. T.; Homölle S. L.; Dana S.; da Silva E. N. Jr; Ackermann L. The Crucial Role of Silver(I)-salts as Additives in C–H Activation Reactions: Overall Analysis of Their Versatility and Applicability. Chem. Soc. Rev. 2023, 52, 6359–6378. 10.1039/D3CS00328K. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Pérez-Ortega I.; Albéniz A. C. Multifaceted Role of Silver Salts as Ligand Scavengers and Different Behavior of Nickel and Palladium complexes: Beyond Halide Abstraction. Dalton Trans. 2023, 52, 1425–1432. 10.1039/D2DT03948F. [DOI] [PubMed] [Google Scholar]; c Bhattacharya T.; Dutta S.; Maiti D. Deciphering the Role of Silver in Palladium-Catalyzed C–H Functionalizations. ACS Catal. 2021, 11, 9702–9714. 10.1021/acscatal.1c02552. [DOI] [Google Scholar]; d Malakar S.; Schaefer H. F.; Sunoj R. B. Is Silver a Mere Terminal Oxidant in Palladium Catalyzed C–H Bond Activation Reactions?. Chem. Sci. 2020, 11, 208–216. 10.1039/C9SC04540F. [DOI] [PMC free article] [PubMed] [Google Scholar]; e Bay K. L.; Yang Y.-F.; Houk K. N. Multiple Roles of Silver Salts in Palladium-catalyzed C–H Activations. J. Organomet. Chem. 2018, 864, 19–25. 10.1016/j.jorganchem.2017.12.026. [DOI] [Google Scholar]; f Funes-Ardoiz I.; Maseras F. Oxidative Coupling Mechanisms: Current State of Understanding. ACS Catal. 2018, 8, 1161–1172. 10.1021/acscatal.7b02974. [DOI] [Google Scholar]; g Anand M.; Sunoj R. B.; Shaefer H. F. III Non-innocent Additives in a Palladium(II)-Catalyzed C–H Bond Activation Reaction: Insights into Multimetallic Active Catalysts. J. Am. Chem. Soc. 2014, 136, 5535–5538. 10.1021/ja412770h. [DOI] [PubMed] [Google Scholar]; h Yang Y.-F.; Cheng G.-J.; Liu P.; Leow D.; Sun T.-Y.; Chen P.; Zhang X.; Yu J.-Q.; Wu Y.-D.; Houk K. N. Palladium-Catalyzed Meta-Selective C–H Bond Activation with a Nitrile-Containing Template: Computational Study on Mechanism and Origins of Selectivity. J. Am. Chem. Soc. 2014, 136, 344–355. 10.1021/ja410485g. [DOI] [PubMed] [Google Scholar]
  34. a Pereiro M.; Baldomir D.; Arias J. E. Unexpected Magnetism of Small Silver Clusters. Phys. Rev. A 2007, 75, 063204 10.1103/PhysRevA.75.063204. [DOI] [Google Scholar]; b Pereiro M.; Baldomir D. Structure of Small Silver Clusters and Static Response to an External Electric Field. Phys. Rev. A 2007, 75, 033202 10.1103/PhysRevA.75.033202. [DOI] [Google Scholar]; c Pereiro M.; Baldomir D. Determination of the Lowest-energy Structure of Ag8 from First-principles Calculations. Phys. Rev. A 2005, 72, 045201 10.1103/PhysRevA.72.045201. [DOI] [Google Scholar]; d Tsuneda T. Theoretical Investigations on Geometrical and Electronic Structures of Silver Clusters. J. Comput. Chem. 2019, 40, 206–211. 10.1002/jcc.25577. [DOI] [PubMed] [Google Scholar]
  35. Ag octamers (n = 8), and octakaidecamers (n = 18) are considered to become more stable than other clusters.
  36. In truth, the energetic span is the difference between the energy of the transition state leading to Pd(THF)2, which is expected to be the turnover frequency determining transition state (TDTS) and that of the resting state. However, as the step of the release of Pd(THF)2 is highly endergonic, according to the Hammond postulate, it should feature a rather late transition state, energetically and geometrically very similar to the state of Pd(THF)2.
  37. a Solel E.; Tarannam N.; Kozuch S. Catalysis: Energy is the Measure of All Things. Chem. Commun. 2019, 55, 5306–5322. 10.1039/C9CC00754G. [DOI] [PubMed] [Google Scholar]; b Besora M.; Maseras F. Microkinetic Modeling in Homogeneous Catalysis. WIREs Comput. Mol. Sci. 2018, 8, e1372 10.1002/wcms.1372. [DOI] [Google Scholar]
  38. Zhang Y.; Cui Z.; Li Z. J.; Liu Z.-Q. Pd(II)-Catalyzed Dehydrogenative Olefination of Vinylic C–H Bonds with Allylic Esters: General and Selective Access to Linear 1,3-Butadienes. Org. Lett. 2012, 14, 1838–1841. 10.1021/ol300442w. [DOI] [PubMed] [Google Scholar]; For a similar type of coupling, see:
  39. a Zhao H.; Ariafard A.; Lin Z. β-Heteroatom versus β-Hydrogen Elimination: A Theoretical Study. Organometallics 2006, 25, 812–819. 10.1021/om050703v. [DOI] [Google Scholar]; b Bogdos M. K.; Stepanović O.; Bismuto A.; Luraschi M. G.; Morandi B. Mechanistically Informed Selection Rules for Competing β-Hydride and β-Heteroatom Eliminations. Nat. Chem. 2022, 1, 787–793. 10.1038/s44160-022-00145-x. [DOI] [Google Scholar]; c Ferber B.; Prestat G.; Vogel S.; Madec D.; Poli G. Synthesis of 3,5-Disubstituted Piperazinones via Palladium (II) -Catalyzed Amination. Synlett 2006, 13, 2133–2135. 10.1055/s-2006-948208. [DOI] [Google Scholar]
  40. a Fabris F.; Illner M.; Repke J.-U.; Scarso A.; Schwarze M. Is Micellar Catalysis Green Chemistry?. Molecules 2023, 28, 4809. 10.3390/molecules28124809. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Roschangar F.; Zhou Y.; Constable D. J. C.; Colberg J.; Dickson D. P.; Dunn P. J.; Eastgate M. D.; Gallou G.; Hayler J. D.; Koenig S. G.; Kopach M. E.; Leahy D. K.; Mergelsberg I.; Scholz U.; Smith A. G.; Henry M.; Mulder J.; Brandenburg J.; Dehli J. R.; Fandrick D. R.; Fandrick K. R.; Gnad-Badouin F.; Zerban G.; Groll K.; Anastas P. T.; Sheldon R. A.; Senanayake C. H. Inspiring Process Innovation via an Improved Green Manufacturing Metric: iGAL. Green Chem. 2018, 20, 2206–2211. 10.1039/C8GC00616D. [DOI] [Google Scholar]
  41. Pontes da Costa A.; Nunes D. R.; Tharaud M.; Oble J.; Poli G.; Rieger J. Palladium(0) Nanoparticles Embedded in Core–shell Nanogels as Recoverable Catalysts for the Mizoroki–Heck Reaction. ChemCatChem. 2017, 9, 2167–2175. 10.1002/cctc.201601645. [DOI] [Google Scholar]
  42. For some recent reviews, see:; a De Martino M. T.; Abdelmohsen L. K. E. A.; Rutjes F. P. J. T.; van Hest J. C. M. Nanoreactors for Green Catalysis. Beilstein J. Org. Chem. 2018, 14, 716–733. 10.3762/bjoc.14.61. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Lipshutz B. H.; Ghorai S.; Cortes-Clerget M. The Hydrophobic Effect Applied to Organic Synthesis: Recent Synthetic Chemistry “in Water. Chem.—Eur. J. 2018, 24, 6672–6695. 10.1002/chem.201705499. [DOI] [PubMed] [Google Scholar]; c Cortes-Clerget M.; Yu J.; Kincaid J. R. A.; Walde P.; Gallou F.; Lipshutz B. H. Water as the Reaction Medium in Organic Chemistry: from our Worst Enemy to our Best Friend. Chem. Sci. 2021, 12, 4237–4266. 10.1039/D0SC06000C. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. For precedents of Sonogashira couplings under micellar conditions, see:; a Lipshutz B. H.; Chung D. W.; Rich B. Sonogashira Couplings of Aryl Bromides: Room Temperature, Water Only. No Copper. Org. Lett. 2008, 10, 3793–3796. 10.1021/ol801471f. [DOI] [PubMed] [Google Scholar]; b Jakobi M.; Gallou F.; Sparr C.; Parmentier M. A General Protocol for Robust Sonogashira Reactions in Micellar Medium. Helv. Chim. Acta 2019, 102, e1900024 10.1002/hlca.201900024. [DOI] [Google Scholar]; c Braje W.; Britze K.; Dietrich J. D.; Jolit A.; Kaschel J.; Klee J.; Lindner T.. Organic Reactions Carried out in Aqueous Solution in the Presence of a Hydroxyalkyl(alkyl)cellulose or an Alkylcellulose. Patent 2017, WO 2017/129796.; d Struwe J.; Ackermann L.; Gallou F. Recent Progress in Copper-free Sonogashira-Hagihara Cross-couplings in Water. Chem. Catal 2023, 3, 100485 10.1016/j.checat.2022.12.002. [DOI] [Google Scholar]
  44. The same coupling carried out under classical conditions on a 1.0 mmol scale with the catalytic system [PdCl2(PPh3)2 (0.2 mol%), CuI (5 mol%), NEt3 (3.0 equiv.)] in DMF for 24 h at 45 °C, afforded 4 in 44% isolated yield.
  45. Poli G.; Pradal A.; Di Matteo M.; Gagliardi A.; Veiros L.; Gallou F. First Pd-catalyzed C(sp2)–H / C(sp2)–H coupling of Limonene. ChemRxiv 2024, 10.26434/chemrxiv-2024-nzkq1. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

jo4c00501_si_001.pdf (5.1MB, pdf)

Data Availability Statement

The data underlying this study are available in the published article, in its Supporting Information, and openly available in Zenodo (10.5281/zenodo.10462156).


Articles from The Journal of Organic Chemistry are provided here courtesy of American Chemical Society

RESOURCES