ACS Publications. Most Trusted. Most Cited. Most Read
Mechanism and Selectivity of Rhodium-Catalyzed 1:2 Coupling of Aldehydes and Allenes
My Activity

Figure 1Loading Img
  • Open Access
Article

Mechanism and Selectivity of Rhodium-Catalyzed 1:2 Coupling of Aldehydes and Allenes
Click to copy article linkArticle link copied!

View Author Information
Department of Organic Chemistry, Arrhenius Laboratory, Stockholm University, SE-106 91 Stockholm, Sweden
Open PDFSupporting Information (1)

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2013, 135, 20, 7647–7659
Click to copy citationCitation copied!
https://doi.org/10.1021/ja4014166
Published April 25, 2013

Copyright © 2013 American Chemical Society. This publication is licensed under these Terms of Use.

Abstract

Click to copy section linkSection link copied!

The rhodium-catalyzed highly regioselective 1:2 coupling of aldehydes and allenes was investigated by means of density functional theory calculations. Full free energy profiles were calculated, and several possible reaction pathways were evaluated. It is shown that the energetically most plausible catalytic cycle is initiated by oxidative coupling of the two allenes, which was found to be the rate-determining step of the overall reaction. Importantly, Rh–allyl complexes that are able to adopt both η3 and η1 configurations were identified as key intermediates present throughout the catalytic cycle with profound implications for the selectivity of the reaction. The calculations reproduced and rationalized the experimentally observed selectivities and provided an explanation for the remarkable alteration in the product distribution when the catalyst precursor is changed from [RhCl(nbd)]2 (nbd = norbornadiene) to complexes containing noncoordinating counterions ([Rh(cod)2X]; X = OTf, BF4, PF6; cod = 1,5-cyclooctadiene). It turns out that the overall selectivity of the reaction is controlled by a combination of the inherent selectivities of several of the elementary steps and that both the mechanism and the nature of the selectivity-determining steps change when the catalyst is changed.

Copyright © 2013 American Chemical Society

1 Introduction

Click to copy section linkSection link copied!

Rhodium(I) complexes are well-established catalysts for three-component cycloadditions of unsaturated compounds, yielding carbocycles with various ring sizes (e.g., six-membered; Scheme 1a). (1) Rhodium(I) also catalyzes hydroacylation reactions in which an acyl unit and the hydrogen atom of an aldehyde are added across a C–C multiple bond (Scheme 1b). (2) A reductive version of the latter process, furnishing the corresponding alcohols, has also been developed (Scheme 1c). (3) In 2006, Krische and co-workers reported a novel reaction that can be considered as a crossover between those presented in Scheme 1a and Scheme 1c. In its course, two molecules of acetylene are coupled with one molecule of aldehyde under a hydrogen atmosphere to form a linear allylic alcohol product (Scheme 2a). (4, 5) Very recently, a combination of the in principle more fundamental couplings from Scheme 1a,b has also been successfully accomplished. (6, 7) In particular, a significant example comes from the laboratory of Murakami and consists of a 1:2 coupling of aldehydes and allenes (Scheme 2b). (6)

Scheme 1

Scheme 2

The reaction in Scheme 2b indeed constitutes a state-of-the-art method in synthetic organic chemistry. It is inherently atom-economical and leads to a large increase in molecular complexity in a single step. A particularly appealing feature of this process is its selectivity, which is a much more complex issue than in the reaction involving acetylene (Scheme 2a). Considering only the couplings of two monosubstituted allene moieties of type 2 with one molecule of aldehyde 1, 36 different isomers of the product are possible, with various connectivities and Z/E arrangements of the double bonds. (8) Experimentally, when [RhCl(nbd)]2 (nbd = norbornadiene) was used as the rhodium source and 1,2-bis(diphenylphosphino)ethane (dppe) as the ligand, product 3 was obtained with 91% selectivity, accompanied by two minor isomers, 4 (6%) and 5 (3%). (6) Very interestingly, when the chloride was replaced by a noncoordinating counterion such as TfO, BF4, or PF6, the selectivity of the reaction was switched, and compound 5 became the dominant product (87–95%). (6) The common features of the products obtained under the two sets of reaction conditions are that the two allene moieties are coupled through their central atoms (C2–C2 linkage) and the aldehyde is attached to one of the terminal carbons (C1 or C3) of one of the allenes.
The reaction mechanism of the related three-component reductive coupling shown in Scheme 2a has been investigated by a combination of experimental and computational approaches. (9) It was established the reaction takes place via an initial oxidative coupling of two acetylene molecules followed by migratory insertion of the aldehyde molecule, hydrogenolysis, and reductive elimination. (9) Hence, except for the extra hydrogenolysis step, this mechanism is similar to the well-established catalytic cycle of the trimerization process depicted in Scheme 1a. (1d)
In the case of the 2:1 coupling of allenes and aldehydes (Scheme 2b), it is not evident that an analogous mechanistic pathway is followed. It has been shown experimentally that allenes of type 2 indeed undergo an initial oxidative coupling in the presence of a Rh(I) catalyst, but instead of the subsequent migratory insertion that would eventually result in the regular [2 + 2 + 2] cycloaddition (Scheme 1a), a β-hydride elimination/reductive elimination sequence occurs, leading overall to a linear dimer. (10, 11) Additionally, it is known that the hydroacylation reactions (Scheme 1b) are initiated by oxidative addition of aldehydes to Rh(I) complexes, (2a) so this process may also be involved in the mechanism of the combined reaction shown in Scheme 2b. Finally, there is also an additional possibility of an initial oxidative coupling between the aldehyde and the allene moieties that has been shown to take place in some catalytic cycles. (12) A mechanism starting with the oxidative coupling of aldehyde and allene was actually suggested by Murakami and co-workers in their report. (6)
These possible mechanisms for the reaction, suggested by the experimental findings or proposed in the literature, are summarized in Scheme 3. They differ in regard to the initial step, which is either C–H oxidative addition, allene–allene oxidative coupling, or aldehyde–allene oxidative coupling. In addition to the general mechanism, there is the issue of the reaction selectivity and its sources. Each of the pathways shown in Scheme 3 may yield each of the 36 possible product isomers, including the experimentally observed 3, 4, and 5, by simply altering the positions at which the new bonds are formed. Moreover, all three pathways are consistent with the results of deuterium labeling studies carried out experimentally. (6) In addition, for almost all of the intermediates in Scheme 3 one can envision alternative reactions that they may undergo (not shown in the scheme) leading to side products with entirely different structures (e.g., 12 or 22 dimers, etc.), which adds to the complexity of the system.

Scheme 3

Scheme 3. Possible General Mechanistic Pathways for the Reaction in Scheme 2b, As Suggested by Experimental Findings or Proposed in the Literaturea

Scheme aEach of the three pathways can yield all of the possible products. Here, only the routes leading to product 3 are shown.

Considering the significance of the coupling reaction depicted in Scheme 2b and the current poor understanding of its mechanism, in particular the origins of the observed selectivity, we decided to investigate it computationally by means of density functional theory (DFT). The calculations showed that the reaction indeed initially follows the pathway starting with oxidative coupling of two allene molecules (Scheme 3, path B). However, we found that a not previously suggested allylation of the aldehyde takes place subsequently with important selectivity implications. It is shown that a key role in the reaction mechanism is played by rhodium complexes containing allyl ligands that can adopt either η3 or η1 coordination, which display intrinsically distinct reactivities. Moreover, it is demonstrated that the reaction catalyzed by rhodium complexes with noncoordinating counterions follows a partially different mechanistic route involving an alkoxide oxidation step. The calculations provide an explanation for the observed selectivity switch when the counterion in the Rh catalyst is altered.

2 Computational Details

Click to copy section linkSection link copied!

All of the calculations were performed using the Gaussian 03 package (13) and the B3LYP functional. (14) Geometry optimizations were done with a combined basis set in which the LANL2DZ basis set with pseudopotential (15) was used for Rh and the 6-31G(d,p) basis set was used for all other atoms. Frequencies were computed analytically at the same level of theory to confirm whether the structures were minima (no imaginary frequencies) or transition states (only one imaginary frequency). Selected transition-state structures were confirmed to connect the correct reactants and products by intrinsic reaction coordinate (IRC) calculations. To obtain better accuracy of the final energy values, the energies of the optimized geometries were recalculated using single-point calculations with a larger basis set, namely, LANL2DZ for Rh and 6-311+G(2d,2p) for the other atoms. The effect of solvation was evaluated by performing single-point self-consistent reaction field calculations with the conductor-like polarizable continuum model (C-PCM) (16) as implemented in Gaussian 03. The parameters for toluene (ε = 2.379), corresponding to the experimental conditions, and the united atom (UA0) radii were used in these calculations.
The final Gibbs energies reported in the article (ΔGTol) are the large-basis-set single-point energies with Gibbs energy corrections (at 298.15 K), solvation corrections, and corrections for dispersion effects using the method of Grimme. (17) Inclusion of the dispersion effects has recently been shown to improve the performance of the B3LYP method significantly. (18)
The calculations were carried out using the full structure of the dppe ligand, while the aldehyde and allene reactants were represented with smaller model structures, namely, benzaldehyde (1, R1 = Ph in Scheme 2b) and buta-1,2-diene (2, R2 = Me), respectively.

3 Results and Discussion

Click to copy section linkSection link copied!

Because of the number of possible mechanistic pathways (Scheme 3), the reaction poses a considerable challenge from a computational point of view. In this section, we first focus on the [RhCl(dppe)]-catalyzed reaction, assuming that chloride is coordinated to the metal center. The mechanism of the reaction is systematically examined step by step by evaluating all of the possible pathways, including potential side reactions. After the overall mechanism is established, the full catalytic cycle and the energy profile are summarized and the origins of the regioselectivity are discussed. In the final part, we present the results of the calculations on the reaction catalyzed by the cationic [Rh(dppe)]+ complex, which is the putative active species in the reactions using catalyst precursors containing noncoordinating counterions. The obtained free energy profile is then analyzed, and the origins of the observed alteration of the selectivity when the catalyst is changed are explained.
All of the free energies are reported relative to the sum of the free energies of the free reactants and catalyst (i.e., 1 + 2 × 2 + [Rh], where [Rh] = [RhCl(dppe)] or [Rh(dppe)]+), which we have chosen to set equal to zero.

3.1 Initial Step of the Catalytic Cycle: Oxidative Coupling versus Oxidative Addition

Before considering the various reaction pathways, we evaluated the different modes of binding between the catalyst and the reactants. The calculations showed that coordination of allene 2 to the Rh catalyst to give INT1 is exergonic by 8.9 kcal/mol, whereas binding of aldehyde 1 to form INT1a is exergonic by 1.6 kcal/mol (Figure 1). The introduction of any additional ligands always resulted in an increase in the free energy, which is in agreement with the known preference of d8 metal complexes to adopt a four-coordinate square-planar geometry (see the Supporting Information). Hence, INT1 containing a single allene molecule, dppe, and chloride as ligands (see Figure 2 for the structure) is the lowest-energy Rh(I) species, to which the barriers for the initial step should be related.

Figure 1

Figure 1. Free energy profiles for the possible initial steps of the reaction. The intermediates connecting INT1 with TS1 and TS1b were found to be higher in free energy than INT1, so they are not shown in the diagram (see the Supporting Information for details).

Figure 2

Figure 2. Optimized structures of selected intermediates and transition states for the initial step (distances in Å and angles in deg). For clarity, the phenyl substituents on the phosphorus centers have been omitted in the figure.

As shown in Scheme 3, three general mechanistic pathways for the investigated reaction are possible, each of which begins with a different step, namely, C–H oxidative addition (path A), allene–allene oxidative coupling (path B), or aldehyde–allene oxidative coupling (path C). We optimized the transition states (TSs) corresponding to these steps and compared them in terms of their relative free energies, taking into account all of the possible coupling patterns of the reactants where applicable.
The C–H oxidative addition of aldehyde (path A in Scheme 3) was found to take place via transition state TS1a (Figure 2). It requires overcoming an overall barrier of 25.2 kcal/mol relative to INT1 and involves an initial interconversion into the aldehyde-containing complex INT1a (Figure 1). The oxidative addition results in the formation of the Rh(III) hydride INT2a, and the process is endergonic by as much as 22.9 kcal/mol.
The oxidative coupling of two allene moieties (path B in Scheme 3) was much more complicated to investigate than the oxidative addition. In the course of this step, the C–C bond formation may occur between any of the three carbon atoms of each of the coupling partners (C1–C1, C1–C2, C1–C3, C2–C1, etc.). Additionally, when the central atom of the allene (C2) is engaged in the C–C bond formation, extra combinations arise from the fact that either of the terminal carbons (C1 or C3) may bind to the rhodium center. When the Z/E arrangements of the double bonds and, in three cases, the cis/trans isomers are also taken into account, there are in total 24 distinct TSs that must be considered (Scheme 4). The free energies for all of these were calculated, and the one with the lowest free energy was found to be TS1, whose structure is shown in Figure 2 (structures and energies of all of the TSs are given in Supporting Information). In this TS, the rhodium center, which constitutes the largest steric bulk in the system because of the proximity of the phenyl-substituted phosphine ligand, is coordinated to the least-substituted carbons (C1) of the allenes. Additionally, the Z,Z arrangement of the double bonds minimizes the repulsion in the other part of the structure. The free energy of TS1 was calculated to be 21.5 kcal/mol higher than that of INT1 (Figure 1). The free energies of the other TSs shown in Scheme 4 were calculated to be higher than that of TS1 by at least 2.2 kcal/mol (see the Supporting Information).

Scheme 4

Scheme 4. Possible Transition States for Oxidative Coupling of the Two Allene Moietiesa

Scheme aIt should be noted that for each of the TS structures there are varying numbers of Z/E or cis/trans isomers, resulting in an overall total of 24 possible TSs.

Interestingly, TS1 does not lead to the expected five-membered metallacycle but instead yields an octahedral Rh(III) complex INT2, containing an η3-allyl-η1-allyl ligand (see Figure 2 for the structure). (19) In what follows, we will refer to this kind of intermediate as η31-bis(allylic), in accordance with the literature. (20) The formation of INT2 is exergonic by 19.6 kcal/mol relative to INT1. This kind of η31-bis(allylic) ligand could be expected to undergo syn/anti isomerization of the methyl substituents. However, intermediate INT2 possesses some unique features that result in the preservation of its stereochemical configuration as formed in the course of the reaction from TS1 (i.e., Z double-bond geometry and anti CH3 group in the η3-allyl system). First, the methyl substituent in the η3-allyl of INT2 cannot move from the anti position to the syn position via an η3–η1–η3 isomerization mechanism, as the required rotation in the transient η1 complex is impossible because the allylic system is a part of metallacycle. Second, it is possible for the η3 and η1 allyl groups in INT2 to exchange places, but this also maintains the overall configuration. During such a process, the anti CH3 group becomes a substituent of the Z-configured double bond and vice versa. Finally, INT2 may transiently be transformed into an isomeric η31-bis(allylic) complex, INT2′, that is calculated to be 8.5 kcal/mol higher in free energy than INT2 (Scheme 5). However, such an interconversion also leaves the stereochemical arrangement of the allene dimer moiety intact. These properties are of importance for the overall selectivity of the reaction, as will be discussed below.

Scheme 5

Scheme 5. Possible η3–η1–η3 Isomerization of INT2a

Scheme aFree energies relative to the global zero are indicated in kcal/mol.

As a final option for the initial step, oxidative coupling between the aldehyde and allene molecules was also examined (path C in Scheme 3). Similar to the case of allene–allene oxidative coupling, the C–C bond may be formed between the carbonyl carbon and positions C1, C2, or C3 of the allene. When the Rh coordination as well as the Z/E and cis/trans isomers are also considered, there are seven possible transition states (see the Supporting Information). The one with the lowest free energy among them is TS1b (Figure 2). The corresponding free energy barrier is 29.7 kcal/mol from INT1, and the formation of the resulting Rh(III) complex INT2b is endergonic by 8.7 kcal/mol relative to INT1 (Figure 1).
These results clearly point to oxidative coupling of the two allenes leading to intermediate INT2 as the favored initial step of the mechanism. The transition state for the C–H oxidative addition (TS1a) is 3.7 kcal/mol higher than TS1 (but lower than many of the transition states depicted in Scheme 4). On the other hand, the oxidative coupling between the aldehyde and allene was found to occur via high-energy transition states, the lowest of which (TS1b) is 8.2 kcal/mol higher than TS1. Importantly, the formation of INT2 is irreversible, and therefore, already at this point it is possible to dismiss paths A and C of Scheme 3 as plausible reaction mechanisms. Using the same argument, one can also discard all of the pathways starting with oxidative coupling of the two allenes in arrangements other than in TS1 (Scheme 4). Hence, from the reaction mixture containing the Rh(I) catalyst (cat), aldehyde 1, and allene 2, intermediate INT2 should be formed practically exclusively. As explained above, this intermediate has a defined stereochemical configuration that strongly narrows the number of possible final products of the reaction. In the following sections, we will show that the experimentally observed product 3, as well as the minor side products 4 and 5, can all be derived from INT2 as a common precursor.

3.2 Formation of the Second C–C Bond

Once the two allene molecules are coupled in the first step of the mechanism to form INT2, the second C–C bond formation must take place, incorporating the aldehyde moiety into the structure and thus assembling the complete backbone of the product. It has been suggested that the second C–C bond formation should occur via migratory insertion of the aldehyde into the Rh–C bond. (6) Such a course of the reaction was also established in mechanistic studies on the above-mentioned related reductive 2:1 coupling of acetylenes with aldehydes (Scheme 2a). (9) In addition to insertion of the aldehyde, INT2 may potentially undergo a competing process, namely, a β-hydride elimination that would eventually lead to the formation of the dimer of allene 2. There is experimental evidence that the latter pathway is followed in the absence of aldehyde 1 in the reaction mixture. (10) In the present study, we discovered an additional alternative for the C–C bond formation that has not been proposed previously and that turns out to have a much lower barrier than the two other processes. This alternative pathway involves direct allylrhodium addition to the aldehyde. In this section, these mechanistic possibilities are investigated and compared. (21)
Irrespective of the specific way the new C–C bond is formed, the reaction begins with the coordination of aldehyde 1 to the rhodium center. To create the necessary coordination site for the aldehyde, the allyl ligand must change its binding mode from η3 to η1. This may occur in two distinct ways, with either C1 or C3 ligating the metal center in the resulting complex (INT3 or INT3′, respectively; Figure 3). This transformation was calculated to be endergonic by 4.1 and 6.7 kcal/mol, respectively.

Figure 3

Figure 3. Free energy profile for the formation of the second C–C bond.

From INT3, the migratory insertion may occur only at the C1 position because of the symmetric structure of this intermediate, while on the other hand, insertion in INT3′ can take place at either C1 or C3. The calculations showed that the lowest-energy transition state leads to the insertion at C1 (TS-INS) and arises from INT3. The second-lowest transition state found was TS-INS′, which involves the insertion at C3 from INT3′. The corresponding barriers for TS-INS and TS-INS′ relative to INT2 are 29.8 and 31.7 kcal/mol, respectively (Figure 3). The free energy difference between them can be explained by the steric repulsion between the proximal methyl group and the aldehyde in the latter transition state.
The alternative C–C bond formation pathway that we discovered to be possible starting from INT3 or INT3′ is a direct attack of the σ-allyl ligand on the aldehyde coordinated to the rhodium center. (22) Such a reaction, which occurs via a six-membered cyclic transition state (TS2 or TS2′, respectively; see Figure 3 for schematic illustrations and Figure 4 for optimized structures), was found to be energetically more favorable than the migratory insertion by as much as 15 kcal/mol.

Figure 4

Figure 4. Optimized structures of the allylation reactants, transition states, and resulting complexes.

In this case, the C–C bond formation from INT3 can occur only at the C3 position (TS2). The corresponding barrier was calculated to be 14.7 kcal/mol relative to INT2. However, allylation from INT3′ may take place at either C1 or C3, but the former scenario is lower in free energy (TS2′ in Figure 3) and requires overcoming an overall barrier of 15.3 kcal/mol relative to INT2. In both TS2 and TS2′, the six-membered ring adopts the chairlike geometry (Figure 4) with the methyl and phenyl groups occupying the axial and equatorial positions, respectively. The small free energy difference between them might be explained by a steric repulsion between the methyl group and the Rh center in TS2′, which is not present in TS2. Transition states TS2 and TS2′ constitute the divergence point from which the reaction proceeds along two separate parallel pathways.
The allylations occurring via TS2 and TS2′ yield octahedral complexes INT4 and INT4′, respectively, containing η2-alkene and alkoxide ligands coordinated to Rh(III) (see Figure 4 for optimized structures). Interestingly, INT4 was found to be 0.9 kcal/mol less stable than INT4′.
Finally, the β-hydride elimination from INT2, the possible side reaction at this stage of the mechanism, was found to have a barrier of 33.0 kcal/mol relative to INT2 (via TS-H, whose optimized structure is given in the Supporting Information), which is 18.3 kcal/mol higher than that found for TS2 (Figure 3). Therefore, in the presence of aldehyde 1 in the reaction mixture, the allylation is expected to occur much more favorably, in agreement with the experimental results.

3.3 β-Hydride Elimination

The next step in the reaction pathway is the β-hydride elimination, which restores the carbonyl group (Scheme 3, path B). In both INT4 and INT4′, the only β-hydrogen available is the one originating from the aldehyde group, and hence, it will be exclusively transferred to the metal center. However, for the β-hydride elimination to take place, a vacant coordination site must be created to accommodate the hydride. This can be generated by dissociation of the η2-alkene from the metal center in INT4 and INT4′ followed by rotation of the alkoxide ligand. The calculations showed that this isomerization takes place through a series of steps, among which the highest-energy transition states are TS3 and TS3′ with barriers of 17.7 and 16.6 kcal/mol relative to INT4 and INT4′, respectively (Figure 5). A detailed description of the isomerization process, together with calculated free energy profiles and optimized structures, can be found in the Supporting Information.

Figure 5

Figure 5. Free energy profile for the β-hydride elimination step.

The isomerization results in intermediates INT5 and INT5′ containing an η3-allyl ligand, which were calculated to be higher in free energy by 6.8 and 4.7 kcal/mol relative to INT4 and INT4′, respectively. From INT5 and INT5′, the β-hydride elimination takes place with a simultaneous change in the coordination mode of the allyl group from η3 to η1 in order to create a vacant site for the hydride at the metal center (TS4 and TS4′; see Figure 6 for optimized structures). The calculated free energies of TS4 and TS4′ relative to the respective intermediates INT5 and INT5′ are 9.0 and 10.1 kcal/mol (Figure 5), respectively. It should be noted that transitions states TS4 and TS4′ are lower in free energy than the preceding TS3 and TS3′, which is of importance for the analysis of the selectivity (see below). The rhodium hydride complexes resulting from the β-elimination (INT6 and INT6′) are octahedral and contain an η1-allyl and a carbonyl oxygen coordinated to the metal.

Figure 6

Figure 6. Optimized structures of the β-hydride elimination transition states and the resulting η1-allyl complexes.

We also considered an alternative process that theoretically may interfere with the β-hydride elimination from INT5 or INT5′, namely, a C–O bond-forming reductive elimination. With such a course, the reaction would yield pyrane derivatives as products and thus would constitute a heteroatom analogue of the [2 + 2 + 2] carbocyclization shown in Scheme 1a. (23) However, the calculated barrier for this process (from INT5′ to TS-RE′) was found to be 26.9 kcal/mol, which is 16.8 kcal/mol higher than for the β-elimination from the same intermediate (Figure 5). Hence, the C–O bond-forming reductive elimination should not take place, which is indeed the result observed experimentally.

3.4 Reductive Elimination

The last step of the mechanism, closing the catalytic cycle, is the C–H bond-forming reductive elimination (Scheme 3). This could occur directly from INT6 and INT6′. The calculated free energy barriers are quite feasible (20.4 and 19.2 kcal/mol from INT6 and INT6′, respectively; see the Supporting Information for the optimized structures of the TSs). However, if the reaction were to follow this pathway, it would yield only compounds 4 and 5 as products rather than the experimentally observed mixture containing predominantly their isomer 3. Prompted by this apparent flaw, we envisioned that the η1-allylic complexes INT6 and INT6′ could be converted into their η3-allylic counterparts INT7 and INT7′, respectively, upon dissociation of the carbonyl oxygen from the rhodium center prior to the reductive elimination. The fact that both terminal carbon atoms of the η3-allyl systems in INT7 and INT7′ can couple with the hydride during the reductive elimination would enable the formation of 3 in such a transformation.
Indeed, as shown in Figure 7, the Rh(III) hydrides containing the η3-allyl ligand, INT7 and INT7′ (see Figure 8 for the structures), are more stable than INT6 and INT6′ by 1.1 and 5.3 kcal/mol, respectively. (24) As mentioned above, from INT7 and INT7′ the reductive elimination can occur at either the C1 or C3 side of the η3-allyl. The calculations showed that irrespective of the starting complex, the reductive elimination engaging the methyl-substituted carbon C3 (TS5a and TS5a′) is more favorable than the one occurring at the unsubstituted carbon C1 (TS5b and TS5b′; Figure 7). This is typical behavior in many Rh-catalyzed reactions involving π-allyl species, which usually afford the branched product. (19a-19m) One reason for this preference is that in TS5a and TS5a′ the sterically crowded rhodium center is positioned closer to the less-substituted double bond (between C1 and C2). For example, in TS5a′, the bond distances of Rh–C1 and Rh–C2 are 2.15 and 2.16 Å, respectively, whereas in TS5b′, because of the steric effect of methyl group, the Rh–C2 and Rh–C3 bond distances are somewhat longer (2.19 and 2.22 Å, respectively) (Figure 8). Importantly, the free energies of all of the TS5 variants are lower than those of the transition states for the direct reductive elimination from INT6 and INT6′, implying that the reaction does not follow the latter pathway. (25) The influence of the relative energies of the reductive elimination transition states on the overall selectivity of the reaction will be discussed in the following section.

Figure 7

Figure 7. Free energy profile for the reductive elimination step.

Figure 8

Figure 8. Optimized structures of the reductive elimination reactants and transition states.

3.5 Overall Catalytic Cycle and Origins of the Regioselectivity in the [RhCl(dppe)]-Catalyzed Reaction

The mechanism of the reaction as established by the present DFT calculations is summarized in Scheme 6, and the corresponding overall free energy profile is given in Figure 9. The catalytic cycle consists of four steps: (1) oxidative coupling of two allene molecules, (2) allylation of the aldehyde, (3) β-hydride elimination, and (4) C–H bond-forming reductive elimination. A prominent role in the mechanism is played by allylrhodium complexes. Because of the flexibility of the allyl ligand binding modes (η3, η1), a variety of reactivity patterns at each step of the catalytic cycle are available. Hence, the mechanism of the reaction involving allenes displays considerable differences compared with the mechanism of the related reductive coupling of acetylene and aldehyde shown in Scheme 2a, (9) in which vinyl ligands limited only to σ bonding are involved.

Scheme 6

Scheme 6. Summary of the Mechanism of the [RhCl(dppe)]-Catalyzed 1:2 Coupling of Aldehydes and Allenes

Figure 9

Figure 9. Free energy profile for the [RhCl(dppe)]-catalyzed 1:2 coupling of aldehydes and allenes.

The oxidative coupling of two allene molecules (TS1) is a common step in all of the pathways leading to the experimentally observed products 35. It has a barrier of 21.5 kcal/mol relative to the most stable Rh(I) complex, INT1. The oxidative coupling is irreversible and leads selectively to the octahedral η31-bis(allylic) Rh(III) complex INT2, which has a well-defined and stable stereochemical configuration. Therefore, in this step of the mechanism, a number of structural features of the products are already determined, such as the C2–C2 linkage and the configuration of the double bonds. It is also important to point out that the overall barrier for the oxidative coupling is the highest among all of the barriers in the catalytic cycle. This step therefore constitutes the rate-determining step of the reaction.
Upon coordination of an aldehyde molecule, INT2 undergoes a transformation to either INT3 or INT3′, depending on which side of the η3-allyl ligand remains bound to the metal (C1 or C3, respectively). Thus, this step constitutes the first divergence point, from which the two pathways proceed through the subsequent steps separately. The following step is the allylation of the aldehyde, which proceeds through the six-membered cyclic transition states TS2 and TS2′. After an isomerization (INT4INT5 via TS3 and INT4′INT5′ via TS3′) that is needed to create an empty coordination site in an appropriate position of the complex, a β-hydride elimination takes place via TS4 and TS4′. Since TS3 and TS3′ are higher in energy than all of the subsequent transition states, passing them can be considered practically irreversible. Therefore, the ratio at which the catalyst is split to follow either one or the other branch of the catalytic cycle is determined by the free energy difference between the transition states TS3 and TS3′. According to the calculations, these free energies differ by 2.0 kcal/mol, and hence, intermediates INT6 and INT6′ are formed in a 5:95 ratio. (26)
The next step of the mechanism is the C–H bond-forming reductive elimination, which starts with conversion of the η1-allylic Rh(III) hydrides INT6 and INT6′ to the η3-allylic species INT7 and INT7′, respectively. The reductive elimination is the final regioselectivity-determining step. This process was found to take place preferentially at the C3 position (TS5a and TS5a′) rather than at C1 (TS5b and TS5b′). In particular, from INT7′, the formation of INT8a′ via TS5a′ is favored over the formation of INT8b′ via TS5b′ by 2.3 kcal/mol. This free energy difference corresponds approximately to a 98:2 ratio. On the other hand, the 5.3 kcal/mol free energy difference between TS5a and TS5b implies the practically exclusive transformation of INT7 into INT8a. Hence, when the previous split of the reaction pathways is taken into account, the overall selectivity predicted by the calculations is 3:4:5 = 93:2:5, which is in very good agreement with the experimentally observed ratio of 91:6:3 (Scheme 2b). (6)
Finally, intermediates INT8a′, INT8b′, and INT8a formed in the reductive elimination regenerate the free catalyst [RhCl(dppe)] by releasing the corresponding products 3, 4, and 5, respectively, closing the catalytic cycle. It should be pointed out that the relative stabilities of the free products 3, 4, and 5 do not correlate at all with their ratios resulting from the catalyzed reaction.
In summary, the above considerations show that the calculated energies can reproduce and rationalize the observed selectivity of the [RhCl(dppe)]-catalyzed reaction, which provides further support for the reaction mechanism suggested in Scheme 6.

3.6 Regioselectivity in the [Rh(dppe)]+-Catalyzed Reaction

As mentioned in the Introduction, the selectivity of the reaction can be altered by replacing the chloride anion in the catalyst precursor with a noncoordinating counterion such as TfO, BF4, or PF6. To rationalize this catalyst-dependent selectivity, we reinvestigated the reaction mechanism with the cationic [Rh(dppe)]+ complex as the active species. The calculations suggested that the cationic catalyst promotes a partially different mechanistic pathway (see Scheme 7 for the mechanism and Figure 10 for calculated energy profile), which in turn causes the alteration of the selectivity.

Scheme 7

Scheme 7. Summary of the Mechanism of the [Rh(dppe)]+-Catalyzed 1:2 Coupling of Aldehydes and Allenes

Figure 10

Figure 10. Free energy profile for the [Rh(dppe)]+-catalyzed 1:2 coupling of aldehydes and allenes.

The first two steps of the mechanism are the same as for the [RhCl(dppe)] catalyst, except for the fact that because of the availability of an extra coordination site in the cationic catalyst, the initial INT1+ complex contains two allene molecules. The reaction starts with oxidative coupling of the two allene moieties (TS1+), resulting in the selective formation of η31-bis(allylic) complex INT2+. The structure of the allylic ligand in INT2+ is identical to that in INT2, thus limiting the possible reaction products to 36 as discussed above for the neutral [RhCl(dppe)] catalyst. It should be noted that this step is also rate-determining for this catalyst. The next step is allylation of the aldehyde via TS2+ and TS2′+, resulting in intermediates INT4+ and INT4′+, respectively. The selectivity of the allylation is the same as before, that is, the allylation occurs preferentially at the C3 position via TS2+.
From INT4+ and INT4′+, the reaction can follow the same isomerization/β-hydride elimination/reductive elimination pathway found for the neutral catalyst. The optimized geometries and calculated energy profile for this scenario are given in the Supporting Information. Similarly to the neutral catalyst, the highest barrier was calculated to occur for the isomerization (TS3+ and TS3′+), amounting to 12.0 and 12.2 kcal/mol relative to INT4+ and INT4′+, respectively. However, although the energy barriers are feasible, if such an analogous pathway were followed, it would lead to the same selectivity as before (i.e., predominantly product 3; see the Supporting Information). This is in conflict with the experimental findings, which thus speaks against such a mechanism for the cationic catalyst.
Instead, we discovered an alternative mechanism that reproduces the reversed selectivity observed with the cationic catalyst. The alkoxide ligand present in the Rh(III) complexes INT4+ and INT4′+ can undergo a feasible direct oxidation to give the corresponding ketones (TS6+ and TS6′+; see Figure 11 for the optimized structures). In the course of the reaction, Rh(III) is reduced to Rh(I) and the final products 3 and 5 are formed directly (coordinated to the metal as INT9′+ and INT9+, respectively). A key role in this process is played by the η1-allyl ligand, which acts as an internal base to abstract a proton from the alkoxide (Figure 11). Although perhaps somewhat surprising in this context, this transformation is analogous to the well-known oxidation of alcohols by high-valent metals, such as Mn(VII) and Cr(VI). (27) The example of related oxidation in Pd(II) complexes has also been reported. (28)

Figure 11

Figure 11. Optimized structures of the alkoxide oxidation transition states TS6+ and TS6′+.

The barriers for TS6+ and TS6′+ relative to INT4+ and INT4′+, respectively, were calculated to be 10.1 and 12.0 kcal/mol, respectively. As shown in the free energy profile in Figure 10, the regioselectivity in this case is controlled by the free energy difference between TS2′+ and TS6+, which constitute the first irreversible steps in the respective pathways leading to the two products. The calculated free energy difference between these transition states is 0.3 kcal/mol, corresponding to a 3:5 ratio of 40:60.
Although the difference in the free energy barriers for the isomerization/β-hydride elimination/reductive elimination pathway (12.0 and 12.2 kcal/mol for TS3+ and TS3′+, respectively) and the alkoxide oxidation pathway (10.1 and 12.0 kcal/mol for TS6+ and TS6′+, respectively) is not so large that the former one can be completely ruled out, the fact that the alkoxide oxidation mechanism can account for the reversal of selectivity in the cationic catalyst is a strong argument in favor of it. Another piece of evidence is that the alkoxide oxidation via TS6+ and TS6′+ leads directly to compounds 3 and 5, which explains why these are the only products observed experimentally in the reaction involving the cationic catalyst, while for the neutral catalyst, on the other hand, three products are observed.
Finally, it should be mentioned that we also located transition states for a similar alkoxide oxidation mechanism in the chloride-containing neutral complexes INT4 and INT4′. However, the barriers were found to be much higher than those in the isomerization/β-hydride elimination/reductive elimination pathway discussed above (see the Supporting Information for details). Hence, the electron-deficient character of rhodium in the cationic complex is crucial for the alkoxide oxidation reaction.

4 Conclusions

Click to copy section linkSection link copied!

The rhodium-catalyzed 1:2 coupling of aldehydes and allenes has been investigated in detail by DFT calculations. The free energy profiles for several possible reaction pathways have been calculated and compared.
For the reaction catalyzed by the neutral [RhCl(dppe)] complex, it has been shown that the energetically most plausible catalytic cycle consists of the following steps: oxidative coupling of the two allenes, allylation of the aldehyde, β-hydride elimination, and finally reductive elimination. The first step was found to be rate-determining for the overall reaction, and it leads to the formation of a bis(allylic) Rh(III) complex. Because of its ability to adopt either an η3 or η1 configuration, the allyl ligand turns out to play a key role throughout the mechanism, with profound consequences for the reaction selectivity.
The calculations showed that the remarkable regioselectivity of the reaction is an overall result of a number of selection events that take place in the catalytic cycle. The first regioselectivity-determining step is the oxidative coupling, in which the C2–C2 linkage between the allene moieties is selectively established. Subsequently, the selectivity of the second C–C bond formation is controlled during the β-hydride elimination. Finally, the reductive elimination was found to yield the branched product preferentially.
For the reaction catalyzed by the cationic [Rh(dppe)]+ complex, on the other hand, the calculations suggested a different reaction mechanism after the allylation step. Namely, oxidation of the alkoxide can take place directly to yield the products. This important difference explains the experimentally observed reversal of selectivity.
The present calculations thus provide important insights into the mechanism of rhodium(I)-catalyzed coupling reactions, in particular those involving allenes, and will have a general bearing on the improvement of existing catalytic systems and the design of new ones.

Supporting Information

Click to copy section linkSection link copied!

Complete ref 13, additional results not shown in the text, and Cartesian coordinates of all optimized structures discussed in the paper. This material is available free of charge via the Internet at http://pubs.acs.org.

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • Fahmi Himo - Department of Organic Chemistry, Arrhenius Laboratory, Stockholm University, SE-106 91 Stockholm, Sweden
  • Authors
    • Genping Huang - Department of Organic Chemistry, Arrhenius Laboratory, Stockholm University, SE-106 91 Stockholm, Sweden
    • Marcin Kalek - Department of Organic Chemistry, Arrhenius Laboratory, Stockholm University, SE-106 91 Stockholm, Sweden
  • Notes
    The authors declare no competing financial interest.

Acknowledgment

Click to copy section linkSection link copied!

We acknowledge financial support from the Swedish Research Council, the Göran Gustafsson Foundation, and the Knut and Alice Wallenberg Foundation. G.H. thanks the Carl Trygger Foundation for a postdoctoral fellowship. Computer time was generously provided by the Swedish National Infrastructure for Computing.

References

Click to copy section linkSection link copied!

This article references 28 other publications.

  1. 1
    (a) Lautens, M.; Klute, W.; Tam, W. Chem. Rev. 1996, 96, 49 92
    (b) Kotha, S.; Brahmachary, E.; Lahiri, K. Eur. J. Org. Chem. 2005, 4741 4767
    (c) Murakami, M. Angew. Chem., Int. Ed. 2003, 42, 718 720
    (d) Inglesby, P. A.; Evans, P. A. Chem. Soc. Rev. 2010, 39, 2791 2805
    (e) Shibata, T.; Tsuchikama, K. Org. Biomol. Chem. 2008, 6, 1317 1323
    (f) Tanaka, K. Chem.—Asian J. 2009, 4, 508 518
    (g) Yu, Z.-X.; Wang, Y.; Wang, Y. Chem.—Asian J. 2010, 5, 1072 1088
  2. 2
    (a) Willis, M. C. Chem. Rev. 2010, 110, 725 748
    (b) Jun, C.-H.; Moon, C. W.; Lee, D.-Y. Chem.—Eur. J. 2002, 8, 2422 2428
    (c) Leung, J. C.; Krische, M. J. Chem. Sci. 2012, 3, 2202 2209
  3. 3
    (a) Jang, H.-Y.; Krische, M. J. Acc. Chem. Res. 2004, 37, 653 661
    (b) Ngai, M.-Y.; Kong, J.-R.; Krische, M. J. J. Org. Chem. 2007, 72, 1063 1072
    (c) Iida, H.; Krische, M. J. Top. Curr. Chem. 2007, 279, 77 104
    (d) Bower, J. F.; Kim, I. S.; Patman, R. L.; Krische, M. J. Angew. Chem., Int. Ed. 2009, 48, 34 46
    (e) Skucas, E.; Ngai, M.-Y.; Komanduri, V.; Krische, M. J. Acc. Chem. Res. 2007, 40, 1394 1401
  4. 4
    Kong, J. R.; Krische, M. J. J. Am. Chem. Soc. 2006, 128, 16040 16041
  5. 5

    The reaction scope was later expanded by inclusion of other aldehydes and imines as coupling partners. See:

    (a) Skucas, E.; Kong, J. R.; Krische, M. J. J. Am. Chem. Soc. 2007, 129, 7242 7243
    (b) Han, S. B.; Kong, J. R.; Krische, M. J. Org. Lett. 2008, 10, 4133 4135
  6. 6
    Toyoshima, T.; Miura, T.; Murakami, M. Angew. Chem., Int. Ed. 2011, 50, 10436 10439
  7. 7
    (a) Oonishi, Y.; Hosotani, A.; Sato, Y. J. Am. Chem. Soc. 2011, 133, 10386 10389
    (b) Hojo, D.; Tanaka, K. Org. Lett. 2012, 14, 1492 1495
  8. 8

    The total of 36 refers only to the number of possible linear compounds containing the reactant molecules coupled in a aldehyde–allene–allene sequence. In fact, two of the above isomers contain two asymmetric carbon atoms and hence can exist as two diastereomers, raising the number of possible products with distinct energies to 38. If cyclic structures and those containing allene–aldehyde–allene sequence are also considered, the number of possible products is even greater.

  9. 9
    Williams, V. M.; Kong, J. R.; Ko, B. J.; Mantri, Y.; Brodbelt, J. S.; Baik, M.-H.; Krische, M. J. J. Am. Chem. Soc. 2009, 131, 16054 16062
  10. 10
    Miura, T.; Biyajima, T.; Toyoshima, T.; Murakami, M. Beilstein J. Org. Chem. 2011, 7, 578 581
  11. 11

    In fact, to the best of our knowledge, all of the transition-metal-catalyzed [n + m + o] cycloadditions involving allenes developed to date incorporate only one or two molecules of allene into the product. See:

    (a) Murakami, M.; Ubukata, M.; Itami, K.; Ito, Y. Angew. Chem., Int. Ed. 1998, 37, 2248 2250
    (b) Shanmugasundaram, M.; Wu, M.-S.; Cheng, C.-H. Org. Lett. 2001, 3, 4233 4236
    (c) Shanmugasundaram, M.; Wu, M.-S.; Jeganmohan, M.; Huang, C.-W.; Cheng, C.-H. J. Org. Chem. 2002, 67, 7724 7729
    (d) Miura, T.; Morimoto, M.; Murakami, M. J. Am. Chem. Soc. 2010, 132, 15836 15838
    (e) Brusoe, A. T.; Alexanian, E. J. Angew. Chem., Int. Ed. 2011, 50, 6596 6600
  12. 12
    (a) McCarren, P. R.; Liu, P.; Cheong, P. H.-Y.; Jamison, T. F.; Houk, K. N. J. Am. Chem. Soc. 2009, 131, 6654 6655
    (b) Liu, P.; McCarren, P.; Cheong, P. H.-Y.; Jamison, T. F.; Houk, K. N. J. Am. Chem. Soc. 2010, 132, 2050 2057
    (c) Liu, P.; Krische, M. J.; Houk, K. N. Chem.—Eur. J. 2011, 17, 4021 4029
  13. 13
    Frisch, M. J.; Gaussian 03, revision D.01; Gaussian, Inc.: Wallingford, CT, 2004.
  14. 14
    (a) Becke, A. D. J. Chem. Phys. 1993, 98, 5648 5652
    (b) Lee, C. T.; Yang, W. T.; Parr, R. G. Phys. Rev. B 1988, 37, 785 789
  15. 15
    Hay, P. J.; Wadt, W. R. J. Chem. Phys. 1985, 82, 270 283
  16. 16
    (a) Barone, V.; Cossi, M. J. Phys. Chem. A 1998, 102, 1995 2001
    (b) Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. J. Comput. Chem. 2003, 24, 669 681
  17. 17
    Grimme, S. J. Comput. Chem. 2006, 27, 1787 1799
  18. 18

    For instance, see:

    (a) Minenkov, Y.; Occhipinti, G.; Jensen, V. R. J. Phys. Chem. A 2009, 113, 11833 11844
    (b) Siegbahn, P. E. M.; Blomberg, M. R. A.; Chen, S.-L. J. Chem. Theory Comput. 2010, 6, 2040 2044
    (c) Harvey, J. N. Faraday Discuss. 2010, 145, 487 505
    (d) McMullin, C. L.; Jover, J.; Harvey, J. N.; Fey, N. Dalton Trans. 2010, 39, 10833 10836
    (e) Lonsdale, R.; Harvey, J. N.; Mulholland, A. J. J. Phys. Chem. Lett. 2010, 1, 3232 3237
    (f) Osuna, S.; Swart, M.; Solà, M. J. Phys. Chem. A 2011, 115, 3491 3496
    (g) Santoro, S.; Liao, R.-Z.; Himo, F. J. Org. Chem. 2011, 76, 9246 9252
    (h) Nordin, M.; Liao, R.-Z.; Ahlford, K.; Adolfsson, H.; Himo, F. ChemCatChem 2012, 4, 1095 1104
    (i) Xu, X.; Liu, P.; Lesser, A.; Sirois, L. E.; Wender, P. A.; Houk, K. N. J. Am. Chem. Soc. 2012, 134, 11012 11025
    (j) Jiménez-Halla, J. O. C.; Kalek, M.; Stawinski, J.; Himo, F. Chem.—Eur. J. 2012, 18, 12424 12436
    (k) Kalek, M.; Himo, F. J. Am. Chem. Soc. 2012, 134, 19159 19169
    (l) Huang, G.; Xia, Y.; Sun, C.; Li, J.; Lee, D. J. Org. Chem. 2013, 78, 988 995
  19. 19

    For selected reports of η3-allyl Rh species, see:

    (a) Evans, P. A.; Lawler, M. J. J. Am. Chem. Soc. 2004, 126, 8642 8643
    (b) Evans, P. A.; Leahy, D. K. Chemtracts 2003, 16, 567 578
    (c) Evans, P. A.; Leahy, D. K.; Slieker, L. M. Tetrahedron: Asymmetry 2003, 14, 3613 3618
    (d) Evans, P. A.; Robinson, J. E.; Moffett, K. K. Org. Lett. 2001, 3, 3269 3271
    (e) Arnold, J. S.; Cizio, G. T.; Nguyen, H. M. Org. Lett. 2011, 13, 5576 5579
    (f) Arnold, J. S.; Cizio, G. T.; Heitz, D. R.; Nguyen, H. M. Chem. Commun. 2012, 48, 11531 11533
    (g) Arnold, J. S.; Stone, R. F.; Nguyen, H. M. Org. Lett. 2010, 12, 4580 4583
    (h) Arnold, J. S.; Nguyen, H. M. J. Am. Chem. Soc. 2012, 134, 8380 8383
    (i) Hayashi, T.; Okada, A.; Suzuka, T.; Kawatsura, M. Org. Lett. 2003, 5, 1713 1715
    (j) Vrieze, D. C.; Hoge, G. S.; Hoerter, P. Z.; Van Haitsma, J. T.; Samas, B. M. Org. Lett. 2009, 11, 3140 3142
    (k) Koschker, P.; Lumbroso, A.; Breit, B. J. Am. Chem. Soc. 2011, 133, 20746 20749
    (l) Lumbroso, A.; Koschker, P.; Vautravers, N. R.; Breit, B. J. Am. Chem. Soc. 2011, 133, 2386 2389
    (m) Nishimura, T.; Hirabayashi, S.; Yasuhara, Y.; Hayashi, T. J. Am. Chem. Soc. 2006, 128, 2556 2557
    (n) Choi, J.-c.; Osakada, K.; Yamamoto, T. Organometallics 1998, 17, 3044 3050
    (o) Barros, H. J. V.; Guimarães, C. C.; dos Santos, E. N.; Gusevskaya, E. V. Organometallics 2007, 26, 2211 2218
    (p) Jiao, L.; Lin, M.; Zhuo, L.-G.; Yu, Z.-X. Org. Lett. 2010, 12, 2528 2531
    (q) Jiao, L.; Lin, M.; Yu, Z.-X. J. Am. Chem. Soc. 2011, 133, 447 461
    (r) Lin, M.; Li, F.; Jiao, L.; Yu, Z.-X. J. Am. Chem. Soc. 2011, 133, 1690 1693
    (s) Lin, M.; Kang, G.-Y.; Guo, Y.-A.; Yu, Z.-X. J. Am. Chem. Soc. 2012, 134, 398 405
    (t) Li, Q.; Yu, Z.-X. Organometallics 2012, 31, 5185 5195
  20. 20
    (a) Szabó, K. J. Chem.—Eur. J. 2004, 10, 5268 5275
    (b) Szabó, K. J. Chem.—Eur. J. 2000, 6, 4413 4421
  21. 21

    There is an additional possibility of a migratory insertion of a third molecule of allene instead of the aldehyde. Such a pathway would eventually lead to [2 + 2 + 2] trimerization of the allene. However, since experimentally allene 2 only dimerizes (see ref 10), this possibility was not considered computationally.

  22. 22

    For selected reviews of allylation of carbonyl compounds, see:

    (a) Denmark, S. E.; Fu, J. Chem. Rev. 2003, 103, 2763 2794
    (b) Kennedy, J. W. J.; Hall, D. G. Angew. Chem., Int. Ed. 2003, 42, 4732 4739
    (c) Yu, C.-M.; Youn, J.; Jung, H.-K. Bull. Korean Chem. Soc. 2006, 27, 463 472
    (d) Marek, I.; Sklute, G. Chem. Commun. 2007, 1683 1691
    (e) Hall, D. G. Synlett 2007, 1644 1655
  23. 23

    For examples of [2 + 2 + 2] cyclization involving carbonyl compounds, see:

    (a) Tsuchikama, K.; Yoshinami, Y.; Shibata, T. Synlett 2007, 1395 1398
    (b) Tanaka, K.; Otake, Y.; Wada, A.; Noguchi, K.; Hirano, M. Org. Lett. 2007, 9, 2203 2206
  24. 24

    Transition states connecting INT6 to INT7 and INT6′ to INT7′ could not be optimized. However, energy scans showed that the barriers are quite low, much lower than those for the direct reductive elimination (see the Supporting Information).

  25. 25

    The methyl group in the η3-allyl complexes INT7 and INT7′, contrary to the case of INT2 discussed above, can undergo a relocation from the anti position to the syn position. This could potentially lead to the formation of isomers of products 4 and 6 containing different configurations at one of the double bonds. However, the barriers for the syn/anti substituent exchange in INT7 and INT7′ were calculated to be much higher than those for the reductive elimination. Therefore, INT7 and INT7′ undergo the reductive elimination before they can isomerize. See the Supporting Information for details.

  26. 26

    The ratios were calculated using the Eyring equation. For example, the ratio of INT6 to INT6′ was calculated as

  27. 27
    Wade, L. G., Jr. Organic Chemistry, 6th ed.; Prentice-Hall: Upper Saddle River, NJ, 2006.
  28. 28
    Nielsen, R. J.; Goddard, W. A., III. J. Am. Chem. Soc. 2006, 128, 9651 9660

Cited By

Click to copy section linkSection link copied!

This article is cited by 22 publications.

  1. Ping Wu, Fanshu Cao, Yu Zhou, Zuqian Xue, Ni Zhang, Lei Shi, Gen Luo. Substrate Facilitating Roles in Rare-Earth-Catalyzed C–H Alkenylation of Pyridines with Allenes: Mechanism and Origins of Regio- and Stereoselectivity. Inorganic Chemistry 2022, 61 (43) , 17330-17341. https://doi.org/10.1021/acs.inorgchem.2c02953
  2. Shihan Liu, Tao Zhang, Lei Zhu, Fenru Liu, Ruopeng Bai, Yu Lan. Layered Chirality Relay Model in Rh(I)-Mediated Enantioselective C–Si Bond Activation: A Theoretical Study. Organic Letters 2020, 22 (6) , 2124-2128. https://doi.org/10.1021/acs.orglett.9b04636
  3. Wei Ding, Yan King Terence Ho, Yasuhiro Okuda, Christopher Kevin Wijaya, Zheng Hao Tan, Naohiko Yoshikai. Cobalt-Catalyzed Hydroacylative Dimerization of Allenes Leading to Skipped Dienes. Organic Letters 2019, 21 (15) , 6173-6178. https://doi.org/10.1021/acs.orglett.9b02465
  4. Mei Zhang, Lingfei Hu, Yanmin Lang, Yang Cao, and Genping Huang . Mechanism and Origins of Regio- and Enantioselectivities of Iridium-Catalyzed Hydroarylation of Alkenyl Ethers. The Journal of Organic Chemistry 2018, 83 (5) , 2937-2947. https://doi.org/10.1021/acs.joc.8b00377
  5. Tao Yang and Masahiro Ehara . Computational Studies on Reaction Mechanism and Origins of Selectivities in Nickel-Catalyzed (2 + 2 + 2) Cycloadditions and Alkenylative Cyclizations of 1,6-Ene–Allenes and Alkenes. The Journal of Organic Chemistry 2017, 82 (4) , 2150-2159. https://doi.org/10.1021/acs.joc.6b02957
  6. Stefano Santoro, Marcin Kalek, Genping Huang, and Fahmi Himo . Elucidation of Mechanisms and Selectivities of Metal-Catalyzed Reactions using Quantum Chemical Methodology. Accounts of Chemical Research 2016, 49 (5) , 1006-1018. https://doi.org/10.1021/acs.accounts.6b00050
  7. Jiajia Chen, Wei Guo, and Yuanzhi Xia . Computational Revisit to the β-Carbon Elimination Step in Rh(III)-Catalyzed C–H Activation/Cycloaddition Reactions of N-Phenoxyacetamide and Cyclopropenes. The Journal of Organic Chemistry 2016, 81 (6) , 2635-2638. https://doi.org/10.1021/acs.joc.6b00003
  8. Genping Huang and Peng Liu . Mechanism and Origins of Ligand-Controlled Linear Versus Branched Selectivity of Iridium-Catalyzed Hydroarylation of Alkenes. ACS Catalysis 2016, 6 (2) , 809-820. https://doi.org/10.1021/acscatal.5b02201
  9. Theresa Sperger, Italo A. Sanhueza, Indrek Kalvet, and Franziska Schoenebeck . Computational Studies of Synthetically Relevant Homogeneous Organometallic Catalysis Involving Ni, Pd, Ir, and Rh: An Overview of Commonly Employed DFT Methods and Mechanistic Insights. Chemical Reviews 2015, 115 (17) , 9532-9586. https://doi.org/10.1021/acs.chemrev.5b00163
  10. Genping Huang . Mechanism and Selectivity in Rhodium-Catalyzed [7 + 2] Cycloaddition and Cyclopropanation/Cyclization of Allenylcyclopentane-alkynes: Metallacycle-Directed C(sp3)-C(sp3) vs C(sp3)-H Activation. The Journal of Organic Chemistry 2015, 80 (15) , 7564-7571. https://doi.org/10.1021/acs.joc.5b01148
  11. Tao Liu, Lingli Han, Shumin Han, and Siwei Bi . Theoretical Investigation on Rhodium(I)-Catalyzed Cycloisomerizations of 4-Allenal Species with Linked Alkyne: Ketone vs Alcohol Products. Organometallics 2015, 34 (1) , 280-288. https://doi.org/10.1021/om501118e
  12. Xin Hong, Matthew C. Stevens, Peng Liu, Paul A. Wender, and K. N. Houk . Reactivity and Chemoselectivity of Allenes in Rh(I)-Catalyzed Intermolecular (5 + 2) Cycloadditions with Vinylcyclopropanes: Allene-Mediated Rhodacycle Formation Can Poison Rh(I)-Catalyzed Cycloadditions. Journal of the American Chemical Society 2014, 136 (49) , 17273-17283. https://doi.org/10.1021/ja5098308
  13. Hongli Wu, Xiaojie Li, Xiangyang Tang, Genping Huang. Mechanism and origins of chemo- and regioselectivities of (NHC)NiH-catalyzed cross-hydroalkenylation of vinyl ethers with α-olefins: a computational study. Organic Chemistry Frontiers 2018, 5 (23) , 3410-3420. https://doi.org/10.1039/C8QO01020J
  14. Xiaojie Li, Hongli Wu, Yanmin Lang, Genping Huang. Mechanism, selectivity, and reactivity of iridium- and rhodium-catalyzed intermolecular ketone α-alkylation with unactivated olefins via an enamide directing strategy. Catalysis Science & Technology 2018, 8 (9) , 2417-2426. https://doi.org/10.1039/C8CY00290H
  15. Hongyan Zou, Zhong‐Liang Wang, Genping Huang. Mechanism and Origins of the Chemo‐ and Regioselectivities in Nickel‐Catalyzed Intermolecular Cycloadditions of Benzocyclobutenones with 1,3‐Dienes. Chemistry – A European Journal 2017, 23 (51) , 12593-12603. https://doi.org/10.1002/chem.201702316
  16. Sérgio Filipe Sousa, António J. M. Ribeiro, Rui P. P. Neves, Natércia F. Brás, Nuno M. F. S. A. Cerqueira, Pedro A. Fernandes, Maria João Ramos. Application of quantum mechanics/molecular mechanics methods in the study of enzymatic reaction mechanisms. WIREs Computational Molecular Science 2017, 7 (2) https://doi.org/10.1002/wcms.1281
  17. Li Wang, Hongyan Zou, Xinwen Zhang, Genping Huang. Mechanism and origins of selectivity in rhodium-catalyzed intermolecular [3 + 2] cycloadditions of vinylaziridines with allenes. Organic Chemistry Frontiers 2017, 4 (4) , 587-596. https://doi.org/10.1039/C6QO00671J
  18. B. A. Murray. Reactions of Aldehydes and Ketones and Their Derivatives. 2016, 1-66. https://doi.org/10.1002/9781118707838.ch1
  19. Xinwen Zhang, Hongyan Zou, Genping Huang. Mechanism and Origins of Ligand‐Controlled Selectivity of Rhodium‐Catalyzed Intermolecular Cycloadditions of Vinylaziridines with Alkynes. ChemCatChem 2016, 8 (15) , 2549-2556. https://doi.org/10.1002/cctc.201600349
  20. Mei Zhang, Genping Huang. Mechanism and Selectivity of Ru II ‐ and Rh III ‐Catalyzed Oxidative Spiroannulation of Naphthols and Phenols with Alkynes through a C−H Activation/Dearomatization Strategy. Chemistry – A European Journal 2016, 22 (27) , 9356-9365. https://doi.org/10.1002/chem.201600884
  21. Mei Zhang, Genping Huang. Mechanism of iridium-catalysed branched-selective hydroarylation of vinyl ethers: a computational study. Dalton Transactions 2016, 45 (8) , 3552-3557. https://doi.org/10.1039/C5DT04973C
  22. Elena Soriano, Israel Fernández. Allenes and computational chemistry: from bonding situations to reaction mechanisms. Chemical Society Reviews 2014, 43 (9) , 3041. https://doi.org/10.1039/c3cs60457h

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2013, 135, 20, 7647–7659
Click to copy citationCitation copied!
https://doi.org/10.1021/ja4014166
Published April 25, 2013

Copyright © 2013 American Chemical Society. This publication is licensed under these Terms of Use.

Article Views

3927

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Scheme 1

    Scheme 2

    Scheme 3

    Scheme 3. Possible General Mechanistic Pathways for the Reaction in Scheme 2b, As Suggested by Experimental Findings or Proposed in the Literaturea

    Scheme aEach of the three pathways can yield all of the possible products. Here, only the routes leading to product 3 are shown.

    Figure 1

    Figure 1. Free energy profiles for the possible initial steps of the reaction. The intermediates connecting INT1 with TS1 and TS1b were found to be higher in free energy than INT1, so they are not shown in the diagram (see the Supporting Information for details).

    Figure 2

    Figure 2. Optimized structures of selected intermediates and transition states for the initial step (distances in Å and angles in deg). For clarity, the phenyl substituents on the phosphorus centers have been omitted in the figure.

    Scheme 4

    Scheme 4. Possible Transition States for Oxidative Coupling of the Two Allene Moietiesa

    Scheme aIt should be noted that for each of the TS structures there are varying numbers of Z/E or cis/trans isomers, resulting in an overall total of 24 possible TSs.

    Scheme 5

    Scheme 5. Possible η3–η1–η3 Isomerization of INT2a

    Scheme aFree energies relative to the global zero are indicated in kcal/mol.

    Figure 3

    Figure 3. Free energy profile for the formation of the second C–C bond.

    Figure 4

    Figure 4. Optimized structures of the allylation reactants, transition states, and resulting complexes.

    Figure 5

    Figure 5. Free energy profile for the β-hydride elimination step.

    Figure 6

    Figure 6. Optimized structures of the β-hydride elimination transition states and the resulting η1-allyl complexes.

    Figure 7

    Figure 7. Free energy profile for the reductive elimination step.

    Figure 8

    Figure 8. Optimized structures of the reductive elimination reactants and transition states.

    Scheme 6

    Scheme 6. Summary of the Mechanism of the [RhCl(dppe)]-Catalyzed 1:2 Coupling of Aldehydes and Allenes

    Figure 9

    Figure 9. Free energy profile for the [RhCl(dppe)]-catalyzed 1:2 coupling of aldehydes and allenes.

    Scheme 7

    Scheme 7. Summary of the Mechanism of the [Rh(dppe)]+-Catalyzed 1:2 Coupling of Aldehydes and Allenes

    Figure 10

    Figure 10. Free energy profile for the [Rh(dppe)]+-catalyzed 1:2 coupling of aldehydes and allenes.

    Figure 11

    Figure 11. Optimized structures of the alkoxide oxidation transition states TS6+ and TS6′+.

  • References


    This article references 28 other publications.

    1. 1
      (a) Lautens, M.; Klute, W.; Tam, W. Chem. Rev. 1996, 96, 49 92
      (b) Kotha, S.; Brahmachary, E.; Lahiri, K. Eur. J. Org. Chem. 2005, 4741 4767
      (c) Murakami, M. Angew. Chem., Int. Ed. 2003, 42, 718 720
      (d) Inglesby, P. A.; Evans, P. A. Chem. Soc. Rev. 2010, 39, 2791 2805
      (e) Shibata, T.; Tsuchikama, K. Org. Biomol. Chem. 2008, 6, 1317 1323
      (f) Tanaka, K. Chem.—Asian J. 2009, 4, 508 518
      (g) Yu, Z.-X.; Wang, Y.; Wang, Y. Chem.—Asian J. 2010, 5, 1072 1088
    2. 2
      (a) Willis, M. C. Chem. Rev. 2010, 110, 725 748
      (b) Jun, C.-H.; Moon, C. W.; Lee, D.-Y. Chem.—Eur. J. 2002, 8, 2422 2428
      (c) Leung, J. C.; Krische, M. J. Chem. Sci. 2012, 3, 2202 2209
    3. 3
      (a) Jang, H.-Y.; Krische, M. J. Acc. Chem. Res. 2004, 37, 653 661
      (b) Ngai, M.-Y.; Kong, J.-R.; Krische, M. J. J. Org. Chem. 2007, 72, 1063 1072
      (c) Iida, H.; Krische, M. J. Top. Curr. Chem. 2007, 279, 77 104
      (d) Bower, J. F.; Kim, I. S.; Patman, R. L.; Krische, M. J. Angew. Chem., Int. Ed. 2009, 48, 34 46
      (e) Skucas, E.; Ngai, M.-Y.; Komanduri, V.; Krische, M. J. Acc. Chem. Res. 2007, 40, 1394 1401
    4. 4
      Kong, J. R.; Krische, M. J. J. Am. Chem. Soc. 2006, 128, 16040 16041
    5. 5

      The reaction scope was later expanded by inclusion of other aldehydes and imines as coupling partners. See:

      (a) Skucas, E.; Kong, J. R.; Krische, M. J. J. Am. Chem. Soc. 2007, 129, 7242 7243
      (b) Han, S. B.; Kong, J. R.; Krische, M. J. Org. Lett. 2008, 10, 4133 4135
    6. 6
      Toyoshima, T.; Miura, T.; Murakami, M. Angew. Chem., Int. Ed. 2011, 50, 10436 10439
    7. 7
      (a) Oonishi, Y.; Hosotani, A.; Sato, Y. J. Am. Chem. Soc. 2011, 133, 10386 10389
      (b) Hojo, D.; Tanaka, K. Org. Lett. 2012, 14, 1492 1495
    8. 8

      The total of 36 refers only to the number of possible linear compounds containing the reactant molecules coupled in a aldehyde–allene–allene sequence. In fact, two of the above isomers contain two asymmetric carbon atoms and hence can exist as two diastereomers, raising the number of possible products with distinct energies to 38. If cyclic structures and those containing allene–aldehyde–allene sequence are also considered, the number of possible products is even greater.

    9. 9
      Williams, V. M.; Kong, J. R.; Ko, B. J.; Mantri, Y.; Brodbelt, J. S.; Baik, M.-H.; Krische, M. J. J. Am. Chem. Soc. 2009, 131, 16054 16062
    10. 10
      Miura, T.; Biyajima, T.; Toyoshima, T.; Murakami, M. Beilstein J. Org. Chem. 2011, 7, 578 581
    11. 11

      In fact, to the best of our knowledge, all of the transition-metal-catalyzed [n + m + o] cycloadditions involving allenes developed to date incorporate only one or two molecules of allene into the product. See:

      (a) Murakami, M.; Ubukata, M.; Itami, K.; Ito, Y. Angew. Chem., Int. Ed. 1998, 37, 2248 2250
      (b) Shanmugasundaram, M.; Wu, M.-S.; Cheng, C.-H. Org. Lett. 2001, 3, 4233 4236
      (c) Shanmugasundaram, M.; Wu, M.-S.; Jeganmohan, M.; Huang, C.-W.; Cheng, C.-H. J. Org. Chem. 2002, 67, 7724 7729
      (d) Miura, T.; Morimoto, M.; Murakami, M. J. Am. Chem. Soc. 2010, 132, 15836 15838
      (e) Brusoe, A. T.; Alexanian, E. J. Angew. Chem., Int. Ed. 2011, 50, 6596 6600
    12. 12
      (a) McCarren, P. R.; Liu, P.; Cheong, P. H.-Y.; Jamison, T. F.; Houk, K. N. J. Am. Chem. Soc. 2009, 131, 6654 6655
      (b) Liu, P.; McCarren, P.; Cheong, P. H.-Y.; Jamison, T. F.; Houk, K. N. J. Am. Chem. Soc. 2010, 132, 2050 2057
      (c) Liu, P.; Krische, M. J.; Houk, K. N. Chem.—Eur. J. 2011, 17, 4021 4029
    13. 13
      Frisch, M. J.; Gaussian 03, revision D.01; Gaussian, Inc.: Wallingford, CT, 2004.
    14. 14
      (a) Becke, A. D. J. Chem. Phys. 1993, 98, 5648 5652
      (b) Lee, C. T.; Yang, W. T.; Parr, R. G. Phys. Rev. B 1988, 37, 785 789
    15. 15
      Hay, P. J.; Wadt, W. R. J. Chem. Phys. 1985, 82, 270 283
    16. 16
      (a) Barone, V.; Cossi, M. J. Phys. Chem. A 1998, 102, 1995 2001
      (b) Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. J. Comput. Chem. 2003, 24, 669 681
    17. 17
      Grimme, S. J. Comput. Chem. 2006, 27, 1787 1799
    18. 18

      For instance, see:

      (a) Minenkov, Y.; Occhipinti, G.; Jensen, V. R. J. Phys. Chem. A 2009, 113, 11833 11844
      (b) Siegbahn, P. E. M.; Blomberg, M. R. A.; Chen, S.-L. J. Chem. Theory Comput. 2010, 6, 2040 2044
      (c) Harvey, J. N. Faraday Discuss. 2010, 145, 487 505
      (d) McMullin, C. L.; Jover, J.; Harvey, J. N.; Fey, N. Dalton Trans. 2010, 39, 10833 10836
      (e) Lonsdale, R.; Harvey, J. N.; Mulholland, A. J. J. Phys. Chem. Lett. 2010, 1, 3232 3237
      (f) Osuna, S.; Swart, M.; Solà, M. J. Phys. Chem. A 2011, 115, 3491 3496
      (g) Santoro, S.; Liao, R.-Z.; Himo, F. J. Org. Chem. 2011, 76, 9246 9252
      (h) Nordin, M.; Liao, R.-Z.; Ahlford, K.; Adolfsson, H.; Himo, F. ChemCatChem 2012, 4, 1095 1104
      (i) Xu, X.; Liu, P.; Lesser, A.; Sirois, L. E.; Wender, P. A.; Houk, K. N. J. Am. Chem. Soc. 2012, 134, 11012 11025
      (j) Jiménez-Halla, J. O. C.; Kalek, M.; Stawinski, J.; Himo, F. Chem.—Eur. J. 2012, 18, 12424 12436
      (k) Kalek, M.; Himo, F. J. Am. Chem. Soc. 2012, 134, 19159 19169
      (l) Huang, G.; Xia, Y.; Sun, C.; Li, J.; Lee, D. J. Org. Chem. 2013, 78, 988 995
    19. 19

      For selected reports of η3-allyl Rh species, see:

      (a) Evans, P. A.; Lawler, M. J. J. Am. Chem. Soc. 2004, 126, 8642 8643
      (b) Evans, P. A.; Leahy, D. K. Chemtracts 2003, 16, 567 578
      (c) Evans, P. A.; Leahy, D. K.; Slieker, L. M. Tetrahedron: Asymmetry 2003, 14, 3613 3618
      (d) Evans, P. A.; Robinson, J. E.; Moffett, K. K. Org. Lett. 2001, 3, 3269 3271
      (e) Arnold, J. S.; Cizio, G. T.; Nguyen, H. M. Org. Lett. 2011, 13, 5576 5579
      (f) Arnold, J. S.; Cizio, G. T.; Heitz, D. R.; Nguyen, H. M. Chem. Commun. 2012, 48, 11531 11533
      (g) Arnold, J. S.; Stone, R. F.; Nguyen, H. M. Org. Lett. 2010, 12, 4580 4583
      (h) Arnold, J. S.; Nguyen, H. M. J. Am. Chem. Soc. 2012, 134, 8380 8383
      (i) Hayashi, T.; Okada, A.; Suzuka, T.; Kawatsura, M. Org. Lett. 2003, 5, 1713 1715
      (j) Vrieze, D. C.; Hoge, G. S.; Hoerter, P. Z.; Van Haitsma, J. T.; Samas, B. M. Org. Lett. 2009, 11, 3140 3142
      (k) Koschker, P.; Lumbroso, A.; Breit, B. J. Am. Chem. Soc. 2011, 133, 20746 20749
      (l) Lumbroso, A.; Koschker, P.; Vautravers, N. R.; Breit, B. J. Am. Chem. Soc. 2011, 133, 2386 2389
      (m) Nishimura, T.; Hirabayashi, S.; Yasuhara, Y.; Hayashi, T. J. Am. Chem. Soc. 2006, 128, 2556 2557
      (n) Choi, J.-c.; Osakada, K.; Yamamoto, T. Organometallics 1998, 17, 3044 3050
      (o) Barros, H. J. V.; Guimarães, C. C.; dos Santos, E. N.; Gusevskaya, E. V. Organometallics 2007, 26, 2211 2218
      (p) Jiao, L.; Lin, M.; Zhuo, L.-G.; Yu, Z.-X. Org. Lett. 2010, 12, 2528 2531
      (q) Jiao, L.; Lin, M.; Yu, Z.-X. J. Am. Chem. Soc. 2011, 133, 447 461
      (r) Lin, M.; Li, F.; Jiao, L.; Yu, Z.-X. J. Am. Chem. Soc. 2011, 133, 1690 1693
      (s) Lin, M.; Kang, G.-Y.; Guo, Y.-A.; Yu, Z.-X. J. Am. Chem. Soc. 2012, 134, 398 405
      (t) Li, Q.; Yu, Z.-X. Organometallics 2012, 31, 5185 5195
    20. 20
      (a) Szabó, K. J. Chem.—Eur. J. 2004, 10, 5268 5275
      (b) Szabó, K. J. Chem.—Eur. J. 2000, 6, 4413 4421
    21. 21

      There is an additional possibility of a migratory insertion of a third molecule of allene instead of the aldehyde. Such a pathway would eventually lead to [2 + 2 + 2] trimerization of the allene. However, since experimentally allene 2 only dimerizes (see ref 10), this possibility was not considered computationally.

    22. 22

      For selected reviews of allylation of carbonyl compounds, see:

      (a) Denmark, S. E.; Fu, J. Chem. Rev. 2003, 103, 2763 2794
      (b) Kennedy, J. W. J.; Hall, D. G. Angew. Chem., Int. Ed. 2003, 42, 4732 4739
      (c) Yu, C.-M.; Youn, J.; Jung, H.-K. Bull. Korean Chem. Soc. 2006, 27, 463 472
      (d) Marek, I.; Sklute, G. Chem. Commun. 2007, 1683 1691
      (e) Hall, D. G. Synlett 2007, 1644 1655
    23. 23

      For examples of [2 + 2 + 2] cyclization involving carbonyl compounds, see:

      (a) Tsuchikama, K.; Yoshinami, Y.; Shibata, T. Synlett 2007, 1395 1398
      (b) Tanaka, K.; Otake, Y.; Wada, A.; Noguchi, K.; Hirano, M. Org. Lett. 2007, 9, 2203 2206
    24. 24

      Transition states connecting INT6 to INT7 and INT6′ to INT7′ could not be optimized. However, energy scans showed that the barriers are quite low, much lower than those for the direct reductive elimination (see the Supporting Information).

    25. 25

      The methyl group in the η3-allyl complexes INT7 and INT7′, contrary to the case of INT2 discussed above, can undergo a relocation from the anti position to the syn position. This could potentially lead to the formation of isomers of products 4 and 6 containing different configurations at one of the double bonds. However, the barriers for the syn/anti substituent exchange in INT7 and INT7′ were calculated to be much higher than those for the reductive elimination. Therefore, INT7 and INT7′ undergo the reductive elimination before they can isomerize. See the Supporting Information for details.

    26. 26

      The ratios were calculated using the Eyring equation. For example, the ratio of INT6 to INT6′ was calculated as

    27. 27
      Wade, L. G., Jr. Organic Chemistry, 6th ed.; Prentice-Hall: Upper Saddle River, NJ, 2006.
    28. 28
      Nielsen, R. J.; Goddard, W. A., III. J. Am. Chem. Soc. 2006, 128, 9651 9660
  • Supporting Information

    Supporting Information


    Complete ref 13, additional results not shown in the text, and Cartesian coordinates of all optimized structures discussed in the paper. This material is available free of charge via the Internet at http://pubs.acs.org.


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.