Next Article in Journal / Special Issue
Synthesis and Characterization of Cerium(IV) Metallocenes
Previous Article in Journal / Special Issue
Magnetic and Photo-Physical Properties of Lanthanide Dinuclear Complexes Involving the 4,5-Bis(2-Pyridyl-N-Oxidemethylthio)-4′,5′-Dicarboxylic Acid-Tetrathiafulvalene-, Dimethyl Ester Ligand
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Dehydrocoupling of Alkenylacetylenes Mediated by Various Samarocene Complexes: A Charming Story of Metal Cooperativity Revealing a Novel Dual Metal σ-Bond Metathesis Type of Mechanism (DM|σ-BM)

CNRS (Centre National de la Recherche Scientifique) & INSA (Institut National des Sciences Appliquées), UPS (Université Paul Sabatier), LPCNO (Laboratoire de Physique et Chimie des Nano-objets), Université de Toulouse, 135 Avenue de Rangueil, Toulouse F-31077, France
*
Authors to whom correspondence should be addressed.
Submission received: 30 September 2015 / Revised: 24 November 2015 / Accepted: 26 November 2015 / Published: 4 December 2015
(This article belongs to the Special Issue Rare Earth and Actinide Complexes)

Abstract

:
The prevailing reductive chemistry of Sm(II) has been accessed and explored mostly by the use of samarocene precursors. The highly reducing character of these congeners, along with their Lewis acidity and predominantly ionic bonding, allows for the relatively facile activation of C–H bonds, as well as peculiar transformations of unsaturated substrates (e.g., C–C couplings). Among other important C–C coupling reactions, the reaction of phenylacetylene with different mono- or bimetallic samarocene complexes affords trienediyl complexes of the type {[(C5Me5)2Sm]2(µ-η22-PhC4Ph)}. In contrast, when t-butylacetylene is used, uncoupled monomers of the type (C5Me5)2Sm(C≡C–tBu) were obtained. Although this type of reactivity may appear to be simple, the mechanism underlying these transformations is complex. This conclusion is drawn from the density functional theory (DFT) mechanistic studies presented herein. The operating mechanistic paths consist of: (i) the oxidation of each samarium center and the concomitant double reduction of the alkyne to afford a binuclear intermediate; (ii) the C–H scission of the acetylinic bond that lies in between the two metals; (iii) a dual metal σ-bond metathesis (DM|σ-SBM) process that releases H2; and eventually (iv) the C–C coupling of the two bridged μ-alkynides to give the final bimetallic trienediyl complexes. For the latter mechanistic route, the experimentally used phenylacetylene was considered first as well as the aliphatic hex-1-yne. More interestingly, we shed light into the formation of the mono(alkynide) complex, being the final experimental product of the reaction with t-butylacetylene.

1. Introduction

The versatility of coupling reactions mediated by single-electron transfer (SET) from Sm(II) complexes is clearly illustrated by the chemistry of SmI2 [1,2]. This coupling chemistry has been expanded by the use of (Cp*)2Sm(THF)n (where Cp* = C5Me5; n = 0–2) [3,4,5,6,7]. Interestingly, both the Lewis acidity and the ionic bonding of these Sm(II) complexes offer ligation of substrates that can undergo facile activation of C–H bonds, or partial reduction of unsaturated compounds [8]. These two elementary steps can combine to achieve dehydrocoupling of alkynes, as was initially reported by Evans in the case of phenylacetylene. This led to the characterization of a new class of dinuclear Sm(III) trienediyl complexes, {[(Cp*)2Sm]2(µ-η22-PhC4Ph)} [9]. These trienediyl species were obtained by four different reaction procedures: (a) the reaction of (Cp*)2Sm[CH(SiMe3)]2 with HC≡CPh; (b) thermolysis of (Cp*)2Sm(C≡CPh)(THF) at 120–145 °C; (c) the reaction of [(Cp*)2Sm(μ-H)]2 with HC≡CPh; and (d) the reaction of (Cp*)2Sm with HC≡CPh. This reactivity was also extended to include aliphatic terminal alkynes, HC≡CR (R = CH2CH2Ph, iPent, and iPr), leading to similar dinuclear trienediyl complexes [7]. However, when a more bulky acetylide is used (R = tBu), the reaction halts at the formation of the alkynide monomers, e.g., the uncoupled dimer of the type [(Cp*)2Sm(C≡C–tBu)]2, which was prepared from the use of two different precursors; [(Cp*)2SmIII(μ-H)]2, and (Cp*)2SmII(THF) [7]. Alkyne dehydrocoupling has also been observed with copper salts [10] and titanium complexes [11].
Around the same period of time as the Evans report, two related contributions appeared in the literature. In particular, Teuben et al. [12] reported a study on early lanthanide carbyls, e.g., (Cp*)2LnCH(SiMe3)2, that promote C–C coupling of terminal alkynes (HC≡CR) to give the corresponding trienediyl bimetallic species, {[(Cp*)2Ln]2(µ-η22-RC4R)} (Ln = Ce, R = Me or tBu; Ln = La, R = Me). At the same time, Marks et al., also reported the reaction of (Cp*)2LaCH(SiMe3) with HC≡CR (R = tBu and Ph) to form the corresponding trienediyl dinuclear complexes [13]. The latter group concluded that the uncoupled dimer [(Cp*)2La(C≡C–R)]2 is the direct kinetic precursor of this reaction, and that neither a redox-active lanthanide ion nor a phenyl substituent on the alkyne is required for its completion. In a recent report from our group, we were able to shed light onto this type of reactivity by means of computational techniques [14]. Nevertheless, up until now, there has not been a clear picture of the alkyne coupling mechanism leading to dinuclear Sm(III) trienediyl complexes starting from low-valent samarocene (Figure 1).
Figure 1. Reaction of samarocene with phenylacetylene to afford {[(C5Me5)2Sm]2(µ-η22-PhC4–Ph)} complex and H2.
Figure 1. Reaction of samarocene with phenylacetylene to afford {[(C5Me5)2Sm]2(µ-η22-PhC4–Ph)} complex and H2.
Inorganics 03 00573 g001
The bimetallic mechanism which was previously proposed in the case of N-anchored tris(aryloxide) uranium(III) can serve as a base [15], but how acetylenic C–H bonds break to release H2 still needs to be answered. Hence, in this paper we attempt to understand the performance of samarocene in the formation of the trienediyl dinuclear complex, {[(Cp*)2Sm]2(µ-η22-PhC4–Ph)}, with plausible energetic reaction profiles being computed and suggested at the DFT (B3PW91) level of theory. The formation of the uncoupled complexes, instead of the trienediyl one where tBuC≡CH was used, is also discussed. Based on these studies, novel transition states which lead to unique reaction pathways in the domain of f-block chemistry are revealed.

2. Results and Discussion

2.1. Dehydrocoupling of PhC≡CH Using Cp*2Sm

Based on chemical intuition and previous knowledge on related mechanisms [15], three different types of reaction sequence are envisioned for this particular reactivity (Figure 2).
Figure 2. Possible reaction paths A, B, and C leading to the formation of the corresponding trienediyl complex using PhC≡CH as the substrate and (Cp*)2Sm as a precursor.
Figure 2. Possible reaction paths A, B, and C leading to the formation of the corresponding trienediyl complex using PhC≡CH as the substrate and (Cp*)2Sm as a precursor.
Inorganics 03 00573 g002
All these sequences share a common initiation step, the double SET one, in which a doubly reduced phenylacetylene unit contained in a dinuclear samarium(III) complex is formed. From this “key intermediate” the reaction can proceed either through: (A) a direct insertion of PhC≡CH that leads to an alkynyl bis-Sm(III) complex; (B) a peculiar direct H–H coupling type of Transition State (TS); (C) a C–H activation leading to a μ-H:μ-C motif. Energetically the most favorable pathway relies on the C–H activation that corresponds to mechanism C and will be discussed in detail afterwards, together with the other routes. The whole energy profile for mechanism C is depicted in Figure 3.
Figure 3. Enthalpy energy ΔH (kcal·mol−1) profile for the formation of the corresponding trienediyl complex using PhC≡CH as substrate and (Cp*)2Sm as a precursor (mechanism C). Cp* ligands are omitted for clarity. Numbers in blue correspond to natural charges.
Figure 3. Enthalpy energy ΔH (kcal·mol−1) profile for the formation of the corresponding trienediyl complex using PhC≡CH as substrate and (Cp*)2Sm as a precursor (mechanism C). Cp* ligands are omitted for clarity. Numbers in blue correspond to natural charges.
Inorganics 03 00573 g003
As already reported by us in a recent theoretical work, the one electron reduction of phenylacetylene, induced by the coordination to the samarium center, is a highly exothermic process [16]. This is the outcome of high level ab-initio calculations (CAS-SCF), but it is also based on the newly introduced theoretical methodology that takes into account the “HOMO-LUMO gap” energy difference. As the overall redox process is known to be exothermic, it contributes to the total energy stabilization of the newly formed “key intermediate” [17]. Hence, the latter species will be the reference point for the whole mechanism. In the 1-Ph intermediate, the α- and β-carbons of the alkylenic substrate bind to each distinct Sm center in an E-configuration, with the Z-isomer being less stable by 3.6 kcal·mol−1 in accordance with earlier structural interpretation [18]. Interestingly, in 1-Ph, the C–C bond distance of the sandwiched alkyne is 1.36 Å which is 0.15 Å longer than in free phenylacetylene. This is a strong indication of the double reduction of the alkyne. In addition, the acetylenic C–H bond points towards one Sm center and is elongated by 0.06 Å with respect to free phenylacetylene, being readily available for a potential bond scission. At that point, the subsequent C–H bond activation (TS12-Ph) yielding a μ-hydride-µ-alkynyl dinuclear complex (2-Ph) can take place with a moderate energy barrier of 17.5 kcal·mol−1 with respect to the “key intermediate”. Hoffmann et al., showed in a seminal work that extra stabilization is achieved by the mixing of the 1s orbital of the hydride with the available empty dz2 and dyz orbitals, of 1a1 and 1b2 symmetry respectively, of each Sm(Cp*)2 fragment [19]. This can nicely explain the relative low energy barrier computed for this step. It is noteworthy that this bridged dimer shares features with geometrically close related yttrium complexes that were previously reported [20,21]. The reaction proceeds via TS23-Ph that corresponds to the activation of the acetylenic proton of a second PhC≡CH molecule by the bridging hydride of complex 2-Ph. This step is better described as a six-member σ-bond metathesis showing a negligible activation energy barrier. The alternation of the natural charge signs between the six atoms in TS23-Ph (See Figure 3) and the respective shortening and lengthening of the bonds involved, confirms its σ-bond metathesis character. This is in line with previous studies [22,23,24]. To the best of our knowledge, this Dual Metal σ-Bond Metathesis transition state (DM|σ-BM) is the first report of such a type that implies two lanthanide centers working in concert. This type of transition state can be also found in other related lanthanide types of reactivity [25]. The latter step affords a μ-alkynyl-η1-alkynyl dihydrogen complex (3-Ph) that is almost isoenergetic with the “key intermediate”. In the subsequent step an isomerization takes place which induces the release of H2, and leads to the formation of bis-µ-phenylacetylenyl complex, 4-Ph. The formation of the latter requires a very low activation barrier (ΔrH# = 3.2 kcal·mol−1). From 4-Ph, the final trienediyl dinuclear complex 5-Ph is obtained by the C–C homocoupling of the two terminal µ-alkynyl moieties via TS45-Ph. Interestingly, similar transition states were postulated to proceed in analogous systems, as for instance the bis-μ-alkynyl titanocene complexes [26]. In addition, analogous structures were structurally isolated and characterized for other lanthanide centers [12,27,28,29]. As already reported from our group, the energy barrier for the C–C coupling step is surprisingly low, being only 10.7 kcal·mol−1 [14]. This is odd since two negatively charged moieties have to be homocoupled. One way to accomplish this is to reduce the electron density of the negatively charged α-carbons of the alkynyls. This is achieved by the nucleophilic assistance from the β-carbon of each triple bond which accumulates enough negative charge, resulting in a stronger interaction with the second samarium center. Finally, the rate determining step of the entire mechanism is the scission of the C–H acetylenic bond (17.5 kcal·mol−1), with the overall exothermicity of the mechanism being 24.7 kcal·mol−1 with respect to 1-Ph intermediate.
The other two pathways depicted in Figure 1 were considered starting from 1-Ph. In the first case (mechanism A), and inspired by our recent mechanistic work [15], the possibility of a direct C–C coupling in 1,1-fashion was investigated. This led to the corresponding Sm(III) bis-vinyl complex 6-Ph (Figure 4). Although this step is strongly exothermic (−31.0 kcal·mol−1), the activation barrier (ΔrH = 24.8 kcal·mol−1) is much higher than the highest barrier of mechanism C, by more than 7 kcal·mol−1. The reaction involves an ionic transition state, TS16-Ph, in which the incoming triple bond is polarized by one samarium center, allowing the formation of a lone-pair at the β-carbon atom and an empty sp2-orbital at its α-carbon that overlaps with the filled sp2-orbital of the α-carbon of the bridged reduced triple bond. This transition state closely resembles the one reported for the C–C coupling of the hex-1-yne using N-anchored tris-aryloxide complexes of U(III) [15]. Then the reaction proceeds through the activation of the C–H bond in close vicinity to the Sm center, surmounting an even higher activation barrier than the previous step (ΔrH = 25.7 kcal·mol−1). The geometry of the intermediate that results, 7-Ph, corresponds to a weakly coordinated complex, in which two fragments develop an H–H interaction (Figure 4). The latter undergoes a proton transfer to give the trienediyl compound upon release of H2, via a transition state of relatively low activation barrier (ΔrH = 6.9 kcal·mol−1). However, this mechanism can be ruled out from further consideration since it suffers kinetically from two relatively high activation barriers of ca. 24 kcal·mol−1, and thermodynamically by the formation of the highly stable bis-vinyl complex 6-Ph. The latter corresponds to the lowest energy point of the mechanism, being 6.3 kcal·mol−1 lower than the final trienediyl complex, as depicted in Figure 4.
Figure 4. Enthalpy energy ΔH (kcal·mol−1) profile for the formation of the corresponding trienediyl complex using PhC≡CH as substrate and (Cp*)2Sm as a precursor (mechanism A). Cp* ligands are omitted for clarity.
Figure 4. Enthalpy energy ΔH (kcal·mol−1) profile for the formation of the corresponding trienediyl complex using PhC≡CH as substrate and (Cp*)2Sm as a precursor (mechanism A). Cp* ligands are omitted for clarity.
Inorganics 03 00573 g004
Finally, from “key intermediate” 1-Ph, a mechanism that eliminates H2 and homocouples the two alkynyls in a concerted homolytic manner (mechanism B) was attempted. Despite our efforts, we were unable to locate any transition state of this type. Instead, we found the direct formation of H2, but without any C–C coupling, leading to 5-Ph after reorganization (Figure 5, mechanism B). By examining the nature of each hydrogen at the transition state, this can be regarded as an H–H coupling type of TS. Interestingly, the corresponding natural charges of the two hydrogens are −0.01 and 0.10 |e| (Figure 5). The first charge corresponds to the acetylenic bridged hydrogen, while the second belongs to the newly inserted alkyne. Nevertheless, the activation energy barrier for this step (ΔrH = 27.2 kcal·mol−1) allows us to discard this pathway from further consideration.
Figure 5. Enthalpy energy ΔH (kcal·mol−1) profile for the formation of the corresponding trienediyl complex using PhC≡CH as substrate and (Cp*)2Sm as a precursor (Mechanism B). Cp* ligands are omitted for clarity. Numbers in blue correspond to natural charges.
Figure 5. Enthalpy energy ΔH (kcal·mol−1) profile for the formation of the corresponding trienediyl complex using PhC≡CH as substrate and (Cp*)2Sm as a precursor (Mechanism B). Cp* ligands are omitted for clarity. Numbers in blue correspond to natural charges.
Inorganics 03 00573 g005
In contrast, when a monometallic mechanism is operating, the C–C coupling can occur in 1,1 or 1,2 fashion (Figure S1 in supplementary information). In particular, addition of a second alkyne molecule to the (Cp*)2SmIII(HC≡CPh) intermediate can take place in two different ways; in 1,1 and 1,2 fashion, affording the corresponding bis-vinyl species. The activation barriers were found to be 9.8 and 5.9 kcal·mol−1 respectively, with the exothermicities of each one being 31.1 and 27.7 kcal·mol−1 with respect to the monoreduced intermediate, (C5Me5)2Sm(HC≡CPh). The computed small energy difference between the two transition states, as well as the preference for 1,2 C–C coupling, cannot account for the observed experimental selectivity which is found to be the reverse. On top of that, another weak point of the monometallic mechanism is the fact that after the formation of bis-vinyl monomeric complex, an oxidation of a second SmII(Cp*)2 has to take place. This will result in the formation of the same energetically buried bis-vinyl intermediate, 6-Ph, as described previously in mechanism A. Since the formation of the latter is undesired, for the reasons described above, the existence of a monometallic pathway can also be ruled out.

2.2. Dehydrocoupling of tBuC≡CH Using Cp*2Sm

We then turned our attention to include the more bulky H–C≡C–tBu alkyne that is found to give monomeric acetylide complexes rather than bimetallic trienediyl complexes [7], with the mechanism C applied to the new alkyne (Figure 6).
Figure 6. Enthalpy energy ΔH (kcal·mol−1) profile for the formation of the uncoupled alkynide complex using tBuC≡CH as substrate and (Cp*)2Sm as a precursor. Cp* ligands are omitted for clarity.
Figure 6. Enthalpy energy ΔH (kcal·mol−1) profile for the formation of the uncoupled alkynide complex using tBuC≡CH as substrate and (Cp*)2Sm as a precursor. Cp* ligands are omitted for clarity.
Inorganics 03 00573 g006
Unlike the previous case, the first step corresponds to SET but with a concomitant dissociation of a THF molecule; the latter molecule being coordinated to the precursor complex Sm(Cp*)2. It is expected that the high energy gain observed in this step will be sufficient for the easy dissociation of the THF molecule. The latter is computed to cost around 10 kcal·mol−1 as it is depicted in Figure 7a. In addition, the isodesmic reaction of the net exchange of THF with the t-butylacetylene costs only 7.3 kcal·mol−1, indicative of the low amount of energy needed for such a reaction (Figure 7b). It should be noted that all the attempts to locate a local minimum that corresponds to the “key intermediate” with at least one THF coordinated to a samarium center were not successful, due to steric reasons. Hence, most likely the THF dissociates to yield the more energetically stable bimetallic intermediate 1-tBu.
Figure 7. (a) Dissociation reaction of THF from the mono-solvated samarocene complex; (b) Exchange reaction of mono-solvated (THF) samarocene complex with tBuC≡CH.
Figure 7. (a) Dissociation reaction of THF from the mono-solvated samarocene complex; (b) Exchange reaction of mono-solvated (THF) samarocene complex with tBuC≡CH.
Inorganics 03 00573 g007
Therefore, as in the phenylacetylene case, the “key intermediate” 1-tBu is considered as the starting point of the whole enthalpy profile. It is noteworthy that the first two steps are found to possess similar energy variations with that of phenylacetylene. This was not expected initially since the bulkier tBuC≡CH will introduce more steric repulsions when it is sandwiched between the two samarocene moieties. This was especially anticipated in the C–H activation step, being in very close proximity to the two pentamethylcyclopentadienyl groups. Nevertheless, the inherent flexibility of the system probably allows the minimization of such repulsive forces by directing the methyl groups of the tBu to lie in an optimal position among the four Cp* ligands. The only, but major differentiation with respect to the phenylacetylene case is the geometry of the product of the DM|σ-BM step. The outcome of this process, computationally, is the disruption of the bimetallic complex into two monomeric alkynyl complexes, with one bearing a weakly coordinated dihydrogen molecule. Hence, instead of the generation of a μ-alkynyl-η1-alkynyl dihydrogen-like complex, as is found in the PhC≡CH case, e.g., 3-Ph, the formation of the two monomers leads to a substantial energy stabilization (being 10.3 kcal·mol−1 lower than the reference point). Then the system releases H2 very easily to gain an extra energy stabilization of 1.4 kcal·mol−1 and to yield consequently the well separated alkynyl complexes 3′-tBu. The latter were experimentally observed as the only product of such reactivity [7]. Following this, the reaction can potentially proceed into the subsequent C–C coupling requiring a considerable amount of energy (ΔrH = 27.7 kcal·mol−1), kinetically hardly accessible [14]. Overall, the formation of the separated alkynyl species corresponds to the lowest point of the whole reaction route, being in perfect agreement with the experimental observations.

2.3. Dehydrocoupling of Hex-1-yne Using Cp*2Sm

To understand the effect induced by an alkyl substituent on the alkyne, we considered computationally the aliphatic hex-1-yne in order to check how this differentiates from the bulky tBuC≡CH, and the PhC≡CH in which there is an extended conjugation of the triple bond with the phenyl ring. Even though the gain in energy is large for the double reduction of the hex-1-yne, as it was computed in a previous work by our group [16], it is half of that obtained for the phenylacetylene, possibly due to the delocalization of the extra electron(s) into the phenyl ring. In Figure 8, the most favorable reaction pathway is presented, reverting again to mechanism C.
Figure 8. Enthalpy energy ΔH (kcal·mol−1) profile for the formation of the trienediyl complex using nBuC≡CH as substrate and (Cp*)2Sm as a precursor. Cp* ligands are omitted for clarity. Numbers in blue correspond to natural charges.
Figure 8. Enthalpy energy ΔH (kcal·mol−1) profile for the formation of the trienediyl complex using nBuC≡CH as substrate and (Cp*)2Sm as a precursor. Cp* ligands are omitted for clarity. Numbers in blue correspond to natural charges.
Inorganics 03 00573 g008
The geometry of the “key intermediate”, 1-nBu, is similar to that found in the two previously described cases with the α- and β-carbons of the doubly reduced triple bond adopting a zig-zag geometry. In the same way, the acetylenic C–H bond is elongated by 0.08 Å (with respect to the free hex-1-yne) being already prepared for the forthcoming intramolecular C–H bond scission. Next, the aforementioned C–H bond activation takes place yielding a µ-hydride-µ-alkynyl dinuclear intermediate with an enthalpy activation barrier of 14.4 kcal·mol−1. This activation barrier is lower than the two previous by almost 4 kcal·mol−1. This is probably attributed to the reduced steric hindrance that substituents as nBu-, and Ph- to a smaller extent, are inducing compared to the tBu- one. The reaction can further proceed with the addition of a second hex-1-yne equivalent to form the bis-µ-alkynyl complex, 4-nBu, after release of H2. This is achieved by passing through the TS23-nBu, which corresponds to a barrierless process. The last step corresponds to the homo C–C coupling of the two µ-alkynyl moieties to afford the final trienediyl dinuclear complex. The activation barrier of this step (ΔrH = 16.9 kcal·mol−1), is the highest of the entire mechanism, being consequently the rate determining step, contrary to the previous cases. Finally, this process is exothermic by 16.4 kcal·mol−1, while the overall reaction exothermicity is found to be 23.6 kcal·mol−1. Therefore, we conclude that the tested alkyne can be possibly used experimentally to afford, upon reaction with the Cp*2Sm complex, the corresponding trienediyl species. Concluding, apart from some small stabilization observed at the C–H bond activation transition state and the formation of the uncoupled dimeric species, 4-nBu, with respect to the other two cases, the nBu- substituent does not play any decisive role in the feasibility of such reactivity.

2.4. Dehydrocoupling of Various Terminal Alkynes Using [Cp*2Sm(μ-H)2]2 Complex

As described in the introduction, two experiments were conducted by Evans et al., for the formation of the potential trienediyl complexes using tBuC≡CH. The first was already investigated in detail in Section 2.2. In the second experiment, the bis-hydride dimer of Sm(III) was used, but as in the first case, it did not lead to the expected homocoupled product but instead to the formation of the uncoupled alkynide complexes as major product. It should also be noted that when PhC≡CH is considered using the same precursor, [Cp*2Sm(μ-H)2]2, it yields the corresponding trienediyl complex. Therefore, a mechanistic investigation at the B3PW91 level of theory was carried out in order to discover the reason for this experimentally observed discrepancy and to propose a rational mechanism. As a reasonable mechanistic scenario we envisioned two consecutive (DM|σ-BM–H2 release) steps (Figure 9a). Interestingly, for the phenylacetelyne case, the first DM|σ-BM type of transition state surmounts a relative low activation barrier (ΔrH = 10.5 kcal·mol−1). The intermediate 2-Ph formed after the H2 release (the isomerization step) is energetically stable enough, being 7.9 kcal·mol−1 lower than the starting material. The following part of the mechanism is essentially the same as proposed in mechanism C and consequently will not be discussed further. Finally, it is worth noting that a potential equilibrium between the dimer 9 and its mono-hydride structure was found computationally in favor of the dimer by 16.5 kcal·mol−1 in terms of enthalpy energy, in direct agreement with the experimentally observed thermal stability of 9 [18]. Hence, any mechanistic scenario starting from the monomeric complex was not considered in our studies.
Figure 9. Enthalpy energy ΔH (kcal·mol−1) profiles for the formation of the corresponding trienediyl complexes using (a) PhC≡CH, and (b) tBuC≡CH as substrates, and [(Cp*)2Sm(μ-H)]2 as a precursor. Cp* ligands are omitted for clarity.
Figure 9. Enthalpy energy ΔH (kcal·mol−1) profiles for the formation of the corresponding trienediyl complexes using (a) PhC≡CH, and (b) tBuC≡CH as substrates, and [(Cp*)2Sm(μ-H)]2 as a precursor. Cp* ligands are omitted for clarity.
Inorganics 03 00573 g009
The same energy reaction pattern was considered in the case of tBuC≡CH. Again, the first step corresponds to a low energy process, even lower than for the phenylacetylene, with sufficient energy stabilization upon H2 release, being almost isoenergetic to that obtained for PhC≡CH. However, unlike PhC≡CH, the second DM|σ-BM step yields the uncoupled alkynyl complex, as was the case in the divalent samarocene, from which the C–C coupling is not kinetically easily accessible (Figure 9b). Hence, such type of two consecutive (DM|σ-BM–H2 release) steps can serve as an explanation of the experimentally observed reactivity, shedding ample light into this peculiar reactivity in lanthanide chemistry.

3. Computational Section

All the quantum calculations, required for the delineation of the intermediate and transition state molecular structures, were performed at the density functional theory level using the B3PW91 [30,31] hybrid functional as implemented in the Gaussian program code [32]. The basis set used for the samarium atoms was the Stuttgart–Dresden–Koln (large-core) [33], augmented by an f polarization function in conjunction with the appropriate scalar relativistic pseudopotential [34]. For all the remaining atoms the all-electron double zeta basis set 6–31 G (d,p) was used [35,36]. Enthalpy energies were obtained at T = 298.15 K based on the harmonic approximation. Intrinsic Reaction Paths (IRPs) were traced from the various transition structures to verify the reactant to product linkage [37,38]. Natural population analysis (NPA) was performed using Weinhold’s methodology [39,40].

4. Conclusions

Although the reaction of a low-valent samarocene with terminal alkynes affording a trienediyl binuclear complex and H2 seems simple, the mechanism underlying this transformation is mechanistically complicated. In the present DFT mechanistic study, the cooperativity of the two lanthanides is highlighted, with a unique mechanistic scenario being proposed, involving novel type of transition states, such as the DM|σ-SBM, for example. In this particular saddle point of the potential energy surface, the two samarium centers are working synergistically to facilitate H2 elimination, by forming a six-member transition state. It is worth noting that this type of transition state consists of a novel paradigm of σ-bond metathesis that can be added to the already well-established related types [41]. Among other mechanistic paths, the operating mechanism consists of (i) the oxidation of each samarium center and the concomitant double reduction of the alkyne to afford a binuclear intermediate; (ii) the C–H scission of the acetylinic bond that lies in between the two metals; (iii) a DM|σ-SBM process that releases the H2 molecule; (iv) the C–C coupling of the two bridged μ-alkynides to give the final experimentally observed bimetallic trienediyl complex. For the latter mechanistic route the experimentally used substrate, phenylacetylene, was considered first, and afterwards the hex-1-yne in order to check the effect of the alkyl substituent. For the aliphatic alkyne, the applied mechanism leads to the same conclusions in terms of enthalpy of reaction, making this process feasible experimentally. On the other hand, the lack of obtaining the trienediyl analogs when the tBuC≡CH is used, as reported experimentally, is due to the thermodynamic preference for disruption of the bimetallic product of the DM|σ-SBM step over the homocoupling of the two bulky alkynyl moieties. We strongly believe that this computational contribution sheds light into the mechanism at work in this peculiar reactivity, and will help in the direction of additional understanding in the future of organolanthanide chemistry.

Supplementary Materials

Supplementary materials can be found at https://0-www-mdpi-com.brum.beds.ac.uk/2304-6740/3/4/0573/s1.

Acknowledgments

Laurent Maron is member of the Institute Universitaire de France. CINES (Centre Informatique National de l’Enseignement Supérieur) and CALMIP (Calcul en Midi-Pyrénées) are acknowledged for a generous grant of computing time. The Humboldt Foundation and The Chinese Academy of Science are also acknowledged. The authors are grateful to Lionel Perrin for fruitful discussions and comments on various aspects of the present work.

Author Contributions

Laurent Maron conceived the project; Christos E. Kefalidis performed the calculations; Laurent Maron and Christos E. Kefalidis analyzed the data and wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Krief, A.; Laval, A.-C. Coupling of organic halides with carbonyl compounds promoted by SmI2, the Kagan reagent. Chem. Rev. 1999, 99, 745–777. [Google Scholar] [CrossRef] [PubMed]
  2. Procter, D.J.; Flowers, R.A., II; Skrydstrup, T. Organic Synthesis Using Samarium Diiodide: A Practical Guide; Royal Society of Chemistry Publishing: Cambridge, UK, 2010. [Google Scholar]
  3. Evans, W.J.; Chamberlain, L.R.; Ulibarri, T.A.; Ziller, J.W. Reactivity of trimethylaluminum with (C5Me5)2Sm(THF)2: Synthesis, structure, and reactivity of the samarium methyl complexes (C5Me5)2Sm[(μ-Me)AlMe2(μ-Me)]2Sm(C5Me5)2 and (C5Me5)2SmMe(THF). J. Am. Chem. Soc. 1988, 110, 6423–6432. [Google Scholar] [CrossRef]
  4. Nolan, S.P.; Stern, C.L.; Marks, T.J. Organo-f-element thermochemistry. Absolute metal-ligand bond disruption enthalpies in bis(pentamethylcyclopentadienyl)samarium hydrocarbyl, hydride, dialkylamide, alkoxide, halide, thiolate, and phosphide complexes. Implications for organolanthanide bonding and reactivity. J. Am. Chem. Soc. 1989, 111, 7844–7853. [Google Scholar]
  5. Evans, W.J.; Ulibarri, T.A.; Ziller, J.W. Reactivity of (C5Me5)2Sm with aryl-substituted alkenes: Synthesis and structure of a bimetallic styrene complex that contains an η2-arene lanthanide interaction. J. Am. Chem. Soc. 1990, 112, 219–223. [Google Scholar] [CrossRef]
  6. Evans, W.J.; Ulibarri, T.A.; Ziller, J.W. Reactivity of (C5Me5)2Sm and related species with alkenes: Synthesis and structural characterization of a series of organosamarium allyl complexes. J. Am. Chem. Soc. 1990, 112, 2314–2324. [Google Scholar] [CrossRef]
  7. Evans, W.J.; Keyer, R.A.; Ziller, J.W. Investigation of organolanthanide-based carbon–carbon bond formation: Synthesis, structure, and coupling reactivity of organolanthanide alkynide complexes, including the unusual structures of the trienediyl complex [(C5Me5)2Sm]2[μ-η22-Ph(CH2)2C=C=C=C–(CH2)2Ph] and the unsolvated alkynide [(C5Me5)2Sm(C≡CCMe3)]2. Organometallics 1993, 12, 2618–2633. [Google Scholar]
  8. Evans, W.J.; Davis, B.L. Chemistry of tris(pentamethylcyclopentadienyl) f-element complexes, (C5Me5)3M. Chem. Rev. 2002, 102, 2119–2136. [Google Scholar] [CrossRef] [PubMed]
  9. Evans, W.J.; Keyer, R.A.; Ziller, J.W. Carbon–carbon bond formation by coupling of two phenylethynyl ligands in an organolanthanide system. Organometallics 1990, 9, 2628–2631. [Google Scholar] [CrossRef]
  10. Wood, G.L.; Knobler, C.B.; Hawthorne, M.F. Synthesis of Cp2Ti[C≡CSi(CH3)3]2 and the synthesis and molecular structure of [Cp2TiC≡CSi(CH3)3]2. Inorg. Chem. 1989, 28, 382–384. [Google Scholar] [CrossRef]
  11. Sekutowski, D.G.; Stucky, G.D. Oxidative coupling of the phenylethynyl group in μ-(1-3η:2-4η-trans,trans-1,4-diphenylbutadiene)-bis(bis(η5-methylcyclopentadienyl)titanium) and the reaction of 1,4-diphenyl-1,3-butadiene with bis(cyclopentadienyl)titanium(II). J. Am. Chem. Soc. 1976, 98, 1376–1382. [Google Scholar] [CrossRef]
  12. Heeres, H.J.; Nijhoff, J.; Teuben, J.H. Reversible carbon–carbon bond formation in organolanthanide systems. Preparation and properties of lanthanide acetylides [Cp*2LnC≡CR]n and their rearrangement products [Cp*2Ln]2-μ-η22-RC4R) (Ln = La, Ce; R = alkyl). Organometallics 1993, 12, 2609–2617. [Google Scholar] [CrossRef]
  13. Forsyth, C.M.; Nolan, S.P.; Stern, C.L.; Marks, T.J.; Rheingold, A.L. Alkyne coupling reactions mediated by organolanthanides. Probing the mechanism by metal and alkyne variation. Organometallics 1993, 12, 3618–3623. [Google Scholar] [CrossRef]
  14. Kefalidis, C.E.; Perrin, L.; Maron, L. Computational insights into carbon–carbon homocoupling reactions mediated by organolanthanide(III) complexes. Dalton Trans. 2014, 43, 4520–4529. [Google Scholar] [CrossRef] [PubMed]
  15. Kosog, B.; Kefalidis, C.E.; Heinemann, F.W.; Maron, L.; Meyer, K. Uranium(III)-mediated C–C-coupling of terminal alkynes: Formation of dinuclear uranium(IV) vinyl complexes. J. Am. Chem. Soc. 2012, 134, 12792–12797. [Google Scholar] [CrossRef] [PubMed]
  16. Kefalidis, C.E.; Essafi, S.; Perrin, L.; Maron, L. Qualitative estimation of the single-electron transfer step energetics mediated by samarium(II) complexes: A “SOMO−LUMO gap” approach. Inorg. Chem. 2014, 53, 3427–3433. [Google Scholar] [CrossRef] [PubMed]
  17. Labouille, S.; Nief, F.; Maron, L. Theoretical treatment of redox processes involving lanthanide(II) compounds: Reactivity of organosamarium(II) and organothulium(II) complexes with CO2 and pyridine. J. Phys. Chem. A 2011, 115, 8295–8301. [Google Scholar] [CrossRef] [PubMed]
  18. Evans, W.J.; Bloom, I.; Hunter, W.E.; Atwood, J.L. Organolanthanide hydride chemistry. 3. Reactivity of low-valent samarium with unsaturated hydrocarbons leading to a structurally characterized samarium hydride complex. J. Am. Chem. Soc. 1983, 105, 1401–1403. [Google Scholar] [CrossRef]
  19. Ortiz, J.V.; Hoffmann, R. Hydride bridges between LnCp2 centers. Inorg. Chem. 1985, 24, 2095–2104. [Google Scholar] [CrossRef]
  20. Schaverien, C.J. Alkoxides as ancillary ligands in organolanthanide chemistry: Synthesis of, reactivity of, and olefin polymerization by the μ-hydride-μ-alkyl compounds [Y(C5Me5)(OC6H3tBu2)]2(μ-H)(μ-alkyl). Organometallics 1994, 13, 69–82. [Google Scholar] [CrossRef]
  21. Duchateau, R.; van Wee, C.T.; Meetsma, A.; van Duijnen, P.T.; Teuben, J.H. Insertion and C–H bond activation of unsaturated substrates by bis(benzamidinato)yttrium alkyl, [PhC(NSiMe3)2]2YR (R = CH2Ph·THF, CH(SiMe3)2), and hydrido, {[PhC(NSiMe3)2]2Y(μ-H)}2, compounds. Organometallics 1996, 15, 2291–2302. [Google Scholar] [CrossRef]
  22. Steigerwald, M.L.; Goddard, W.A. The 2s + 2s reactions at transition metals. 1. The reactions of D2 with Cl2TiH+, Cl2TiH, and Cl2ScH. J. Am. Chem. Soc. 1984, 106, 308–311. [Google Scholar] [CrossRef]
  23. Ziegler, T.; Folga, E.; Berces, A. A density functional study on the activation of hydrogen–hydrogen and hydrogen–carbon bonds by Cp2Sc–H and Cp2Sc–CH3. J. Am. Chem. Soc. 1993, 115, 636–646. [Google Scholar] [CrossRef]
  24. Maron, L.; Eisenstein, O. DFT study of H−H activation by Cp2LnH d0 complexes. J. Am. Chem. Soc. 2001, 123, 1036–1039. [Google Scholar] [CrossRef] [PubMed]
  25. Ren, J.; Hu, J.; Lin, Y.; Xing, Y.; Shen, Q. The reactivity of lanthanide alkyl compounds with phenylacetylene: Synthesis and structure of [(ButCp)2LnC≡CPh]2 (Ln= Nd, Gd). Polyhedron 1996, 15, 2165–2169. [Google Scholar] [CrossRef]
  26. Cuenca, T.; Gómez, R.; Gómez-Sal, P.; Rodriguez, G.M.; Royo, P. Reactions of titanium- and zirconium(III) complexes with unsaturated organic systems. X-ray structure of {[(η5-C5H5)Zr(CH3)]2[μ-η12-CN(Me2C6H3)] (μ-η55-C10H8)}. Organometallics 1992, 11, 1229–1234. [Google Scholar] [CrossRef]
  27. Atwood, J.L.; Hunter, W.E.; Evans, W.J. Synthesis and crystallographic characterization of a dimeric alkynide-bridged organolanthanide: [(C5H5)2ErC≡CC(CH3)3]2. Inorg. Chem. 1981, 20, 4115–4119. [Google Scholar] [CrossRef]
  28. Evans, W.J.; Bloom, I.; Hunter, W.E.; Atwood, J.L. Synthesis of organosamarium complexes containing Sm–C and Sm–P bonds. Crystallographic characterization of [(MeC5H4)2SmC≡CCMe3]2. Organometallics 1983, 2, 709–714. [Google Scholar] [CrossRef]
  29. Shen, Q.; Zheng, D.; Lin, L.; Lin, Y. Synthesis of (t-C4H9C5H4)2Sm(DME) and its reactivity with phenylacetylene: Synthesis and structure of ((t-C4H9C5H4)2SmC≡CPh)2. J. Organomet. Chem. 1990, 391, 307–312. [Google Scholar] [CrossRef]
  30. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  31. Perdew, J.P.; Wang, Y. Accurate and simple analytic representation of the electron-gas correlation energy. Phys. Rev. B 1992, 45, 13244–13249. [Google Scholar] [CrossRef]
  32. Gaussian 09, Revision A.02. Gaussian Inc.: Wallingford, CT, USA, 2009.
  33. Dolg, M.; Stoll, H.; Savin, A.; Preuss, H. Energy-adjusted pseudopotentials for the rare earth elements. Theor. Chim. Acta 1989, 75, 173–194. [Google Scholar] [CrossRef]
  34. Dolg, M.; Stoll, H.; Preuss, H. A combination of quasirelativistic pseudopotential and ligand field calculations for lanthanoid compounds. Theor. Chim. Acta 1993, 85, 441–450. [Google Scholar] [CrossRef]
  35. Hehre, W.J.; Ditchfield, R.; Pople, J.A. Self-consistent molecular orbital methods. XII. Further extensions of Gaussian-type basis sets for use in molecular orbital studies of organic molecules. J. Chem. Phys. 1972, 56, 2257–2261. [Google Scholar] [CrossRef]
  36. Hariharan, P.C.; Pople, J.A. The influence of polarization functions on molecular orbital hydrogenation energies. Theor. Chim. Acta 1973, 28, 213–223. [Google Scholar] [CrossRef]
  37. Gonzalez, C.; Schlegel, H.B. An improved algorithm for reaction path following. J. Chem. Phys. 1989, 90, 2154–2161. [Google Scholar] [CrossRef]
  38. Gonzalez, C.; Schlegel, H.B. Reaction path following in mass-weighted internal coordinates. J. Phys. Chem. 1990, 94, 5523–5527. [Google Scholar] [CrossRef]
  39. Reed, A.E.; Curtiss, L.A.; Weinhold, F. Intermolecular interactions from a natural bond orbital, donor–acceptor viewpoint. Chem. Rev. 1988, 88, 899–926. [Google Scholar] [CrossRef]
  40. Weinhold, F. Natural bond orbital methods. In Encyclopedia of Computational Chemistry; Schleyer, P.V.R., Allinger, N.L., Clark, T., Eds.; John Wiley & Sons: Chichester, UK, 1998; Volume 3, pp. 1792–1811. [Google Scholar]
  41. Waterman, R. σ-Bond metathesis: A 30-year retrospective. Organometallics 2013, 32, 7249–7263. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Kefalidis, C.E.; Maron, L. On the Dehydrocoupling of Alkenylacetylenes Mediated by Various Samarocene Complexes: A Charming Story of Metal Cooperativity Revealing a Novel Dual Metal σ-Bond Metathesis Type of Mechanism (DM|σ-BM). Inorganics 2015, 3, 573-588. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics3040573

AMA Style

Kefalidis CE, Maron L. On the Dehydrocoupling of Alkenylacetylenes Mediated by Various Samarocene Complexes: A Charming Story of Metal Cooperativity Revealing a Novel Dual Metal σ-Bond Metathesis Type of Mechanism (DM|σ-BM). Inorganics. 2015; 3(4):573-588. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics3040573

Chicago/Turabian Style

Kefalidis, Christos E., and Laurent Maron. 2015. "On the Dehydrocoupling of Alkenylacetylenes Mediated by Various Samarocene Complexes: A Charming Story of Metal Cooperativity Revealing a Novel Dual Metal σ-Bond Metathesis Type of Mechanism (DM|σ-BM)" Inorganics 3, no. 4: 573-588. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics3040573

Article Metrics

Back to TopTop