• Keine Ergebnisse gefunden

5 T HEORY

5.1 C ARBENES

Carbenes have emerged as a powerful ligand class, outrivalling the previously used phosphines in many aspects[8,26]. The fact that carbenes have almost replaced phosphines as spectator ligands in numerous relevant transition metal catalyzed reactions, such as palladium-catalyzed coupling reactions and ruthenium-palladium-catalyzed olefin metathesis, is mostly due to their superior σ-donor properties. They are defined as neutral species with a carbon atom featuring only six valence electrons. A special class of carbenes that has gained a lot of attention are N-heterocyclic carbenes (NHCs). NHCs bear at least one nitrogen atom directly adjacent to the carbene center.

Scheme 1: (A) First evidence for the formation of carbenes by Wanzlick[27] and (B) first isolated carbene by Arduengo[28].

First evidence for the existence of carbenes was provided by Wanzlick through the isolation of 1,3-diphenylimidazolin-2-on[27a] and the respective carbene dimers[27b] (Wanzlick equilibrium) from heated solutions of 1,3-diphenyl-2-trichloromethylimidazolin (Figure 5, (A)). Also, Wanzlick[29] and Öfele[30] were able to trap carbenes with metal fragments, leading to metal carbene complexes. However, it was the first isolation of an NHC with sterically demanding adamantyl substituents by Arduengo et al. in 1991[28] (Figure 5, (B)) that triggered an increasing application of NHCs as ligands in transition metal catalysis and in organocatalysis. The depicted syntheses, deprotonation of imidazolium salts with strong bases (KOtBu, KHMDS) as well as abstraction of small molecules (CHCl3, MeOH, CO2) from carbene adducts, still represent the most common routes to carbenes. Many routes to their precursor imidazolium salts and NHC-adducts are published.[31] Often, bulky groups on the nitrogen atoms (e.g.

adamantyl or mesityl substituents) provide kinetic stability to the carbon center by preventing dimerization. The main contribution to the stability of NHCs however, lies in their special

electronic structure. Carbenes can either exist in a triplet (two unpaired electrons) or a singlet (all electrons paired) ground state.[32] Which ground state a certain carbene prefers mainly depends on its chemical environment. NHCs are mostly singlet carbenes, which renders them especially stable. Four different electronic configurations can be envisioned for NHCs (Figure 5, top).[32] A singlet 1A1 state with both electrons in the σ-orbital (HOMO: highest occupied molecular orbital), a singlet 1A1 state with both electrons in the p-orbital (LUMO, lowest unoccupied molecular orbital), a singlet 1B1 state and a triplet 3B1 state, each with one electron in the σ- and the p-orbital, respectively (Figure 5, top). Most NHCs are in the 1A12) singlet ground state because the σ-electron-withdrawing properties of the nitrogen atom(s) stabilize the HOMO and lower its energy (ES-T, Figure 5) resulting in an increase in the HOMO-LUMO gap (ΔE, Figure 5). The higher ΔE and ES-T get, the more favored is the 1A12) singlet ground state. The singlet 1B11pπ1) and the triplet 3B11pπ1) configuration become competitive when the energy difference between the HOMO and the LUMO (ΔE, Figure 5, bottom) is small. The

1A1(pπ2) configuration can usually be neglected under standard conditions. The stability of NHCs therefore correlates with ΔE between the σ(sp2)- (HOMO) and the pπ-orbital (LUMO). If the energy gap between the triplet and the singlet configuration of the carbene, ES-T (Figure 5, bottom), exceeds approximately 40 kcal/mol, the singlet ground state is favored.[32]

Figure 5: Top: Possible electronic configurations of NHCs. Singlet ground state configuration 1A12) and singlet configurations 1B11pπ1) and 1A1(pπ2) as well as rivalling triplet ground state 3B11pπ1). Bottom: Stabilization of HOMO (ES-T) by electron-withdrawing properties of nitrogen atoms, leading to an increase in the HOMO-LUMO gap (ΔE), therefore favoring the singlet ground state 1A12) over the triplet ground state 3B11pπ1). [32]

Another factor favoring the singlet ground state in NHCs is the cyclic structure, forcing the

imidazolidin-2-ylidenes) and can therefore be isolated with less bulky substituents on the nitrogen atoms. Further electronic stabilization is provided by π-donation of the nitrogen lone pairs into the empty p-orbital.

Figure 6: Special classes of NHC: (A) Normal vs. abnormal (mesoionic) NHCs with charge separation.[32] (B) Comparison of HOMO and LUMO energies in “normal” carbenes, cyclic alkyl amino carbenes (CAAC)[33] and anti-Bredt carbenes[34] with pyramidalized nitrogen atoms. (C) Multiple heteroatom containing carbenes (triazol- and thiazol-ylidenes).[35]

Lately, also other classes of stable carbenes have been prepared. Abnormal (mesoionic, MIC)[36] carbenes, cyclic alkyl amino carbenes (CAAC)[33] and anti-Bredt carbenes[34], and their respective metal complexes have been isolated (Figure 6). MICs, in contrast to “normal”

carbenes, are carbenes that have no neutral mesomeric structure but can only be written with negative and positive partial charges (Figure 6, (A)). They are often synthesized by blocking the C2-carbon to prevent deprotonation at the more acidic C2-position. In some cases, abnormal carbenes form by rearrangements when the corresponding carbene is coordinated to a sterically encumbered metal center.[36] CAAC are carbenes with only one nitrogen atom adjacent to the carbene carbon. Instead, one of the electronegative nitrogen atoms is replaced by a σ-donating (not π-donating) alkyl group, which renders CAACs better σ-donors and better π-acceptors than normal NHCs with two nitrogen atoms next to the carbene (Figure 6, (B)). In addition, the quaternary carbon atom in direct proximity to the carbon center provides them with a unique steric profile.[33] Anti-Bredt carbenes are cyclic diamino carbenes featuring a pyramidalized nitrogen atom.[34] The pyramidalization prohibits the nitrogen atom from donating

electron density to the p-orbital of the carbene carbon, therefore only lowering the energy of the LUMO and not effecting the HOMO energy (Figure 6). The resulting carbene therefore shows increased electrophilicity in comparison to a “normal” NHC but only negligible changes in nucleophilicity[34]. Also, carbenes with more than two heteroatoms and heteroatoms other than nitrogen and their metal complexes have been synthesized. Amongst them are 1,2,4-triazol-5-ylidenes[35a] and 1,2,3-triazol-5-ylidenes (MIC)[35b] as well as thiazol-2-ylidenes[35c,35d]

(Figure 6).

Scheme 2: Synthesis of metal NHC complexes.[37] (A) Reactions with free carbenes either proceed by substitution of neutral donor ligands or by cleavage of dimeric complexes. (B) The carbene is generated in situ, either by addition of an external base or by internal deprotonation. (C) Transmetallation with e.g. silver NHC complexes.

NHC (transition) metal complexes find a variety of applications throughout organic chemistry.

Many different approaches to their syntheses have been explored (Scheme 2).[37] In many cases, complexes can be synthesized by replacement of another neutral donor ligand (e.g.

phosphine, THF) by the free carbene or by cleavage of a dimeric metal complex with the free

transfer of carbenes from silver or gold complexes to other transition metals is a viable approach (Scheme 2, (C)).

Scheme 3: Determination of TEP of a carbene by measurement of the A1 C≡O vibrational mode of the respective iridium, rhodium[38] or nickel[24] carbonyl complex. The stronger the σ-donation from the carbene to the metal center, the higher the amount of π back donation from the metal into the anti-bonding π*-orbital is. A decrease in C≡O bond strength leads to a decrease in vibration frequency, translating into a lower TEP. Strong σ-donors have a low TEP and vice versa.[39]

Since the σ-donor strength of carbenes is highly important for their reactivity and the reactivity of the corresponding carbene complexes, several ways to measure this property have been developed.[8] The Tolman electronic parameter (TEP, Scheme 3) has emerged as a prominent parameter to determine and compare σ-donor properties of phosphines as well as carbenes.[39]

TEP is derived from the frequency of the A1 C≡O vibrational mode of mostly nickel[24], iridium or rhodium[38] complexes (Scheme 3). C≡O is a σ-donor/ π-acceptor ligand, the metal C≡O bond consist of a σ-bond from the carbonyl ligand to the metal center and a π back bond from the metal center into a low lying π*-orbital of C≡O. If the overall electron density on the metal center increases, for example through σ-donation from another ligand, π back donation from the metal into the anti-bonding orbital is increased, resulting in a weaker C≡O bond and a lower frequency. The lower the frequency and the lower the TEP, the higher the σ-donor strength of the corresponding carbene is. TEP can also be retrieved from computational methods. Some general trends can be concluded from the measurements made so far, although, as always, there are exceptions.[8,24] In terms of heteroatoms adjacent to the carbene center, the σ-donor strength decreases from cyclic alkyl amino carbenes (CAAC) over NHCs and oxazol-ylidenes to thiazol-ylidenes. Also, TEP increases from 5- to 7-membered NHCs and is highly dependent

on the substituents on the nitrogen atoms. Evidently, not only electronic but also steric influences of the carbene have a great impact on the catalytic properties of the carbene metal complex. Sterics in NHCs can be described with the cone angle that has been developed for phosphines[39b,40]; but today sterics are mainly characterized in terms of buried volume (Vbur).[41]

Vbur has been defined as the space a ligand takes up in the first ligand sphere of a central (metal) atom. The coordination sphere is constructed with a given radius R around the central atom (Figure 7). The coordinating atom of the ligand is then placed in the distance d from the metal center on the z-axis (Figure 7). For the determination of Vbur the geometry of the ligand must be known, either from computational methods such as density functional theory (DFT) optimizations or from single-crystal X-ray analysis.

Figure 7: Determination of buried volume (Vbur) by calculating the space the ligand takes up in the first coordination sphere (radius R) of the metal center in the distance d.[41a,41b]

Furthermore, the distance between the coordinating atom and the central atom (d, Figure 7) has a high impact on the buried volume of a ligand. Of course, the buried volume of one NHC can differ greatly in different metal complexes. Sambvca2 (former Sambvca), developed by Cavallo et al., presents an online tool for the determination of buried volume.[41a,41b] The software requires the geometry of the ligand and the distance between the coordinating atom and the metal atom to determine the buried volume. The coordination sphere is set 3.5 Å in the default settings but can be changed by the user. The buried volume is then determined by dividing the coordination sphere into voxels (3D pixels, Figure 7). Every voxel within the van-der-Waals radius of a ligand atom belongs to the Vbur. Buried volume is usually given as percentage of the complete coordination sphere (%V ). Buried volume in combination with

is defined as the sum of an electronic contribution NHCelectronic and a steric contribution NHCsteric. The electronic contribution has been reduced to the singlet-triplet energy gap ES-T, whereas the steric contribution has been coupled to the buried volume. This allows for calculations of the theoretical values for the dimerization energy (Efit,dim) of a range of NHCs.

Good accordance with experimental dimerization energies Edim could be obtained and an operating window of Efit,dim for stable monomeric NHCs was postulated. For Efit,dim>-3 kcal/mol carbenes are proposed to be stable in their monomeric form, whereas for carbenes with Efit,dim<-22 kcal/mol dimerization will occur.[41e]

Figure 8: The metal carbene bond. Left: σ-donation from the sp2 hybrid orbital into the dz-orbital (arbitrarily chosen).

Right: π-donation from the carbene to the metal as well as π back donation from metal orbitals into π*-orbitals of the carbene. [32a]

The NHC metal bond is mainly characterized by a σ-bond from the carbene-carbon into empty d-orbitals at the metal center. Nevertheless, also π back donation from the metal into an empty π*-orbital at the carbene and π-donation from the carbene into empty d-orbitals at the metal are to be discussed[32a]. In fact, studies on a multitude of NHC-metal complexes with different NHCs and different d-electron count at the metal center revealed, that π-contributions mainly depend on the d-electron count. The more d-electrons on the metal, the higher the π back donation into the carbene orbitals (average π-contribution of 20% in d10 systems, average of 10% in d0 systems; 100%: all orbital interactions). On the other hand, σ- and π-donation from carbene to metal increase with decreasing d-electron count. Of course, also electrostatic interactions contribute to the carbene metal bond, although they are not easily accessible. For neutral d0-complexes and cationic complexes, electrostatic interactions play a more crucial role. As already discussed for the dimerization energy of carbenes, steric repulsion also has an impact on the stability of the metal carbene bond.

Figure 9: Exemplary chiral carbenes with stereocenters (A) and axial chirality as well as planar chirality (B).[42]

Apart from electronic stabilization, NHCs also offer the possibility to introduce chirality. NHCs with one or several stereocenters, axial chirality and planar chirality have been synthesized and coordinated to metal centers.[42] Imidazolium salts and their respective carbenes containing chiral stereocenters on the nitrogen substituents as well as in the backbone of the NHC have been isolated (Figure 9, (A)). Chiral residues in the backbone, even though not in proximity to the central metal atom, influence the orientation of the NHC wingtips, which in turn have direct impact on the stereochemistry of the metal center.[42a] Amongst others, 3,3´-substituted biphenols and 1,1´-binaphthyl-2,2´-diamine or 2-amino-2’-hydroxy-1,1’-binaphthalene[43] have been used to introduce axial chirality to NHCs (Figure 9, (B)). Especially chiral ferrocene derivatives[44] have been utilized to gain planar chirality (Figure 9, (B)), comparable to commonly used phosphine ligands (eg. Josiphos[45]). Furthermore, trans-1,2-diamino cyclohexane has been widely applied as an element of chirality in NHCs.[42]

Another means to dramatically influence the coordination chemistry of NHC metal complexes is the incorporation of chelating bidentate or pincer carbenes (Figure 10, top).[46] Bidentate carbenes bear a second coordinating group, which is tethered to the imidazol-2-ylidene core.

The second donor can either be charged or neutral. Pincer-type carbenes feature two additional chelating groups, thereby having the ability to block three coordination sites on a metal center. Famous examples for chelating carbenes with anionic tethers are alcoholates[47], thiolates[48], amines[49] and sulfonates[48,50] (Figure 10, exemplary anionic donors). Neutral

[51] [49] [48] [52]

additional donors, the concept of hemilability emerges.[54] In hemilabile complexes, the donor can easily dissociate and recoordinate, thereby blocking or opening coordination sites. This becomes particularly interesting in the case of instable complexes and transition states. A special class of chelating carbenes are bis-carbenes.[55] Apart from influencing the coordination sphere, bidentate and pincer carbenes enhance stability through the chelate effect.[7]

Figure 10: Bidentate and pincer carbenes bearing additional anionic or neutral donor functionalities.[46]