Next Article in Journal
Recent Advances on 2D Materials towards 3D Printing
Next Article in Special Issue
On the Aromaticity and 13C-NMR Pattern of Pentagonal-Pyramidal Hexamethylbenzene Dication [C6(CH3)6]2+: A {C5(CH3)5}–{CCH3}3+ Aggregate
Previous Article in Journal / Special Issue
Is Aromatic Nitration Spin Density Driven?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Are Metallacyclopentadienes Always Non-Aromatic?

by
Ricardo Casiano-González
and
José Enrique Barquera-Lozada
*
Instituto de Química, Universidad Nacional Autónoma de México, Circuito Exterior, Ciudad Universitaria Coyoacán, Ciudad de México 04510, Mexico
*
Author to whom correspondence should be addressed.
Submission received: 10 September 2021 / Revised: 30 September 2021 / Accepted: 4 October 2021 / Published: 4 November 2021

Abstract

:
Even though metallacyclopentadienes (MCPs) are among the most common metallacycles, their electron delocalization (aromaticity) has received far less attention than other metallacycles, such as metallabenzenes. We systematically studied the aromaticity of MCPs with energetic (isomerization stabilization energy), density (delocalization index) and magnetic (current density) aromaticity indices. The indices agree that metallacyclopentadienes are, in general, weakly aromatic at most. The 18 e complexes showed the expected weak aromaticity, and only the d 8 molecules are somewhat anti-aromatic. However, the theoretical account of the aromaticity of the 16 e MCPs is more convoluted. We find that the aromatic criteria for a 16 e d 4 ruthenacyclopentadiene disagree. The lack of agreement shows that significant electron delocalization is not always related to great stability or to strong diatropic currents.

Graphical Abstract

1. Introduction

One of the first families of metallocycles discussed in the literature were the metallacyclopentadienes (MCPs). In fact, in the famous paper of Thorn and Hoffmann [1] in which the metallabenzenes (MBs) were predicted, the principal topic of the manuscript was the metallacyclopentadienes. At that time, several crystal structures of MCPs were already known [2,3,4,5,6,7,8,9], many years before the first metallabenzene was synthesized [10]. Furthermore, a simple search in the Cambridge Structural Database (CSD) shows far more MCP than MB structures. However, the electronic structure of the former has received much less attention than the latter. The aromaticity of metallabenzenes has been thoroughly discussed in the literature [11,12,13,14,15,16,17,18,19,20,21,22] and some general conclusions have been drawn. Their C–C bonds have a significant degree of equalization, and most of them are planar [12,13,22], although some of them strongly deviate from planarity [23,24]. In general, the MBs’ aromatic stabilization energy (ASE) are around 10 kcal/mol below the value for benzene [14,15,25,26]. Moreover, their nuclear independent chemical shifts (NICS) are negative [27,28,29,30,31,32], a common characteristic of aromatic compounds. Therefore, all these results indicate that metallabenzenes are at least slightly aromatic. However, some contradictory results have also been published. For example: Iron et al. and Periyasamy et al. found that the NICS and magnetic susceptibility anisotropy values of some complexes are positive, indicating a certain degree of anti-aromaticity [16,27]. These contradictions could be because these two magnetic indices depend not only on the magnetically induced electron current around the ring, but also on all the currents circulating around each atom. In aromatic organic molecules the contribution of these atomic currents to NICS is negligible compare to the ring current, but in metallacycles, the current around the metal is not negligible. The magnetically induced current density ( j ( r ) ) avoids this problem and has also been calculated for some metallabezenes. The calculations have shown that the 18 e rings have diatropic currents, indicating aromaticity, while the 16 e rings have paratropic currents, which suggest the opposite [16,18].
In the case of metallacyclopentadienes, there are not so many papers that discuss their aromaticity, maybe because, since the beginning, they were classified as non-aromatic [1], although there is a large number of published MCP structures [33], which suggest that they are quite stable. From the published studies, we draw some tendencies about the aromaticity of the MCPs. Their equalization of C–C bonds is significantly lower than for MBs. It has been shown that the maximum difference between C–C distances is around 0.09 Å [1,34,35,36] for metallacyclopentadienes, while for metallabenzenes this difference is approximately 0.03 Å [12,13,36]. There is an exception for this low C–C equalization: in bimetallic complexes all C–C bond distances are very similar [37,38,39,40,41,42], as Thorn and Hoffman described in their seminal work [1]. The vast majority of reported crystal structures are almost planar (see CSD or Supplementary Materials), although there a few examples that are not [43,44,45,46,47,48]. This bending is less common than in metallabenzenes (we found 18 significantly bent structures for MBs and only 6 for MCPs). The ASE reported until now for metallacyclopentadiens are very low, around 4 kcal/mol, compared to metallabenzenes (~15 kcal/mol) [36]. Moreover, it has also been shown that half sandwich nickela-, pallada- and platinacyclopentadiene complex are less stable than their corresponding metallabenzene isomers [49]. The NICS has been the most widely used index to evaluate the aromaticity of the metallacyclopentadienes. Jemmis et al. found very small negative NICS values for Ti and Zr complexes [34]. Islas et al. also found small negative values for a series of Os, Ru Rh and Ir compounds [35]. On the other hand, Zeng et al. found small positive values for a Pd and a Pt metallacyclopentadiene (large positive values are normally associated with anti-aromatic rings) [49]. The discussed results seem to indicate a non-aromatic or a small aromatic character of these metallacycles. However, metallacyclopentadienes are very stable and more common than metallabenzenes, but their aromaticity and its electronic structure has only been studied in a handful of complexes. In addition, the NICS were used as the primary tool to evaluate its aromaticity, although, as mentioned, their values can be affected by the strong metal currents [50,51]. This effect can be more significant for weak ring currents, such as the currents that these compounds presumably have. Therefore, it is necessary to study systematically the aromaticity of these types of important metallacycles and with different aromaticity indexes that do not suffer the same drawbacks as NICS. In this contribution, we studied the aromaticity of a series of d 0 , d 2 , d 4 , d 6 and d 8 metallacyclopentadienes using j ( r ) , isomerization stabilization energy (ISE) and the delocalization indices (DI).

2. Computational Methods

All the geometries were fully optimized, unless otherwise stated in the main text, at the M062x/Def2TZVP level of theory. The M062x functional has broad applicability and has shown good results for π conjugated systems [52], which is a common characteristic of the studied molecules. A frequency calculation was performed to ensure that the stationary points were true minima. Because of the use of core potentials in the Def2TZVP basis set for the second row transition metals, we prefer to use a full electron basis set for all the calculations of the magnetic properties (Sapporo-TZP-2012 [53]). The magnetic perturbed wave functions were calculated with gauge-including atomic orbitals, GIAO, at the M062x/Sapporo-TZP-2012 level of theory (For the analysis of the contribution of independent MO to j ( r ) the CSGT method was used, see below). All these calculations were performed with Gaussian 16 [54].
The DI and j ( r ) were calculated with AIMALL [55]. The triple product of j ( r ) (tp j ( r ) = B · × j ( r ) ) was calculated numerically using python scripts and the pvpython module included in Paraview 5.2 as has been previously described [56,57,58]. The calculations of the single orbital contributions to j ( r ) were performed with the CTOCD-DZ (continuous transformations of the origin of the current density-diamagnetic zero) or ipsocentric method [59,60], because only this method can guarantee that the orbital contributions do not include magnetic transitions between occupied molecular orbitals (MOs) [61]. The calculations of the MO contributions were performed with AIMALL and SYSMOIC [62].

3. Results

The 18 metallacyclopentadienes of the first and second transition series that we studied are shown in Scheme 1. The studied metallacycles can be divided into two main groups: 16 e (110) and 18 e (1118). Moreover, the two groups can be subdivided into d 0 (1,2), d 2 (3,4,11,12), d 4 (5,6,13,14), d 6 (7,8,15,16) and d 8 (9,10,17,18) complexes.
The optimized structures of the studied compounds are similar to what has been previously reported (for some examples see [4,5,8,63,64,65,66,67,68,69,70,71]). The C–C bonds are poorly equalized: the C2-C3 and C4-C5 bonds are around 1.35 Å and the C3-C4 distances are approximately 1.46 Å, and the bond distance differences between C2-C3 or C4-C5 with C3-C4 are ~0.11 Å (Table 1). All these differences are larger than 0.09 Å except for the d 4 complex 6 with a bond distance differences between C2-C3 and C3-C4 of only 0.06 Å. Another 16 e complex than stands apart from the C–C distance trend is 8, in which the C–C distances are inverted. The C3-C4 bond is the shortest (1.34 Å) and C2-C3 and C4-C5 are longer (1.48 Å). Moreover, the M-C bonds are particularly short, only 1.84 Å. This is indicative that 8 resembles the structure of a metallacyclopentatriene. In the literature, there are some Ru complexes with this type of structure, although with a different d electron count [65,72,73,74,75,76]. Another geometrical feature related to aromaticity is the planarity. We estimated it with the M-C2-C3-C4 dihedral. Only d 0 and d 8 rings have a perfectly planar structure. The others present a slight deviation from the plane, except for the d 2 complexes 3 and 4. The fully optimized structure of 3 is significantly bent, with a M-C2-C3-C4 dihedral of 36° and 4 does not even converge with 5-membered ring structure. For these two molecules, we force the planarity to be able to compare them with the all the other MCPs. In the literature, we could not find any stable d 2 16 e metallacyclopentadiene. The lack of stable d 2 16 e metallacyclopentadienes is compatible with our results.
Energetically, the aromaticity is normally evaluated with the ASE. To estimate this energy, we use the ISE proposed by Schleyer and Pühlhofer et al. [77]. De Proft and Geerlings have used the these energies to evaluate the aromaticity of 5-membered heterocycles [25]. The ISE values depicted in Table 2 are predominantly small compared to well-known aromatic molecules like thiophene or the cyclopentadienyl anion, which have an ISE at the same level of theory of 18.68 and 27.59 kcal/mol, respectively. This confirms the low aromatic or anti-aromatic character of these metallacycles. Most of 16 e metallacyclopentadienes have slightly negative energies (i.e., the non-aromatic isomer is favoured, see Supplementary Materials) and again, an exception appears with the Ru complex 8, a large negative ISE, which indicates that its aromatic isomer is strongly disfavoured. The d 0 and d 4 16 e complexes are the ones with the energy values slightly positive or very close to zero. In contrast, the most negative ISE values are for the d 6 complexes, although we should not forget that the planar d 2 complexes are not even a minimum in the potential energy surface. A CSD search shows that the most common 16 e complexes are d 0 and d 8 (see Supplementary Materials). Therefore, we were surprised that the ISEs were clearly negative for both d 8 molecules. A closer look at the CSD search shows that all the reported d 8 structures have strong electron withdrawing groups around the ring or/and secondary ligands that are not π acceptors, which could help in the stability of the aromatic structure. Indeed, the metallacyclopentadiene complexes with the best C–C bond equalization have strong electron withdrawing groups, as COOH or CF 3 [2,78,79,80]. Moreover, there are some reports on metallabenzenes that indicate that the aromaticity of the ring is enhanced when there are electron withdrawing groups attached to the ring [28,30,31]. On the other hand, the 18 e complexes are slightly positive except for the d 8 compounds. This result shows that, apart from the d 8 complexes, all the others are slightly aromatic.
The DIs measure the number of electron pairs that are shared between two atoms, and they have been used as a bond order index [81,82,83,84]. Then, the DIs of a single and double C–C bond are slightly below 1.0 and 2.0 a.u., respectively, and an aromatic bond has a value between these two [83,84]. Poater et al. have studied the aromaticity of five-member heterocycles by subtracting the DI of the formal C–C double bond from the DI of the single bond ( Δ DI) [85], because in a totally aromatic pentacyclic ring, these two should be equal (we define the Δ DI as DI C 2 C 3 DI C 3 C 4 ). They found that for classical aromatic heterocycles such as pyrrol or thiophene, Δ DI < 0.4 a.u., and for non-aromatic rings such as cyclopentadiene, Δ DI > 0.6 a.u. The Δ DI values (Table 2) show that all the studied 18 e complexes, except the d 8 molecules again, can be classified as only slightly aromatic. Most of the 16 e metallacyclopentadienes are non-aromatic according to this criteria. However, there are two exceptions, again, complexes 6 and 8. In the latter, the C–C DI are inverted, DI C 2 C 3 indicates a single bond character while DI C 3 C 4 a double bond nature. Moreover, the DI M C 2 also suggests that the Ru-C is a double bond. These results coincide with the geometrical parameters discussed above. We can confirm that 8 is a metallacyclopentatriene. On the other hand, the Δ DI of 6 is significantly smaller than the values of all the other complexes. Indeed, its value is similar to the Δ DI of thiophene, calculated at the same level of theory (0.304 a.u.). This results contradicts the ISE, which suggests that 6 is non-aromatic. Therefore, the significant electron delocalization found in this complex does not imply an important stabilization.
Instead of calculating the NICS that have contributions of all the present currents, it is possible to calculate directly j ( r ) and integrate it to calculate the current flux of the ring of interest. If gauge independent atomic orbitals are used, this method is known as gauge-including magnetically induced currents (GIMIC) [86]. The current density has been used to study multiple aromatic compounds [87,88] including metallacyclobenzenes [16,18]. We used j ( r ) and its triple product (tp j ( r ) = B · × j ( r ) ) [56,57], which reveals the local direction of the rotation of j ( r ) (negative/diatropic, positive/paratropic), to evaluate the aromaticity of the metallacyclopentadienes. As in case of the other aromaticity indices, the ring current values are predominately smaller (Table 2) than in case of well-known aromatic rings, as benzene, which support a induced current density of around 12.0 nA/T [86,89,90,91]. The 18 e complexes have low diatropic currents (positive), except for the d 8 rings, which have negative values (paratropic currents). As mentioned in the introduction, strong diatropic currents are related to aromatic rings while strong paratropic currents to anti-aromatic compounds. Then, the small positive j ( r ) values of 18 e metallacyclopentadienes also indicates a low aromaticity, excepting d 8 complexes that are slightly anti-aromatic (small negative j ( r ) values). The 16 e molecules do not show a clear trend. The d 0 rings show a very low current, which confirms their non-aromatic character. The d 8 complexes have a low paratropic current, which coincides with the low anti-aromatic character predicted by the ISE and the Δ DI values. The d 2 molecules have strong paratropic currents, which is compatible with their lack of stability. The d 6 compound 8 have an extremely strong paratropic current, which agrees with the strong anti-aromatic nature indicated by its very negative ISE. Finally, the two 16 e d 4 metallacyclopentadienes present a very odd behaviour. Both have currents comparable with the benzene but one has a diatropic current (5) and the other a paratropic current (6), even though both have very similar structures. Even more intriguing is that several crystal structures of Ru complexes similar to 6 have been reported in the literature [65,72,73,74,75,76], which suggests that this type of molecule is seemingly stable. Moreover, the j ( r ) value of 6 does not coincide with the non-aromatic nature predicted by its near-zero ISEs or the aromatic character predicted by its low Δ DI value. This lack of coincidence between aromatization energy a ring currents has been previously reported in the literature. The currents for 9b-boraphenalene and 9b-azaphenalene are strongly paratropic but their ASE values are positive [92]. To explain the mismatch, the authors analyse the currents with the ipsocentric method to show that the strong paratropic currents were a consequence of symmetry of magnetic virtual excitation and not due to a lack of stability. The ipsocentric method has shown that the symmetry of the virtual excitations determine the strength and direction of the current density and not the electron delocalization [60,93,94].
To further investigate the opposed behaviour of j ( r ) in very similar complexes, we studied the contributions of the frontier molecular orbitals to the total current density. As mentioned above the direction of the rotation (tropicity) of the currents depends on the symmetry of the magnetic virtual excitation, as has been described by the Steiner–Fowler selection rules [61]. The diamagnetic contribution to j ( r ) is determined by the accessibility of transitionally allowed transitions, while the paramagnetic contribution by the accessibility of rotationally allowed transitions. Figure 1 shows the current and tp j ( r ) of 5 and 6 at 1.0 Å from the molecular plane. In this region, the current is mainly diatropic for 5 (Figure 1a) and mainly paratropic for 6 (Figure 1b), as the currents strength suggest (Table 2). Figure 1 also shows the contributions of some frontier MOs to the current. The HOMO contribution for both complexes is very small (Figure 1c,d), even though the transition HOMO (a”) –> LUMO (a”) is transitionally allowed (5 and 6 have a C s symmetry). However, the HOMO and LUMO are spatially different (Figure 2), the HOMO is mainly located in C 4 H 4 ligand, while the LUMO is at the metal centre. These spatial differences reduce the probability of the transition [61]. The j ( r ) maps show that the transitions from the HOMO-1 are significantly different between the Fe and Ru complexes (Figure 1e,f, respectively). The HOMO-1 current for 5 is very weak around the ring, while for 6 it has a strong paratropic nature. In case of the Ru complex, the energy of the HOMO-1 is very close to the HOMO, and it has a spatial distribution similar to the LUMO (Figure 2). Both are mainly at the metal and have a π nature at the C 4 H 4 moiety. Then, HOMO-1 (a’) –> LUMO (a”), which is rotationally and transitionally allowed, should be very relevant. Contrary, the HOMO-1 of the Fe complex is significantly lower in energy and the C 4 H 4 ligand has a σ character. These two factors reduce the relevance of the HOMO-1 (a’) –> LUMO (a”) transition in complex 5. Above the molecular plane, it is the HOMO-2 which contributes the most to the current for 5 (Figure 1g). This MO has a π nature as the LUMO, which is favourable for a HOMO-2 –> LUMO transition. Then, the opposite j ( r ) tropicity between the Fe and the Ru complexes are mainly the result of the much higher energy of the HOMO-1 in 6 and due to the differences in the spatial distribution of the two HOMO-1s.

4. Conclusions

We found that the electron delocalization in metallacyclopentadienes is more interesting than was previously thought. All the types of aromaticity indices (energetic, electron density, magnetic) coincide in predicting that all the 18 e metallacyclopentadienes are slightly aromatic, except for the d 8 complexes that are slightly anti-aromatic. However, the picture of the 16 e complexes is more complicated. We found that the d 0 compounds are non-aromatic, as previous studies have reported. From the lack of stability of the rings and their strong paratropic electron currents, we conclude that d 2 complexes are anti-aromatic, these results also explain the lack of 16 e d 2 metallacyclopentadienes crystal structures. The d 8 are, according to all the aromaticity indices, slightly anti-aromatic. At first, this result was a surprise because many structures have been reported for this type of complex, although all these structures have electron withdrawing groups. We anticipate that these groups help to stabilize the ring, but more research is needed in that direction. The two d 6 16 e metallacyclopentadienes that we studied are anti-aromatic. Indeed, complex 8 shows the most anti-aromatic behaviour of all the studied complexes. To gain some stability, it forms localized double bonds with the carbons of the C 4 H 4 moiety to become a formally 18 e metallacyclopentatriene. However, as mentioned above, there are some examples of d 6 16 e metallacyclopentadienes in the literature. The structure of these complexes involve again electron withdrawing groups or non π acceptor ligands. This could help in the stabilization of the complexes. This observation suggests that the low aromaticity or anti-aromaticity character of MCPs could be reversed by the nature of the groups attached to the ring or by the ligands bonded to the metal centre. We leave 16 e d 4 complexes at the end, because they present the largest discrepancies between the aromaticity indices. The extreme is the compound 6: the energetic criterion (ISE) predicts that it is non-aromatic, the density index ( Δ DI) indicates that it is aromatic, while the magnetic fields ( j ( r ) and tp j ( r ) ) suggests that it is anti-aromatic. Moreover, there are reports of some structures similar to 6, which suggest that the 16 e d 4 complexes are stable. Thus, the magnetic anti-aromaticity (paratropicity) of 6 seems not to be related to its stability. The paratropic currents of this complex are not a consequence of highly localized electrons, but rather due to the symmetry, the energy and the spatial distribution of the frontier MOs, as the ipsocentric description of the current postulates. The 16 e d 4 metallacyclopentadienes deserves more attention because their large electron delocalization does not turn into large stabilization and diatropic currents, as in case of most aromatic compounds.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/chemistry3040094/s1, Figure S1: Isomeration reaction for the calculation of ISE, Figure S2: Maps of contributions to j ( r ) of HOMO-1 of complex 6 and its rotational and translational part, Table S1: XYZ coordinates of the studied metallacyclopentadienes.

Author Contributions

Conceptualization, J.E.B.-L.; methodology, J.E.B.-L.; validation, J.E.B.-L. and R.C.-G.; formal analysis, R.C.-G.; investigation, R.C.-G.; resources, J.E.B.-L.; data curation, R.C.-G.; writing—original draft preparation, J.E.B.-L.; writing—review and editing, J.E.B.-L. and R.C.-G.; visualization, J.E.B.-L. and R.C.-G.; supervision, J.E.B.-L.; project administration, J.E.B.-L.; funding acquisition, J.E.B.-L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Consejo Nacional de Ciencia y Tecnología (project: FC 2328) and Dirección General de Asuntos del Personal Académico, Universidad Nacional Autónoma de México (project: IN208820).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors thanks DGTIC-UNAM for the computer time. They also thank Gladys E. Cortés, Marıá M. Aguilar and José D.Vázquez for the computer support.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript:
MCPMetallacyclopentadiene
MBMetallabenzene
CSDCambridge Structural Database
ASEAromatic stabilization energy
NICSNuclear independent chemical shifts
ISEIsomerization stabilization energy
DIDelocalization index
MOMolecular orbital
CTOCD-DZContinuous transformations of the origin of the current density-diamagnetic zero
GIAOGauge-including atomic orbital

References

  1. Thorn, D.L.; Hoffmann, R. Delocalization in Metallocycles. Nouv. J. Chem. 1979, 3, 39–45. [Google Scholar]
  2. Mague, J.T. Crystal and molecular structure of chlorobis(triphenylstibine)tetrakis(trifluoromethyl)rhodiacyclopentadaiene-dichloromethane solvate, RhCl(Sb(C6H5)3)2C4(CF3)4.CH2Cl2. Inorg. Chem. 1970, 9, 1610–1618. [Google Scholar] [CrossRef]
  3. Mogue, J.T. Crystal and molecular structure of chloroaquobis(trimethylarsine)tetrakis)trifluoromethyl)rhodiacyclopentadiene. Inorg. Chem. 1973, 12, 2649–2654. [Google Scholar] [CrossRef]
  4. Atwood, J.L.; Hunter, W.E.; Alt, H.; Rausch, M.D. The molecular structure of 1,1-bis(η5-cyclopentadienyl)-2,3,4,5-tetraphenyltitanole and its hafnium analogue. J. Am. Chem. Soc. 1976, 98, 2454–2459. [Google Scholar] [CrossRef]
  5. Gastinger, R.G.; Rausch, M.D.; Sullivan, D.A.; Palenik, G.J. The synthesis and molecular structure of 1-(η5-cyclopentadienyl)-1-triphenylphosphine-2,3,4,5-tetrakis(pentafluorophenyl)rhodole. J. Organomet. Chem. 1976, 117, 355–364. [Google Scholar] [CrossRef]
  6. Gastinger, R.G.; Rausch, M.D.; Sullivan, D.A.; Palenik, G.J. Synthesis and molecular structure of 1-(π-cyclopentadienyl)-1-triphenylphosphine-2,3,4,5-tetrakis(pentafluorophenyl)cobaltole. J. Am. Chem. Soc. 1976, 98, 719–723. [Google Scholar] [CrossRef]
  7. Suzuki, H.; Itoh, K.; Ishii, Y.; Simon, K.; Ibers, J.A. Preparation structure, and role of tetrakis(methoxycarbonyl)palladiacyclopentadiene cyclic diolefin complexes in selective palladium-catalyzed cyclocotrimerization of acetylenes with olefins. J. Am. Chem. Soc. 1976, 98, 8494–8500. [Google Scholar] [CrossRef]
  8. Pierpont, C.G.; Downs, H.H.; Itoh, K.; Nishiyama, J.; Ishii, Y. Synthesis and structure of the triphenylphosphonium cyclopentadienylide adduct of oligomeric tetrakis(methoxycarbonyl)palladiacyclopentadiene. J. Organomet. Chem. 1977, 124, 93–101. [Google Scholar] [CrossRef]
  9. Brown, L.D.; Itoh, K.; Suzuki, H.; Hirai, K.; Ibers, J.A. Effects of donor molecules on the palladium-catalyzed cyclocotrimerization of acetylenes with olefins. Preparation of dimeric tetrakis(methoxycarbonyl)palladiacyclopentadiene(base) complexes and structure with base = 2,6-lutidine. J. Am. Chem. Soc. 1978, 100, 8232–8238. [Google Scholar] [CrossRef]
  10. Elliott, G.P.; Roper, W.R.; Waters, J.M. Metallacyclohexatrienes or ‘metallabenzenes.’ Synthesis of osmabenzene derivatives and X-ray crystal structure of [Os(CSCHCHCHCH)(CO)(PPh3)2]. J. Chem. Soc. Chem. Commun. 1982, 811–813. [Google Scholar] [CrossRef]
  11. Fernández, I.; Frenking, G. Aromaticity in Metallabenzenes. Chem. Eur. J. 2007, 13, 5873–5884. [Google Scholar] [CrossRef]
  12. Dalebrook, A.F.; Wright, L.J. Chapter Three-Metallabenzenes and Metallabenzenoids. In Advances in Organometallic Chemistry; Hill, A.F., Fink, M.J., Eds.; Academic Press: Cambridge, MA, USA, 2012; Volume 60, pp. 93–177. [Google Scholar] [CrossRef]
  13. Landorf, C.W.; Haley, M.M. Recent Advances in Metallabenzene Chemistry. Angew. Chem. Int. Ed. 2006, 45, 3914–3936. [Google Scholar] [CrossRef]
  14. Fernández, I.; Frenking, G.; Merino, G. Aromaticity of metallabenzenes and related compounds. Chem. Soc. Rev. 2015, 44, 6452–6463. [Google Scholar] [CrossRef]
  15. Fernández, I.; Frenking, G. Theoretical Studies of Metallabenzenes: From Bonding Situation to Reactivity. In Metallabenzenes; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2017; pp. 267–304. [Google Scholar] [CrossRef]
  16. Periyasamy, G.; Burton, N.A.; Hillier, I.H.; Thomas, J.M.H. Electron Delocalization in the Metallabenzenes: A Computational Analysis of Ring Currents. J. Phys. Chem. A 2008, 112, 5960–5972. [Google Scholar] [CrossRef]
  17. Mauksch, M.; Tsogoeva, S.B. Demonstration of “Möbius” Aromaticity in Planar Metallacycles. Chem. Eur. J. 2010, 16, 7843–7851. [Google Scholar] [CrossRef] [PubMed]
  18. Havenith, R.W.A.; Proft, F.D.; Jenneskens, L.W.; Fowler, P.W. Relativistic ring currents in metallabenzenes: An analysis in terms of contributions of localised orbitals. Phys. Chem. Chem. Phys. 2012, 14, 9897–9905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Feixas, F.; Matito, E.; Poater, J.; Solà, M. Metalloaromaticity. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2013, 3, 105–122. [Google Scholar] [CrossRef]
  20. Dávila, K.L.; Contreras, R.R.; Fontal, B.; Torres, F.J.; Rincón, L. An alternative description of aromaticity in metallabenzenes. J. Mex. Chem. Soc. 2017, 61. [Google Scholar] [CrossRef]
  21. Frogley, B.J.; Wright, L.J. Recent Advances in Metallaaromatic Chemistry. Chem. Eur. J. 2018, 24, 2025–2038. [Google Scholar] [CrossRef]
  22. Chen, D.; Hua, Y.; Xia, H. Metallaaromatic Chemistry: History and Development. Chem. Rev. 2020, 120, 12994–13086. [Google Scholar] [CrossRef]
  23. Zhu, J.; Jia, G.; Lin, Z. Understanding Nonplanarity in Metallabenzene Complexes. Organometallics 2007, 26, 1986–1995. [Google Scholar] [CrossRef]
  24. Chen, Z.N.; Fu, G.; Zhang, I.Y.; Xu, X. Understanding the Nonplanarity in Aromatic Metallabenzenes: A sigma-Control Mechanism. Inorg. Chem. 2018, 57, 9205–9214. [Google Scholar] [CrossRef]
  25. Proft, F.D.; Geerlings, P. Relative hardness as a measure of aromaticity. Phys. Chem. Chem. Phys. 2004, 6, 242–248. [Google Scholar] [CrossRef]
  26. Lin, R.; Lee, K.H.; Poon, K.C.; Sung, H.H.Y.; Williams, I.D.; Lin, Z.; Jia, G. Synthesis of Rhenabenzenes from the Reactions of Rhenacyclobutadienes with Ethoxyethyne. Chem. Eur. J. 2014, 20, 14885–14899. [Google Scholar] [CrossRef]
  27. Iron, M.A.; Lucassen, A.C.B.; Cohen, H.; van der Boom, M.E.; Martin, J.M.L. A Computational Foray into the Formation and Reactivity of Metallabenzenes. J. Am. Chem. Soc. 2004, 126, 11699–11710. [Google Scholar] [CrossRef] [PubMed]
  28. Ghiasi, R.; Pasdar, H. Computational study of substituent effect in para substituted platinabenzene complexes. Russ. J. Phys. Chem. A 2013, 87, 973–978. [Google Scholar] [CrossRef]
  29. Zhang, H.; Lin, R.; Li, J.; Zhu, J.; Xia, H. Interconversion between Ruthenacyclohexadiene and Ruthenabenzene: A Combined Experimental and Theoretical Study. Organometallics 2014, 33, 5606–5609. [Google Scholar] [CrossRef]
  30. Ghiasi, R.; Amini, E. Substituent and solvent effects on geometric and electronic structure of C5H5Ir(PH3)3 iridabenzene: A theoretical insight. J. Struct. Chem. 2015, 56, 1483–1494. [Google Scholar] [CrossRef]
  31. Ghiasi, R.; Boshak, A. Substituent Effect in para Substituted Osmabenzene Complexes. J. Mex. Chem. Soc. 2017, 57. [Google Scholar] [CrossRef] [Green Version]
  32. Han, F.; Wang, T.; Li, J.; Zhang, H.; Xia, H. m-Metallaphenol: Synthesis and Reactivity Studies. Chem. Eur. J. 2014, 20, 4363–4372. [Google Scholar] [CrossRef]
  33. Ma, W.; Yu, C.; Chen, T.; Xu, L.; Zhang, W.X.; Xi, Z. Metallacyclopentadienes: Synthesis, structure and reactivity. Chem. Soc. Rev. 2017, 46, 1160–1192. [Google Scholar] [CrossRef]
  34. Jemmis, E.D.; Phukan, A.K.; Jiao, H.; Rosenthal, U. Structure and Neutral Homoaromaticity of Metallacyclopentene, -pentadiene, -pentyne, and -pentatriene:  A Density Functional Study. Organometallics 2003, 22, 4958–4965. [Google Scholar] [CrossRef]
  35. Islas, R.; Poater, J.; Solà, M. Analysis of the Aromaticity of Five-Membered Heterometallacycles Containing Os, Ru, Rh, and Ir. Organometallics 2014, 33, 1762–1773. [Google Scholar] [CrossRef]
  36. Huang, Y.Z.; Yang, S.Y.; Li, X.Y. An investigation of the aromaticity of transition metal heterocyclic complexes by conventional criteria and indices of aromaticity. J. Organomet. Chem. 2004, 689, 1050–1056. [Google Scholar] [CrossRef]
  37. Jeffreys, J.A.D.; Willis, C.M. Crystal and molecular structure of tricarbonyl-π-[1,1,1-tricarbonyl-2,3-dimethoxy-5-(diphenylmethyl) ferracyclopentadiene]iron(Fe–Fe), a product from the reaction between diphenyldiazomethane and tricarbonyl-π-[1,1,1-tricarbonyl-2,5-dimethoxyferracyclopentadiene]iron(Fe–Fe). J. Chem. Soc. Dalton Trans. 1972, 2169–2173. [Google Scholar] [CrossRef]
  38. King, M.; Holt, E.M.; Radnia, P.; McKennis, J.S. Metallametallocenes: Ferracobaltocene and ferrarhodocene. New aromatic species. Organometallics 1982, 1, 1718–1720. [Google Scholar] [CrossRef]
  39. Hong, F.E.; Lue, I.R.; Lo, S.C.; Lin, C.C. Reactions of molybdenum-cobalt complex with phenylacetylene: X-ray crystal structure of [MoCo(CO)4{CPhCHCHCPh}(η5-C5H5)]. J. Organomet. Chem. 1995, 495, 97–101. [Google Scholar] [CrossRef]
  40. Baxter, R.J.; Knox, G.R.; Pauson, P.L.; Spicer, M.D. Synthesis of Dicarbonyl(η4-tricarbonylcobaltacyclopentadiene)cobalt Complexes from Co2(CO)8. A General Route to Intermediates in Cobalt Carbonyl Mediated Alkyne Trimerization. Organometallics 1999, 18, 197–205. [Google Scholar] [CrossRef]
  41. Kishimoto, Y.; Eckerle, P.; Miyatake, T.; Kainosho, M.; Ono, A.; Ikariya, T.; Noyori, R. Well-Controlled Polymerization of Phenylacetylenes with Organorhodium(I) Complexes:  Mechanism and Structure of the Polyenes. J. Am. Chem. Soc. 1999, 121, 12035–12044. [Google Scholar] [CrossRef]
  42. Dennett, J.N.L.; Knox, S.A.R.; Anderson, K.M.; Charmant, J.P.H.; Orpen, A.G. The synthesis of [FeRu(CO)2(μ-CO)2(η-C5H5)(η-C5Me5)] and convenient entries to its organometallic chemistry. Dalton Trans. 2005, 63–73. [Google Scholar] [CrossRef] [PubMed]
  43. Nugent, W.A.; Thorn, D.L.; Harlow, R.L. Cyclization of diacetylenes to E,E exocyclic dienes. Complementary procedures based on titanium and zirconium reagents. J. Am. Chem. Soc. 1987, 109, 2788–2796. [Google Scholar] [CrossRef]
  44. Hill, J.E.; Fanwick, P.E.; Rothwell, I.P. Formation, fragmentation, and isomerization of titanacycle rings supported by aryloxide ligation. Organometallics 1990, 9, 2211–2213. [Google Scholar] [CrossRef]
  45. Kaleta, K.; Strehler, F.; Hildebrandt, A.; Beweries, T.; Arndt, P.; Rüffer, T.; Spannenberg, A.; Lang, H.; Rosenthal, U. Synthesis and Characterization of Multiferrocenyl-Substituted Group 4 Metallocene Complexes. Chem. Eur. J. 2012, 18, 12672–12680. [Google Scholar] [CrossRef] [PubMed]
  46. Àrias, O.; Petrov, A.R.; Bannenberg, T.; Altenburger, K.; Arndt, P.; Jones, P.G.; Rosenthal, U.; Tamm, M. Titanocene and Zirconocene Complexes with Diaminoacetylenes: Formation of Unusual Metallacycles and Fulvene Complexes. Organometallics 2014, 33, 1774–1786. [Google Scholar] [CrossRef]
  47. Kiel, G.R.; Ziegler, M.S.; Tilley, T.D. Zirconacyclopentadiene-Annulated Polycyclic Aromatic Hydrocarbons. Angew. Chem. Int. Ed. 2017, 56, 4839–4844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Urrego-Riveros, S.; Medina, I.M.R.Y.; Duvinage, D.; Lork, E.; Sönnichsen, F.D.; Staubitz, A. Negishi’s Reagent Versus Rosenthal’s Reagent in the Formation of Zirconacyclopentadienes. Chem. Eur. J. 2019, 25, 13318–13328. [Google Scholar] [CrossRef] [Green Version]
  49. Zeng, Y.; Feng, H.; King, R.B.; Schaefer, H.F. Metallocene versus Metallabenzene Isomers of Nickel, Palladium, and Platinum. Organometallics 2014, 33, 7193–7198. [Google Scholar] [CrossRef]
  50. Ebrahimi, A.A.; Ghiasi, R.; Foroutan-Nejad, C. Topological characteristics of the Ring Critical Points and the aromaticity of groups IIIA to VIA hetero-benzenes. J. Mol. Struct. THEOCHEM 2010, 941, 47–52. [Google Scholar] [CrossRef]
  51. Foroutan-Nejad, C. Is NICS a reliable aromaticity index for transition metal clusters? Theor. Chem. Accounts 2015, 134, 8. [Google Scholar] [CrossRef]
  52. Zhao, Y.; Truhlar, D.G. Density Functionals with Broad Applicability in Chemistry. Accounts Chem. Res. 2008, 41, 157–167. [Google Scholar] [CrossRef]
  53. Noro, T.; Sekiya, M.; Koga, T. Segmented contracted basis sets for atoms H through Xe: Sapporo-(DK)-nZP sets (n = D, T, Q). Theor. Chem. Accounts 2012, 131, 1124. [Google Scholar] [CrossRef]
  54. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16 Revision B.01; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  55. Keith, T.A. AIMAll, TK Gristmill Software. Overland, Park KS, USA, 2019. [Google Scholar]
  56. Barquera-Lozada, J.E. The vorticity of the current density tensor and 3D-aromaticity. Int. J. Quantum Chem. 2019, 119, e25848. [Google Scholar] [CrossRef]
  57. Barquera-Lozada, J.E. Vorticity: Simplifying the analysis of the current density. J. Comput. Chem. 2019, 40, 2602–2610. [Google Scholar] [CrossRef] [PubMed]
  58. Rico-Sotomayor, E.M.; Barquera-Lozada, J.E. Triangulenes and their ions: Reaching the limits of Clar’s rule. Phys. Chem. Chem. Phys. 2020, 22, 24704–24711. [Google Scholar] [CrossRef]
  59. Keith, T.A.; Bader, R.F.W. Calculation of magnetic response properties using a continuous set of gauge transformations. Chem. Phys. Lett. 1993, 210, 223–231. [Google Scholar] [CrossRef]
  60. Lazzeretti, P.; Malagoli, M.; Zanasi, R. Computational approach to molecular magnetic properties by continuous transformation of the origin of the current density. Chem. Phys. Lett. 1994, 220, 299–304. [Google Scholar] [CrossRef]
  61. Steiner, E.; Fowler, P.W. Patterns of Ring Currents in Conjugated Molecules:  A Few-Electron Model Based on Orbital Contributions. J. Phys. Chem. A 2001, 105, 9553–9562. [Google Scholar] [CrossRef]
  62. Monaco, G.; Summa, F.F.; Zanasi, R. Program Package for the Calculation of Origin-Independent Electron Current Density and Derived Magnetic Properties in Molecular Systems. J. Chem. Inf. Model. 2021, 61, 270–283. [Google Scholar] [CrossRef] [PubMed]
  63. Hunter, W.E.; Atwood, J.L.; Fachinetti, G.; Floriani, C. The crystal structure of 1,1-bis(η5-cyclopentadienyl)-2,3,4,5-tetraphenylzirconole. J. Organomet. Chem. 1981, 204, 67–74. [Google Scholar] [CrossRef]
  64. Sala-Pala, J.; Amaudrut, J.; Guerchais, J.E.; Mercier, R.; Douglade, J.; Theobald, J.G. Chimie organométallique du niobium VIII. Etude détaillée de la réaction d’acétyléniques perfluorés avec (η5C5H5)2NbH3: Mise en evídence de la coupure de liaison C-F, de la formation de liaison Nb-F et de la reduction du niobium. Structure cristalline d’un dérivé niobiacyclopentadiène. J. Organomet. Chem. 1981, 204, 347–359. [Google Scholar] [CrossRef]
  65. Albers, M.O.; Waal, D.J.A.D.; Liles, D.C.; Robinson, D.J.; Singleton, E.; Wiege, M.B. The novel cyclodimerization of phenylacetylene at a ruthenium(II) centre. The synthesis and X-ray structural characterization of the first metallacyclopentatriene, [(η-C5H5)Ru(C4Ph2H2)Br], and its facile conversion into metallacyclopentadienes. J. Chem. Soc. Chem. Commun. 1986, 1680–1682. [Google Scholar] [CrossRef]
  66. Sánchez, G.; Vives, J.; Serrano, J.L.; Pérez, J.; López, G. New palladacyclopentadiene complexes containing an N,P-donor setting. Crystal structure of [Pd{C4(COOMe)4}(o-Ph2PC6H4CHNiPr)]. Inorganica Chim. Acta 2002, 328, 74–80. [Google Scholar] [CrossRef]
  67. Canovese, L.; Visentin, F.; Levi, C.; Santo, C.; Bertolasi, V. Facile synthesis and reactivity study of mixed phosphane–isocyanide Pd(II) and Pd(0) complexes. Inorganica Chim. Acta 2011, 378, 239–249. [Google Scholar] [CrossRef]
  68. Yi, C.S.; Torres-Lubian, J.R.; Liu, N.; Rheingold, A.L.; Guzei, I.A. Selective Linear Coupling Reaction of Acetylene and Acrylonitrile Catalyzed by the Well-Defined Metallacyclopentadiene Complex C5Me5(PPh3)(Cl)RuCHCHCHCH. Organometallics 1998, 17, 1257–1259. [Google Scholar] [CrossRef]
  69. Bruce, M.I.; Hall, B.C.; Skelton, B.W.; Tiekink, E.R.T.; White, A.H.; Zaitseva, N.N. Some Pentamethylcyclopentadienyl-Ruthenium Derivatives of Methyl Propiolate. Aust. J. Chem. 2000, 53, 99–107. [Google Scholar] [CrossRef]
  70. Yamazaki, H.; Wakatsuki, Y. Cobalt metallocycles: XIII. Preparation and x-ray crystallography of cobaltacyclopentadiene and dinuclear cobalt complexes. J. Organomet. Chem. 1984, 272, 251–263. [Google Scholar] [CrossRef]
  71. Nguyen, H.V.; Yeamine, M.R.; Amin, J.; Motevalli, M.; Richards, C.J. Synthesis and 1H NMR spectroscopic properties of substituted (η4-tetraarylcyclobutadiene)(η5-cyclopentadienyl)cobalt metallocenes. J. Organomet. Chem. 2008, 693, 3668–3676. [Google Scholar] [CrossRef]
  72. Ernst, C.; Walter, O.; Dinjus, E.; Arzberger, S.; Görls, H. Structural characterization of Cp*Ru-intermediates of phenylacetylene cyclotrimerization. J. FüR Prakt. Chem. 1999, 341, 801–804. [Google Scholar] [CrossRef]
  73. Ernst, C.; Walter, O.; Dinjus, E. Cp*Ru-allylcarbene complexes by nucleophilic attack of cyclic Cp*Ru-dicarbenes. J. Organomet. Chem. 2001, 627, 249–254. [Google Scholar] [CrossRef]
  74. Yamada, Y.; Mizutani, J.; Kurihara, M.; Nishihara, H. Synthesis of a new bis(ferrocenyl)ruthenacyclopentatriene compound with a significant inter-metal electronic communication. J. Organomet. Chem. 2001, 637–639, 80–83. [Google Scholar] [CrossRef]
  75. Yamamoto, Y.; Arakawa, T.; Ogawa, R.; Itoh, K. Ruthenium(II)-Catalyzed Selective Intramolecular [2 + 2 + 2] Alkyne Cyclotrimerizations. J. Am. Chem. Soc. 2003, 125, 12143–12160. [Google Scholar] [CrossRef] [PubMed]
  76. Dutta, B.; Curchod, B.F.E.; Campomanes, P.; Solari, E.; Scopelliti, R.; Rothlisberger, U.; Severin, K. Reactions of Alkynes with [RuCl(cyclopentadienyl)] Complexes: The Important First Steps. Chem. Eur. J. 2010, 16, 8400–8409. [Google Scholar] [CrossRef]
  77. Schleyer, P.V.R.; Pühlhofer, F. Recommendations for the Evaluation of Aromatic Stabilization Energies. Org. Lett. 2002, 4, 2873–2876. [Google Scholar] [CrossRef]
  78. O’Connor, J.M.; Merwin, R.; Rheingold, A.L.; Adams, M.L. Synthesis and Structural Characterization of a Diiridium .mu.-Acyl Complex. Organometallics 1995, 14, 2102–2105. [Google Scholar] [CrossRef]
  79. O’Connor, J.M.; Hiibner, K.; Rheingold, A.L.; Liable-Sands, L.M. New transition metal binding modes for creatinine: Molecular structures of [(C4R4)Ir(C4H7N3O)(PPh3)2Cl] and [(C4R4)Ir(C4H7N3O)(PPh3)2]BF4, (R = CO2CH3). Polyhedron 1997, 16, 2029–2035. [Google Scholar] [CrossRef]
  80. Wang, T.; Zhang, H.; Han, F.; Long, L.; Lin, Z.; Xia, H. cine-Substitution Reactions of Metallabenzenes: An Experimental and Computational Study. Chem. Eur. J. 2013, 19, 10982–10991. [Google Scholar] [CrossRef]
  81. Bader, R.F.W.; Stephens, M.E. Spatial localization of the electronic pair and number distributions in molecules. J. Am. Chem. Soc. 1975, 97, 7391–7399. [Google Scholar] [CrossRef]
  82. Poater, J.; Solà, M.; Duran, M.; Fradera, X. The calculation of electron localization and delocalization indices at the Hartree–Fock, density functional and post-Hartree–Fock levels of theory. Theor. Chem. Accounts 2002, 107, 362–371. [Google Scholar] [CrossRef]
  83. Firme, C.L.; Antunes, O.A.C.; Esteves, P.M. Relation between bond order and delocalization index of QTAIM. Chem. Phys. Lett. 2009, 468, 129–133. [Google Scholar] [CrossRef]
  84. Outeiral, C.; Vincent, M.A.; Pendás, A.M.; Popelier, P.L.A. Revitalizing the concept of bond order through delocalization measures in real space. Chem. Sci. 2018, 9, 5517–5529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Poater, J.; Fradera, X.; Duran, M.; Solà, M. The Delocalization Index as an Electronic Aromaticity Criterion: Application to a Series of Planar Polycyclic Aromatic Hydrocarbons. Chem. Eur. J. 2003, 9, 400–406. [Google Scholar] [CrossRef] [PubMed]
  86. Jusélius, J.; Sundholm, D.; Gauss, J. Calculation of current densities using gauge-including atomic orbitals. J. Chem. Phys. 2004, 121, 3952–3963. [Google Scholar] [CrossRef] [PubMed]
  87. Gershoni-Poranne, R.; Stanger, A. Magnetic criteria of aromaticity. Chem. Soc. Rev. 2015, 44, 6597–6615. [Google Scholar] [CrossRef] [PubMed]
  88. Sundholm, D.; Fliegl, H.; Berger, R.J.F. Calculations of magnetically induced current densities: Theory and applications. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2016, 6, 639–678. [Google Scholar] [CrossRef]
  89. Fliegl, H.; Sundholm, D.; Taubert, S.; Jusélius, J.; Klopper, W. Magnetically Induced Current Densities in Aromatic, Antiaromatic, Homoaromatic, and Nonaromatic Hydrocarbons. J. Phys. Chem. A 2009, 113, 8668–8676. [Google Scholar] [CrossRef] [PubMed]
  90. Bast, R.; Jusélius, J.; Saue, T. 4-Component relativistic calculation of the magnetically induced current density in the group 15 heteroaromatic compounds. Chem. Phys. 2009, 356, 187–194. [Google Scholar] [CrossRef]
  91. Monaco, G.; Zanasi, R.; Pelloni, S.; Lazzeretti, P. Relative Weights of σ and π Ring Currents in a Few Simple Monocycles. J. Chem. Theory Comput. 2010, 6, 3343–3351. [Google Scholar] [CrossRef] [PubMed]
  92. Cyrański, M.; Havenith, R.; Dobrowolski, M.; Gray, B.; Krygowski, T.; Fowler, P.; Jenneskens, L. The Phenalenyl Motif: A Magnetic Chameleon. Chem. Eur. J. 2007, 13, 2201–2207. [Google Scholar] [CrossRef] [PubMed]
  93. Havenith, R.W.A.; Fowler, P.W. The origin of the ring current in the all-metal aromatic, Al42-. Phys. Chem. Chem. Phys. 2006, 8, 3383–3386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Steiner, E.; Fowler, P.W. On the orbital analysis of magnetic properties. Phys. Chem. Chem. Phys. 2004, 6, 261–272. [Google Scholar] [CrossRef]
Scheme 1. Structures of the studied metallacyclopentadienes.
Scheme 1. Structures of the studied metallacyclopentadienes.
Chemistry 03 00094 sch001
Figure 1. j ( r ) and tp j ( r ) maps of complexes 5 (a) and 6 (b). The contributions to j ( r ) of HOMO (c,d), HOMO-1 (e,f) and HOMO-2 (g,h) are also shown. Colour code: ≤ 3.0 × 10 4 a.u. (dark blue), 0.0 a.u. (white) and ≥ 3.0 × 10 4 a.u. (red). The plane is parallel to the ring at 1.0 Å from it. The magnetic field, B, is pointing toward the observer, then clockwise/anticlockwise currents are diatropic/paratropic.
Figure 1. j ( r ) and tp j ( r ) maps of complexes 5 (a) and 6 (b). The contributions to j ( r ) of HOMO (c,d), HOMO-1 (e,f) and HOMO-2 (g,h) are also shown. Colour code: ≤ 3.0 × 10 4 a.u. (dark blue), 0.0 a.u. (white) and ≥ 3.0 × 10 4 a.u. (red). The plane is parallel to the ring at 1.0 Å from it. The magnetic field, B, is pointing toward the observer, then clockwise/anticlockwise currents are diatropic/paratropic.
Chemistry 03 00094 g001
Figure 2. Frontier MO and their energies of complexes 5 and 6. Energy in a.u.
Figure 2. Frontier MO and their energies of complexes 5 and 6. Energy in a.u.
Chemistry 03 00094 g002
Table 1. Selected geometrical parameters.
Table 1. Selected geometrical parameters.
Molecule d M C 2 (Å) d C 2 C 3 (Å) d C 3 C 4 (Å) ϕ MC 2 C 3 C 4 (°)
12.121.341.470.1
22.241.351.480.1
32.251.331.471.0
42.131.331.470.2
51.921.341.453.4
61.941.361.425.0
72.001.341.495.7
81.841.481.340.0
91.971.341.470.0
102.051.341.460.0
112.061.331.462.2
122.161.341.460.0
132.141.341.451.4
142.161.351.442.3
152.011.351.463.3
162.031.351.452.5
171.921.341.470.0
181.981.331.460.0
Table 2. ISE, DI and ring’s j ( r ) .
Table 2. ISE, DI and ring’s j ( r ) .
MoleculeISE (kcal/mol) DI M C 2 (a.u.) DI C 2 C 3 (a.u.) DI C 3 C 4 (a.u.) Δ DI (a.u.) j ( r ) (nA/T)
10.940.6431.7391.0910.647−0.14
2−0.240.6151.7691.0820.6870.28
3−3.700.6541.7581.0600.698−20.58
4−5.150.5131.7531.0620.691−7.86
5−1.200.8941.6831.1180.56413.93
60.571.1471.5391.2140.324−13.58
7−7.820.7931.8011.0470.7530.70
8−26.601.6901.0751.670−0.594−55.99
9−5.710.7741.7841.0660.718−2.91
10−5.940.8651.7641.0760.688−5.43
114.940.7391.6901.1060.5857.43
123.400.7751.7021.1040.5983.86
132.040.6101.6841.1200.5643.94
144.770.7321.6311.1530.4785.41
150.910.7681.7001.1060.5947.02
163.380.9451.6651.1170.5485.12
17−5.710.8841.7831.0690.715−2.50
18−4.990.9931.7481.0820.666−4.89
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Casiano-González, R.; Barquera-Lozada, J.E. Are Metallacyclopentadienes Always Non-Aromatic? Chemistry 2021, 3, 1302-1313. https://0-doi-org.brum.beds.ac.uk/10.3390/chemistry3040094

AMA Style

Casiano-González R, Barquera-Lozada JE. Are Metallacyclopentadienes Always Non-Aromatic? Chemistry. 2021; 3(4):1302-1313. https://0-doi-org.brum.beds.ac.uk/10.3390/chemistry3040094

Chicago/Turabian Style

Casiano-González, Ricardo, and José Enrique Barquera-Lozada. 2021. "Are Metallacyclopentadienes Always Non-Aromatic?" Chemistry 3, no. 4: 1302-1313. https://0-doi-org.brum.beds.ac.uk/10.3390/chemistry3040094

Article Metrics

Back to TopTop