Mechanism of Redox-Active Ligand-Assisted...

10
& Quantum Chemistry Mechanism of Redox-Active Ligand-Assisted Nitrene-Group Transfer in a Zr IV Complex: Direct Ligand-to-Ligand Charge Transfer Preferred Soumya Ghosh and Mu-Hyun Baik* [a] Abstract: The mechanism of the nitrene-group transfer reac- tion from an organic azide to isonitrile catalyzed by a Zr IV d 0 complex carrying a redox-active ligand was studied by using quantum chemical molecular-modeling methods. The key step of the reaction involves the two-electron reduction of the azide moiety to release dinitrogen and provide the ni- trene fragment, which is subsequently transferred to the iso- nitrile substrate. The reducing equivalents are supplied by the redox-active bis(2-iso-propylamido-4-methoxyphenyl)- amide ligand. The main focus of this work is on the mecha- nism of this redox reaction, in particular, two plausible mechanistic scenarios are considered: 1) the metal center may actively participate in the electron-transfer process by first recruiting the electrons from the redox-active ligand and becoming formally reduced in the process, followed by a classical metal-based reduction of the azide reactant. 2) Al- ternatively, a non-classical, direct ligand-to-ligand charge- transfer process can be envisioned, in which no appreciable amount of electron density is accumulated at the metal center during the course of the reaction. Our calculations in- dicate that the non-classical ligand-to-ligand charge-transfer mechanism is much more favorable energetically. Utilizing a series of carefully constructed putative intermediates, both mechanistic scenarios were compared and contrasted to ra- tionalize the preference for ligand-to-ligand charge-transfer mechanism. Introduction Recently, redox-active ligands have become increasingly recog- nized as an important part of the remarkable reactivity dis- played by transition-metal complexes [1–6] that promote chal- lenging chemical reactions. [7–18] For example, Heyduk and co- workers prepared a zirconium-based catalyst with a redox- active ligand, [NNN cat ]Zr IV Cl(CNtBu) 2 ; [NNN cat ] = bis(2-iso-propyl- amido-4-methoxyphenyl)-amide that transfers a nitrene group from an organic azide to an isonitrile moiety forming a carbo- diimide species. [19] The principle of combining the redox prop- erties of the metal center and non-innocent ligand is well ap- preciated. In some cases, the oxidation state of the metal center has been established by using sophisticated spectro- scopic techniques; [20] however, much less is established as to how the metal center modulates the reactivity of the redox- active ligands bound to it to promote catalysis. [21] To highlight the difference between classical, non-innocent and redox- active ligand-based mechanisms, we examined the mechanism of nitrene transfer by Heyduk’s catalyst and contrast it to the mechanism encountered with Bergman’s [Ta III Cp 2 (CH 3 )] com- plex that has previously been studied experimentally. [22] Classical metal-based versus non-innocent and redox-active ligand-based reactions Scheme 1 conceptualizes classical mononuclear metal-based reactions, in which a substrate R is reduced by first binding to the metal center (a), followed by a metal-to-ligand charge transfer (b), and product release (c). Redox-active ligands can serve as the source for reducing equivalents giving rise to an analogous sequence of events illustrated in Scheme 1 as d-e-c. After reactant binding, the electrons that originally reside on the redox-active ligand L transfer to R in a ligand-to-ligand Scheme 1. [a] Dr. S. Ghosh, Prof. Dr. M.-H. Baik Department of Chemistry, Indiana University 800 E. Kirkwood Avenue, Bloomington, IN 47405 (USA) Department of Materials Chemistry, Korea University Jochiwon-eup, Sejong-si, 339-700 (South Korea) E-mail : [email protected] Supporting information for this article is available on the WWW under http://dx.doi.org/10.1002/chem.201405738. It contains isodensity plots of the vacant d orbitals on Ta of [TaCp 2 (N 3 Ph)(CH 3 )], Cartesian coordinates, vi- brational frequencies for all structures, and energy components. Chem. Eur. J. 2015, 21, 1780 – 1789 # 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 1780 Full Paper DOI: 10.1002/chem.201405738

Transcript of Mechanism of Redox-Active Ligand-Assisted...

Page 1: Mechanism of Redox-Active Ligand-Assisted Nitrene-Groupstorage.googleapis.com/wzukusers/user-16009293/documents/... · 2015-09-16 · &Quantum Chemistry Mechanism of Redox-Active

& Quantum Chemistry

Mechanism of Redox-Active Ligand-Assisted Nitrene-GroupTransfer in a ZrIV Complex: Direct Ligand-to-Ligand ChargeTransfer Preferred

Soumya Ghosh and Mu-Hyun Baik*[a]

Abstract: The mechanism of the nitrene-group transfer reac-tion from an organic azide to isonitrile catalyzed by a ZrIV d0

complex carrying a redox-active ligand was studied by usingquantum chemical molecular-modeling methods. The keystep of the reaction involves the two-electron reduction ofthe azide moiety to release dinitrogen and provide the ni-trene fragment, which is subsequently transferred to the iso-nitrile substrate. The reducing equivalents are supplied bythe redox-active bis(2-iso-propylamido-4-methoxyphenyl)-amide ligand. The main focus of this work is on the mecha-nism of this redox reaction, in particular, two plausiblemechanistic scenarios are considered: 1) the metal centermay actively participate in the electron-transfer process by

first recruiting the electrons from the redox-active ligandand becoming formally reduced in the process, followed bya classical metal-based reduction of the azide reactant. 2) Al-ternatively, a non-classical, direct ligand-to-ligand charge-transfer process can be envisioned, in which no appreciableamount of electron density is accumulated at the metalcenter during the course of the reaction. Our calculations in-dicate that the non-classical ligand-to-ligand charge-transfermechanism is much more favorable energetically. Utilizinga series of carefully constructed putative intermediates, bothmechanistic scenarios were compared and contrasted to ra-tionalize the preference for ligand-to-ligand charge-transfermechanism.

Introduction

Recently, redox-active ligands have become increasingly recog-nized as an important part of the remarkable reactivity dis-played by transition-metal complexes[1–6] that promote chal-lenging chemical reactions.[7–18] For example, Heyduk and co-workers prepared a zirconium-based catalyst with a redox-active ligand, [NNNcat]ZrIVCl(CNtBu)2; [NNNcat] = bis(2-iso-propyl-amido-4-methoxyphenyl)-amide that transfers a nitrene groupfrom an organic azide to an isonitrile moiety forming a carbo-diimide species.[19] The principle of combining the redox prop-erties of the metal center and non-innocent ligand is well ap-preciated. In some cases, the oxidation state of the metalcenter has been established by using sophisticated spectro-scopic techniques;[20] however, much less is established as tohow the metal center modulates the reactivity of the redox-active ligands bound to it to promote catalysis.[21] To highlightthe difference between classical, non-innocent and redox-active ligand-based mechanisms, we examined the mechanism

of nitrene transfer by Heyduk’s catalyst and contrast it to themechanism encountered with Bergman’s [TaIIICp2(CH3)] com-plex that has previously been studied experimentally.[22]

Classical metal-based versus non-innocent and redox-activeligand-based reactions

Scheme 1 conceptualizes classical mononuclear metal-basedreactions, in which a substrate R is reduced by first binding to

the metal center (a), followed by a metal-to-ligand chargetransfer (b), and product release (c). Redox-active ligands canserve as the source for reducing equivalents giving rise to ananalogous sequence of events illustrated in Scheme 1 as d-e-c.After reactant binding, the electrons that originally reside onthe redox-active ligand L transfer to R in a ligand-to-ligand

Scheme 1.[a] Dr. S. Ghosh, Prof. Dr. M.-H. Baik

Department of Chemistry, Indiana University800 E. Kirkwood Avenue, Bloomington, IN 47405 (USA)Department of Materials Chemistry, Korea UniversityJochiwon-eup, Sejong-si, 339-700 (South Korea)E-mail : [email protected]

Supporting information for this article is available on the WWW underhttp ://dx.doi.org/10.1002/chem.201405738. It contains isodensity plots ofthe vacant d orbitals on Ta of [TaCp2(N3Ph)(CH3)] , Cartesian coordinates, vi-brational frequencies for all structures, and energy components.

Chem. Eur. J. 2015, 21, 1780 – 1789 � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim1780

Full PaperDOI: 10.1002/chem.201405738

Page 2: Mechanism of Redox-Active Ligand-Assisted Nitrene-Groupstorage.googleapis.com/wzukusers/user-16009293/documents/... · 2015-09-16 · &Quantum Chemistry Mechanism of Redox-Active

electron-transfer process, and the reduced product is released.One key question that remains open in this plausible concep-tual scheme is whether or not the metal center plays an explic-it role in the electron-transfer process. Another closely relatedmechanism involves non-innocent ligands. In this case, theelectron density is delocalized over the metal ligand frame-work, preventing the assignment of a distinct oxidation stateto the metal center, as shown schematically by g-h-c. The dis-tinction between “redox-active” and “non-innocent” ligands isoften subtle and depends heavily on the extent of metal–ligand covalency.[23] In the case of redox-active ligands, do theelectrons transfer directly from the L to R (step e in Scheme 1), ordo they first reduce the metal center and then transfer to R(step f in Scheme 1)?

The mechanism becomes further complicated if the elec-trons are transferred in a stepwise fashion, one electron ata time. Single electron transfer may result in several radicaloidspecies, depending on where the two unpaired electrons arelocated. The several possible intermediates are shown inScheme 2. Once the two electrons are separated into two un-

paired electrons located at two distinct centers, spin crossoverto a higher spin state may take place depending on the extentof spin–orbit coupling between the two magnetic centers.These radicaloid species have previously been shown toengage in remarkable reactivity.[17, 24]

Intuitively, we expect the d-e-c (Scheme 1)/d-e1-e2-c (Scheme 2) pathway to be feasible when: 1) the local redoxpotentials of the L and R ligands are matched as to make theelectron transfer energetically viable; and 2) the structural ar-rangement of the two ligands L and R in the metal complexallow for a donor–acceptor coupling. The donor orbital on theredox-active supporting ligand L and the acceptor orbital onthe substrate R need to overlap either directly or througha metal d orbital for the electron transfer to take place. Other-wise, pathway d-e-c/d-e1-e2-c will become improbable. Path-way d-f-b-c/d-f1-f2-b1-b2-c/d-f1-f4-e2-c, on the other hand,does not require direct communication between ligands L andR. The donor and the acceptor orbitals on L and R may for in-stance overlap with two orthogonal metal d orbitals. Internalelectronic reorganization within the metal d orbitals can leadto electron transfer from L to R in a stepwise mechanism. Thetwo one-electron transfer pathways d-f1-f2-b1-b2 and d-f1-f4-e2 are very similar and differ only in the relative affinity of themetal center to hold an extra electron. The most unusual sce-

nario is depicted by pathway d-f1-f3-b2-c that requires theoverlap between L and R ligands to increase significantly uponone electron reduction of the intervening metal center. If theredox potentials of L and M are close enough, we will encoun-ter pathway g-h-c/g-h1-h2-c, in which case it will be impossibleto pinpoint whether the electron has been transferred from Lor M. It is fundamentally important for catalyst design to un-derstand whether the metal center is reduced during the reac-tion. If a stable intermediate with a reduced metal center is in-volved, the catalytic cycle can potentially be optimized bychanging the redox properties of the metal center. However, ifsuch an intermediate is absent, the main function of the metalcenter may be to hold the two ligands in a structural arrange-ment that maximizes the communication between them. If so,the redox catalytic process will be mostly controlled by theproperties of the L and R moieties, and the most viable strat-egies of reaction control will lie in functionalizing the ligands.The nitrene-group transfer reaction studied in this work pro-vides a unique opportunity to address this mechanistic ques-tion in detail.

Computational details

All calculations were carried out using DFT as implemented inthe Jaguar 7.0 suite[25] of ab initio quantum chemistry pro-grams. Geometry optimizations were performed with theB3LYP[26–30] functional and the 6-31G** basis set. Zr was repre-sented using the Los Alamos LACVP basis[31, 32] that includes rel-ativistic effective core potentials. The energies of the optimizedstructures were reevaluated by additional single-point calcula-tions on each optimized geometry using Dunning’s correlationconsistent triple-z basis set[33] cc-pVTZ(-f) that includesa double set of polarization functions. For Zr, we used a modi-fied version of LACVP, designated as LACV3P, in which the ex-ponents were decontracted to match the effective core poten-tial with triple-z quality. Vibrational/rotational/translational en-tropies of the solute(s) were included using standard thermo-dynamic approximations. Solvation energies were evaluated bya self-consistent reaction field (SCRF)[34–36] approach based onaccurate numerical solutions of the linearized Poisson–Boltz-mann equation.[37] Solvation calculations were carried out atthe optimized gas phase geometry employing the dielectricconstant of e= 7.6 (tetrahydrofuran). As is the case for all con-tinuum models, the solvation energies are subject to empiricalparametrization of the atomic radii that are used to generatethe solute surface. We employ the standard set of optimizedradii for H (1.150 �), C (1.900 �), N (1.600 �), and O (1.600 �).The scaled van der Waals radius used for Zr is 1.562 �.

Convergence to plausible electronic states that correspondto conceptually meaningful electronic configurations wasmonitored by carefully observing the Mulliken spin densitiesand visualizing the frontier molecular orbitals. When multipleminima were encountered, we compared the total energiesand chose the structure with the lowest energy. Antiferromag-netically (AF) coupled states were modeled by using Noodle-man’s broken symmetry (BS) formalism without spin projec-tion,[38–40] because the spin projection corrections are uniformly

Scheme 2.

Chem. Eur. J. 2015, 21, 1780 – 1789 www.chemeurj.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim1781

Full Paper

Page 3: Mechanism of Redox-Active Ligand-Assisted Nitrene-Groupstorage.googleapis.com/wzukusers/user-16009293/documents/... · 2015-09-16 · &Quantum Chemistry Mechanism of Redox-Active

found to be negligibly small in these systems. Energy compo-nents have been computed following the protocol of our pre-vious work.[41]

DGðsolÞ ¼ DGðgasÞ þ DDGsolv ð1Þ

DGðgasÞ ¼ DHðgasÞ�TDSðgasÞ ð2Þ

DHðgasÞ ¼ DEðSCFÞ þ DZPE ð3Þ

in which DG(gas) = free energy in gas phase; DDGsolv = freeenergy of solvation, as was computed by using the continuumsolvation model ; DH(gas) = enthalpy in gas phase; T = temper-ature (298.15 K); DS(gas) = entropy difference in gas phase;DE(SCF) = self-consistent field energy, that is, “raw” electronicenergy, as was computed from the SCF procedure; DZPE = vi-brational zero-point energy difference.

Results and Discussion

The overall mechanism we found to be most plausible is sum-marized in Scheme 3.[42] This mechanism agrees well with what

was proposed previously by Heyduk,[31] but several unanticipat-ed intermediates have been located. The catalytic cycle beginswith the dissociation of one of the isonitrile ligands from reac-tant complex 1 to generate the five-coordinate Zr complex 2,which requires 4.0 kcal mol�1 and is therefore expected to bereversible in good agreement with previous experimental ob-servations.[19] Intermediate 2 binds the organic azide substrateutilizing the nitrogen adjacent to the alkyl group in an h1-alkyl-amino fashion. This step, 2!3, is energetically uphill by17.0 kcal mol�1. Transferring a single electron from [NNNcat]3� to

the azide ligand leads to an energetically viable rearrangementof 3 to 4. Dissociation of N2 from 4 is also feasible with an acti-vation barrier of 27.0 kcal mol�1. Release of dinitrogen placesthe isonitrile ligand and the nitrene group trans to each otheraffording intermediate 5, which is not helpful to complete thereaction. To transfer the nitrene moiety from the metal centerto the isonitrile ligand, both groups must be cis to each other.Dissociation of the isonitrile ligand is calculated to be downhillby 6.9 kcal mol�1. Reattachment of the ligand to the Zr com-plex cis to the nitrene, 7, is uphill by 7.6 kcal mol�1. The conver-sion of 7!8 regenerates the catecholate form of the redoxactive supporting ligand.

Binding of azide to the metal center

The azide substrate can either bind to the metal center in h1-terminal fashion to generate intermediate 3 a or in h1-alkylami-no mode, which leads to 3 (Scheme 4). The former is 3.5 kcalmol�1 (2.2 kcal mol�1 electronically) lower in energy than the

latter. Intuitively,[43–45] h1-alkylamino binding should be energet-ically more favorable than h1-terminal binding, because thelone pair electrons on the amino nitrogen are expected to besignificantly higher in energy than the analogous electrons lo-cated on the terminal nitrogen.[46] However, terminal binding isfavored sterically and slightly outweighs the electronic prefer-ence in this case, giving rise to the marginally favored h1-termi-nal binding in 3 a over h1-alkylamino binding in 3. Can the rela-tive stability of these binding modes be influenced by metal oxi-dation states? Previously, Proulx and Bergman showed that theazide binds exclusively through the terminal nitrogen.[22] A sim-ilar observation was made by Cummins and co-workers.[47]

Scheme 4 compares the free-energy changes associated withthe azide binding for Heyduk’s catalyst and Bergman’s Ta com-plex. The azide binding to the Zr complex 2 is uphill by 9.6and 13.1 kcal mol�1 for the terminal and alkylamino bindingmodes, respectively. Curiously, the azide binding to the Tacomplex 9 is 34.7 kcal mol�1 downhill. Figure 1 illustrates thecomputed structures of the azide bound Ta and Zr complexes3 a and 10. The most clear difference is that the azide liganddisplays a nearly linear N-N-N bond angle of 1748 in 3 a, where-

Scheme 3.

Scheme 4. Binding of tBuN3 to ZrIV and TaIII.

Chem. Eur. J. 2015, 21, 1780 – 1789 www.chemeurj.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim1782

Full Paper

Page 4: Mechanism of Redox-Active Ligand-Assisted Nitrene-Groupstorage.googleapis.com/wzukusers/user-16009293/documents/... · 2015-09-16 · &Quantum Chemistry Mechanism of Redox-Active

as a much more acute 1178 is adopted in 10.[48] These stark dif-ferences in structure and free energy of binding point to fun-damentally different metal–ligand interactions in these twocomplexes: The binding of the azide to the TaIII center in 9 isformally an oxidative addition process, in which the metal oxi-dation state increases to V with the concomitant reduction ofthe azide ligand. The reduction of the ligand disrupts the p

framework resulting in the bending of the azide moiety. The Tacomplex shows five vacant metal d-based frontier orbitals, aswas expected for a TaV d0 complex. As a result, the electrostaticinteraction between the TaV center and the negative ligand ismuch stronger in 10 than in 3 a, in which a formally neutralazide ligand is bound to a ZrIV center. The azide binding in 3/3 a is best characterized as a classical Lewis acid/base type ofbonding without any change in the oxidation of the metal. Inaddition, the azido ligand in 10 becomes formally a four-elec-tron donor, giving rise to a significant Ta�N double-bond char-acter,[49] as illustrated in Scheme 4, which increases the Ta�Nbond strength even more.

Different pathways for the release of dinitrogen

Several mechanistic proposals exist for dinitrogen extrusionfrom an organic azide moiety.[50, 51] Most relevantly, the classicalStaudinger reaction[52–56] between phosphines and azides isthought to involve a transition state that consists of a four-membered ring.[57–59] By using 15N-labeled azides, Bergmanet al. showed that the nitrogen atom that is directly bound tothe organic fragment becomes the imido nitrogen that re-mains bound to the metal upon extrusion of dinitrogen,[22]

which is a direct consequence of the metallacyclic transitionstate, as illustrated in Scheme 5. The fundamentally differentelectronic structures of the Ta and Zr complexes translate to

significantly different mechanisms for dinitrogen extrusion inthese two seemingly related complexes. The two most plausi-ble mechanisms of dinitrogen release for the ZrIV complex arepresented in Scheme 6.

Pathway 1 is analogous to the mechanism discussed above,but requires an intramolecular electron transfer to trigger thenuclear rearrangements leading to the imido product. In theazide-bound Zr complex 3 a, a single electron-transfer eventfrom the redox-active support ligand [NNNcat]3� to the alkylazide to afford intermediate 4 a. Our calculations suggest thatthis step is energetically feasible, because the intermediates3 a and 4 a are practically isoenergetic at 9.6 and 8.2 kcal mol�1,respectively. Intermediate 4 a rearranges to form the four-membered metallacycle 4 b at 23.1 kcal mol�1. The lowestenergy barrier for this step is 38.0 kcal mol�1, which is muchtoo high to be mechanistically relevant. In the Ta complex dis-cussed above, the reduction of the azide moiety was facilitatedby the oxidative addition process, and the resonance-stabilizedp-bond network was already disrupted in the intermediate 10.Consequently, the metallacycle formation was energeticallyeasy to accomplish. In the Zr complex 3 a, the p-bond networkin the azide fragment is intact, as was discussed above, andadditional energy penalties must be paid if the four-memberedmetallacycle is to be formed.

In pathway 2, intermediate 3 is envisioned to liberate dini-trogen directly in a single step, identified as process A inScheme 6. This step requires the transfer of two electrons fromthe ligand fragment [NNNcat]3� to the bound azide reactant.Concomitant to this azide reduction, N2 can be liberated ina concerted fashion affording the imido product, complex 5.This concerted, single-step mechanism is associated with a sig-nificantly high barrier of 34.5 kcal mol�1. Alternatively, the twoelectrons can be transferred in a stepwise fashion, designatedby pathway B in Scheme 5, along with a significant structuralreorganization of the azide fragment prior to the final step ofN2 loss. This pathway is found to be more relevant energetical-ly, as shown in Figure 2 for two different substituents on theazide ligand. The presence of stable intermediates like 4 and 4’in the free-energy landscape is characteristics of redox active li-gands. However, in both pathways, one needs to transfer over-all two electrons to trigger dinitrogen extrusion. The significant

Figure 1. Computed structures of 3 a and 10. Only a few selected hydrogenatoms are shown for clarity.

Scheme 5. Release of dinitrogen from the Ta complex.[22]

Scheme 6. Different pathways for dinitrogen release from the ZrIV complex.

Chem. Eur. J. 2015, 21, 1780 – 1789 www.chemeurj.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim1783

Full Paper

Page 5: Mechanism of Redox-Active Ligand-Assisted Nitrene-Groupstorage.googleapis.com/wzukusers/user-16009293/documents/... · 2015-09-16 · &Quantum Chemistry Mechanism of Redox-Active

difference between the transition states for N2 release for path-ways A and B poses a couple of interesting questions:

1) Why is the electron transfer to the reduced and rearrangedazide fragment energetically easier than reducing the unmodifiedalkyl azide twice ? To gain further insight into this structuralchange coupled to charge redistribution, we decoupled theelectron affinity of the azide molecule into the vertical elec-tron-attachment energy and the structural relaxation energyby first calculating the energies of reduced species without al-lowing structural relaxation to take place, and then obtainingthe change of energy that the structural relaxation affords.These artificial energy components are summarized as a dia-gram in Scheme 7. We call these diagrams theoretical squareschemes,[60] after the square schemes used in electrochemistry,in which intrinsically coupled chemical and electrochemicalsteps are decoupled from each other.[61–64]

Scheme 7 shows that reducing A to A*� at the equilibriumgeometry of A is downhill by 23.1 kcal mol�1 (step Ia inScheme 7),[65] and the structural relaxation of A*� to A1

�affords�24.8 kcal mol�1 (step Ib). Interestingly, these energies arewithin the same order of magnitude. Typically, electron-attach-ment energies are expected to be much greater in magnitudethan the structural changes that accompany them. The unusu-ally large structural relaxation energy originates from the par-tial localization of the singly occupied molecular orbital(SOMO) on the two N atoms at the terminal end, as illustratedin Figure 3, which decreases the anti-bonding character signifi-cantly as the N�N bond elongates. Thus, the electron transferand structural changes are strongly coupled to each other. Al-ternatively, A can be distorted first to the equilibrium geome-try of the one-electron reduced species A1

� . This structurally

distorted species is represented by A1* in the lower half ofScheme 7. Analyses of these analogous steps may be meaning-ful for active sites in enzymes and more complex catalytic sys-tems where conformational changes can be imposed on thesubstrate prior to the redox events. For our purposes, thesesteps, shown in grey in Scheme 7, are associated with unrealis-tic energies and are not enlightening. Therefore, they are onlyshown for completeness. The square scheme allows estimatingthe energies required for reducing an isolated alkyl azide bytwo electrons in concerted or stepwise fashion. As was dis-cussed above, the addition of the first electron to the azidefragment triggers substantial structural reorganization, whichgives rise to a significant stabilization of the reduced species.Quantifying how this structural change impacts the second re-duction is important for understanding the relative energeticsof pathways A and B outlined in Scheme 6.

Our calculations indicate that the second reduction at thegeometry of A, A*�! A*2�, is uphill by 17.9 kcal mol�1 imply-ing that the electron is unbound, that is, it is energetically notpossible to add two electrons to A without structural change.Figure 4 shows isodensity plots of the redox-active orbitals ofthe azide fragment. Also shown in the same figure are the cor-responding redox active orbitals of the Zr complex. The LUMOof A is a N-N-N p* and N�C s-bonding orbital. Thus, injectingan electron into this molecular orbital (MO) will cause structur-al stress across the N�N bond. When the structural relaxation

Figure 2. Free-energy profile for dinitrogen extrusion for tert-butyl and para-(tert-butyl) phenyl substituents. Transition states marked with * are not ex-plicitly calculated and are only shown for illustration.

Scheme 7. Theoretical square scheme for the reduction of tBuN3 by twoelectrons.

Figure 3. Isodensity plots (isodensity value=0.05 au) of SOMOs of A*�1 andA1�1.

Chem. Eur. J. 2015, 21, 1780 – 1789 www.chemeurj.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim1784

Full Paper

Page 6: Mechanism of Redox-Active Ligand-Assisted Nitrene-Groupstorage.googleapis.com/wzukusers/user-16009293/documents/... · 2015-09-16 · &Quantum Chemistry Mechanism of Redox-Active

is not allowed, as in our case, the N-N-N p* orbital is furtherdestabilized and becomes higher in energy than the fragmentMO based on the alkyl substituent resulting in N�C s* anti-bonding interaction, illustrated by 29b in Figure 4. By invokinga structural relaxation after the transfer of the first electron,the p* orbital of the azide fragment localizes on the two Natoms at the terminal end analogous to the SOMO discussedabove. At this relaxed geometry, A1

� can accept an electroneven without any additional structural change (A1

�!A1*2�)with an electron attachment energy of 15.4 kcal mol�1 (step IIain Scheme 6). The spatial localization of the p* orbital results inthe unequal distribution of spin and charge on the different Natoms. The unpaired spin density resides on the two N atomsat the terminal end, whereas the negative charge is mostlyconcentrated on the alkylamino nitrogen. In the metal com-plex, this electron-density distribution enhances the favorableCoulombic interaction between the ZrIV center and the azidefragment and compensates partially for the energy penalty as-sociated with the destruction of the p resonance.

2) Does the metal center change its redox state during N2 re-lease ? A significant charge and/or spin buildup on the metalcenter as the transition state is traversed is expected if themetal is electronically involved in electron transfer. The partialcharge[66] on Zr decreases from 1.63 in 3 to 1.23 at the transi-tion state, and increases back to 1.60 in 5 in pathway A, indi-cating a significant accumulation of charge density on themetal center at the transition state. Note that the computedpartial charge value of 1.63 can be correlated to the formal oxi-dation state of IV on Zr. Although oxidation states are ofcourse only a formal concept, they continue to be useful forvisualizing the electronic structure. A shift of the partial chargeby as much as 0.4 can be interpreted as a significant changeof the electrostatic environment at the metal center, which isconsistent with a change of oxidation state. Thus, we assignthe formal oxidation state of Zr in the transition state to be III.In pathway B, in which each electron is transferred in a step-wise and energetically more favorable fashion, there is practi-cally no spin density (Mulliken spin-density value is 0.019) on

the Zr center in 4-TS, suggestingthat the electron moves fromthe [NNN]2� ligand to the azidefragment directly without for-mally changing the oxidationstate of the metal. In conclusion,our calculations suggest that themost viable single-electron path-way involves no change in oxi-dation state at the metal centerduring ligand-to-ligand chargetransfer.

This result is somewhat sur-prising, because we may haveexpected a more active role ofthe transition-metal center inthe redox event. A better under-standing could be derived, if wecould locate the hypothetical

open-shell singlet state 3OS, in which the metal has becomea ZrIII d1 center, as illustrated in Figure 5. Obtaining sucha state proved difficult, even when we approximated it at thegeometry of 3, to afford what we label as 3OS* in 5, becausethe electron can easily move back to the redox active tris(ami-do) ligand (IIIa-rev in Figure 5) or can be transferred to the in-coming substrate accompanied by structural change, illustrat-ed as step IV in Figure 5. However, the triplet analogue of 3OS*,which we label as 3T* is much easier to locate in our calcula-tions. Because the two unpaired electrons have the same spin,the reverse charge-transfer process, step IIIa-rev, becomes im-possible. The antiferromagnetic coupling in the hypotheticalcomplex 3OS* is expected to be weak, which implies that theenergy difference between 3OS* and 3T* is equally small, a fewkcal mol�1 at best. Therefore, the energy of 3T* provides a rea-sonable estimate for the energy of the hypothetical intermedi-ate 3OS*. Our calculations show that 3T* lies 36.4 kcal mol�1

higher in energy than 3. Because we did not allow the struc-ture to change, this energy represents an upper bound esti-

Figure 4. Isodensity plots (isodensity value=0.05 au) of redox-active orbitals of A, A*�1, A1�1, 3, and 4.

Figure 5. Energy of the triplet state with reference to the closed-shell singletstate for isonitrile and tert-butyl azide ligands.

Chem. Eur. J. 2015, 21, 1780 – 1789 www.chemeurj.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim1785

Full Paper

Page 7: Mechanism of Redox-Active Ligand-Assisted Nitrene-Groupstorage.googleapis.com/wzukusers/user-16009293/documents/... · 2015-09-16 · &Quantum Chemistry Mechanism of Redox-Active

mate. Not surprisingly, a very similar energy difference of35.2 kcal mol�1 is found for 1T*, in which the azide moiety wasreplaced by isonitrile.

To estimate the lower limit for the metal-based reductionenergy, we must relax the geometries of 1T* and 3T*. Uponstructural relaxation, the electron on the metal center has thechoice to either be transferred to the substrate according tostep IV in Figure 5 or to remain on the metal center. Our calcu-lations show that the metal-to-ligand charge transfer onlytakes place for L = tBu�N3, whereas the metal retains the elec-tron when L = isonitrile. This difference in electron distributioncan be understood by considering the relative energies of thep-type orbitals on the metal center and the substrate withinthe complex, as illustrated in Figure 6. To enable spontaneous

metal-to-ligand charge transfer, the metal dp orbital must behigher in energy than the ligand-p* orbital (case 1 in Figure 6).Conversely, if the ligand p* orbital is higher in energy than themetal dp orbital, (case 2 in Figure 6), the metal dp/ligand p*bonding orbital is metal dominated, and the electron willremain on the metal center upon occupation of this MO.

Species 3, in which L is the azide fragment, is a suitable rep-resentative of case 1, because the delocalized N3 p* is relativelylow in energy. Species 1, on the other hand, is a good candi-date for case 2, because the p* orbital of isonitrile will bemuch higher in energy. Figure 7 shows the relevant MOs basedon the metal center and the ligands along with percentagecomposition in each MO for a) species 3, in which the p* orbi-tal on the azide fragment is accessible; and b) species 1, inwhich the p* orbital on the isonitrile fragment is much higherin energy. As was expected, the bonding orbitals in 1(<128> , <130> ) have more metal character than the equiv-alent orbitals in 3 (<144> , <145> , <147> ) confirming ourclassification of these two complexes. The calculated Mulliken

spin density of Zr in 1T is 0.73 confirming that there is a singleunpaired electron located on the metal, consistent with a ZrIII

d1 center. This putative intermediate is 25.8 kcal mol�1 higherin energy than 1, providing a lower limit for the energy re-quired to access a ZrIII complex within the given ligand envi-ronment. Thus, we estimate that the energy required to tra-verse a one-electron reduced intermediate consistent witha fully reduced ZrIII center ranges from 25 to 35 kcal mol�1. Be-cause 3 is already at a relative energy of 17 kcal mol�1 in thereaction energy profile (Figure 2), this additional energy makesthe single-electron pathway involving the ZrIII d1 species mech-anistically irrelevant.

Based on the results discussed above, it appears that thesingle-electron ligand-to-ligand charge transfer (LLCT) route isthe mechanism by which multiple electron transfer takesplace. Although we must examine many other systems ata similar level of detail to draw a general conclusion, we aretempted to speculate that what we presented above may beone of the general mechanisms by which early transitionmetals bound to redox-active ligands can promote redox reac-tions: The mediation of electron transfer without gaining anydiscernable electron density may be a special feature that char-

Figure 6. Conceptual MO diagram illustrating the interaction betweenmetal-based d and * on L fragment in the complex.

Figure 7. Isodensity plots (isodensity value=0.05 au) of relevant MOs of spe-cies 3 and 1. We have also included conceptual picture of the MOs, MOnumber in angled brackets and percentage of metal character in squarebrackets.

Chem. Eur. J. 2015, 21, 1780 – 1789 www.chemeurj.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim1786

Full Paper

Page 8: Mechanism of Redox-Active Ligand-Assisted Nitrene-Groupstorage.googleapis.com/wzukusers/user-16009293/documents/... · 2015-09-16 · &Quantum Chemistry Mechanism of Redox-Active

acterizes early transition metals. One interesting conceptualquestion is how this mechanism changes, if late transitionmetals are employed. In general, we expect the p-type metal–ligand interactions to weaken when we move to late transitionmetals due to poor overlap between metal and ligand p orbi-tals. This diminished metal–ligand interaction and increasinglymore positive redox potentials will affect the relative orderingof the singlet and triplet states. Figure 8 illustrates qualitatively

the relative ordering of the different states when we movealong the periodic table. Because the metal–ligand p interac-tion becomes less important when we move from early transi-tion metals to late transition metals, the multireference open-shell singlet state becomes dominant for late-transition-metalcomplexes with redox active ligands. Thus, we expect theligand-to-metal charge-transfer pathways to dominate for late-transition-metal complexes in general.[67–69] This is a new con-ceptual proposal that attempts to classify and distinguish theearly- and late-transition-metal complexes bound to redoxactive ligands according to their mechanistic characteristics.This mechanistic hypothesis is consistent with catalytic hydroa-mination reactions of early-transition-metal–imido com-plexes[70–73] and olefin aziridination reactions of late-transition-metal–nitrene systems.[74–76] From a practical perspective, theabove-discussed hypothesis is simplistic for the following twomajor reasons: 1) the ligand under study needs to bind bothearly- and late-transition metals ; and 2) the coordination ge-ometry around the metal center usually changes upon movingfrom electron-deficient to electron-rich metals. Clearly, muchmore comparative work must be done to test this hypothesisin the future.

Transfer of nitrene group to isonitrile ligand

The transition state for nitrene transfer, 7!8, can be envi-sioned as a nucleophilic attack of the nitrene group on carbon

center of the isonitrile ligand. This two-electron reduction canoccur in a concerted or a stepwise fashion. A one-step barrierof 25.8 kcal mol�1 relative to 6, as shown in Figure 10, is easilyaccessible at 55 8C. Despite significant efforts, we were unableto locate any open-shell intermediate analogous to species 4shown in Scheme 2 for this process. As illustrated in Figure 9,the computed structure of the transition state 7-TS shows sig-nificant bending of the C-N-C angle of the isonitrile ligand indi-cating substantial charge transfer from the imido moiety tothe p* orbital of the isonitrile fragment. When the reactionproceeds beyond the transition state, the two electrons in theZr�C bond are transferred to the redox-active ligand alongwith C�N coupling. In this case, the reduction of the quininoidto the catecholate form results in the favorable aromatizationof the redox-active ligand. Illustrating the features outlined inthe discussion above, the concerted pathway is feasible due to1) accessibility of the redox-active orbital on the supportingligand; and 2) large thermodynamic driving force for C�Nbond formation. Replacement of the product by another mole-

Figure 8. Relative ordering of different states resulting from the interactionbetween the redox-active ligand, L and the metal center in the manifold.The triplet state is taken as the reference. The actual energy differences arearbitrary.

Figure 9. Computed structure of 7-TS ; H atoms are removed for clarity.

Figure 10. Free-energy profile for dinitrogen cleavage for tert-butyl andpara-(tert-butyl) phenyl substituents.

Chem. Eur. J. 2015, 21, 1780 – 1789 www.chemeurj.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim1787

Full Paper

Page 9: Mechanism of Redox-Active Ligand-Assisted Nitrene-Groupstorage.googleapis.com/wzukusers/user-16009293/documents/... · 2015-09-16 · &Quantum Chemistry Mechanism of Redox-Active

cule of isonitrile regenerates the active form of the catalystand closes the catalytic cycle.

Nitrene-group transfer versus dimer formation

Interestingly, substitution of the tert-butyl group with para-(tert-butyl) phenyl moiety decreases the activation barrier of ni-trene-group transfer from the azide to the metal center by3.7 kcal mol�1 making it facile at room temperature, as illustrat-ed in Figure 2. Intuitively, this energy difference must originatefrom the different steric requirements of the two substituents.The initial binding of the two organic azides is energeticallyequivalent with 3 and 3’ being 17.0 versus 16.2 kcal mol�1, re-spectively. Upon the transfer of one electron, the Zr�N dis-tance decreases from 2.5 to 2.2 �, which brings the organicgroup of the azide closer to the substituents on the terminalnitrogen atoms of the auxiliary ligand. Hence, the complexwith tert-butyl group experiences a higher degree of stericclash, and the energy difference between 4 and 4’ becomes4.4 kcal mol�1. This energy difference is carried over to their re-spective transition states 4-TS and 4’-TS, as illustrated inFigure 2. Although the barrier for nitrene-group transfer is1.2 kcal mol�1 lower in energy than N2 extrusion from 4 for thetert-butyl substituent, the equivalent process is 4.5 kcal mol�1

higher when the substituent is the para-(tert-butyl) phenylgroup as illustrated in Figure 10. The relative difficulty in thenitrene-group transfer step from the metal center to the isoni-trile ligand correlates well with the nucleophilicity of the tert-butyl and para-(tert-butyl) phenyl groups. The N�C(phenyl)bond length is 1.36 � versus 1.44 � for N�C(tBu) bond. Theshortening of N�C bond length implies delocalization of thelone pair on the imido nitrogen into the phenyl ring. This phe-nomenon decreases the nucleophilicity of the imido nitrogenthat results in the increase of the activation barrier for the ni-trene-group transfer to the isonitrile ligand. A barrier of27.8 kcal mol�1 is difficult to overcome at room temperature.Experimentally, the complex with the para-(tert-butyl) phenylgroup undergoes dimerization instead of nitrene transfer. Ourcomputations suggest that electronically, this dimer formationis favored by 6.6 kcal mol�1. The analogous step for the tert-butyl group is disfavored by 22.8 kcal mol�1 due to steric inter-action with the [NNNq]� supporting ligand. The trans isomer, inwhich the two Cl� ligands are on the opposite sides, wasfound to be higher in energy than the corresponding cisisomer in both cases due to steric interaction between the iso-propyl groups on the [NNNq]� ligand.

Conclusion

We demonstrated how the presence of the redox-active sup-porting ligand [NNNcat]3� changes the mechanism of nitrene-group transfer from an organic azide to a metal center. The tra-ditional metal-to-ligand charge-transfer pathway, a-b-c inScheme 1 as was demonstrated by Bergman’s Ta complex, hasbeen replaced by direct ligand-to-ligand electron transfer, inwhich the metal center acts only as a scaffold. We found thatthe transfer of two electrons from the catecholate ligand to

the azide fragment leading to the ZrIV–amido complex is ac-complished in a stepwise fashion analogous to pathway d-e1-e2-c in Scheme 2. In contrast, subsequent transfer of the ni-trene moiety to the isonitrile ligand occurs in a single step,similar to pathway d-e-c in Scheme 1, to regenerate the cate-cholate ligand. Consequently, any exceedingly exothermic stepis avoided, making the nitrene-group-transfer reaction catalyticunder moderate conditions.

Acknowledgements

We thank the NSF(0116050, CHE-0645381, CHE-1001589), theResearch Corporation (Scialog Award to M.H.B.) and the Na-tional Research Foundation of Korea for a WCU Award toKorea University (R31-2012-000-10035-0).

Keywords: charge transfer · density functional calculations ·quantum chemistry · zirconium

[1] C. G. Pierpont, C. W. Lange, Prog. Inorg. Chem. 1994, 41, 331.[2] J. L. Boyer, J. Rochford, M.-K. Tsai, J. T. Muckerman, E. Fujita, Coord.

Chem. Rev. 2010, 254, 309.[3] J. Rochford, M.-K. Tsai, D. J. Szalda, J. L. Boyer, J. T. Muckerman, E. Fujita,

Inorg. Chem. 2009, 48, 4372 – 4383.[4] K. J. Blackmore, J. W. Ziller, A. F. Heyduk, Inorg. Chem. 2005, 44, 5559.[5] M. R. Haneline, A. F. Heyduk, J. Am. Chem. Soc. 2006, 128, 8410.[6] C. Gunanathan, D. Milstein, Acc. Chem. Res. 2011, 44, 588.[7] S. Mukherjee, T. Weyhermuller, E. Bothe, K. Wieghardt, P. Chaudhuri,

Dalton Trans. 2004, 3842.[8] W. Kaim, Dalton Trans. 2003, 761.[9] M. R. Ringenberg, S. L. Kokatam, Z. M. Heiden, T. B. Rauchfuss, J. Am.

Chem. Soc. 2008, 130, 788.[10] X. van der Vlugt, J. Ivar, J. N. H. Reek, Angew. Chem. Int. Ed. 2009, 48,

8832; Angew. Chem. 2009, 121, 8990.[11] K. Ray, T. Petrenko, K. Wieghardt, F. Neese, Dalton Trans. 2007, 1552.[12] D. G. H. Hetterscheid, B. d. Bas, J. Mol. Catal. A 2006, 251, 291.[13] M. D. Ward, J. A. McCleverty, J. Chem. Soc. Dalton Trans. 2002, 275.[14] T. Wada, K. Tsuge, K. Tanaka, Angew. Chem. Int. Ed. 2000, 39, 1479;

Angew. Chem. 2000, 112, 1539.[15] J. T. Muckerman, D. E. Polyansky, T. Wada, K. Tanaka, E. Fujita, Inorg.

Chem. 2008, 47, 1787.[16] X. Yang, M.-H. Baik, J. Am. Chem. Soc. 2006, 128, 7476.[17] Y. Dede, X. Zhang, M. Schlangen, H. Schwarz, M.-H. Baik, J. Am. Chem.

Soc. 2009, 131, 12634.[18] S. Ghosh, M.-H. Baik, Angew. Chem. Int. Ed. 2012, 51, 1221; Angew.

Chem. 2012, 124, 1247.[19] A. I. Nguyen, R. A. Zarkesh, D. C. Lacy, M. K. Thorson, A. F. Heyduk,

Chem. Sci. 2011, 2, 166.[20] J. Rittle, M. T. Green, Science 2010, 330, 933.[21] O. R. Luca, R. H. Crabtree, Chem. Soc. Rev. 2013, 42, 1440.[22] G. Proulx, R. G. Bergman, Organometallics 1996, 15, 684.[23] P. J. Chirik, Inorg. Chem. 2011, 50, 9737.[24] M. Schlangen, D. Schrçder, H. Schwarz, Angew. Chem. Int. Ed. 2007, 46,

1641; Angew. Chem. 2007, 119, 1667.[25] Jaguar 7.0; Schrçdinger, LLC, New York, NY, 2007.[26] A. D. Becke, Phys. Rev. A 1988, 38, 3098.[27] A. D. Becke, J. Chem. Phys. 1993, 98, 5648.[28] S. H. Vosko, L. Wilk, M. Nusair, Can. J. Phys. 1980, 58, 1200.[29] C. T. Lee, W. T. Yang, R. G. Parr, Phys. Rev. B 1988, 37, 785.[30] P. J. Stephens, F. J. Devlin, C. F. Chabalowski, M. J. Frisch, J. Phys. Chem.

1994, 98, 11623.[31] P. J. Hay, W. R. Wadt, J. Chem. Phys. 1985, 82, 270.[32] W. R. Wadt, P. J. Hay, J. Chem. Phys. 1985, 82, 284.[33] T. H. Dunning, Jr. , J. Chem. Phys. 1989, 90, 1007.

Chem. Eur. J. 2015, 21, 1780 – 1789 www.chemeurj.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim1788

Full Paper

Page 10: Mechanism of Redox-Active Ligand-Assisted Nitrene-Groupstorage.googleapis.com/wzukusers/user-16009293/documents/... · 2015-09-16 · &Quantum Chemistry Mechanism of Redox-Active

[34] B. Marten, K. Kim, C. Cortis, R. A. Friesner, R. B. Murphy, M. N. Ringnalda,D. Sitkoff, B. Honig, J. Phys. Chem. 1996, 100, 11775.

[35] M. Friedrichs, R. H. Zhou, S. R. Edinger, R. A. Friesner, J. Phys. Chem. B1999, 103, 3057.

[36] S. R. Edinger, C. Cortis, P. S. Shenkin, R. A. Friesner, J. Phys. Chem. B1997, 101, 1190.

[37] A. A. Rashin, B. Honig, J. Phys. Chem. 1985, 89, 5588.[38] L. Noodleman, J. Chem. Phys. 1981, 74, 5737.[39] L. Noodleman, T. Lovell, W.-G. Han, J. Li, F. Himo, Chem. Rev. 2004, 104,

459.[40] L. Noodleman, E. R. Davidson, Chem. Phys. 1986, 109, 131.[41] M.-H. Baik, R. A. Friesner, J. Phys. Chem. A 2002, 106, 7407.[42] The numbers shown in parenthesis in all the figures and schemes are

solution-phase free energies in kcal mol�1.[43] K. Fukui, Science 1982, 218, 747.[44] T. A. Albright, J. K. Burdett, M.-H. Whangbo, Orbital Interactions in

Chemistry, Wiley, New York, 1985.[45] I. Fleming, Frontier Orbitals and Organic Chemical Reactions, Wiley,

London, 1976.[46] Our calculations confirm this expectation and indicate that the occu-

pied lone-pair orbitals have energies of �7.9 and �12.2 eV.[47] M. G. Fickes, W. M. Davis, C. C. Cummins, J. Am. Chem. Soc. 1995, 117,

6384.[48] Our calculated angle of 117 is in good agreement with the experimen-

tal value of 115 found in the crystal structure of 10 (Ref. [53]).[49] See the Supporting Information for a Mulliken–Mayer bond-order analy-

sis that confirms this assignment.[50] H. Wu, M. B. Hall, J. Am. Chem. Soc. 2008, 130, 16452.[51] N. E. Travia, Z. Xu, J. M. Keith, E. A. Ison, P. E. Fanwick, M. B. Hall, A.-M. M.

Omar, Inorg. Chem. 2011, 50, 10505.[52] Y. G. Gololobov, L. F. Kasukhin, Tetrahedron 1992, 48, 1353.[53] J. E. Leffler, R. D. Temple, J. Am. Chem. Soc. 1967, 89, 5235.[54] H. Goldwhite, P. Gysegem, S. Schow, C. Swyke, J. Chem. Soc. Dalton

Trans. 1975, 16.[55] W. L. Mosby, M. L. Silva, J. Chem. Soc. 1965, 1003.[56] L. Horner, A. Gross, Justus Liebigs Ann. Chem. 1955, 591, 117.[57] M. Alajarin, C. Conesa, S. Rzepa, J. Chem. Soc. Perkin Trans. 2 1999, 1811.

[58] C. Widauer, H. Gr�tzmacher, I. Shevchenko, V. Gramlich, Eur. J. Inorg.Chem. 1999, 1659.

[59] W. Q. Tian, Y. A. Wang, J. Org. Chem. 2004, 69, 4299.[60] M.-H. Baik, T. Ziegler, C. K. Schauer, J. Am. Chem. Soc. 2000, 122, 9143.[61] D. H. Evans, K. L. OConnell, Electroanalytical Chemistry : A Series of Ad-

vances, Vol. 14, Marcel Dekker, New York, 1986.[62] J. Jacq, J. Electroanal. Chem. 1971, 29, 149.[63] G. Balducci, G. Costa, J. Electroanal. Chem. 1993, 348, 355.[64] E. Laviron, J. Electroanal. Chem. 1981, 124, 1.[65] DG(sol)* is defined as DG(SCF) + DDG(solv). This expression represents

the relative solution-phase energies for hypothetical structures. Thesestructures are not fully optimized and do not correspond to properminima on the potential energy surfaces. Hence, zero-point energiesand entropies hold no real meaning for these compounds. However,these species are exceedingly useful for the conceptual understandingof relevant chemical transformations.

[66] We utilized the electrostatic potential fit charge (CHELP) as introducedby X. Breneman, X. Wiberg, J. Comput. Chem. 1990, 11, 361.

[67] T. R. Cundari, A. Dinescu, A. B. Kazi, Inorg. Chem. 2008, 47, 10067.[68] S. Kundu, E. Miceli, E. Farquhar, F. F. Pfaff, U. Kuhlmann, P. Hildebrandt,

B. Braun, C. Greco, K. Ray, J. Am. Chem. Soc. 2012, 134, 14710.[69] L. Maestre, W. M. C. Sameera, M. M. D�az-Requejo, F. Maseras, P. J. P�rez,

J. Am. Chem. Soc. 2013, 135, 1338.[70] J. S. Johnson, R. G. Bergman, J. Am. Chem. Soc. 2001, 123, 2923.[71] A. L. Odom, Dalton Trans. 2005, 225.[72] C. M�ller, W. Saak, S. Doye, Eur. J. Org. Chem. 2008, 2731.[73] D. C. Leitch, P. R. Payne, C. R. Dunbar, L. L. Schafer, J. Am. Chem. Soc.

2009, 131, 18246.[74] P. M�ller, C. Fruit, Chem. Rev. 2003, 103, 2905.[75] L. D. Amisial, X. Dai, R. A. Kinney, A. Krishnaswamy, T. H. Warren, Inorg.

Chem. 2004, 43, 6537.[76] J. A. Halfen, Curr. Org. Chem. 2005, 9, 657.

Received: October 20, 2014

Published online on December 2, 2014

Chem. Eur. J. 2015, 21, 1780 – 1789 www.chemeurj.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim1789

Full Paper