New Insights into Mechanism of Molybdenum(VI)–Dioxo Complex

May 31, 2016 - Recently, a series of oxo/nitrido-ReV/MoVI/RuVI/MnV complexes were demonstrated to be efficient catalysts in activating silanes and cat...
2 downloads 17 Views 693KB Size
Subscriber access provided by West Virginia University | Libraries

Article

New Insights into Mechanism of Molybdenum(VI)-dioxo Complex Catalyzed Hydrosilylation of Carbonyls: An Alternative Model for Activating Si-H Bond Xiaoshuang Ning, Jiandi Wang, and Haiyan Wei J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.6b01978 • Publication Date (Web): 31 May 2016 Downloaded from http://pubs.acs.org on June 6, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry A is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

New Insights into Mechanism of Molybdenum(VI)dioxo Complex Catalyzed Hydrosilylation of Carbonyls:An Alternative Model for Activating Si−H Bond Xiaoshuang Ning, Jiandi Wang, Haiyan Wei* Jiangsu Collaborative Innovation Center of Biomedical Functional Materials, Jiangsu Provincial Key Laboratory for NSLSCS, College of Chemistry and Materials Science, Nanjing Normal University, Nanjing 210046, China * Supporting Information

ABSTRACT. Recently, a series of oxo/nitrido-ReV/MoVI/RuVI/MnV complexes have been demonstrated to be efficient catalysts in activating silanes, and catalyzing hydrosilylations of unsaturated organic substrates. In the present study, the high-valent molybdenum(VI)-dioxo complex MoO2Cl2 catalyzed hydrosilylations of carbonyls was re-investigated using DFT method. Previous experimental and theoretical investigations suggested a [2+2] addition pathway for MoO2Cl2 catalyzed hydrosilylations of ketones. In the present study, we propose an ionic outer-sphere mechanistic pathway to be the most favorable pathway. The key step in the ionic outer-sphere pathway is oxygen atom of C=O bonds nucleophilically attacking the silicon atom in an η1-silane molybdenum adduct. The Si−H bond is then cleaved heterolytically. This process features a novel SN2@Si transition state, which then generates a loosely bound ion pair: anionic molybdenum hydride paired with silylcarbenium ion ([MoO2Cl2H]− ACS Paragon Plus Environment

1

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 30

[SiR3(OCR'R'')]+) in solvent. The last step is silylcarbenium ion abstracting the hydride on molybdenum hydride to yield silyl ether. The calculated activation free energy barrier of the rate-determing step was 24.1 kcal/mol for diphenylketone (PhC=OPh) and silane of PhMe2SiH. Furthermore, the ionic outersphere pathway is calculated to be ~10.0 kcal/mol lower than the previously proposed [2+2] addition pathway for a variety of silanes and aldehyde/ketone substrates. This preference arises from stronger electrophilicity of the high-valent molybdenum(VI) metal center toward a hydride. Here, we emphasize MoO2Cl2 behaves similar to Lewis acidic trispentafluorophenyl borane B(C6F5)3 in activating Si–H bond.

Introduction Hydrosilylation has been widely used in reducing organic functional groups, carbonyl compounds, alkenes, imines, amides, nitrile, and esters, and, etc,1-16 and is a valuable synthetic technique in organic chemistry for constructing silicon-based products both in industrial and academic settings.17-20 Transition metals ranging from heavy metals Ru21-25, Ir26-29, Rh30-33, Re34-37, and Mo38-42 to first-row transition metals Ti43-48, Fe149-62, Cu63-68, Mn69-72, and Zn73-80, etc81-93 are frequently developed as powerful reducing catalysts to mediate numerous hydrosilylation reactions. Several mechanistic schemes have been proposed based on experimental and theoretical investigations.15,16,94-101 In most cases, a metal hydride intermediate is generated in the catalytic cycle and assumed to be a reactive metal center for reducing the multiple C=X (X=C, O, N, etc) bonds.102-115 For example, in Chalk−Harrod and modified Chalk-Harrod cycles (Scheme 1a), the active metal hydride was proposed to be generated via Si–H bond oxidative addition to low-valent transition-metal atoms.116-118 In the [2+2] addition mechanism (Scheme 1b), the active metal hydride was proposed to be generated via Si−H bonds 1,2 cycloaddition to M=O bond. Moreover, the [2+2] addition mechanism has recently been proposed for high-valent rhenium/molybdenum complexes bearing multiply bound oxo/nitrido ligands catalyzing the hydrosilylation reaction of aldehydes and ketones.119-124 In both mechanisms (Scheme 1a, b), the reaction proceeds via unsaturated substrates coordinating to the transition metal hydride intermediates. ACS Paragon Plus Environment

2

Page 3 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Then the double bond of unsaturated substrates migrate into the metal-hydride bond, to complete reduction of unsaturated double bond. The most recent mechanistic proposals include the Glaser−Tilley mechanism and the ionic outersphere mechanism. Glaser−Tilley (Scheme 1c) mechanism was proposed for cationic rhodium complexes catalyzing olefin hydrosilylation.125-133 It involves generation of a silylene dihydride as the key intermediate. Then the ketone coordinates to the silicon center through its oxygen atom. Finally, the hydride on the metal center transfers to the carbonyl carbon atom. The ionic outer-sphere mechanism was proposed for several cationic iridium and ruthenium complexes in catalytic carbonyls hydrosilylation.133-137 It involves ketone substrate attacking the silicon center in transition metal silane σcomplexes, resulting the Si−H bond cleavaged heterolytically. Then, the activated ketone accepts a hydride from the metal. Scheme 1. Description of (a) Chalk−Harrod mechanism and modified Chalk-Harrod mechanism, (b) [2+2] addition mechanism, (c) Glaser−Tilley mechanism (GT), and (d) the ionic outer-sphere mechanism. (a) Chalk−Harrod and modified Chalk-Harrod mechanism

(b) [2+2] addition mechanism n

M

O

OCHR'R''

O CR'R'' H O CR'R'' Mn M H OSiR3 OSiR3

R3Si H

Mn

n

Mn

O

OCHR'R''

OSiR3

SiR3

(c) Glaser −Tilley (GT) mechanism M

R2Si H2

H M H

H

R Si

O CR'R'' R

R

M

Si

H

O

R

H R

M

CR'R''

ACS Paragon Plus Environment

Si

R

O CR'R''H

3

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 30

(d) Ionic out-sphere mechanism R3Si M

R3Si H

M

H

O CR'R'' SiR3

M

H + R'

O SiR3OCHR'R'' M C R''

Due to the fact that reactive intermediates have not yet been able to be isolated or clearly detected experimentally, mechanistic investigations on these hypotheses for the transition metal complexes catalyzing reactions have been somewhat hampered. Recently, high-valent transition metals of Re(V)138145

/Mo(VI)146-154/Ru(VI)155-158/Mn(V)159 have emerged as efficient catalysts in activating silanes,

boranes and hydrogens, thus brought increasing interest in this field. Considering the reversal role of high-valent transition metal complexes catalyzing reduction reactions, in contrast to their traditional use as oxidative catalysts160, it is important to define in more detail the specific pathways for those complexes in reduction-type catalysis. Here, we present our computational results to clarify the mechanism of carbonyl hydrosilylation catalyzed by molybdenum(VI)-dioxo complex MoO2Cl2. By investigating energy profiles, it was shown that the previously proposed [2+2] addition pathway featuring Si−H bond of silane adding to one of the Mo=O bond requires high barrier of ~30.0 kcal/mol. In contrast, the ionic outer-sphere pathway is a relatively low energy process, requiring overcoming barrier of ~20.0 kcal/mol. The ionic outer-sphere pathway features the Si−H bond heterolytically cleavaged upon C=O bond of carbonyl substrates attacking at silicon atom coordinated at Mo center in an η1-H(Si) metal adduct. Therefore, in contrast to previous theoretical studies of Strassner

123

and

Calhorda124, our calculational results indicate that the ionic outer-sphere catalytic cycle is the most favored pathway for MoO2Cl2 catalyzed carbonyl hydrosilylations. To the best of our knowledge, ionic outer-sphere pathways are rarely reported.161-164 Most noticeably, heterolytic activation of silanes on highly electrophilic transition metal complexes has been focused on cationic transition metal complexes. For example, Brookhart and co-workers have demonstrated a highly electrophilic iridium complex that could activate Si−H bonds and cleavages Si−H bonds in a heterolytical model.

134,165-169

In addition,

Piers and co-workers have proposed the ionic outer-sphere pathway to account for hydrosilylation of ACS Paragon Plus Environment

4

Page 5 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

ketones by strongly Lewis acidic trispentafluorophenyl borane B(C6F5)3.170-179 The ion outer-sphere catalytic cycle involves Si−H bonds cleavaged heterolytically, with transferring silyl cation to unsaturated substrates and silane hydrogen to electrophilic metal or borane center. The current findings are particularly interesting in light of high-valent oxo/nitrido-transition metal complexes which act as Lewis acids to catalyze reduction of organic substrates. Notably, the multiply-bonded oxo does not play a role in activating Si−H bonds along the ionic outer-sphere mechanistic pathway. The theoretical investigation presented in this study provides a novel understanding on high-valent transition metal complexes activating X−H (X= Si, B, P and H) bonds, to aid in development of new hydrosilylation reactions and improved catalysts. Computational Details All molecular geometries were performed using B3LYP functional.180-184 Gaussian 09 program was employed for all calculations.185 The triple-zeta basis sets plus polarization functions, 6-311g(d,p) were used for carbon, nitrogen, silicon, chloride and oxygen atoms. The LanL2DZ basis set were used for molybdenum atom,186-188 with polarization functions added (Mo, ζf = 1.043) (BS1). Frequency calculations were preformed at the same level of theory to identify transition states with only one imaginary frequency and intermediates with zero imaginary frequency. Furthermore, intrinsic reaction coordinates (IRC) were calculated on some important transition states to confirm them toward the corresponding minima. The SMD solvation model has been employed to optimize all geometry structures under solvent conditions.189 The final Gibbs energies were single-point energies calculations based on a higher-level basis set, BS2, with quintuple zeta set with extra polarization basis sets (ccQZVP) for molybdenum atom and 6-311++G(2d,p) for carbon, nitrogen, silicon, chloride and oxygen atoms.190-191 Corrections for dispersion effects using Grimme’s empirical dispersion correction: B3LYPD3 was employed to refine the final energies.192-196 Comparisons of results on DFT methods and several basis sets are presented in Supporting Information. The geometries are all displayed employing CYLview package.197

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 30

Results and Discussion The model reaction, hydrosilylation of diphenylketone with PhMe2SiH catalyzed bymolybdenumdioxo complex MoO2Cl2 was studied. In the following section, two possible reaction cycles were considered. Pathway I involves ketone substrate nucleophilically attacking on the silicon atom in a η1H(Si) metal adduct, taken as the ionic outer-sphere pathway. Pathway II is the [2+2] addition pathway, involving the Si–H bond adding to one of the Mo=O bond in a [2+2] manner, forming a hydride complex. Pathway I: Ionic outer-sphere pathway for MoO2Cl2 (1)-catalyzed hydrosilylation. The ionic outer-sphere pathway begins with silanes coordinating to the catalyst MoO2Cl2. Silane of PhMe2SiH is initially coordinated to Mo center to give η1-silane adduct 2, and the step is endergonic by 9.4 kcal/mol. The transition-metal η1-silane complexes have become the target of immense research interest for their role as important intermediates in metal-catalyzed hydrosilylation reactions.198-199 The Si−H bond distance in adduct 2 is calculated to be 1.57 Å (Fig. 1), showing an elongation of +0.09 Å compared to free silanes (ca. 1.48 Å). And the Mo−H bond in adduct 2 is 1.94 Å, largely lengthened compared with the normal Mo−H bond (ca. 1.68 Å). And angle of the Mo−H−Si is 164.3°. This geometry indicates a weak end-on η1-H(Si) mode in adduct 2. Adduct 2 subsequently undergoes nucleophilic addition of diphenylketone substrate to cleave the Si−H bond. Initially, van der Waals complex 3 is formed (Fig. 1), featuring a long Si−O3(PhC=OPh) distance (3.68 Å). From van der Waals complex 3, a transition state for the Si−H bond of silane cleavaged is determined. TS4 is calculated to possess a barrier of 19.3 kcal/mol. In TS4 (Fig. 1), the Si−H bond distance is increased to 1.76 Å, indicating the Si−H bond is partially breaking. At the same time, silane (PhMe2SiH) hydrogen moves close to the molybdenum atom, forming a Mo−H bond (1.79 Å). In addition, the silyl ion moves close to diphenylketone oxygen, with Si−O3 (Ph2C=O) distance decreased considerably to be 2.28 Å. The transition state TS4 features a SN2@Si mode, and the normal mode of imaginary frequency of this TS corresponds to the Si−H bond being cleavaged, the Mo−H bond

ACS Paragon Plus Environment

6

Page 7 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

and Si−O3 (C=O) bond being formed. The geometry around the silicon atom shows a trigonal bipyramidal structure, in which the silicon atom shares a plane with three carbon atoms of methyl and phenyl groups in equatorial positions, the silane hydrogen lies above the plane and the ketone oxygen lies below the plane, both in apical positions. Moreover, ∠Mo−H−Si=172.5º and ∠H−Si−O3=174.8º indicates that the four atoms of Mo−H···Si−O3 are roughly in a line, with. This SN2@Si transition state is analogous with hydrosilylation of ketones by Lewis acidic trispentafluorophenyl borane B(C6F5)3 according to Piers’ proposal 170-179: the Si−H bond is likewise broken with transfer of silane hydrogen to boron center of B(C6F5)3 and the silyl group transfer to a ketone substrate. TS4 evolves with formation of intermediate 4. The optimized structure of 4 (Fig. 1) exhibits a large Si···H separation (3.66 Å), which indicates that the metal-coordinated Si−H bond is cleavaged. The intermediate 4 is an ion pair comprised of an anionic molybdenum hydride [MoO2Cl2H]− and a silylcarbenium ion [PhMe2Si−OCPh2]+. Furthermore, the dissociated state of the ion pair, ([MoO2Cl2H]− + [PhMe2SiOCPh2]+), is 4.3 kcal/mol lower than ion pair 4 , and therefore the two moieties are probably spontaneously separated into two ionic species in solvent.

2

4

3

TS4

5

ACS Paragon Plus Environment

TS6

7

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 30

Figure 1. The optimized geometric structures 2→3→TS4→4→5→TS6 along the ionic outer-sphere pathway for molybdenum(VI)-dioxo complex MoO2Cl2-catalyzed hydrosilylation of diphenylketone. (The bond distances are shown in Å). To accommodate the desired hydride transfer, the silyl cation rotates in a direction perpendicular to the Mo−H bond, leading to another ion pair, 5 (Fig. 1). The Mo−H−O3−Si arrangement in intermediate 5 is bent with a dihedral angle of 98.3º, in contrast to linear geometry of 4. The two isomers of intermediates 4 and 5 are located within only 0.6 kcal/mol. Fig. 1 shows that the calculated distance between hydrogen atom of the anionic molybdenum hydride and carbon atom of the silylcarbenium ion in 5 is 3.70 Å. A transition state (TS6) for carbon atom in silylated diphenylketone abstracting the hydride on molybdenum center was located, and calculated to require a significantly low barrier of 1.2 kcal/mol (relative to intermediate 5), a barrier that is readily accessible. At this transition state (Fig. 1), the Mo−H distances is increased to 1.79 Å and the C−H distance is decreased to be 1.63 Å. TS6 leads to product-coordinated complex 6, which dissociates into a silyl ether product and catalyst MoO2Cl2. This step is favorable with ∆G (5→1+silylether) = 29.4 kcal/mol. The computed ionic outer-sphere reaction profiles for molybdenum(VI)-dioxo complex MoO2Cl2 catalyzed hydrosilylation (PhMe2SiH) of diphenylketone is summarized in Scheme 2. Note that the ratedeterming step is SN2@Si transition state (TS4), featuring the Si−H bond of silane heterolytically cleavaged. The barrier of 19.3 kcal/mol indicates the ionic outer-sphere pathway to be a facile reaction pathway for MoO2Cl2 catalyzed the carbonyl hydrosilylation.

ACS Paragon Plus Environment

8

Page 9 of 30

Cl

Ph Me

O

Mo Cl

Relative Free Energy (kcal/mol)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

O

Si O C Me Ph TS4

H

O

Cl

Ph

Ph

Mo Cl

H

O

19.3 C O Si H

3 16.9

4 14.6

2 9.4

TS6 15.2

5 14.0

O Me Si

Ph

Me

10.3

[MoO2Cl2H]+ [PhMe2Si-OCPh2]+

0.0 1 silane addition

Ph

C

Si-H bond cleavage

hydride abstraction

6 -4.7

1+ silyl ether

-15.4

Reaction coordinate

Scheme 2. The potential free-energy profile for MoO2Cl2-catalyzed hydrosilylation of diphenylketone along the ionic outer-sphere catalytic cycle. Pathway II: [2+2] addition pathway for MoO2Cl2 (1)-catalyzed hydrosilylation. The first step of the [2+2] addition pathway is Si−H bond of silane adding to one of Mo=O bond to give a five-coordinated tetragonal pyramid hydrido-oxomolybdenum complex. The [2+2] addition transition state TS7 features a Mo···O2···Si···H four-membered cyclic structure (Fig. 2). In TS7, the Si−H bond is breaking with distance of 1.99 Å, the Mo−H bond is forming with distance of 2.24 Å and Si−O2 bond is forming with distance of 1.73 Å, respectively. The barrier of TS7 is calculated to be 29.7 kcal/mol above silane and catalyst. Step 2 of the [2+2] addition pathway involves diphenylketone C=O double bond insertion into molybdenum-hydride bond in intermediate 7 (MoO(OSiPhMe2)Cl2H). As previously reported, intially, diphenylketone molecule coordinates to Mo center, forming an adduct, 8 (MoO(OSiPhMe2)Cl2H(Ph2C=O)). Subsequently, C=O double bond inserted into the Mo−H bond. The corresponding transition state is located, TS9. The migratory insertion step (TS9) is calculated to possedd a barrier of 23.6 kcal/mol above the catalyst. In the transition state TS9 (Fig. 2), diphenylketone carbon atom moves close Mo–H bond, d(C−H) = 1.51 Å and d(Mo−H)=1.81 Å, showing an η2-carbonyl mode, d(Mo−O3) = 2.25 Å andd(Mo−C) = 2.64 Å, respectively. The imaginary vibrational frequency of TS9 clearly shows the Mo−H bond being broken and C−H bond being formed. Thus, a metal alkoxide complex 9 (MoO(OSiPhMe2)Cl2(OCHPh2)) is generated. From the metal alkoxide complex 9, a retro-[2 9 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 30

+ 2] addition reaction corresponds to silyl silicon nucleophically attacking the alkoxide oxygen atom to generate the product-coordinated complex and complete the catalytic cycle. The corresponding transition state TS10 is calculated to associate with a barrier of 25.6 kcal/mol. In TS10 (Fig. 2), the Si−O3(C=O) bond is decreased to 2.12 Å, and the Si−O2 is increased to 1.98 Å.

TS7

TS9

TS10

Figure 2. The optimized geometrical structures of TS7, TS9 and TS10 for MoO2Cl2-catalyzed hydrosilylation of diphenylketone, (The bond distances are shown in Å). The computed [2+2] addition catalytic cycle for molybdenum(VI)-dioxo complex MoO2Cl2 catalyzing hydrosilylation (PhMe2SiH) of diphenylketone is summarized in Scheme 3. Note that the ratedeterming step in the [2+2] addition pathway is the [2+2] addition transition state (TS7), being significantly high of 29.7 kcal/mol. Remarkablely, we find that the transition state TS7 is 10.4 kcal/mol higher than the rate-determining step in the ionic outer-sphere pathway (Scheme 2, TS4). Therefore, in contrast to previous studies of Strassner123 and Calhorda124, we suggested that the ionic outer-sphere mechanistic pathway involving formation of η1-silane molybdenum adduct is kinetically more favorable than the previously proposed [2+2] addition mechanism for MoO2Cl2 catalyzed ketones hydrosilylations.

ACS Paragon Plus Environment

10

Page 11 of 30 [2+2] addition Relative Free Energy (kcal/mol)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry migration insertion

TS7 29.7

O

Cl Me3SiO Mo Cl H O C Ph Ph

Si H

0.0 1

C O 7

O

Me3SiO Mo H

TS10 25.6

TS9 23.6

8 0.6 O

Cl -3.9 Cl

retro-[2+2] addition

Cl

Me3SiO Mo

Cl

O C Ph

H

1+ silyl ether -15.4

9 -14.9

Ph

Reaction coordinate

Scheme 3.The potential free-energy profile for MoO2Cl2-catalyzed hydrosilylation of diphenylketone via the [2+2] addition catalytic cycle. We next explored the scope of this preference for molybdenum(V)-dioxo complex MoO2Cl2 catalyzed hydrosilylation reactions using a variety of silanes and aldehyde/ketone substrates. Our results for the transition state TS7 and the transition state TS4 are summarized in Table 1. Table 1. Activation free energies (kcal/mol) for transition states of TS4 and TS7 with various carbonyls and silanes for molybdenum(V)-dioxo complex MoO2Cl2 catalyzed the hydrosilylation reaction. Silanes

Ketones

TS4

TS7

PhMe2SiH

PhC=OPh

19.3

29.7

Ph2MeSiH

PhC=OPh

23.0

30.6

Me3SiH

PhC=OPh

23.9

30.6

SiH4

PhC=OPh

20.8

36.9

SiH4

PhCH=O

23.1

36.9

As depicted in Table 1, hydrosilylation reactions of diphenylketone with a number of silanes, PhMe2SiH, Ph2MeSiH, Me3SiH, SiH4 under the catalyst of MoO2Cl2 all prefer the ionic outer-sphere pathway. The corresponding transition states TS4 are ~7−13 kcal/mol lower than transition states TS7. Moreover, it is worth noting that with silane of SiH4, the ionic outer-sphere pathway is characterized

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 30

with a triple-well potential energy surface compared to the double-well potential energy surface. In the optimized geometry of the central intermediate, 41-SiH4, as shown in Fig. 3, the coordination around the silicon atom is pentavalent, with Si−H bond is elongated to 2.08 Å, and Mo−H bond is partially formed, d(Mo−H) = 1.71 Å. and diphenylketone oxygen atom is bound to the silyl group with distance of 1.81 Å. To proceed, the Mo−H is cleavaged, and the ion pair intermediate, 42-SiH4 is formed in which the Si··· H bond is separated with distance of 2.26 Å, forming the anionic molybdenum hydide (d(Mo−H) = 1.70 Å)), and the silylcarbenium ion (d(Si−O3)= 1.78 Å).

41-SiH4

42-SiH4

Figure 3. The optimized geometric structures of 4-SiH4 and 42-SiH4 for MoO2Cl2-catalyzed hydrosilylation (silane of SiH4) of diphenylketone. Along the ionic outer-sphere catalytic cycle, first, a free silane coordinates to to molybdenum(V)dioxo complex MoO2Cl2 forming the η1-silane molybdenum adduct 2, passing the SN2@Si transition state (TS4), to from the ion pair intermediate 4, the Si−H bond strength is weaken progressively. As indicated by Wiberg bond index (WBI) associated with Si−H interaction, WBI decreased along 0.912 (Ph2MeSiH) → 0.536 (2) → 0.514 (3) → 0.306 (TS4) → 0.001 (4), thus indicating a weaker, i.e., less stabilizing Si···H interaction along the process. The rate determining step in the ionic outer-sphere pathway is the SN2@Si transition state TS4. An analysis of frontier orbitals of van der Waals complex 3, the transition state TS4 and the ion pair intermediate 4 is presented in Fig. 4. The interacting frontier orbital of van der Waals complex 3 consists of molybdenum dxz orbital, and the Si−H bond σ bonding orbital. Upon attacking of carbonyl substrates, the interacting frontier orbital in the transition state TS4 mainly consists of molybdenum dxz orbital, σ bonding orbital of Si−H bond, together with π orbital from ACS Paragon Plus Environment

12

Page 13 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

carbonyl oxygen atom. In the ion pair 4, the molybdenum dxz orbital clearly forms a bonding orbital with the hydride. This pair of interaction orbitals presented revealed the molybdenum metal can attract electron from silane hydrogen. Subsequently, from van der Waals complex 3 to TS4, the Mo−H interaction is enhanced (1.91 Å→1.79 Å) while the Si−H interaction is considerably weakened (1.58 Å →1.76 Å). This argument is supported by our calculations of NBO charges in van der Waals complex 3, and the transition state TS4 (Fig. 4). Inspection of NBO charge indicates that with stretching of Si-H, NBO charges of silane hydrogen become less negative, along −0.243 e (2)→ −0.233 e (3)→ −0.175 e (TS4)→-0.007 e (4), respectively. Accordingly, NBO charges of silicon becomes more positive, along 1.637 e (2) → 1.661 e (3) →1.783 e (TS4) → 1.877 e (4). Accompanying with the Si−H bond distance increasing, silane hydrogen donated its electron to Mo atom progressively. Thus, NBO charges on Mo becomes less positively, along 0.917 e (2)→ 0.867 e (3)→ 0.834 e (TS4) → 0.812 e (4), as well as Wiberg bond index associated with Mo−H interaction increased along 0.296 (2) → 0.320 (3) →0.463 (TS4) → 0.823 (4), indicating a stronger, i.e., more stabilizing, Mo···H interaction in the process.

3

TS4 ACS Paragon Plus Environment

4 13

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 30

Figure 4. Relevant orbitals and NBO charges in van der Waals species 3, transition state TS4 and ion pair 4. Therefore, upon interaction with molybdenum unoccupied d orbitals, the Si−H σ bonding orbital is effectivelypolarized. Such polarization would induce the Si−H bond heterolytic cleavaged upon nucleophilically attacking of unsaturated substrates. Until now, neutral transition metal complexes mediating reduction reactions favoring the ionic outer-sphere pathway has been rarely reported. 200 In summary, the electron-accepting ability, or the Lewis acidity, of molybdenum center in the molybdenum(VI)-dioxo complex MoO2Cl2 is stronger in induing the heterolytic cleavage of Si−H bond. Conclusion We have reinvestigated the high-valent molybdenum(VI)-dioxo complex MoO2Cl2 catalyzed hydrosilylation reactions of carbonyls applying DFT method. Previously, the multiply bound oxo ligand is previously suggested to directly participate in activating the Si−H bond of silane. And the [2+2] addition pathway featuring Si−H bond adding to one of Mo=O bond is proposed to be primarily responsible for hydrosilylation of carbonyls using catalyst MoO2Cl2 by Strassner

123

and Calhorda124.

However, our in-depth studies indicate that a different mechanism − the ionic outer-sphere pathway involving carbonyl substrates back-side attacking the silicon center of η1 –silane molybdenum adduct, to heterolytically cleavage Si−H bonds is the preferable path. The ionic outer-sphere pathway is calculated to possess a relatively low barrier (the rate-determining step of 19.3 kcal/mol) compared to the [2+2] addition pathway (the rate-determining step of 29.7 kcal/mol). The ionic outer-sphere pathway for hydrosilylation reactions with molybdenum(VI)-dioxo catalysts MoO2Cl2 closely resembles the catalytic cycle proposed for Lewis acid B(C6F5)3 catalyzed hydrosilylation of carbonylc compounds, featuring silane hydrogen transfering to B(C6F5)3 and simultaneously silyl cation (R3Si+) transfering to the organic substrates. The main characters of the ionic outer-sphere mechanistic pathway are as follows: (1) the multiply bound oxo atom does not play role in activating silanes, (2) the unsaturated substrates does not need coordinates to metal centers. Our ACS Paragon Plus Environment

14

Page 15 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

calculation results demonstrates that molybdenum(VI)-dioxo complex MoO2Cl2 favoring the ionic outer-sphere mechanism is due to high electrophilicity of the molybdenum(VI) atom. The electron density of silane is thus drawn to central metal through the bridging hydrogen atom, which promotes Si−H bond of silanes heterolytically cleavaged. ASSOCIATED CONTENT Supporting Information. This material is available free of charge via the Internet at http://pubs.acs.org. Complete Ref. 185, , comparison of different methods and basis set, and the selected geometrical parameters for all calculated structures of reactants, intermediates, transition states, and products. AUTHOR INFORMATION Corresponding Author *E-mail: [email protected] Phone: +86 15301584462. ACKNOWLEDGMENT The Computer resources for theoretical calculations were provided by Jiangsu Provincial Key Laboratory for NSLSCS in Nanjing Normal University. We acknowledge the National Natural Science Foundation of China No. 21103093, and Chair Professor of Jiangsu Province to Start Funds for the financial support of this research and a Project Funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions. REFERENCES (1) Ojima, I.; Li, Z.; Zhu, J. Recent Advances in Hydrosilylation and Related Reactions, in: Z. Rappoport, Y. Apeloig (eds) The Chemistry of Organic Silicon Compounds, Wiley Chichester, 1998, vol. 2, Chapter 29. (2) Rappoport, Z.; Apeloig, Y. The Chemistry of Organic Silicon Compounds; Wiley: Chichester, U.K., 1998.

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 30

(3) Brook, M. A. Silicon in Organic, Organometallic and Polymer Chemistry, Wiley, New York, 2000. (4) Nishiyama, H.; Itoh, K. Asymmetric Hydrosilylations and Related Reactions, in: Ojima, I. (ed) Catalytic Asymmetric Synthesis, 2nd Ed, Wiley-VCH, Weinheim, 2000, Chapter 2. (5) Nishiyama, H. Hydrosilylations of Carbonyl and Imino Groups, in: Jacobsen, E. N.; Pfaltz, A.; Yamamoto, H. (eds) Comprehensive Asymmetric Catalysis, Springer, Berlin, 1999, vol. 1, Chapter 6.3. (6) Nishiyama, H. Hydrosilylations of Carbonyl and Imine Compounds, in: Beller, M.; Bolm, C. (eds) Transition Metals for Organic Synthesis: Building Blocks and Fine Chemicals, WileyVCH, Weinheim, 2004, 2, Chapter 1.4.2. (7) Marciniec, B.; Gulinski, J.; Urbaniak, W.; Kornetka, Z. W. Comprehensive Handbook on Hydrosilylation, Marciniec, B. (ed), Pergamon Press, Oxford, 1992. (8) Carpentier, J. F.; Bette, V. Chemo- and Enantioselective Hydrosilylation of Carbonyl and Imino Groups. An Emphasis on Non-Traditional Catalyst Systems. Curr. Org. Chem. 2002, 6, 913-936. (9) Riant, O.; Mostefai, N.; Courmarcel, J. Recent Advances in the Asymmetric Hydrosilylation of Ketones, Imines and Electrophilic Double Bonds.Synthesis 2004, 2943-2958. (10) Rendler, S.; Oestreich, M. Polishing a Diamond in the Rough: "Cu-H" Catalysis with Silanes. Angew. Chem., Int. Ed. 2007, 46, 498-504. (11) Diez-Gonzales, S.; Nolan, S. Transition Metal-Catalyzed Hydrosilylation of Carbonyl Compounds and Imines. A Review. Org. Prep. Proced. Int. 2007, 39, 523-559. (12) Diez-Gonzalez, S.; Nolan, S. P. Copper, Silver, and Gold Complexes in Hydrosilylation Reactions. Acc. Chem. Res. 2008, 41, 349-358. (13) Schneider, N.; Finger, M.; Haferkemper, C.; Bellemin-Laponnaz, S.; Hofmann, P.; Gade, L. H. Metal Silylenes Generated by Double Silicon-Hydrogen Activation: Key Intermediates in the RhodiumCatalyzed Hydrosilylation of Ketones. Angew. Chem., Int. Ed. 2009, 48, 1609-1613. (14) Gigler, P.; Bechlars, B.; Kuhn, F. E. Hydrosilylation with Biscarbene Rh(I) Complexes: Experimental Evidence for a Silylene-Based Mechanism. J. Am. Chem. Soc. 2011, 133, 1589-1596. (15) Beddie, C.; Hall, M. B. A Theoretical Investigation of Ruthenium-Catalyzed Alkene Hydrosilation: Evidence to Support an Exciting New Mechanistic Proposal. J. Am. Chem. Soc. 2004, 126, 1356413565. (16) Mai, V. H.; Korobkov, I. ; Nikonov, G. I. Transfer Hydrogenation of Nitriles, Olefins, and NHeterocycles Catalyzed by an N-Heterocyclic Carbene-Supported Half-Sandwich Complex of Ruthenium. Organometallics, 2016, 35, 943-949. (17) Corey, J. Y. Cheminform Abstract: Reactions of Hydrosilanes with Transition-Metal Complexes: Formation of Stable Transition-Metal Silyl Compounds. Chem. Rev. 1999, 99, 175-292.

ACS Paragon Plus Environment

16

Page 17 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(18) Corey, J. Y. Reactions of Hydrosilanes with Transition Metal Complexes and Characterization of the Products. Chem. Rev. 2011, 111, 863-1071. (19) Roy, A. K. A Review of Recent Progress in Catalyzed Homogeneous Hydrosilation (Hydrosilylation). Adv. Organomet. Chem. 2007, 55, 1-59. (20) Troegel, D.; Stohrer, J. Recent Advances and Actual Challenges in Late Transition Metal Catalyzed Hydrosilylation of Olefins From An Industrial Point of View. Coord.Chem. Rev. 2011, 255, 1440-1459. (21) Zhu, G.; Terry, M.; Zhang, X. Asymmetric Hydrosilylation of Ketones Catalyzed by Ruthenium Complexes with Chiral Tridentate Ligands. J. Organomet. Chem. 1997, 547, 97-101. (22) Glaser, P. B.; Tilley, T. D. Catalytic Hydrosilylation of Alkenes by a Ruthenium Silylene Complex. Evidence for New Hydrosilylation Mechanism. J. Am. Chem. Soc. 2003, 125, 13640-13641. (23) Igarashi, M.; Mizuno, R.; Fuchikami, T. Ruthenium Complex Catalyzed Hydrosilylation of Esters: A Facile Transformation of Esters to Alkyl Silyl Acetals and Aldehydes. Tetrahedron Lett. 2001, 42, 2149-2151. (24) Nishibayashi, Y.; Takei, I.; Uemura, S. Ruthenium-Catalyzed Asymmetric Hydrosilylation of Ketones and Imine. Organometallics 1998, 17, 3420-3422. (25) Nagashima, H.; Suzuki, A.; Iura, T.; Ryu, K.; Matsubara, K. Stoichiometric and Catalytic Activation of Si-H Bonds by a Triruthenium Carbonyl Cluste (µ3,η2:η3:η5-acenaphthylene)Ru3(CO)7: Isolation of the Oxidative Adducts, Catalytic Hydrosilylation of Aldehydes, Ketones, and Acetals, and Catalytic Polymerization of Cyclic Ethers. Organometallics 2000, 19, 3579-3590. (26) Calimano, E.; Tilley, T. D. Synthesis and Structure of PNP-Supported Iridium Silyl and Silylene Complexes: Catalytic Hydrosilation of Alkenes. J. Am. Chem. Soc. 2009, 131, 11161-11173. (27) Klei, S. R.; Tilley, T. D.; Bergman, R. G. Stoichiometric and Catalytic Behavior of Cationic Silyl and Silylene Complexes. Organometallics 2002, 21, 4648-4641. (28) Calimano, E.; Tilley, T. D. Alkene Hydrosilation by a Cationic Hydrogen-Substituted Iridium Silylene Complex. J. Am. Chem. Soc. 2008, 130, 9226-9227. (29) Niyomura, O.; Iwasawa, T.; Sawada, N.; Tokunaga, M.; Obora, Y.; Tsuji, Y. A Bowl-Shaped Phosphine as a Ligand in Rhodium-Catalyzed Hydrosilylation: Rate Enhancement by a Mono(Phosphine) Rhodium Species. Organometallics 2005, 24, 3468-3475. (30) Gade, L. H.; Cesar, V.; Bellemin-Laponnaz, S. A Modular Assembly of Chiral OxazolinylcarbeneRhodium Complexes: Efficient Phosphane-Free Catalysts for the Asymmetric Hydrosilylation of Dialkyl Ketones. Angew. Chem., Int. Ed. 2004, 43, 1014-1017. (31) Riener, K.; Hogerl, M. P.; Gigler, P.; Kuhn, F. E. Rhodium-Catalyzed Hydrosilylation of Ketones: Catalyst Development and Mechanistic Insights. ACS Catal. 2012, 2, 613-621.

ACS Paragon Plus Environment

17

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 30

(32) Tao, B.; Fu, G. C. Application of A New Family of P, N Ligands to the Highly Enantioselective Hydrosilylation of Aryl Alkyl and Dialkyl Ketones. Angew. Chem., Int. Ed. 2002, 41, 3892-3894. (33) Sawamura, M.; Ryoichi, K.; Ito, Y. Trans-Chelating Chiral Diphosphane Ligands Bearing Flexible P-Alkyl Substituents (Alkyltraps) and Their Application to the Rhodium-Catalyzed Asymmetric Hydrosilylation of Simple Ketones. Angew. Chem., Int. Ed. 1994, 33, 111-113. (34) Smeltz, J. L.; Boyle, P. D.; Ison, E. A. Role of Low-Valent Rhenium Species in Catalytic Hydrosilylation Reactions with Oxorhenium Catalysts. Organometallics 2012, 31, 5994-5997. (35) Choualeb, A.; Maccaroni, E.; Blacque, O.; Schmalle, H. W.; Berke, H. Rhenium Nitrosyl Complexes for Hydrogenations and Hydrosilylations. Organometallics 2008, 27, 3474-3481. (36) Dong, H.; Berke, H. A Convenient and Efficient Rhenium-Catalyzed Hydrosilylation of Ketones and Aldehydes. Adv. Synth. Catal. 2009, 351, 1783-1788. (37) Huang, W. J.; Berke, H. Rhenium Complexes as Highly Active Catalysts for the Hydrosilylation of Carbonyl Compounds. Chimia. 2005, 59, 113-115. (38) Khalimon, A. Y.; Ignatov, S. K.; Simionescu, R.; Kuzmina, L. G.; Howard, J. A. K.; Nikonov, G. I. An Unexpected Mechanism of Hydrosilylation by a Silyl Hydride Complex of Molybdenum. Inorg. Chem. 2011, 51, 754-756. (39) Shirobokov, O. G.; Simionescu, R.; Kuzmina, L. G.; Nikonov, G. I. The Unexpected Mechanism of Carbonyl Hydrosilylation Catalyzed by (Cp)(ArN=)Mo(H)(PMe3). Chem. Commun. 2010, 46, 78317833. (40) Khalimon, A. Y.; Shirobokov, O. G.;Peterson, E.; Simionescu, R.; Kuzmina, L. G.; Howard, J. A. K.; Nikonov, G. I. Mechanistic a Spects of Hydrosilylation Catalyzed by (ArN=)Mo(H)(Cl)(PMe3)3. Inorg. Chem. 2012, 51, 4300-4313. (41) Shirobokov, O. G.; Kuzmina, L. G.; Nikonov, G. I. Nonhydride Mechanism of Metal-Catalyzed Hydrosilylation. J. Am. Chem. Soc. 2011, 133, 6487-6489. (42) Peterson, E.; Khalimon, A. Y.; Simionescu, R.; Kuzmina, L. G.; Howard, J. A. K.; Nikonov, G. I. Diversity of Catalysis by an Imido-Hydrido Complex of Molybdenum. Mechanism of Carbonyl Hydrosilylation and Silane Alcoholysis. J. Am. Chem. Soc. 2009, 131, 908-909. (43) Carter, M. B.; Schiott, B.; Gutierrez, A.; Buchwald, S. L. Enantioselective Hydrosilylation of Ketones with a Chiral Titanocene Catalyst. J. Am. Chem. Soc. 1994, 116, 11667-11670. (44) Imma, H.; Mori, M.; Nakai, T. Asymmetric Catalytic Hydrosilylation of Ketones with Triethoxysilane Using a Chiral Binaphthol-Titanium Complex. Synlett 1996, 1229-1230. (45) Nakano, T.; Nagai, Y. Bis(η5-Cyclopentadienyl)Diphenyltitanium-Catalyzed Hydrosilylation of Ketones. Chem. Lett. 1988, 3, 481-484.

ACS Paragon Plus Environment

18

Page 19 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(46) Halterman, R.; Ramsey, T. M.; Chen, Z. Catalytic Asymmetric Hydrosilation of Aryl Alkyl Ketones with C-2-Symmetrical Chiral Metallocene Complexes. J. Org. Chem. 1994, 59, 2642-2644. (47) Bandini, M.; Bernardi, F.; Bottoni, A.; Cozzi, P. G.; Miscione, G. P.; Umani-Ronchi, A. Eur. Bis(Oxazoline)Titanium Complexes as Chiral Catalysts for Enantioselective Hydrosilylation of Ketones - A Combined Experimental and Theoretical Investigation. J. Org. Chem. 2003, 15, 2972-2984. (48) Yun, J.; Buchwald, S. L. Titanocene-Catalyzed Asymmetric Ketone Hydrosilylation: The Effect of Catalyst Activation Protocol and Additives on the Reaction Rate and Enantioselectivity. J. Am. Chem. Soc. 1999, 121, 5640-5644. (49) Shaikh, N. S.; Enthaler, S.; Junge, K.; Beller, M. Iron-Catalyzed Enantioselective Hydrosilylation of Ketones. Angew. Chem., Int. Ed. 2008, 47, 2497-2501. (50) Morris, R. H. Asymmetric Hydrogenation, Transfer Hydrogenation and Hydrosilylation of Ketones Catalyzed by Iron Complexes. Chem. Soc. Rev. 2009, 38, 2282-2291. (51) Addis, D.; Shaikh, N.; Zhou, S.; Das, S.; Junge, K.; Beller, M. Chemo- and Stereoselective IronCatalyzed Hydrosilylation of Ketones. Chem. Asian. J. 2010, 5, 1687-1691. (52) Bezier, D.; Jiang, F.; Roisnel, T.; Sortais, J.-B.; Darcel, C. Cyclopentadienyl-NHC Iron Complexes for Solvent-Free Catalytic Hydrosilylation of Aldehydes and Ketones. Eur. J. Inorg. Chem. 2012, 13331337. (53) Brunner, H.; Fisch, K. Catalytic Hydrosilylation or Hydrogenation at one Coordination Site of [Cp'Fe(Co)(X)] Fragments. Angew. Chem., Int. Ed. 1990, 29, 1131-1132. (54) Langlotz, B. K.; Wadepohl, H.; Gade, L. H. Chiral Bis(Pyridylimino)Isoindoles: A Highly Modular Class of Pincer Ligands for Enantioselective Catalysis. Angew. Chem., Int. Ed. 2008, 47, 4670-4674. (55) Nishiyama, H.; Furuta, A. An Iron-Catalysed Hydrosilylation of Ketones. Chem. Commun. 2007, 7, 760-762. (56) Tondreau, A. M.; Lobkovsky, E.; Chirik, P. Bis(Imino)Pyridine Iron Complexes for Aldehyde and Ketone Hydrosilylation. J. Org. Lett. 2008, 10, 2789-2792. (57) Buitrago, E.; Zani, L.; Adolfsson, H. Selective Hydrosilylation of Ketones Catalyzed by in SituGenerated Iron NHC Complexes. Appl. Organomet. Chem. 2011, 25, 748-752. (58) Bleith,T.; Gade, L.H. Mechanism of The Iron(II)-Catalyzed Hydrosilylation of Ketones: Activation of Iron Carboxylate Precatalysts and Reaction Pathways of the Active Catalyst. J. Am. Chem. Soc. 2016, 138, 4972-4983 (59) Bleith, T.; Wadepohl, H.; Gade, L. H. Iron Achieves Noble Metal Reactivity and Selectivity: Highly Reactive and Enantioselective Iron Complexes as Catalysts in the Hydrosilylation of Ketones. J. Am. Chem. Soc. 2015, 137, 2456-2459.

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 30

(60) Bhattacharya, P.; Krause, J. A.; Guan, H. Iron Hydride Complexes Bearing Phosphinite-Based Pincer Ligands: Synthesis, Reactivity, and Catalytic Application in Hydrosilylation Reactions. Organometallics 2011, 30, 4720-4729. (61) Buitrago, E.; Tinnis, F.; Adolfsson, H. Efficient and Selective Hydrosilylation of Carbonyl Compounds Catalyzed by Iron Acetate and N-Hydroxyethylimidazolium Salts. Adv. Synth. Catal. 2012, 354, 217-222. (62)Metsänen, T. T.; Gallego, D.; Szilvási, T.; Driess, M.; Oestreich, M. Peripheral Mechanism of a Carbonyl Hydrosilylation Catalysed by an SiNSi Iron Pincer Complex. Chem. Sci. 2015, 6, 7143-7149. (63) Brunner, H.; Miehling, W. Asymmetrische Katalysen: XXII. Enantioselektive Hydrosilylierung Von Ketonen Mit Cui-Katalysatoren. J. Organomet. Chem. 1984, 275, C17-C21. (64) Wu, J.; Ji, J. X.; Chan, A. S. C. A Remarkably Effective Copper(II)-Dipyridylphosphine Catalyst System for the Asymmetric Hydrosilylation of Ketones in Air. Proc. Nat. Acad. Sci. 2005, 102, 35703575. (65) Lipshutz, B. H.; Chrisman, W.; Noson, K. Hydrosilylation of Aldehydes and Ketones Catalyzed by [Ph3P(CuH)]6. J. Organomet. Chem. 2001, 624, 367-371. (66) Lee, D. W.; Yun, J. Copper-Catalyzed Asymmetric Hydrosilylation of Ketones Using Air- and Moisture- Stable Precatalyst Cu(OAc)2H2O. Tetrahedron Lett. 2004, 45, 5415-5417. (67) Issenhuth, J. T.; Dagorne, S.; Bellemin-Laponnaz, S. Efficient Enantioselective Hydrosilylation of Aryl Ketones Catalyzed by a Chiral BINAP-Copper(I) Catalyst-Phenyl(Methyl)Silane System. Adv. Synth. Catal. 2006, 348, 1991-1994. (68) Deutsch, C.; Krause, N.; Lipshutz, B. H. CuH-Catalyzed Reactions. Chem. Rev. 2008, 108, 29162927. (69) Son, S. U.; Paik, S. J.; Lee, I. S.; Lee, Y. A.; Chung, Y. K. Chemistry of [(1HHydronaphthalene)Mn(CO)3]: The Role of Ring-Slippage in Substitution, Catalytic Hydrosilylation, and Molecular Crystal Structure of [(η3-C10H9)Mn(CO)3P(OMe)3]. Organometallics 1999, 18, 4114-4118. (70) Cavanaugh, M. D.; Gregg, B. T.; Cutler, A. R. Manganese Carbonyl Complexes as Catalysts for the Hydrosilation of Ketones: Comparison with RhCl(PPh3)3. Organometallics 1996, 15, 2764-2769. (71) Son, S. U.; Paik, S. J.; Chung, Y. K. Hydrosilylation of Ketones Catalyzed by Tricarbonyl(Naphthalene)Manganese Cation. J. Mol. Catal. A: Chem. 2000, 151, 87-90. (72) Mao, Z. B.; Gregg, B. T.; Cutler, A. R. Catalytic Hydrosilation of Organic Esters Using Manganese Carbonyl Acetyl Complexes. J. Am. Chem. Soc. 1995, 117, 10139-10140. (73) Mimoun, H. Selective Reduction of Carbonyl Compounds by Polymethylhydrosiloxane in The Presence of Metal Hydride Catalysts. J. Org. Chem. 1999, 64, 2582-2589. (74) Bette, V.; Mortreux, A.; Savoia, D.; Carpentier, J. F. New Developments in Zinc-Catalyzed ACS Paragon Plus Environment

20

Page 21 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Asymmetric Hydrosilylation of Ketones with PMHS. Tetrahedron 2004, 60, 2837-2842. (75) Pang, S.; Peng, J.; Li, J.; Bai, Y.; Xiao, W.; Lai, G. Asymmetric Zinc-Catalyzed Hydrosilylation of Ketones and the Effect of Carboxylate on the Enantioselectivity. Chirality 2013, 25, 275-280. (76) Gajewy, J.; Gawronski, J.; Kwit, M. Mechanism and Enantioselectivity of [Zinc(diamine)(diol)]Catalyzed Asymmetric Hydrosilylation of Ketones: DFT, NMR and ECD Studies. Eur. J. Org. Chem. 2013, 2, 307-318 (77) Mimoun, H.; de Saint Laumer, J. Y.; Giannini, L.; Scopelliti, R.; Floriani, C. Enantioselective Reduction of Ketones by Polymethylhydrosiloxane in the Presence of Chiral Zinc Catalysts. J. Am. Chem. Soc. 1999, 121, 6158-6166. (78) Bette, V.; Mortreux, A.; Savoia, D.; Carpentier, J.-F. [Zinc-Diamine]-Catalyzed Hydrosilylation of Ketones in Methanol. New Developments and Mechanistic Insights. Adv. Synth. Catal. 2005, 347, 289302. (79) Junge, K.; Moller, K.; Wendt, B.; Das, S.; Gordes, D.; Thurow, K.; Beller, M. Enantioselective Zinc-Catalyzed Hydrosilylation of Ketones using Pybox or Pybim Ligands. Chem. Asian J. 2012, 7, 314320. (80) Liu, S.; Peng, J.; Yang, H.; Bai, Y.; Li, J.; Lai, G. Highly Efficient and Convenient Asymmetric Hydrosilylation of Ketones Catalyzed with Zinc Schiff Base Complexes. Tetrahedron 2012, 68, 13711375. (81) Fontaine, F.-G.; Nguyen, R.V.; Zargarian, D. Hydrosilylation of alkenes and ketones catalyzed by Nickel(II) indenyl complexes. Can. J. Chem. 2003, 81, 1299-1306. (82) Chakraborty, S.; Krause, J. A.; Guan, H. Hydrosilylation of Aldehydes and Ketones Catalyzed by Nickel PCP-Pincer Hydride Complexes. Organometallics 2009, 28, 582-586. (83) Tran, B. L.; Pink, M.; Mindiola, D. J. Catalytic Hydrosilylation of the Carbonyl Functionality via a Transient Nickel Hydride Complex. Organometallics 2009, 28, 2234-2243. (84) Cano, R.; Yus, M.; Ramón, D. J. Impregnated Platinum on Magnetite as an Efficient, Fast, and Recyclable Catalyst for the Hydrosilylation of Alkynes. ACS Catal. 2012, 2,1070-1078. (85) Scheuermann, M. L.; Semproni, S. P.; Pappas, I.; Chirik, P. J. Carbon Dioxide Hydrosilylation Promoted by Cobalt Pincer Complexes. Inorg. Chem. 2014, 53, 9463-9465. (86) Niu, Q.; Sun, H. J.; Li, X. Y.; Klein, H. F.; Flörke, U. Synthesis and Catalytic Application in Hydrosilylation of the Complex mer-Hydrido(2-mercaptobenzoyl)tris(trimethylphosphine)cobalt(III). Organometallics 2013, 32, 5235-5238. (87) Mo, Z. B.; Xiao, J.; Gao, Y. F.; Deng, L. Regio- and Stereoselective Hydrosilylation of Alkynes Catalyzed by Three-Coordinate Cobalt(I) Alkyl and Silyl Complexes. J. Am. Chem. Soc. 2014, 136, 17414-17417. ACS Paragon Plus Environment

21

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 30

(88) Murai, T.; Sakane, T.; Kato, S. Cobalt Carbonyl Catalyzed Hydrosilylation of Nitriles: A New Preparation of N,N-Disilylamines. J. Org. Chem. 1990, 55, 449-453. (89) Stahl, T.; Hrobárik, Peter.; Königs, ; Ohki, Y.; Tatsumi, K. Kemper, S.; Kaupp, M.; Klare, H. F. T.; Oestreich, M. Mechanism of the Cooperative Si–H Bond Activation at Ru–S Bonds. Chem. Sci., 2015, 6, 4324-4334. (90) Bähr, S.; Simonneau, A.; Irran, E.; Oestreich, M.; An Air-Stable Dimeric Ru−S Complex with an NHC as Ancillary Ligand for Cooperative Si−H Bond Activation. Organometallics, 2016, 35, 925-928 (91) Sattler, W.; Ruccolo, S.; Chaijan, M.R.; Allah, T. N.; Parkin, G. Hydrosilylation of Aldehydes and Ketones Catalyzed by a Terminal Zinc Hydride Complex, [κ3-Tptm]ZnH. Organometallics, 2015, 34, 4717-4731. (92) Ghosh, C. ; Mukhopadhyay, T. K.; Flores, M.; Groy, Thomas. L.; Trovitch, R. J. A Pentacoordinate Mn(II) Precatalyst That Exhibits Notable Aldehyde and Ketone Hydrosilylation Turnover Frequencies. Inorganic Chemistry 2015, 54, 10398-10406 . (93) Corre, Y.; Iali,W.; Hamdaoui, M.; Trivelli, X.; Djukic, J. P.; Agbossou-Niedercorn, F.; Michon, C. Efficient Hydrosilylation of Imines Using Catalysts Based on Iridium(III) Metallacycles. Catal. Sci. Technol. 2015, 5, 1452-1458. (94) Ojima, I.; Nihonyanagi, M.; Kogure, T.; Kumagai, M.; Horiuchi, S.; Nakatsugawa, K. Reduction of Carbonyl Compounds via Hydrosilylation: I. Hydrosilylation of Carbonyl Compounds Catalyzed by Tris(Triphenylphosphine)Chlororhodium. J. Organomet. Chem. 1975, 94, 449-461. (95) Ojima, I.; Kogure, T.; Kumagai, M.; Horiuchi, S.; Sato, T. Reduction of Carbonyl Compounds via Hydrosilylation : II. Asymmetric Reduction Of Ketones via Hydrosilylation Catalyzed by a Rhodium(I) Complex with Chiral Phosphine Ligands. J. Organomet. Chem. 1976, 122, 83-97. (96) Zheng, G. Z.; Chan, T. H. Regiocontrolled Hydrosilation of α,β-Unsaturated Carbonyl Compounds Catalyzed by Hydridotetrakis(triphenylphosphine)rhodium(I). Organometallics 1995, 14, 70-79. (97) Schneider, N.; Finger, M.; Haferkemper, C.; Bellemin-Laponnaz, S.; Hofmann, P.; Gade, L. H. Multiple Reaction Pathways in Rhodium-Catalyzed Hydrosilylations of Ketones. Chem.—Eur. J. 2009, 15, 11515-11529. (98) Beddie, C.; Hall, M. B. Do B3LYP and CCSDT Predict Different Hydrosilylation Mechanisms? Influences of Theoretical Methods and Basis Sets on Relative Energies in Ruthenium-SilyleneCatalyzed Ethylene Hydrosilylation. J. Phys. Chem. A 2006, 110, 1416-1425. (99) Tuttle, T.; Wang, D. Q.; Thiel, W. Kohler, J.; Hofmann, M.; Weis, J. Mechanism of Olefin Hydrosilylation Catalyzed by RuCl2(CO)2(PPh3)2: A DFT Study. Organometallics 2006, 25, 4504-4513.

ACS Paragon Plus Environment

22

Page 23 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(100) Tuttle, T.; Wang, D. Q.; Thiel, W.; Kohler, J.; Hofmann, M.; Weis, J. Mechanism of Olefin Hydrosilylation Catalyzed by [RuCl(NCCH3)5](+): A DFT Study. J. Organomet. Chem. 2007, 692, 2282-2290. (101) Tuttle, T.; Wang, D. Q.; Thiel, W.; Kohler, J.; Hofmann, M.; Weis, J. Ruthenium Based Catalysts for Olefin Hydrosilylation: Dichloro(P-Cymene)Ruthenium and Related Complexes. Dalton Trans. 2009, 30, 5894-5901. (102) Sark, N.; Bruno, J. W. Thermodynamic and Kinetic Studies of Hydride Transfer for a Series of Molybdenum and Tungsten Hydrides. J. Am. Chem. Soc. 1999, 121, 2174-2180. (103) Darensbourg, M. Y.; Ash, C. E. Advances in Organometallic Chemistry. Adv. Organomet. Chem. 1987, 27, 1-50. (104) Parshall, G. W.; Ittel, S. D. Homogeneous Catalysis, 2nd ed.; Wiley: New York, 1992. (105) Hembre, R. T.; McQueen, S. Hydrogenase Enzyme Reactivity Modeling with a Transition-Metal Dihydrogen Complex. J. Am. Chem. Soc. 1994, 116, 2141-2142. (106) Hembre, R. T.; McQueen, J. S. “Redox-Switch” Catalysis of C-C Bond Formation with H2: OneElectron Reduction of the Trityl Cation. Angew. Chem., Int. Ed. 1997, 36, 65-67. (107) Bunting, J. W. Merged Mechanisms for Hydride Transfer from 1,4-dihydronicotinamides. Bioorg Chem. 1991, 19, 456-491. (108) Kreevoy, M. M.; Kotchevar, A. T. Dynamics of Hydride Transfer between NAD+ Analogs. J. Am. Chem. Soc. 1990, 112, 3579-3583. (109) Kreevoy, M. M.; Ostovic, D.; Lee, I. S. H.; Binder, D. A.; King, G. W. Structure Sensitivity of The Marcus λfor Hydride Transfer between NAD+ Analogs. J. Am. Chem. Soc. 1988, 110, 524-530. (110) Lee, I. S. H.; Ostovic, D.; Kreevoy, M. Marcus Theory of Hydride Transfer from an Anionic Reduced Deazaflavin to NAD+ Analogs. J. Am. Chem. Soc. 1988, 110, 3989-3993. (111) Labinger, J. A.; Komadina, K. H. Hydridic Character of Early Transition Metal Hydride Complexes. J. Organomet. Chem. 1978, 155, C25-C28. (112) Mock, M. T.; Potter, R. G.; Camaioni, D. M.; Li, J. Thermodynamic Studies and Hydride Transfer Reactions from a Rhodium Complex to BX3 Compounds. J. Am. Chem. Soc. 2009, 131, 14454-14465. (113) Bolm, C.; Legros, J.; Le Paih, J.; Zani, L. Iron-Catalyzed Reactions in Organic Synthesis. Chem. Rev. 2004, 104, 6217-6254. (114) Junge, K.; Schroder, K.; Beller, M. Homogeneous Catalysis Using Iron Complexes: Recent Developments in Selective Reductions. Chem. Commun. 2011, 47, 4849-4859. (115) Metsänen,T. T.; Hrobárik, P.; Klare, H. F. T.; Kaupp, M.; Oestreich, M. Insight into the Mechanism of Carbonyl Hydrosilylation Catalyzed by Brookhart's Cationic Iridium(III) Pincer Complex. J. Am. Chem. Soc. 2014, 136, 6912-6915. ACS Paragon Plus Environment

23

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 30

(116) Chalk, A. J.; Harrod, J. F. Homogeneous Catalysis. II. The Mechanism of the Hydrosilation of Olefins Catalyzed by Group VIII Metal Complexes. J. Am. Chem. Soc. 1965, 87, 16-21. (117) Chalk, A. J.; Harrod, J. F.; Reactions between Dicobalt Octacarbonyl and Silicon Hydrides. J. Am. Chem. Soc. 1965, 87, 1133-1135. (118) Duckett, S. B.; Perutz, R. H. Mechanism of Homogeneous Hydrosilation of Alkenes by (η5Cyclopentadienyl)Rhodium. Organometallics 1992, 11, 90-98. (119) Sousa, S. C. A.; Cabrita, I.; Fernandes, A. C. High-Valent Oxo-Molybdenum and Oxo-Rhenium Complexes as Efficient Catalysts for X-H (X = Si, B, P and H) Bond Activation and for Organic Reductions. Chem. Soc. Rev. 2012, 41, 5641-5653. (120) Ison, E. A.; Trivedi, E. R.; Corbin, R. A.; Abu-Omar, M. M. Mechanism for Reduction Catalysis by Metal Oxo: Hydrosilation of Organic Carbonyl Groups Catalyzed by a Rhenium(V) Oxo Complex. J. Am. Chem. Soc. 2005, 127, 15374-15375. (121) Chung, L. W.; Lee, H. G.; Lin, Z.; Wu, Y. D. Computational Study on the Reaction Mechanism of Hydrosilylation of Carbonyls Catalyzed by High-Valent Rhenium(V)−Di-oxo Complexes. J. Org. Chem. 2006, 71, 6000-6009. (122) Calhorda, M. J.; Costa, P. J. Expanding the Role of Oxo-Molybdenum(VI) Catalysts: A DFT Interpretation of X–H Activation Leading to Reduction or Oxidation. Dalton Trans. 2009, 8155-8161. (123) Drees, M.; Strassner, T. Mechanism of the MoO2Cl2-catalyzed Hydrosilylation: a DFT study. Inorg. Chem. 2007, 46, 10850-10859. (124) Costa, P. J.; Romão, C. C.; Fernandes, A. C.; Roya, B.; Reis, P. M.; Calhorda, M. J. Catalyzing Aldehyde Hydrosilylation with a Molybdenum(VI) Complex: A Density Functional Theory Study. Chem.–Eur. J. 2007, 13, 3934-3941. (125) Mork, B. V.; Tilley, T. D. High Oxidation-State (Formally d0) Tungsten Silylene Complexes via Double Si-H Bond Activation. J. Am. Chem. Soc. 2001, 123, 9702-9703. (126) Feldman, J. D.; Peters, J. C.; Tilley, T. D. Activations of Silanes with [PhB(CH2PPh2)3]Ir(H)(η3C8H13). Formation of Iridium Silylene Complexes via the Extrusion of Silylenes from Secondary Silanes R2SiH2. Organometallics 2002, 21, 4065-4075. (127) Glaser, P. B.; Tilley, T. D. Catalytic Hydrosilylation of Alkenes by a Ruthenium Silylene Complex. Evidence for a New Hydrosilylation Mechanism. J. Am. Chem. Soc. 2003, 125, 13640-13641. (128) Hayes, P. G.; Beddie, C.; Hall, M. B.; Waterman, R.; Tilley, T. D. Hydrogen-Substituted Osmium Silylene Complexes: Effect of Charge Localization on Catalytic Hydrosilation. J. Am. Chem. Soc. 2006, 128, 428-429. (129) Waterman, R.; Hayes, P. G.; Tilley, T. D. Synthetic Development and Chemical Reactivity of Transition-Metal Silylene Complexes. Acc. Chem. Res. 2007, 40, 712-719. ACS Paragon Plus Environment

24

Page 25 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(130) Schneider, N.; Finger, M.; Haferkemper, C.; BelleminLaponnaz, S.; Hofmann, P.; Gade, L. H. Metal Silylenes Generated by Double Silicon-Hydrogen Activation: Key Intermediates in the RhodiumCatalyzed Hydrosilylation of Ketones. Angew. Chem., Int. Ed. 2009, 48, 1609-1613. (131) Lipke, M. C.; Tilley, T. D. Hypercoordinate Ketone Adducts of Electrophilic η3-H2SiRR' Ligands on Ruthenium as Key Intermediates for Efficient and Robust Catalytic Hydrosilation. J. Am. Chem. Soc. 2014, 136, 16387-16398. (132) Schneider, N.; Finger, M.; Haferkemper, C.; Bellemin-Laponnaz, S.; Hofmann, P.; Gade, L. H. Multiple Reaction Pathways in Rhodium-Catalyzed Hydrosilylations of Ketones. Chem.-Eur. J. 2009, 15, 11515-11529. (133) Gigler, P.; Bechlars, B.; Kuhn, F. E. Hydrosilylation with Biscarbene Rh(I) Complexes: Experimental Evidence for A Silylene-Based Mechanism. J. Am. Chem. Soc. 2011, 133, 1589-1596. (134) Park, S.; Brookhart, M. Hydrosilylation of Carbonyl-Containing Substrates Catalyzed by an Electrophilic η1-Silane Iridium(III) Complex. Organometallics 2010, 29, 6057-6064. (135) Bézier, D.; Park, S.; Brookhart, M. Selective Reduction of Carboxylic Acids to Aldehydes Catalyzed by B(C6F5)3. Org. Lett. 2013, 15, 496-9. (136) Wang W. M.; Gu, P.; Wang, Y. O.; Wei, H. Y. Theoretical Study of POCOP-Pincer Iridium(III)/Iron(II) Hydride Catalyzed Hydrosilylation of Carbonyl Compounds: Hydride Not Involved in the Iridium(III) System but Involved in the Iron(II) System. Organometallics 2014, 33, 847-857. (137) Gutsulyak, D. V.; Vyboishchikov, S. F.; Nikonov, G. I. Cationic Silane Sigma-Complexes of Ruthenium with Relevance to Catalysis. J. Am. Chem. Soc. 2010, 132, 5950-1. (138) Ison, E. A.; Trivedi, E. R.; Corbin, R. A.; Abu-Omar, M. M. Mechanism for Reduction Catalysis by Metal Oxo: Hydrosilation of Organic Carbonyl Groups Catalyzed by a Rhenium(V) Oxo Complex. J. Am. Chem. Soc. 2005, 127, 15374-5. (139) Du, G.; Abu-Omar, M. M. Catalytic Hydrogen Production from Hydrolytic Oxidation of Organosilanes. Organometallics. 2006, 25, 4920-4923. (140)

Royo,

B.;

Romão,

C.

C. Synthesis,

Bis(Chloro)Dioxomolybdenum(VI)-Chiral

Characterization

Diimine

Complexes.

and

Catalytic

J.

Mol.

Studies

of

Catal.

A:

Chem. 2005, 236, 107-112. (141) Reis, P. M.; Royo, B. Perrhenic Acid as Catalyst for Hydrosilylation of Aldehydes and Ketones and Dehydrogenative Silylation of Alcohols. Catal. Commun. 2007, 8, 1057-1059. (142) Kennedy-Smith, J. J.; Nolin, K. A.; Gunterman, H. P.; Toste, F. D. Reversing the Role of the Metal—Oxygen π-Bond. Chemoselective Catalytic Reductions with a Rhenium(V)-Dioxo Complex. J. Am. Chem. Soc. 2003, 125, 4056-4057. (143) Nolin, K. A.; Ahn, R. W.; Toste, F. D. Enantioselective Reduction of Imines Catalyzed by a ACS Paragon Plus Environment

25

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 30

Rhenium(V)-Oxo Complex. J. Am. Chem. Soc. 2005, 127, 12462-3. (144) Nolin, A. K.; Krumper, J. R.; Pluth, M. D.; Bergman, R. G.; Toste, F. D. Analysis of an Unprecedented Mechanism for the Catalytic Hydrosilylation of Carbonyl Compounds. J. Am. Chem. Soc. 2007, 129, 14684-96. (145) Fernandes, A. C.; Fernandes, J. A.; Almeida Paz, F. A.’ Romão, C. C. Activation of B-H Bonds by an Oxo-Rhenium Complex. Dalton Trans. 2009, 252, 6686-8. (146) Fernandes, A. C.; Fernandes, R.; Romão, C. C.; Royo, B. [MoO2Cl2] as Catalyst for Hydrosilylation of Aldehydes and Ketones. Chem. Commun. 2005, 36, 213-214. (147) Reis, P. M.; Romão, C. C.; Royo, B. Dioxomolybdenum(VI) Complexes as Catalysts for the Hydrosilylation of Aldehydes and Ketones. Dalton Trans. 2006, 6, 1842-6. (148) Costa, P. J.; Romão, C. C.; Fernandes, A. C.; Royo, B.; Reis, P. M.; Calhorda, M. J. Catalyzing Aldehyde Hydrosilylation with a Molybdenum(VI) Complex: A Density Functional Theory Study. Chem. Eur. J. 2007, 13, 3934-3941. (149) Fernandes, A. C.; Romão, C. C. Reduction of Sulfoxides with Boranes Catalyzed by MoO2Cl2 Tetrahedron lett. 2007, 48, 9176-9179. (150) Calhorda, M. J.; Costa, P. J. Expanding the Role of Oxo-Molybdenum(VI) Catalysts: A DFT Interpretation of X–H activation Leading to Reduction or Oxidation. Dalton Trans. 2009, 8155. (151) Noronha, R. G.; Costa, P. J.; Romão, C. C.; Calhorda, M. J.; Fernandes, A. C. MoO2Cl2 as a Novel Catalyst for C—P Bond Formation and for Hydrophosphonylation of Aldehydes. Organometallics 2009, 28, 6206-6212. (152) Fernandes, A. C.; Romão, C. C. A Novel Method for the Reduction of Imines Using the System Silane/MoO2Cl2. Tetrahedron Lett. 2005, 46, 8881-8883. (153) Fernandes, A. C.; Romão, C. C. Reduction of Amides with Silanes Catalyzed by MoO2Cl2. J. Mol. Catal. A: Chem. 2007, 272, 60-63. (154) Noronha, R. G.; Fernandes, A. C.; Romão, C. C. ChemInform Abstract: MoO2Cl2 as a Novel Catalyst for Friedel—Crafts Acylation and Sulfonylation. Tetrahedron Lett. 2009, 50, 1407-1410. (155) Truong, T. V.; Kastl, E. A.; Du, G. Cationic Nitridoruthenium(VI) Catalyzed Hydrosilylation of Ketones and Aldehydes. Tetrahedron Lett. 2011, 52, 1670-1672. (156) Man, W. L.; Lam, W. W. Y.; Yiu, S. M.; Lau, T. C.; Peng, S. M. Direct Aziridination of Alkenes by a Cationic (Salen)Ruthenium(VI) Nitrido Complex. J. Am. Chem. Soc. 2004, 126, 15336-15337. (157) Man, W. L.; Lam, W. W. Y.; Kwong, H. K.; Peng, S. M.; Wong, W. T.; Lau, T. C. Reaction of a (Salen)Ruthenium(VI) Nitrido Complex with Thiols. C-H Bond Activation by (Salen)Ruthenium(IV) Sulfilamido Species. Inorg. Chem. 2010, 49, 73-81. (158) Abbina, S.; Bian, S.; Oian, C.; Du, G. Scope and Mechanistic Studies of Catalytic Hydrosilylation ACS Paragon Plus Environment

26

Page 27 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

with a High-Valent Nitridoruthenium(VI). ACS Catal. 2013, 3, 678-684. (159) Chidara, V. K.; Du, G. An Efficient Catalyst Based on Manganese Salen for Hydrosilylation of Carbonyl Compounds. Organometallics. 2013, 32, 5034-5037. (160) Thiel, W. R. On the Way to a New Class of Catalysts -High-Valent Transition-Metal Complexes that Catalyze Reductions. Angew. Chem., Int. Ed. 2003, 42, 5390-5392. (161) Dioumaev, V. K.; Bullock, R. M. A Recyclable Catalyst that Precipitates at the End of the Reaction. Nature. 2003, 424, 530-532. (162) Voges, M. H.; Bullock, R. M. Catalytic Ionic Hydrogenations of Ketones Using Molybdenum and Tungsten Complexes. Dalton Trans. 2002, 41,759-770. (163) Bullock, R. M.; Voges, M. H. Homogeneous Catalysis with Inexpensive Metals: Ionic Hydrogenation of Ketones with Molybdenum and Tungsten Catalysts. J. Am. Chem. Soc. 2000, 122, 12594-12595. (164) Yang, J.; Brookhart, M. Reduction of Alkyl Halides by Triethylsilane Based on a Cationic Iridium Bis(phosphinite) Pincer Catalyst: Scope, Selectivity and Mechanism. Adv. Synth. Catal. 2009, 351, 175187. (165) Park, S.; Brookhart, M. An Efficient Iridium Catalyst for Reduction of Carbon Dioxide to Methane with Trialkylsilanes. J. Am. Chem. Soc. 2012, 134, 11404-11407. (166) Yang, J.; White, P. S.; Brookhart, M. A New Family f Multiferrocene Complexes with Enhanced Control of Structure and Stoichiometry via Coordination-Driven Self-Assembly and Their Electrochemistry. J. Am. Chem. Soc. 2008, 130, 839-841. (167) Park, S.; Brookhart, M. Phosphidation of Li4Ti5O12 Nanoparticles and their Electrochemical and Biocompatible Superiority for Lithium Rechargeable Batteries. Chem. Commun. 2011, 47, 1147411476. (168) Yang, J.; Brookhart, M. Iridium-Catalyzed Reduction of Alkyl Halides by Triethylsilane. J. Am. Chem. Soc. 2007, 129, 12656-12657. (169) Yang, J.; White, P. S.; Schauer, C.; Brookhart, M. Pd-Catalyzed Polymerization of Dienes that Involves Chain-Walking Isomerization of the Growing Polymer End: Synthesis of Polymers Composed of Polymethylene and Five-Membered-Ring Units. Angew. Chem., Int. Ed. 2007, 119, 6253. (170) Parks, D. J.; Piers, W. E. Tris(pentafluorophenyl)boron-Catalyzed Hydrosilation of Aromatic Aldehydes, Ketones, and Esters. J. Am. Chem. Soc. 1996, 118, 9440-9441. (171) Blackwell, J. M.; Sonmor, E. R.; Scoccitti, T.; Piers, W. E. B(C6F5)3-Catalyzed Hydrosilation of Imines via Silyliminium Intermediates. Org. Lett. 2000, 2, 3921-3923. (172) Tutusaus, O.; Chengbao, N.; Szymczak, N. K. A Transition Metal Lewis Acid/Base Triad System for Cooperative Substrate Binding. J. Am. Chem. Soc. 2013, 135, 3403-3406.

ACS Paragon Plus Environment

27

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 30

(173) Chapman, A. M.; Haddow, M. F.; Wass, D. F. Frustrated Lewis Pairs Beyond the Main Group: Synthesis, Reactivity, and Small Molecule Activation with Cationic Zirconocene-Phosphinoaryloxide Complexes. J. Am. Chem. Soc. 2011, 133, 18463-18478. (174) Geier, S. J.; Stephan, D. W. Lutidine/B(C6F5)3: At the Boundary of Classical and Frustrated Lewis Pair Reactivity. J. Am. Chem. Soc. 2009, 131, 3476-3477. (175) Parks, D. J.; Blackwell, J. M.; Piers, W. E. Studies on the Mechanism of B(C6F5)3-Catalyzed Hydrosilation of Carbonyl Functions. J. Org. Chem. 2000, 65, 3090-3098. (176) Parks, D. J.; Piers, W. E.; Parvez, M.; Atencio, R.; Zaworotko, M. J. Synthesis and Solution and Solid-State Structures of Tris(pentafluorophenyl)borane Adducts of PhC(O)X (X = H, Me, OEt, NPri2). Organometallics 1998, 17, 1369-1377. (177) Berkefeld, A.; Piers, W. E.; Parvez, M. T. Frustrated Lewis Pair/Tris(Pentafluorophenyl)BoraneCatalyzed Deoxygenative Hydrosilylation of Carbon Dioxide. J. Am. Chem. Soc. 2010, 132, 1066010661. (178) Bertini, F.; Lyaskovskyy, V.; Timmer, B. J. J.; Lutz, M.; Ehlers, A. W.; Slootweg, J. C.; Lammertsma, K. Preorganized Frustrated Lewis Pairs. J. Am. Chem. Soc. 2012, 134, 201-204. (179) Xiaoxi, Z.; Stephan, D. W. Olefin-Borane "Van Der Waals Complexes": Intermediates in Frustrated Lewis Pair Addition Reactions. J. Am. Chem. Soc. 2011, 133, 12448-12450. (180) Becke, A. D. Density‐Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648-5653. (181) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785-789. (182) Miehlich, B.; Savin, A.; Stoll, H.; Preuss, H. Results Obtained with the Correlation Energy Density Functionals of Becke and Lee, Yang and Parr. Chem. Phys. Lett. 1989, 157, 200-206. (183) Stephens, P. J.; Devlin, F. J.; Chaobalowski, C. F. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98, 11623-11627. (184) Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. Self-Consistent Molecular Orbital Methods. XX. A Basis Set for Correlated Wave Functions. J. Chem. Phys. 1980, 72, 650-654. (185) Frisch, M. J., et al. Gaussian 09, Revision A.02; Gaussian, Inc., Wallingford, CT, 2009. (186) Hay, P. J.; Wadt, W. R. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials For K to Au Including the Outermost Core Orbitals. J. Chem. Phys. 1985, 82, 299-310. (187) Wadt, W. R.; Hay, P. J. Ab Initio Effective Core Potentials For Molecular Calculations. Potentials for Main Group Elements Na to Bi. J. Chem. Phys. 1985, 82, 284-298. (188) For Mo, the triple-ζ SDD basis set with Stuttgart–Dresden ECP was also employed, the results ACS Paragon Plus Environment

28

Page 29 of 30

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

show that the dependence on the basis set is insignificant, shown in the supporting Information. (189) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Chem. Phys. B 2009, 113, 6378-6396. (190) Hehre, W. J.; Radom, L.; Schleyer, P. V. R.; Pople. J. A. Ab Initio Molecular Orbital Theory, Wiley, New York, 1986. (191) Andrae, D.; Häussermann, U.; Dolg, M.; Stoll, H.; Preuss, H. Energy-Adjusted Abinitio Pseudopotentials for the 2nd and 3rd Row Transition-Elements. Theor. Chim. Acta. 1990, 77, 123-141. (192) Grimme, S. Accurate Description Of Van Der Waals Complexes by Density Functional Theory Including Empirical Corrections. J. Comput. Chem. 2004, 25, 1463-1473. (193) Grimme, S. Semiempirical GGA-Type Density Functional Constructed with a Long-Range Dispersion Correction. J. Comput. Chem. 2006, 27, 1787-1799. (194) Grimme, S. A Theoretical Study of the Chiroptical Properties of Molecules with Isotopically Engendered Chirality. J. Chem. Phys. 2006, 124, 280-289. (195) Schwabe, T.; Grimme, S. Double-Hybrid Density Functionals with Long-Range Dispersion Corrections: Higher Accuracy and Extended Applicability. Phys. Chem. Chem. Phys. 2007, 9, 33973406. (196)

Kryspin,

I.

H.;

Grimme,

S.

The

Cr-Mn

Interaction

in

syn-Facial

[Tricarbonyl(benzyl)chromium]manganesetricarbonyl Complexes: A Non-Covalent Metal-Metal Bond. Organometallics 2009, 28, 1001-1013. (197) Legault, C. Y. CYLview, 1.0b; Université de Sherbrooke, 2009. (198) Yang, J.; White, P. S.; Schauer, C. K.; Brookhart, M. Structural and Spectroscopic Characterization of an Unprecedented Cationic Transition-Metal η¹-Silane Complex. Angew. Chem., Int. Ed. 2008, 47, 4141-4143. (199) Lin, Z. Y. Structural and Bonding Characteristics in Transition Metal–Silane Complexes. Chem. Soc. Rev. 2002, 31, 239-245. (200) Huang, L. F.; Wang, W. M.; Wei, X. Q.; Wei, H. Y. New Insights into Hydrosilylation of Unsaturated Carbon–Heteroatom (C═O, C═N) Bonds by Rhenium(V)–Dioxo Complexes. J. Chem. Phys. A 2015, 119, 3789-3799.

ACS Paragon Plus Environment

29

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 30

SYNOPSIS TOC (Word Style “SN_Synopsis_TOC”).

O Cl Cl PhMe2SiH

Mo

H

Si

O

O ∆G = 19.3 kcal/mol

C

+ C O +

the ionic outer-sphere pathway is

10.4 kcal/mol prefered than

H

O

PhMe2Si

the [2+2] addition pathway

MoO2Cl2

O Cl Cl

Mo

O

H Si

∆G = 29.7 kcal/mol

ACS Paragon Plus Environment

30