19
S1 SUPPORTING INFORMATION Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can Explain the Hydroxyl Activating Effect and the Diastereoselectivity Glenn R. Morello a , Hongyu Zhong b , Paul J. Chirik b , and Kathrin H. Hopmann a a Hylleraas Centre for Quantum Molecular Sciences, Department of Chemistry, University of Tromsø – The Arctic University of Norway, N-9037 Tromsø, Norway. b Department of Chemistry, Princeton University, New Jersey 08544, United States. CONTENT: 1. COMPUTAIONAL DETAILS 2. COBALT DIHYDRIDE VS COBALT DIHYDROGEN INTERMEDIATE 3. COMPUTATION OF ALTERNATE MECHANISMS FOR TERPINEN-4-OL 4. COMPUTATIONAL ANALYSIS OF PRECATALYST ACTIVATION 5. COMPUTED FORMATION OF 1 R FROM 1 M 6. MECHANISTIC PATHWAYS FOR α-METHYLSTYRENE 7. MECHANISTIC ANALYISIS OF FORMATION OF DIASTEREOMERS FROM TERPINEN-4-OL 8. COMPUTED SELECTIVITIES 9. EXPERIMENTAL DETAILS 10. NMR SPECTRA Electronic Supplementary Material (ESI) for Chemical Science. This journal is © The Royal Society of Chemistry 2018

Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

  • Upload
    others

  • View
    2

  • Download
    2

Embed Size (px)

Citation preview

Page 1: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S1

SUPPORTING INFORMATION

Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can Explain the

Hydroxyl Activating Effect and the Diastereoselectivity

Glenn R. Morelloa, Hongyu Zhongb, Paul J. Chirikb, and Kathrin H. Hopmanna

aHylleraas Centre for Quantum Molecular Sciences, Department of Chemistry, University of Tromsø – The

Arctic University of Norway, N-9037 Tromsø, Norway.

bDepartment of Chemistry, Princeton University, New Jersey 08544, United States.

CONTENT:

1. COMPUTAIONAL DETAILS

2. COBALT DIHYDRIDE VS COBALT DIHYDROGEN INTERMEDIATE

3. COMPUTATION OF ALTERNATE MECHANISMS FOR TERPINEN-4-OL

4. COMPUTATIONAL ANALYSIS OF PRECATALYST ACTIVATION

5. COMPUTED FORMATION OF 1R FROM 1M

6. MECHANISTIC PATHWAYS FOR α-METHYLSTYRENE

7. MECHANISTIC ANALYISIS OF FORMATION OF DIASTEREOMERS FROM TERPINEN-4-OL

8. COMPUTED SELECTIVITIES

9. EXPERIMENTAL DETAILS

10. NMR SPECTRA

Electronic Supplementary Material (ESI) for Chemical Science.This journal is © The Royal Society of Chemistry 2018

Page 2: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S2

1. COMPUTATIONAL DETAILS

Full molecular models of the catalyst C1 and the terpinen-4-ol substrate were employed (Figure 1

in the main text). All calculations were performed with the Gaussian 09 package Rev. D.01i and

the B3LYP hybrid density functional.ii Dispersion corrections were added to all calculations

through use of the Grimme empirical dispersion correction GD3.iii To obtain a better description

of the reaction in solution, the IEFPCM solvation modeliv was used for all calculations, including

geometry optimizations. Toluene was chosen as the solvent to match experimental parameters.

All calculations were performed assuming a doublet electronic state for the cobalt complex. For

geometry optimizations, the basis set 6-311G(d,p) (BS1) was employed.v Single point electronic

energies were computed with the basis set 6-311++G(2df,2pd) (BS2).

All optimised geometries and transition states (TS) were confirmed by the presences of zero and

one imaginary frequency, respectively, in the calculated Hessian. Gibbs free energies were

determined at 298.15 K and 1 atm utilizing unscaled vibrational frequencies. The final free

energies were determined as ΔGfinal = ΔGraw – ΔEBS1 + ΔEBS2.

2. COBALT DIHYDRIDE COMPARED TO COBALT DIHYDROGEN INTERMEDIATE

Relative Energy = 0.0 +2.6

Figure S1. Geometric and energetic differences between the Co0 species, 2R, (left) and the CoII version of

the same intermediate (right), where the H2 atom has been split into a dihydride. Hydrogen atoms of

(dppe) ligand removed for clarity. Bond lengths reported in Å, energies given in kcal/mol.

Page 3: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S3

3. ALTERNATE MECHANISMS

3.1. ALTERNATE REDOX MECHANISM 1: HYDRIDE TRANSFER TO THE NON-METHYLATED CARBON

Rate-determining step

OH

OH

H2 attack Hydride transfer

Proton transfer

Co

RO

1R

0

HH

TS1-2R

Co

RO

2R HH

Co

RO0

HH

Co

RO

CoII

H

RO

3R_24R_2Co0

ROTS3-4R_2

Co

H

RO

TS2-3R_222.8

24.4

Selectivity-determining step

10.49.1

H

12.3

12.6

Figure S2. Computed free energies (kcal/mol) for hydrogenation of terpinen-4-ol to the cis-product via a

mechanism in which the hydride transfer occurs to the non-substituted sp2 carbon atom of the substrate

(TS2-3R_2) resulting in cobalt coordination to the methylated carbon atom. This is followed by proton

transfer to the methylated carbon (TS4-5R_2). The overall barrier for this mechanism (24.4 kcal/mol)

exceeds the redox pathway described in Figure 3 of the main text (barrier of 23.6 kcal/mol).

3.2. ALTERNATE REDOX MECHANISM 2: BETA-HYDRIDE ELIMINATION

CoH

RO

3R

CoH

RO

TS3-4R_3

H

Co

H

0

4R_3H

CoH

RO

H

CoH

RO

5R_3

CoH

RO

TS5-6R_3

-H elimination

TS4-5R_3 RO

Hydride transfer

Co

RO

1R

0

Proton transfer

6R_3Co

H

0

RO

OH

OH

26.9

22.125.4

2R HH

Co

RO0

HH

TS1-2R

Co

RO

Selectivity-determining step

HH

Co

RO

TS2-3R23.6

18.0

23.49.1

II

II

Rate-determiningstep

12.3

12.6

Figure S3. Computed free energies (kcal/mol) for hydrogenation of terpinen-4-ol to the cis-product via a

mechanism in which hydride transfer occurs to the methyl-substituted sp2 carbon atom of the substrate

(TS2-3R) followed by β-hydride elimination from the carbon atom adjacent to the OR substituted carbon,

TS3-4R_3. The overall barrier for this mechanism (26.9 kcal/mol) exceeds the redox pathway described in

Fig. 3 of the main text (barrier 23.6 kcal/mol). For optimised coordinates, see XYZ file.

Page 4: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S4

3.3 ALTERNATE NON-REDOX MECHANISM 3: σ-BOND METATHESIS

Selectivity-determiningstep

OH

OH

H2 attack Hydride transfer

Proton transfer

Co

RO

1R

0

HH

TS1-2R

Co

RO

2R HH

Co

RO0

HH

Co

RO

CoII

H

RO

3R

TS2-3R23.6

28.5CoII

H

RO

H H

4R_4

Co

H

RO

TS3-4R_4

H HH2attack

24.5Co

H

ROH

H

TS4-5R_4

CoII

H

RO

H5R_4

Rate-determiningstep

24.6

H

12.3

12.6 6.2

Figure S4. Computed free energies (kcal/mol) for hydrogenation of terpinen-4-ol to the cis-product via a

non-redox σ-bond metathesis mechanism, in which the hydride transfer occurs to the methyl-

substituted sp2 carbon atom of the substrate (TS2-3R), followed by H2 coordination (TS3-4R_4) and proton

transfer to the substrate (TS4-5R_4). The overall barrier for this mechanism (28.5 kcal/mol) exceeds the

redox pathway described in Figure 3 of the main text (barrier of 23.6 kcal/mol). For optimised

coordinates, see XYZ file.

Page 5: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S5

4. GENERATION OF THE ACTIVE CATALYST

We predict that generation of the active species 1M occurs through σ-bond metathesis from the

precatalyst C1 (Fig. S5). The starting trimethylsilyl (TMS) complex can undergo ligand substitutions by

either splitting H2 or by activating the O-H bond of a substrate molecule, each liberating one TMS group.

H2 splitting (barrier of 29.3 kcal/mol) followed by splitting of an OH group (barrier 7.7 kcal/mol) leads to

formation of 1M, with a relative energy of -32.3 kcal/mol. Formation of the dihydride complex has a

higher barrier for the second step (12.0 kcal/mol) and results in a species of higher energy (-23.6

kcal/mol). This indicates that in presence of OH-bearing substrates, formation of 1M is preferred over

formation of the previously proposed dihydride species. Formation of a bis-alkoxide species (-28.9

kcal/mol) has a higher barrier (32.5 kcal/mol) than initial H2 splitting (29.3 kcal/mol).

Ph2P

PPh2

CoIITMS

TMS

Ph2P

PPh2

CoIIH

TMS

Ph2P

PPh2

CoIIORsub

TMS

Ph2P

PPh2

CoIIH

H

Ph2P

PPh2

CoIIH

ORsub

Ph2P

PPh2

CoII

ORsub

ORsub

+ H2 + HORsub

-13.4 -11.7

0.0

-23.6 -32.3 -28.9

+ HORsub

+ H2 + HORsub

1M

G‡= +29.3

G‡= +7.7G‡= +12.0

Ph2P

PPh2

CoII TMS

TMS

H H

sigma bondmetathesis

Ph2P

PPh2

CoIIH

TMS

ORsub

H

sigma bondmetathesis

sigma bondmetathesis

Ph2P

PPh2

CoIIH

TMS

H

H

- HTMS

- HTMS

- HTMS

- HTMS - HTMS

Ph2P

PPh2

CoII TMS

TMS

H

RsubO

G‡= +32.5sigma bondmetathesis

Co-dihydride Co-bisalkoxide

C1

Figure S5. Energetic flow chart (kcal/mol) comparing transformation of the precatalyst C1 to either a

dihydride, the mono-alkoxide 1M or a bis-alkoxide.

Page 6: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S6

5. REDUCTIVE ELIMINATION OF ALCOHOL SUBSTRATE TO FORM COBALT(0) SPECIES 1R

Co

O

1MH

Co

HO0

1R

TS1R-1MII

Figure S6. Left: Reductive elimination of the oxygen-coordinated substrate in complex 1M to form the

Co(0) species 1R. Right: Optimised geometry of the transition state TS1R-1M, with a computed activation

energy (∆G‡) of +35.5 kcal/mol relative to 2M (the lowest point of the metallacycle mechanism, Figure 4,

main text). Hydrogen atoms of the pincer ligand removed for clarity. Bond angles reported in degrees,

bonds in Å.

Page 7: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S7

6. MECHANISTIC PATHWAYS FOR α-METHYLSTYRENE

H2 attack Hydride transfer

Proton transfer

1RTS1-2R

2R

3R4R

TS3-4R

TS2-3R

Co

H H

Co0Co0

H HCo

H H

Co

HII

CoH

Co0

Co C2C1

0Co C2C1

0

down up

H HH H

2,1up = 19.92,1down = 21.91,2up = 23.51,2down = 22.3

2,1up = 8.82,1down = 7.51,2up = 8.81,2down = 7.3

2,1up = 15.22,1down = 17.31,2up = 26.11,2down = 25.2

Figure S7. Computed free energies (kcal/mol) for hydrogenation of α-methylstyrene via 2,1- or 1,2-

insertion, with two different substrate orientations (up or down).

Page 8: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S8

7. MECHANISTIC ANALYISIS OF FORMATION OF DIASTEREOMERS FROM TERPINEN-4-OL

OH

OH

H2 attack Hydride transfer

Proton transfer

Co

RO

1R

0

HH

TS1-2R

Co

RO

2R HH

Co

RO0

HH

Co

RO

CoH

2+

RO

3R4RCo

H

0

ROTS3-4R

Co

H

RO

TS2-3R

OH

OH

23.618.9

19.126.5

6.214.1

0.0

12.3

Figure S8. Comparison of computed free energies (kcal/mol) for hydrogenation of terpinen-4-ol to the

cis- or trans-product, respectively, via a mechanism in which the hydride transfer occurs to the methyl-

substituted sp2 carbon atom of the substrate. Here black represents the cis-product, while blue

represents the trans-product. Note that this is the most favorable pathway computed for formation of

the cis-product (barrier 23.6 kcal/mol, TS2-3R). The schematic drawings reflect the cis-geometries. For

optimised coordinates, see XYZ file.

OH

OH

H2 attack Hydride transfer

Proton transfer

Co

RO

1R

0H

HTS1-2R_2

Co

RO

2R_2 HH

Co

RO0

HH

Co

RO

CoH

2+

RO

3R_24R_2Co0

ROTS3-4R_2

Co

H

RO

TS2-3R_222.820.8

24.429.3

OH

OH

H

Figure S9. Comparison of computed free energies (kcal/mol) for hydrogenation of terpinen-4-ol to the

cis- or trans-product, respectively, via a mechanism in which the hydride transfer occurs to the non-

substituted sp2 carbon atom of the substrate (TS2-3R_2). This is followed by proton transfer to the

methylated carbon (TS4-5R_2). Here black represents the cis-product, while blue represents the trans

product. The schematic drawings reflect the cis-geometries. For optimised coordinates, see XYZ file.

Page 9: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S9

CoH

RO

3R

CoH

RO

TS3-4R_3

H

Co

H

0

4R_3H

CoH

RO

H

CoH

RO

5R_3

CoH

RO

TS5-6R_3

-H elimination

TS4-5R_3 RO

Hydride transfer

Co

RO

1R

0

Proton transfer

6R_3Co

H

0

RO

OH

OH

26.9

22.125.4

2R HH

Co

RO0

HH

TS1-2R

Co

RO

24.2

19.4

21.6

OH

OH

HH

Co

RO

TS2-3R23.618.9

18.012.5

23.411.7

II

II

Figure S10. Comparison of computed free energies (kcal/mol) for hydrogenation of terpinen-4-ol to the

cis- or trans-product, respectively, via a mechanism in which the hydride transfer occurs to the methyl-

substituted sp2 carbon atom of the substrate (TS2-3R) followed by β-H elimination from the carbon atom

adjacent to the OR substituted carbon, TS3-4R_3. Black represents the cis-product, blue the trans-product.

Note that this is the most favorable pathway computed for formation of the trans-product (barrier 24.2

kcal/mol relative to 1R). The drawings reflect the cis-geometries. For optimised coordinates, see XYZ file.

OH

OH

H2 attack Hydride transfer

Proton transfer

Co

RO

1R

0

HH

TS1-2R

Co

RO

2R HH

Co

RO0

HH

Co

RO

CoII

H

RO

3R

TS2-3R

OH

OH

23.618.9

28.536.0

CoII

H

RO

H H

4R_4

Co

H

RO

TS3-4R_4

H H

H2attack

24.5Co

H

ROH

H

TS4-5R_4

CoII

H

RO

H

H

5R_4

29.5

24.627.3

12.6

12.3 6.214.1

-5.6-7.9

Figure S11. Comparison of computed free energies (kcal/mol) for hydrogenation of terpinen-4-ol to the

cis- or trans-product, respectively, via a non-redox σ-bond metathesis mechanism in which the hydride

transfer occurs to the methyl-substituted sp2 carbon atom of the substrate (TS2-3R), followed by H2

coordination (TS3-4R_4) and proton transfer to the substrate (TS4-5R_4). Black represents the cis-product,

blue represents the trans-product. The drawings reflect the cis-geometries. For optimised coordinates,

see XYZ file.

Page 10: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S10

8. COMPUTED SELECTIVITIES

The selectivities were computed on basis of barrier differences between different pathways (ΔΔG≠)

employing equation 1: [A]/[B] = exp(ΔΔG‡B-A/RT)

Table S1. Substrate selectivity for terpinen-4-ol versus its methoxy derivative

ΔG‡ for formation of cis-

product from terpinen-4-ol

via metallacycle pathway

(Fig.4 main text)

ΔG‡ for formation of cis-

product from OMe-

derivative of terpinen-4-ol

via redox pathway (Fig.4

main text)

ΔΔG‡

[ΔG‡OMe-

ΔG‡OH]

Product

ratio

(eq 1)

Formation of experimen-

tally observed cis-

product from terpinen-4-

ol or OMe derivative

21.3 kcal/mol 23.3 kcal/mol 2 kcal/mol97(OH):

3(OMe)

Table S2. Diastereoselectivity for redox pathway

ΔG‡ for formation of cis-

product from terpinen-4-ol

via redox pathway (Fig.4

main text)

ΔG‡ for formation of trans-

product from terpinen-4-ol

via redox β-H pathway (Fig.

S10)

ΔΔG‡

[ΔG‡trans-

ΔG‡cis]

Product

ratio

(eq 1)

Formation of cis- or trans-

product from terpinen-4-ol

23.6 kcal/mol

(relative to 1R)

24.2 kcal/mol

(relative to 1R)0.6 kcal/mol

75(cis):

25(trans)

Table S3. Diastereoselectivity for metallacycle pathway

ΔG‡ for formation of cis-

product from terpinen-4-ol

via metallacycle pathway

(Fig.4 main text)

ΔG‡ for formation of

trans-product from

terpinen-4-ol via

metallacycle pathway

ΔΔG‡

[ΔG‡trans-

ΔG‡cis]

Product

ratio

(eq 1)

Formation of cis-product

or trans product from

terpinen-4-ol

21.3 kcal/mol 49.6 kcal/mol28.3

kcal/mol

100(cis):

0(trans)

Page 11: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S11

9. EXPERIMENTAL DETAILS

General Information: All air- and moisture-sensitive manipulations were carried out using standard high

vacuum line or in an M. Braun glovebox under N2 atmosphere. The M. Braun glovebox was equipped

with a cold well designed for freezing samples at liquid nitrogen temperature. Solvents for air- and

moisture-sensitive manipulations were dried and deoxygenated using literature procedures.vi Deuterated

solvents for NMR spectroscopy were distilled from sodium metal under an atmosphere of argon prior to

use. Anhydrous cobalt dichloride was purchased from Sigma-Aldrich and used as received. 1,2-

Bis(diphenylphosphino)ethane ligand was purchased from Strem Chemicals and used as received.

Terpinen-4-ol was purchased from Sigma-Aldrich and purified by dynamic vacuum distillation followed

by passage through a pad of calcium hydride and celite and storage under an inert atmosphere under

room temperature.

1H-NMR and 13C-NMR were recorded on a Bruker NanoBay-300(300 MHz 1H), or Bruker Avance-

500(500 MHz 1H and 125 MHz 13C). 1H chemical shifts are reported relative to SiMe4 using 1H chemical

shifts of the residual protio solvent as a secondary standard. 13C chemical shifts are reported relative to

SiMe4 using chemical shifts of the solvent as a secondary standard.

Gas chromatography for the alkane products was performed on a Shimadzu GC-2010 gas

chromatograph. GC analyses were performed using a Restek 15 m x 0.25 mm RTX-5 5% diphenyl/95%

dimethyl polysiloxane column with a film thickness of 0.25 μm.

Single crystals suitable for X-ray diffraction were coated with polyisobutylene oil in a glovebox,

transferred to a nylon loop and then quickly transferred to the goniometer head of a Bruker D8 APEX3

Venture diffractometer equipped with a molybdenum X-ray tube (λ = 0.71073 Å) and a Cu X-ray tube (λ =

1.54178 Å). Preliminary data revealed the crystal system. The data collection strategy was optimised for

completeness and redundancy using the Bruker COSMO software suite. The space group was identified,

and the data were processed using the Bruker SAINT+ program and corrected for absorption using

SADABS. The structures were solved using direct methods (SHELXS) completed by subsequent Fourier

synthesis and refined by full-matrix least-squares procedures.

Experimental Procedures: The catalyst dppeCo(CH2SiMe)3 (C1) was synthesised using a previously

reported method.vii Catalytic hydrogenation reactions were carried out in thick-walled glass vessels

under 1 or 4 atm H2 using the previously reported procedures.vii

Synthesis of deuterium-labeled terpinen-4-ol: In a nitrogen-filled glovebox, a 20 mL

scintillation vial was charged with 1.00 g (6.48 mmol) of terpinen-4-ol, 10 mL of THF and a stir

bar. The solution was cooled at 77 K. nBuLi solution (7.13 mmol, 2.8 mL, 2.5 M in hexane) was

diluted using 10 mL hexanes and cooled at 77 K. Into the thawing solution of terpinen-4-ol,

the nBuLi solution was added dropwise. The mixture was stirred at room temperature for 1 hour, and 1

mL of CD3OD (99.8% atom %D) was added dropwise to the mixture. The resulting solution was stirred at

r.t. for 15 minutes, and the volatiles were removed en vacuo. The residue was reconstituted in toluene,

filtered through celite followed by passage through a thin pad of alumina. The volatiles were removed en

vacuo to afford (0.94 g, 93% yield) of deuterium-labeled terpinen-4-ol (C10H8-OD). 1H-NMR(500 MHz,

CDCl3): δ 5.21 (m, 1H, Hb), δ 2.03 (m, 2H), δ 1.75 (m, 2H), δ 1.58 (s, 3H, Ha), δ 1.52 (m, 2H), δ 1.34 (m, 1H),

δ 1.09 (s, 0.04H, residual OH), δ 0.93 (d, 3H, Hg), δ 0.88 (d, 3H, Hg’). 96% deuterium incorporation (96:4

C10H8-OD : C10H8-OH) was identified based on 1H-NMR integration.

OD

Page 12: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S12

Labeling experiments supports 1,2-H2 addition across C=C bond without incorporation of alcohol

proton consistent with proposed mechanism: Hydrogenation of deuterium-labeled terpinen-4-ol was

carried out using standard procedure.vii

OD OD

Ph2P

PPh2

Co

SiMe3

SiMe32.5 mol%

4 atm H2

0.1 M toluene, r.t., 12 h

HH

> 99% conv. 1,2-H2 alkane

Hydrogenation of deuterium-labeled alcohol:

(96% D-incorporation)

Figure S12. Hydrogenation of deuterium-labeled terpinen-4-ol. 1H-NMR integration suggests 1,2-H2

alkane product. No observable deuterium-attached carbons were found in 13C-NMR.

Evidence for protonolysis of cobalt dialkyl by terpinen-4-ol and formation of cobalt alkoxide: In a

nitrogen-filled glovebox, a 20 mL scintillation vial was charged with 0.015 g of dppeCo(CH2SiMe3)2, 0.5

mL of benzene-d6. The resulting solution was transferred into a J-Young tube and cooled at 77 K. Into

another 20 mL scintillation vial was charged with 0.018 g of terpinen-4-ol and 0.5 mL of benzene -d6. The

resulting solution was cooled at 77 K, and upon thawing, transferred into the J-Young tube. The tube was

quickly sealed, brought outside the box, and let stand at room temperature for 12 hours. Vacuum

distillation of the volatiles of the reaction content (benzene-d6, SiMe4, unreacted terpinen-4-ol) under

high vacuum into another J-Young tube was performed, and 1H-NMR of the volatiles identifies both

SiMe4 product and unreacted excess terpinen-4-ol. The non-volatile components were brought into the

glovebox, transferred into a 20 mL-scintillation vial in THF. The volatiles were removed, affording 0.016 g

of residue. Into the residue was added 5 mL THF and 0.095 g (2.0 equiv, assuming mono-alkoxy product

based on 1H-NMR integration of volatiles) of TMSI. The mixture was stirred at r.t. for 12 h, and the

volatiles (THF, unreacted TMSI) was removed en vacuo. The residue (dppeCo(CH2SiMe3)x(I)n, TMS-

terpinen-4-ol) was dissolved in benzene-d6 into a J-Young tube, and the contents were distilled on high

vacuum line into another J-Young tube. 13C-NMR of the volatiles confirmed the identity of TMS-terpinen-

4-ol.

Ph2P

PPh2

Co

SiMe3

SiMe3 OH

5 equiv.

+vacuum distillation

benzene-d6, r.t.OH

+

1 equiv.

SiMe4Volatile components

Ph2PPPh2

Co (OR)nn = 1, 2x = 2-n

Non-volatile components

R =2 equiv. TMSI

vacuum distillation

THF, r.t.

Ph2PPPh2

Co(CH2SiMe3)xIn

OTMS

(CH2SiMe3)x

unreacted S.M.

then benzene-d6

n = 1,2x = 2-n

Non-volatile components

Volatile components

Figure S13. Reaction between dppeCo(CH2SiMe3)2 and substrate occurs at room temperature to

generate SiMe4 and a proposed cobalt alkoxide species

Page 13: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S13

Catalyst deactivation occurs from reacting with H2 in the absence of substrate: In a nitrogen-filled

glovebox, a thick-walled glass vessel was charged with 0.100 g of dppeCoCH2SiMe3, 5 mL Et2O, a stir bar

and sealed. The vessel was bought outside the box and attached to a high vacuum line. The contents of

the vessel were frozen in liquid nitrogen, and the head space of N2 was removed. Then 4 atm H2 was

added to the vessel and sealed. The mixture was stirred for 15 minutes at room temperature and the

contents were frozen in liquid nitrogen again. The headspace of H2 was removed and the vessel was

brought back into the glovebox. The mixture was concentrated en vacuo, affording a red solid in 0.086 g

yield and was sampled for 1H-NMR analysis. The solid was tested for catalytic activity, and hydrogenation

of terpinen-4-ol using the sample solid at 2.5 mol% loading afforded no desired alkane product.

Crystallization from a saturated pentane solution of the red solid at -35 °C overnight afforded crystals

suitable for x-ray diffraction.

Ph2P

PPh2

Co

SiMe3

SiMe3

4 atm H2

Et2O, r.t. 15 min

HCo(dppe)2Minor

Ph2PPPh2

Co

Ph2P

PPh2

H

dppeCo(CH2SiMe3)2

- 2 SiMe4

+

Mixture is catalytically inactive fordirected hydrogenation of 4-terpineol

Co(dppe)2Major

Ph2PPPh2

Co

Ph2P

PPh2

Figure S14. Treating dppeCo(dialkyl) with H2 in the absence of substrate leads to rapid formation of

catalytically inactive bis(ligand)-cobalt species.

The mixture of (dppe)2Co and HCo(dppe)2 was isolated and observed by 1H-NMR. (dppe)2Co was also

crystallographically characterised (Figure S15). The complex (dppe)2Co was first reported by Sacco.viii The

complex HCo(dppe)2 was first reported by Saccoix and was synthesised from addition of NaBH4 to

dppeCoBr2. The complex HCo(dppe)2 was crystallographically characterised by DuBois.x (dppe)2Co was

reported to reduce H2 at atmospheric pressure in benzene solution to give the Co(I) monohydride

HCo(dppe)2.viii

Figure S15. X-ray structure of (dppe)2Co at 30% probability ellipsoids with H atoms omitted for clarity

Page 14: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S14

10. NMR SPECTRA

0.00.51.01.52.02.53.03.54.04.55.05.56.06.57.07.5f1 (ppm)

6.00

0.04

1.08

1.75

3.00

2.07

2.01

0.85

0.87

0.90

0.92

0.94

1.09

1.34

1.52

1.58

1.70

1.75

1.80

1.97

2.03

5.21

Figure S16. 1H-NMR spectrum of deuterium-labeled terpinen-4-ol.

1H-NMR(500 MHz, CDCl3): δ 5.21 (m, 1H, Hb), δ 2.03 (m, 2H), δ 1.75 (m, 2H), δ 1.58 (s, 3H, Ha),

OD

a

bcd

e

fg

δ 1.52 (m, 2H), δ 1.34 (m, 1H), δ 1.09 (s, 0.04H, residual OH), δ 0.93 (d, 3H, Hg), δ 0.88 (d, 3H, Hg’). 96:4

C10H8-OD : C10H8-OH based on 1H-NMR integration.

05101520253035404550556065707580859095100105110115120125130135140145f1 (ppm)

17.0

717

.08

23.4

727

.37

31.2

735

.32

37.1

2

71.2

1

119.

10

133.

54

Figure S17. 13C-NMR spectrum of deuterium-labeled terpinen-4-ol.

Page 15: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S15

13C-NMR(125 Hz, CDCl3): δ 133.5 (b), δ 119.1 (c), δ 71.2 (e), δ 37.1, δ 35.3, δ 31.3, δ 23.5, δ17.1 OD

a

bcdef

g

h i

(i, i’).

0.00.51.01.52.02.53.03.54.04.55.05.56.06.57.07.5f1 (ppm)

9.00

2.07

2.09

0.93

2.02

1.98

1.03

0.88

0.89

0.90

0.91

1.10

1.29

1.54

1.64

1.80

1.85

Figure S18. 1H-NMR spectrum of hydrogenation product of deuterium-labeled terpinen-4-ol.

1H-NMR (500 MHz, CDCl3): δ 1.85 (hept, 1H, He, overlapping), δ 1.80 (m, 2H), δ 1.64 (m, 2H), δ

OD

a

bcd

ef

1.54 (m, 1H, Hb), δ 1.29 (m, 2H), δ 1.10 (m ,2H), δ 0.90 (d, 3H, Ha, overlapping), δ 0.89 (d, 6H, Hf,

overlapping). Proton integration matches 1,2-H2 alkane with no deuterium incorporation.

Page 16: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S16

141618202224262830323436384042444648505254565860626466687072747678f1 (ppm)

16.2

6

20.6

8

29.8

530

.55

30.7

531

.38

34.4

3

73.3

8

Figure S19. 13C-NMR spectrum of hydrogenation product of deuterium-labeled terpinen-4-ol.

13C-NMR (500 MHz, CDCl3): δ 73.4 (Ce), δ 34.4 (Cb), (δ 31.4 grease), δ 30.8 (Cf), δ 30.6 (Cc), δ

OD

a

bcde

f

g

29.9 (Cd), δ 20.7 (Ca), δ 16.3 (Cg).

Page 17: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S17

-1.0-0.50.00.51.01.52.02.53.03.54.04.55.05.56.06.57.07.58.0f1 (ppm)

1.19

6.00

0.04

1.14

3.09

2.05

2.01

0.73

0.00

0.87

0.90

0.92

0.95

1.06

1.35

1.58

1.75

2.02

5.20

OH

SiMe4

(unreacted alcohol)

+

(protonolysis pdt)

Volatile component (benzene-d6):

Figure S20. 1H-NMR evidence for reaction between dppeCo(CH2SiMe3)2 and terpinen-4-ol and generation

of SiMe4. 49% of -CH2SiMe3 group in dppeCo(CH2SiMe3)2 reacted with terpinen-4-ol based on 1H-NMR

integration of generated SiMe4 and residual terpinen-4-ol.

Ph2P

PPh2

Co

SiMe3

SiMe3 OH

5 equiv.

+vacuum distillation

benzene-d6, r.t.OH

+

1 equiv.

SiMe4Volatile components

Ph2PPPh2

Co (OR)nn = 1, 2x = 2-n

Non-volatile components

R =2 equiv. TMSI

vacuum distillation

THF, r.t.

Ph2PPPh2

Co(CH2SiMe3)xIn

OTMS

(CH2SiMe3)x

unreacted S.M.

then benzene-d6

n = 1,2x = 2-n

Non-volatile components

Volatile components

Page 18: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S18

Figure S21. 13C-NMR evidence for formation of dppeCo(OR)n(CH2SiMe3)x and its reaction with TMSI

generating TMS-terpinen-4-ol

Ph2P

PPh2

Co

SiMe3

SiMe3 OH

5 equiv.

+vacuum distillation

benzene-d6, r.t.OH

+

1 equiv.

SiMe4Volatile components

Ph2PPPh2

Co (OR)nn = 1, 2x = 2-n

Non-volatile components

R =2 equiv. TMSI

vacuum distillation

THF, r.t.

Ph2PPPh2

Co(CH2SiMe3)xIn

OTMS

(CH2SiMe3)x

unreacted S.M.

then benzene-d6

n = 1,2x = 2-n

Non-volatile components

Volatile components

Page 19: Cobalt-Catalysed Alkene Hydrogenation: A Metallacycle Can ... · 3-4R_2 Co H RO TS 2-3R_2 22.8 24.4 Selectivity-determining step 10.4 9.1 H 12.3 12.6 FigureS2. Computed free energies

S19

Ph2P

PPh2

Co

SiMe3

SiMe3

4 atm H2

Et2O, r.t. 15 min

HCo(dppe)2Minor

Ph2PPPh2

Co

Ph2P

PPh2

H

dppeCo(CH2SiMe3)2

- 2 SiMe4

+

Mixture is catalytically inactive fordirected hydrogenation of 4-terpineol

Co(dppe)2Major

Ph2PPPh2

Co

Ph2P

PPh2

Figure S22. 1H-NMR for a mixture of (dppe)2Co and H(dppe)2Co from catalyst deactivation

References:

i Gaussian 09, Revision D.01, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian, Inc., Wallingford CT, 2009.ii (a) Becke, A. D. Phys. Rev. A 1988, 38, 3098-3100, (b) Lee, C.; Yang, W.; Parr, R.G., Phys. Rev. B 1988, 37, 785-789.iii Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. J. Chem. Phys. 2010, 132, 154104.iv a) Tomasi, J.; Mennucci, B.; Cammi, R. Chem. Rev. 2005, 105, 2999-3093, b) Tomasi, J.; Mennucci, B.; Cancès E. J. Mol. Struct.: THEOCHEM 1999, 464, 211-216, c) Cancès, E.; Mennucci, B.; Tomasi, J. J. Chem. Phys. 1997, 107, 3032-3041.v Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. J. Chem. Phys. 1980, 72, 650-654.vi Pangborn, A. B.; Giardello, M. A.; Grubbs, R. H.; Rosen, R. K.; Timmers, F. J. Organometallics, 1996, 15, 1518.vii Friedfeld, M. R.; Margulieux, G. W.; Schaefer, B. A.; Chirik, P. J. J. Am. Chem. Soc., 2014, 136, 13178-13181.viii Sacco, A; Rossi, M. Chem. Commun. (London) 1965, 23, 602-603.ix Sacco, A.; Ugo, R. J. Chem. Soc. 1964, 3274-3278.x Ciancanelli, R.; Noll, B. C.; DuBois, D. L.; DuBois, M. R. J. Am. Chem. Soc. 2002, 124, 2984.