quantum chemistry in organic reactions: investigations of

140
UCLA UCLA Electronic Theses and Dissertations Title Quantum Chemistry in Organic Reactions: Investigations of Mechanism and Stereoselectivity in C-H Bond Functionalizations and Cycloaddition Reactions Permalink https://escholarship.org/uc/item/3pj1n318 Author ZOU, LUFENG Publication Date 2013-01-01 Peer reviewed|Thesis/dissertation eScholarship.org Powered by the California Digital Library University of California

Upload: vubao

Post on 14-Feb-2017

224 views

Category:

Documents


2 download

TRANSCRIPT

UCLAUCLA Electronic Theses and Dissertations

TitleQuantum Chemistry in Organic Reactions: Investigations of Mechanism and Stereoselectivity in C-H Bond Functionalizations and Cycloaddition Reactions

Permalinkhttps://escholarship.org/uc/item/3pj1n318

AuthorZOU, LUFENG

Publication Date2013-01-01 Peer reviewed|Thesis/dissertation

eScholarship.org Powered by the California Digital LibraryUniversity of California

UNIVERSITY OF CALIFORNIA

Los Angeles

Quantum Chemistry in Organic Reactions: Investigations of Mechanism and Stereoselectivity in

C-H Bond Functionalizations and Cycloaddition Reactions

A dissertation submitted in partial satisfaction of the

requirements for the degree Doctor of Philosophy

in Chemistry

by

Lufeng Zou

2013

© Copyright by

Lufeng Zou

2013

ii

ABSTRACT OF THE DISSERTATION

Quantum Chemistry in Organic Reactions: Investigations of Mechanism and Stereoselectivity in

C-H Bond Functionalizations and Cycloaddition Reactions

by

Lufeng Zou

Doctor of Philosophy in Chemistry

University of California, Los Angeles, 2013

Professor Kendall N. Houk, Chair

C−H bond functionalization is a powerful method in organic synthesis, since it avoids

prefunctionalization of substrates to introduce desired functionality. However, differentiating

between numerous C−H bonds and effecting site specific and stereoselective chemical

modifications is an ongoing challenge. In the first part of thesis, I will focus on the applications

of quantum mechanical calculations to study C−H activations by organic and metallic reagents.

Dimethyldioxirane (DMDO) is a non-metal C−H oxidation reagent. In Chapter 1, site

selectivities and reactivities in C−H activations by DMDO were studied in substituted

cyclohexanes and trans-decalins, which are models of natural steroids. It is found that the release

of 1,3−diaxial strain in the transition state contributes to the site selectivities and enhanced

iii

equatorial C−H bond reactivities for tertiary C−H bonds. In Chapter 2, C−H activation reactions

that occur with the Grubbs metathesis catalyst were found to be controlled by closed−shell

repulsions between the groups that chelate the ruthenium metal.

In the second part of the thesis, quantum mechanical computations with density functional

theory (DFT) are employed to investigate the stereoselectivities and reactivities in the reactions.

In Chapter 3, the origins of asymmetric conjugative addition of of arylboronic acid to

-substituted enone. In Chapter 4, I have demonstrated the hyperconjugative aromaticity of

5-substituted cyclopentadienes and its effect on the -facial selectivity in Diels-Alder

cycloadditions.

iv

The dissertation of Lufeng Zou is approved.

Miguel A. García-Garibay

Peter M. Felker

Albert M Thomas

Kendall N. Houk, Committee Chair

University of California, Los Angeles

2013

v

To my family and friends.

vi

TABLE OF CONTENTS

ABSTRACT OF THE DISSERTATION ....................................................................... ii

TABLE OF CONTENTS .................................................................................................vi

LIST OF SCHEMES ......................................................................................................... x

LIST OF FIGURES ..........................................................................................................xi

LIST OF TABLES ......................................................................................................... xiii

ACKNOWLEDGEMENTS ...........................................................................................xiv

VITA.................................................................................................................................. xv

Chapter 1. Enhanced Reactivity in Dioxirane C-H Oxidations via Strain Release: A

Computational and Experimental Study ......................................................................... 1

1.1 Abstract ................................................................................................................. 1

1.2 Introduction ........................................................................................................... 1

1.3 Computational Methods ........................................................................................ 4

1.4 Results and Discussions ........................................................................................ 5

1.4.1 Modeling the Dioxirane Oxidation of C-H Bonds .................................... 5

1.4.2 Oxidation of Equatorial and Axial C-H Bonds of Cyclohexane ............. 10

1.4.3 Oxidation of Methylcyclohexane ............................................................ 14

1.4.4 Influence of axial methyl substitution on reactivity ................................ 18

1.4.5 Strain Release and Enhanced Reactivity towards Oxidations in Steroidal Systems

20

1.4.6 Selective Oxidations of Sclareolide ........................................................ 22

1.5 Conclusions ......................................................................................................... 28

1.6 Experimental Section .......................................................................................... 29

vii

1.6.1 General Procedures.................................................................................. 29

1.6.2 General Procedure for Oxidation using TFDO Generated in situ: .......... 29

1.7 References ........................................................................................................... 30

Chapter 2. Formation of Chelated Z-selective Ruthenium Metathesis Catalysts:

Computational Study on C–H Activation Mechanism, Reactivity, and Selectivity ... 35

2.1 Abstract ............................................................................................................... 35

2.2 Introduction ......................................................................................................... 35

2.3 Computational Details ........................................................................................ 38

2.4 Results and Discussions ...................................................................................... 38

2.4.1 Transition state geometries for the C-H activation ................................. 38

2.4.2 The free energy profile for the formation of the chelated ruthenium catalyst 40

2.4.3 Regio- and stereoselectivity in CH activations ....................................... 41

2.4.4 Effects of chelating group and N-substituents on reactivity ................... 43

2.5 Conclusions ......................................................................................................... 48

2.6 References ........................................................................................................... 48

Chapter 3. Mechanism and Enantioselectivity in Palladium-Catalyzed Conjugate Addition

of Arylboronic Acids to -Substituted Cyclic Enones: Insights from Computation and

Experiment ....................................................................................................................... 50

3.1 Abstract ............................................................................................................... 50

3.2 Introduction ......................................................................................................... 51

3.3 Results ................................................................................................................. 53

3.3.1 Effects of Water on Catalyst Turnover.................................................... 53

3.3.2 Effects of Salt Additives on Reaction Rate ............................................. 55

3.3.3 Non-Linear Effect Correlation of Catalyst and Product Enantioenrichment 59

viii

3.3.4 Computational Investigations of the Reaction Mechanism ..................... 61

3.3.5 Experimental and Computational Investigations of the Enantioselectivities 65

3.3.6 Experimental Investigation of Arylpalladium(II) Intermediates and Formation of

the Key C–C Bond ........................................................................................... 74

3.4 Summary and Discussion .................................................................................... 77

3.5 Conclusions ......................................................................................................... 81

3.6 Materials and Methods ........................................................................................ 82

3.7 References ........................................................................................................... 84

Chapter 4. Stereoselectivities of Diels-Alder Reactions of 5-Substituted Cyclopentadienes:

Effects of Hyperconjugative Aromaticity on Reactivity and Stereoselectivity .......... 90

4.1 Abstract ............................................................................................................... 90

4.2 Introduction ......................................................................................................... 90

4.3 Background ......................................................................................................... 92

4.4 Computational Details ........................................................................................ 94

4.5 Results and Discussions ...................................................................................... 95

4.5.1 Comparison of various theoretical levels ................................................ 95

4.5.2 π-Facial stereoselectivity and geometrical distortions of reactants and transition

states, characterized by the C5-X out-of-plane angle ..................................... 100

4.5.3 Distortion/Interaction energies in π-facial stereoselectivity .................. 106

4.5.4 Reaction energies and the reaction pathways ........................................ 109

4.5.5 The origin of substituent-induced bending in the reactants – second order orbital

interaction ....................................................................................................... 114

4.5.6 Diels-Alder reaction energies and hydrogenation energies ................... 115

4.5.7 Hammett plots on the substitution effects- induction and resonance effects 116

4.5.8. Comparison with Torquoselectivity ..................................................... 118

4.6 Conclusions ....................................................................................................... 119

ix

4.7 References ......................................................................................................... 120

x

LIST OF SCHEMES

Scheme 1.1. Selective C-H oxidation. Only one of the five tertiary C-H bonds is

oxidized……………………………………………………………………………………………3

Scheme 1.2. Reaction products in a competition reaction……………………………………….20

Scheme 2.1. Generation of the Z-selective catalyst 3 through C-H activation………………….36

Scheme 2.2. Generation of the Z-selective catalyst 6 through C-H activation………………….37

Scheme 2.3. Three possible sites for C-H activation of 1a, leading to three possible chelated

catalysts 1g……………………………………………………………………………………….41

Scheme 3.1. Asymmetric conjugate addition with (S)-t-BuPyOx ligand…………….………….52

Scheme 3.2. (a) Examination of reaction mass balance. (b) Absence of water prohibits scale-up.

(c) Addition of water facilitates larger scale reactions……………………………………………54

Scheme 3.3. Enantioselectivity-determining step in asymmetric conjugate addition of arylboronic

acids to cyclic enones…………………………………………………………………………….62

Scheme 3.4. (a) Direct formation of C–C bond from arylpalladium(II) cation. (b)

Triphenylphosphine trapping experiments demonstrates configurational stability of arylpalladium

cation………………………………………………………………………………………….….77

Scheme 4.1. The π- facial stereoselectivity………………………………………………………91

Scheme 4.2. Utilization of the π-facial stereoselectivity in total synthesis by David Gin……….91

Scheme 4.3. Benchmarking reactions in anti/syn fashion………………………………………..96

Scheme 4.4. The C5-X out-of-plane angle, θ…………………………………………………..100

Scheme 4.5. Staggered conformation in the transition state for the cyclopentadiene-ethene

reaction...………………………………………………………………………………………..101

Scheme 4.6. Substituent effects on the reactant geometries leading to π- facial

stereoselectivity…………………………………………………………………………………109

Scheme 4.7. Comparison of 5-substituted cyclopentadiene Diels-Alder reactions and

hydrogenations………………………………………………………………………………….115

xi

LIST OF FIGURES

Figure 1.1. Transition state geometries for the reaction of DMDO and isobutene …………...….6

Figure 1.2. Reaction pathway in DMDO oxidations. ……………………………………...…….9

Figure 1.3. Optimized TSs for the reactions of DMDO with equatorial and axial C-H bonds.…10

Figure 1.4. Optimized TSs for the reactions of DMDO with (a) cis-1,4-dimethylcyclohexanes at

the equatorial and axial C-H bonds; (b) cis-1,2-dimethylcyclohexanes at the equatorial and axial

C-H bonds…………………………………………………………………………..……………14

Figure 1.5. Breakdown of distortion/interaction energies along the internal reaction coordinates

(IRC) for the reactions of DMDO with (a) 1-methylcyclohexanes and (b) cyclohexanes………16

Figure 1.6. Optimized reactant and transition state geometries for the reaction of DMDO with

substituted cyclohexanes…………………………………………………………………………18

Figure 1.7. Optimized reactant and transition state geometries for the reaction of TFDO with (a)

8 and (b) 9. R = -C(O)NHCH2CF3…………………………………………………….…………21

Figure 1.8. Optimized reactant and transition state geometries for the reaction of TFDO with

sclareolide 10…………………………………………………………………………………….25

Figure 1.9. Optimized reactant and transition state geometries for the reaction of tBu-TFDO with

sclareolide 10. …………………………………………………………………………..……….27

Figure 2.1. Four possible transition state geometries for C-H activation of 4…………………..39

Figure 2.2. The free energy profile of the C-H activation of 4(1a) to form the C-H activated

catalyst (1g)………………………………………………………………………………………40

Figure 2.3. Substrate and TS geometries for C-H activation on three different sites of 1c……..42

Figure 2.4. Steric effects on stereoselectivity for the cis- and trans-TS for C-H activation of

1c…………………………………………………………………………………………………43

Figure 2.5. Substrate and TS geometries for C-H activation on three different sites of 5………46

Figure 3.1. (a) deuterium incorporation using PhB(OH)2. (b) deuterium incorporation using

(PhBO)3. (c) 1H NMR data measuring deuterium incorporation by integral comparison of

-protons relative to H5, control: treatment of ketone 3 to deuterium incorporation conditions..55

Figure 3.2. Determination of linearity between catalyst ee and product ee…………………….60

Figure 3.3. Computed potential energy surface of the catalytic cycle, the alternative direct

nucleophilic addition pathway, and the isomeric carbopalladation pathway …………………...63

xii

Figure 3.4. The optimized geometries of transition states in the enantioselectivity-determining

alkene insertion step of the reaction of 3-methyl-2-cyclohexenone and phenyl boronic acid with

(S)–t-BuPyOx ligand……………………………………………………………………………..67

Figure 3.5. The optimized geometries of transition states in the enantioselectivity-determining

alkene insertion step of the reaction of (a) 3-methyl-2-cyclohexenone and phenyl boronic acid

with (S)-i-PrPyOx ligand; (b) 3-methyl--2-pentenolide and phenyl boronic acid with

(S)-t-BuPyOx ligand…………………………………………………………..…………………70

Figure 3.6. Proposed catalytic cycle for Pd/PyOx-catalyzed conjugate addition of arylboronic

acids to cyclic enones…………………………………………………………………………….80

Figure 4.1. Calculated activation free energies (ΔGX‡) in 35 reactions (kcal/mol)……………..97

Figure 4.2. Comparison of various theories against G4 in predicting π-facial stereoselectivities

ΔΔGX‡ (anti-syn, kcal/mol)………………………………………………………………………99

Figure 4.3. The Diels-Alder transition state geometries for dienophile ethene with

cyclopentadiene, anti/syn to 5-methoxycyclopentadiene, and 5-methborylcyclopentadiene…..101

Figure 4.4. The geometries of reactant diene. Cyclopentadiene, 5-methoxycyclopentadiene. and

5-methborylcyclopentadiene……………………………………………………………………103

Figure 4.5. The correlation between activation energies and distortion energies……………...107

Figure 4.6. The correlation between π- facial stereoselectivities (ΔEa‡(a) - ΔE a

‡(s)) with

differences in anti and syn distortion energies………………………………………………….107

Figure 4.7. The correlation between distortion energies and the reactant C5-X out-of-plane

angle…………………………………………………………………………………………….108

Figure 4.8. The correlation between activation energies (Y axis) and reaction energies……...112

Figure 4.9. The reaction pathways for ethene with cyclopentadiene, methoxycyclopentadiene

and methborylcyclopentadiene ………………………………………………………………...113

Figure 4.10. Orbital interactions on model systems (X= H, –OMe, –BHMe)…………………115

Figure 4.11. Comparison of 5-substituted cyclopentadiene Diels-Alder reaction energies (ΔErxn)

and hydrogenation energies (ΔEHY)…………………………………………………………….116

Figure 4.12. Hammett plot of the selectivity dependence on the inductive effect……………..117

Figure 4.13. Hammett plot of the selectivity dependence on the resonance effect…………….118

Figure 4.14. π-facial stereoselectivity Ea‡(a-s) vs. torquoselectivity Ea

‡(i-o)………………….119

xiii

LIST OF TABLES

Table 1.1. Natural orbital occupancies of frontier orbitals, and activation energies in the

transition states ……………………………………………………………………………………6

Table 1.2. Summary of activation energies and distortion/interaction energies (kcal/mol)…….12

Table 1.3 Experimental and calculated ratios for C2/C3 products from 10……………………..23

Table 1.4 Analysis of site selectivities on 10 for TFDO vs tBu-TFDO…………………………23

Table 2.1. Activation energies for C-H activation of various N-alkyl groups…………………..43

Table 2.2. Activation energies for C-H activation with different N-aryl groups………………..44

Table 2.3. Activation energies for C-H activation of 5- and 6-membered NHC rings………….46

Table 2.4. Activation energies for C-H activation with different carboxylate ligands………….47

Table 3.1. Effect of salt additives on reaction rate………………………………………………56

Table 3.2. Effect of water and NH4PF6 on reaction rate………………………………………...57

Table 3.3. Increased reaction yields with different arylboronic acid substrates under new reaction

conditions………………………………………………………………………………………...59

Table 3.4. Activation energies and enantioselectivities of alkene insertion with (S)-t-BuPyOX,

(S)-i-PrPyOX, (S)-i-BuPyOX, and (S)-PhPyOx ligands…………………………………………68

Table 3.5. Remote ligand substituent effects on activation energies and enantioselectivities of

alkene insertion…………………………………………………………………………………..69

Table 3.6. Activation energies and enantioselectivities of alkene insertion with substrates varying

at the '-position…………………………………………………………………………………71

Table 3.7. Activation energies and enantioselectivities of alkene insertion with various boronic

acids……………………………………………………………………………………………...72

Table 3.8. Enantioselectivity trends in pyridinooxazoline and related ligand frameworks…….72

Table 3.9. Arylpalladium(II) catalyzed conjugate addition……………………………………..75

Table 4.1. Distortion/Interaction energies for various substituents and out-of-plane angle θ…104

Table 4.2. Reaction energies and C5-X out-of-plane angles (θrxn) for various substituents

(kcal/mol), and the relative (anti - syn) reaction energies (ΔΔErxn) and activation energies

(ΔΔEa‡)………………………………………………………………………………………….110

xiv

ACKNOWLEDGEMENTS

Chapter 1

This chapter is a version of:

Zou, L.; Paton, R. S.; Eschenmoser, A.; Newhouse, T. R.; Baran, P. S.; Houk, K. N. J. Org.

Chem. 2013, 78, 4037.

Permission was given for use of this work in this thesis.

Kendall N. Houk was the PI for this work.

Chapter 3

This chapter is a version of:

Holder, J. C.; Zou, L.; Marziale, A. N.; Liu, P.; Lan, Y.; Gatti, M.; Kikushima, K.; Houk, K. N.;

Stoltz, B. M. J. Am. Chem. Soc. 2013, accepted.

Permission was given for use of this work in this thesis.

Kendall N. Houk and Brian M. Stoltz were the PIs for this work.

Sections 3.3.1, 3.3.2, 3.3.3 and 3.3.6 were primarily performed by Brian M. Stoltz and his group.

Sections 3.3.4 and 3.3.5 were primarily performed by Kendall N. Houk and his group. Brian M.

Stoltz and his group wrote the introduction, section 3.2, and the experimentals, section 3.6. Both

groups contribute to the writing of Sections 3.1, 3.4 and 3.5, and various discussions and

proof-reading of the whole manuscript.

Funding

This work was supported by the National Institute of General Medical Sciences, National

Institutes of Health (GM 36700), National Science Foundation (CHE 0614591). Computer time

was provided in part by the UCLA Institute for Digital Research and Education (IDRE)

Hoffman2 cluster, and the Extreme Science and Engineering Discovery Environment (XSEDE).

xv

VITA

LUFENG ZOU

EDUCATION

March 2009 M. S. Chemistry: University of California, Los Angeles

July 2007 B. S. Chemistry: University of Science and Technology of China

HONORS & AWARDS

November 2008 Excellence in First Year Academics and Research, UCLA

PRESENTATIONS

March 2012 ACS Meeting, San Diego

“-Facial Stereoselectivities with 5-Substituted Cyclopentadienes”

PUBLICATIONS

Holder, J. C.; Zou, L.; Marziale, A. N.; Liu, P.; Lan, Y.; Gatti, M.; Kikushima, K.; Houk, K. N.;

Stoltz, B. M. “Mechanism and Enantioselectivity in Palladium-Catalyzed Conjugate Addition of

Arylboronic Acids to -Substituted Cyclic Enones: Insights from Computation and Experiment”

J. Am. Chem. Soc. 2013, accepted

Zou, L.; Paton, R. S.; Eschenmoser, A.; Newhouse, T. R.; Baran, P. S.; Houk, K. N. “Enhanced

reactivity in dioxirane C-H oxidations via strain release: a computational and experimental

study” J. Org. Chem. 2013, 78: 4037.

Bigi, M. A.; Liu, P.; Zou, L.; Houk, K. N.; White, M. C. “Cafestol to Tricalysiolide B and

Oxidized Analogues: Biosynthetic and Derivatization Studies Using Non-heme Iron Catalyst

xvi

Fe(PDP)” Synlett 2012, 2768.

Lan, Y.; Zou, L.; Cao, Y.; Houk, K. N. “Computational Methods To Calculate Accurate

Activation and Reaction Energies of 1,3-Dipolar Cycloadditions of 24 1,3-Dipoles” J. Phys.

Chem. A 2011, 115, 13906.

.

1

Chapter 1. Enhanced Reactivity in Dioxirane C-H Oxidations via Strain Release: A

Computational and Experimental Study

1.1 Abstract

The site- and stereo-selectivities of C-H oxidations of substituted cyclohexanes and

trans-decalins by dimethyldioxirane (DMDO) were investigated computationally with quantum

mechanical density functional theory (DFT). The multi-configuration CASPT2 method was

employed on model systems to establish the preferred mechanism and transition state geometry.

The reaction pathway involving a rebound step is established to account for the retention of

stereochemistry. The oxidation of sclareolide with dioxirane reagents is reported, including the

oxidation by the in situ generated tBu-TFDO, a new dioxirane that better discriminates between

C-H bonds based on steric effects. The release of 1,3-diaxial strain in the transition state

contributes to the site selectivity and enhanced equatorial C-H bond reactivity for tertiary C-H

bonds, a result of the lowering of distortion energy. In addition to this strain release factor, steric

and inductive effects contribute to the rates of C-H oxidation by dioxiranes.

1.2 Introduction

The functionalization of unactivated sp3 C-H bonds is of considerable interest in

contemporary synthetic organic and organometallic chemistry.1-3

Compared to traditional

functional group manipulations and interconversions employed in synthesis, C-H bond

functionalization offers a direct route avoiding pre-functionalization of substrates. However,

differentiating between numerous C-H bonds and effecting site-specific and stereoselective

2

chemical modifications is an ongoing challenge.4 A range of reagents are known for C-H to

C-OH conversion, such as Fe, Cu, Ru, Cr and Pd based systems.5 The pioneering work of

Murray6 and Curci

7 pointed to the use of dimethyldioxirane (DMDO) and

methyl(trifluoromethyl)dioxirane (TFDO) for sp3 C-H oxidation.

8 The dioxiranes are potent

oxidants that lead to conservation of stereochemistry but no intermediate was characterized.9 The

mechanism is still under an ongoing debate and radical products are observed in some

experiments.10-12

Due to their high reactivity, dioxiranes are sometimes generated in situ during

synthesis.13

Recently, a two-phase strategy for terpene total synthesis was invented to holistically

mimic the way such molecules are constructed in nature. In the first phase, the so-called "cyclase

phase", simple terpenes at low oxidation states are stitched together as a prelude to the "oxidase

phase" wherein these molecules are systematically oxidized. In the case of the eudesmane

terpene class this was done using both innate and guided C-H functionalization logic.4 The

purpose of this strategy for retrosynthetic analysis is not to necessarily recapitulate biosynthesis

but rather to uncover new basic reactivity principles and invent new reaction methodology.

The factors that contribute to site selectivity in C-H oxidation reactions has been a question

of interest for some time and continues today.14

Terpenoids contain a host of cyclohexane

derivatives, and the rationalization of reactivities in such systems traces back to Barton’s

fundamental work in the fifties.15,16

Reactivity differences, such as observed for the oxidation of

axial and equatorial secondary alcohols with chromic acid, were explained by steric effects.

Eschenmoser interpreted the enhanced oxidation rates of axial alcohols as compared to their

equatorial epimers as resulting from the release of 1,3-diaxial strain in the transition state of the

3

oxidation of the axial epimers.17

Experimental measurements of the rates and regioselectivites of

C-H oxidation of substituted cyclohexanes by dioxiranes were performed by Curci,18,19

and a

preference for equatorial attack was observed. Furthermore, Curci8 and others

20 delineated a

variety of factors that contribute to site selectivity in oxidations by dioxiranes. Similar factors

have also been reported by M. C. White in C-H oxidation reactions of complex molecules by a

non-heme, iron-based system.21,22

This field has a storied history and has been reviewed

extensively recently.23

In the context of Eschenmoser’s hypothesis based on strain release, a recent example of

site-selective C-H oxidation of a substituted decalin from Baran’s laboratory has been

rationalized.24

Of five unactivated tertiary C-H bonds in the substrate, only equatorial C-H1 is

oxidized to an alcohol by the dioxirane reagent. (Scheme 1.1) Strain release in the transition state

was proposed to account for the preference.25

Based on computational studies from the Houk26

and Bach27

labs, a flattening at the carbon undergoing oxidation is expected in the transition

structure, which alleviates the 1,3-diaxial strain between two axial methyl groups in the

transition state for H1 abstraction and lowers the activation energy for C-H

1 oxidation.

Scheme 1.1. Selective C-H oxidation. Only one of the five tertiary C-H bonds is oxidized24

4

Given the intensive efforts underway to develop substrate controlled selective C-H

activation processes, we undertook calculations at the density functional and

multi-configurational ab initio levels of theory to gain quantitative insights into the exquisite

selectivities reported by Baran. Here we describe computational studies of the transition

structures for C-H oxidation of Baran’s eudesmane decalins and a variety of other simpler

cyclohexanes by dioxirane reagents. We also analyze various substituted cyclohexyl substrates.

For several such cases, we report transition structures and present explanations for the enhanced

reactivity of equatorial C-H bonds in these strained systems. Additionally, we include a new

synthesis of trifluoromethyl, tert-butyl ketone and its use as a reagent in C-H oxidation by in situ

formation of the corresponding dioxirane, tBu-TFDO. A quantitative understanding of the

factors contributing to selective C-H activation is provided, in the context of the

distortion/interaction theory that has been successfully applied to other types of reactions.28-31

1.3 Computational Methods

All density functional theory (DFT) calculations were performed with Gaussian 09.32

Mimina and transition structures were optimized using the unrestricted DFT method, UB3LYP,

with the 6-31G(d) basis set.32

Frequency analyses were carried out on stationary points to verify

that they are minima or saddle points (transition structures - TSs). Energies were recalculated

with UB3LYP/6-311++G(d,p) on these optimized geometries, unless otherwise stated. To ensure

that the correct unrestricted wavefunctions were obtained, a stability test was carried out with

Gaussian keyword stable=opt. Solvation corrections were calculated with the conductor-like

5

polarizable continuum model (CPCM) using the UAHF atomic radii. Multi-determinant

wavefunction theory, Complete Active Space with Second-order Perturbation Theory (CASPT2)

calculations were conducted with MOLCAS 7.4.33

A 10-electron, 10-orbital active space was

employed for CASPT2 calculations with a large cc-pVTZ basis set on the optimized geometries.

1.4 Results and Discussions

1.4.1 Modeling the Dioxirane Oxidation of C-H Bonds

Modeling of dioxirane oxidation has led to different conclusions about the TS geometries.

Goddard34,35

and Bach36

investigated the electronic structure of dioxirane. The O-O bond lengths

in both DMDO and TFDO are 1.51 Å, readily broken due to the ring strain. Early calculations

show that a spiro geometry is preferred over a planar geometry for epoxidations.26,27,37

Houk

explored the geometry and selectivities of dioxirane and cyanodioxirane, a model for TFDO,

with restricted DFT calculations.26

Due to the complexity of the potential energy surface (PES),

and possible intersystem crossing of the singlet and triplet PES, later studies proposed several

different first-order saddle points, associated with three different transition structures.38,39

(Fig.

1.1) We have explored these in more detail in order to determine which is likely to be the most

important transition structure for the reaction.

6

Figure 1.1. Transition state geometries for the reaction of DMDO and isobutane

Table 1.1. Natural orbital occupancies of frontier orbitals, and activation energies in the

transition states a

HOMO LUMO CASPT2 UB3LYPb

TS1-A 1.42 0.58 21.5 19.9c

TS1-B 1.24 0.76 22.5 15.8

TS1-C 1.64 0.37 19.5 20.0

a CASPT2(10,10)/cc-pVTZ, energies in kcal/mol

b UB3LYP/6-311++G(d,p)

c 24.2 kcal/mol with RB3LYP/6-311++G(d,p)

The first transition state, TS1-A, is a spiro transition state with all electrons paired, located

by restricted DFT calculations.26,40

This corresponds to a concerted process in which the t-butyl

alcohol and acetone are produced without the formation of an intermediate. A wavefunction

stability test, however, reveals that this wavefunction is unstable with respect to an unrestricted

wavefunction. That is, an open-shell wavefunction is lower in energy. The closed-shell treatment

7

was further invalidated with the multi-determinant CASPT2, as strong diradical character is

observed in the transition state, indicated by the HOMO/LUMO occupation number of 1.52/0.49.

(Table 1.1)

The second TS, TS1-B, has the O-O bond perpendicular to the breaking C-H bond. This is

found to be a radical hydrogen abstraction process beginning from the 1,3-dioxy radical. The

resulting radical intermediates could then recombine to transfer the OH group to form the

products. For this mechanism, the rate limiting step is the dissociation of the DMDO O-O bond

from the singlet diradical, which requires an activation enthalpy of 23.1 kcal/mol according to

Cremer.41

The bond-dissociation rate determining step has a higher barrier than all three C-H

activation barriers in Table 1.1, so it is disfavored under kinetic control.

The third TS, TS1-C, has the O-O bond aligned with the breaking C-H bond. We use

stable=opt to obtain the correct wavefunction at the initial geometry, then perform each

geometry optimization for transition state using the optimized wavefunction as an initial guess

with the Gaussian keyword guess=read. To make sure the optimized geometries have the correct

wavefunction, the same procedure was repeated on the optimized geometries. Both the UB3LYP

and CASPT2 indicated slight diradical character for the transition state, with <S2> value =

0.5312 with B3LYP, and HOMO/LUMO occupation numbers 1.64/0.37 with CASPT2.

Therefore TS1-C is the more reliable transition state. As CASPT2 becomes formidably

expensive with larger systems, UB3LYP with the stable open-shell wavefunction was employed

for further studies.

8

The mechanism involving TS1-C is analogous to the “oxygen rebound” mechanism

common in iron-oxo oxidations.42,43

The complete reaction pathway was then explored and is

summarized in Figure 1.2. The reactant complex goes through C-H activation transition state

TS1 and forms a weakly-bound radical pair intermediate, which rebounds and forms the final

product. The reaction can be tracked with the forming C-O bond length. As the two reactants

approaches each other, the C-O bond distance decreases until it reaches a minimum at 2.50 Å in

TS1, then increases back to 3.15 Å after hydrogen abstraction, leaving two radical centers in the

intermediate. The C-O bond distance then decreases again, (“rebounds”) through 2.51 Å in TS2

to eventually 1.43 Å in the product. As the second transition state has essentially no barrier, the

radical pair intermediate is expected to have too short a lifetime to escape from the solvent cage,

leading to the retention of stereochemistry as observed in experiments.

The same conclusions hold for the more reactive oxidant, TFDO. The C-H activation free

energy is around 5 kcal/mol lower with TFDO than DMDO, but the preferred transition state is

again obtained with unrestricted and optimized wavefunctions. The overall reaction pathway is

similar and the rate-limiting step is again the C-H activation step with diradical character (not

shown). Based on the similarity of the DMDO and TFDO reaction pathways, the in situ

generated tBu-TFDO is expected to follow the same mechanism. The computational results are

in agreement with Curci’s finding that TFDO is more reactive than DMDO without diminished

selectivity, as an exception to the reactivity-selectivity rule.9

9

Figure 1.2. Reaction pathway in DMDO oxidations. Energies in kcal/mol.

10

1.4.2 Oxidation of Equatorial and Axial C-H Bonds of Cyclohexane

Figure 1.3. Optimized TSs for the reactions of DMDO with equatorial (2-eq-TS) and axial

(2-ax-TS) C-H bonds. Energies in kcal/mol with respect to the corresponding reactants.

Individual components of distortion energies (alkane distortion, dioxirane distortion) are given in

parentheses.

The TSs for axial and equatorial C-H oxidation by DMDO are shown in Figure 1.3. No

significant axial/equatorial preference is found. The activation energy for oxidizing the

cyclohexane equatorial hydrogen is 21.6 kcal/mol, while for the axial hydrogen it is 21.4

kcal/mol.

11

To evaluate the factors governing relative reactivities of C-H bonds quantitatively, a

distortion/interaction energy analysis was conducted. The distortion energy (ΔEd‡) is the energy

required to distort reactants, hydrocarbon plus the oxidant, dimethyldioxirane (DMDO), into the

transition state (activated) geometry. The interaction energy (ΔEi‡) is the energy lowering due to

the interaction of the two distorted reactants. Distortion energies favor attack on the axial

hydrogen by 1.1 kcal/mol, while interaction energies favor equatorial attack by 0.9 kcal/mol.

These two components largely cancel out, leaving essentially no preference (0.2 kcal/mol). The

nearly identical activation energies for 2-eq-TS and 2-ax-TS indicate very weak axial/equatorial

selectivity, if any, as observed experimentally for C-H activations of different types.25

The distortion of the DMDO moiety is about the same for both transition states, and the

difference in the distortion energies mainly comes from the cyclohexane distortion.

Hyperconjugation is a likely factor to the 0.9 kcal/mol difference in interaction energies. (Fig.

1.3) In the 2-eq-TS, there are two C-C bonds anti to the breaking C-H bond, while there are two

antiperiplanar C-H bonds in the 2-ax-TS. The hyperconjugative stabilization of the partial

positive charge in the polarized transition states is larger in 2-eq-TS.20

A slight lengthening of

the antiperiplanar bonds is observed in the optimized transition structures, indicative of this

hyperconjugative interaction.

The distortion energy difference can be attributed to torsional interactions. The Newman

projections of these two transition states are given in Fig. 1.3. The internal C-C-C-C dihedral

angle is 56° in the 2-eq-TS, slightly more distorted than the one in cyclohexane (55°). This angle

is 48° in 2-ax-TS, closer to the more relaxed angle in the cyclohexyl radical (44°). The

relaxation is reflected as a 1.0 kcal/mol lowering in the distortion energy. Hydrogen-hydrogen

12

repulsion also plays a part in the relative stability. As hydrogen is being abstracted from a

tetrahedral sp3 carbon, the carbon flattens and becomes sp

2 in character. As shown in the

Newman projection for 2-eq-TS, this lead to slightly greater eclipsing of the vicinal C-H bonds,

as indicated in the 45° dihedral angle, while that in 2-ax-TS is 50°. Therefore, distortion energy

for 2-eq-TS is higher. This torsional effect favoring axial attack was observed in the case of

nucleophilic attack on cyclohexanone (Felkin’s rule)44,45

and reactions of cyclohexyl radicals,46

in which the axial TS has a more staggered conformation than the equatorial TS. In the reactions

studied here, the cancellation of distortion and interaction energy preferences leads to no intrinsic

selectivity for the secondary C-H bonds of cyclohexane.

This type of analysis has been performed on a variety of cyclohexanes and decalins. The

results are tabulated in Table 1.2.

Table 1.2. Summary of activation energies and distortion/interaction energies (kcal/mol)

Compound

C-H bond for

oxidation ΔH‡ ΔE

ΔEd‡

(alkane)

ΔEd‡

(DMDO)

ΔEd‡

(total)

ΔEi‡

1 tBu-H 16.2 20.0 11.1 21.7 32.7 -12.8

2-eq 17.7 21.6 13.1 22.1 35.2 -13.6

2-ax 17.5 21.4 12.1 22.0 34.1 -12.7

3-eq 15.4 18.9 10.5 21.6 32.1 -13.2

3-ax 15.8 19.5 10.5 21.6 32.1 -12.6

13

3’-ax 16.1 19.6 10.5 21.6 32.2 -12.6

4-eq 15.3 18.7 10.2 21.5 31.7 -13.0

4-ax 15.8 19.2 10.2 21.5 31.7 -12.5

4’-ax 16.4 19.8 10.5 21.7 32.2 -12.4

5-eq 15.4 18.9 10.5 21.6 32.1 -13.2

6 14.7 18.2 9.9 21.3 31.2 -13.0

7 14.2 17.7 9.4 21.1 30.4 -12.8

8a 13.3 16.4 10.1 20.2 30.3 -13.9

9a 13.5 16.8 10.3 20.6 30.9 -14.1

10-eq-C2 18.3 22.2 12.8 22.1 34.9 -12.7

10-eq-C3 18.8 22.7 13.8 23.8 36.0 -13.3

10-ax-C3 18.5 22.2 12.5 22.1 34.5 -12.3

14

a R = -C(O)NHCH2CF3, iPr on the decalin ring is omitted

1.4.3 Oxidation of Methylcyclohexane

Figure 1.4. Optimized TSs for the reactions of DMDO with (a) cis-1,4-dimethylcyclohexanes at

the equatorial (3-eq-TS) and axial C-H bonds (3-ax-TS); (b) cis-1,2-dimethylcyclohexanes at the

equatorial (4-eq-TS) and axial (4-ax-TS) C-H bonds. Energies in kcal/mol with respect to the

corresponding reactants. Individual components of distortion energies (alkane distortion,

dioxirane distortion) are in parentheses.

15

The TSs for tertiary axial/equatorial C-H oxidation of cis-1,4-dimethylcyclohexane by

DMDO are shown in Fig. 1.4(a), as a eudesmane model system with both axial and equatorial

tertiary C-H bonds. The activation energy for the reaction of the equatorial C-H bond in 3-eq-TS

is 0.6 kcal/mol lower than for the axial C-H bond in 3-ax-TS. While interaction energies again

favor equatorial hydrogen abstraction, the distortion energies are essentially the same. The

axial/equatorial selectivity comes from the interaction energies, which are -13.2 kcal/mol for

3-eq-TS and -12.6 kcal/mol for 3-ax-TS. This preference, coming from interaction energies, is

found again in 4-eq-TS and 4-ax-TS, where both axial and equatorial C-H bonds are present in

the same molecule. (Fig. 1.4(b))

The distortion energies are identical for the two transition states, while the interaction

energies are -13.0 kcal/mol and -12.5 kcal/mol, respectively. Just as the case of secondary

hydrocarbons, the lengthening of antiperiplanar C-C bonds is shown in the equatorial TS, leading

to better hyper- conjugation. A more detailed study of how distortion energies change along the

reaction coordinate is summarized in Figure 1.5. Distortion energies are the same since the

tertiary C-H oxidation transition states are earlier than the transition states for secondary C-H

oxidation. With the same distortion energy and favored interaction energy for abstraction of

equatorial hydrogens, the latter is slightly more susceptible to oxidation.

16

Figure 1.5. Breakdown of distortion/interaction energies along the internal reaction coordinates

(IRC) for the reactions of DMDO with (a) 1-methylcyclohexanes and (b) cyclohexanes. The C-H

bonds are functionalized at the equatorial position (for 1-axial-1-methyl-cyclohexane) or axial

position (for 1-equatorial-1-methyl-cyclohexane). Energies in kcal/mol with respect to there

corresponding reactants. The x axis is the “internal reaction coordinate” from calculation, and the

origin is the transition state.

The stereochemistry of the remote methyl group has no effect on the reactivity, only the

stereochemistry of the carbon on which hydrogen is abstracted matters. As in Table 1.2, axial

C-H oxidations give almost identical distortion and interaction energies for

cis-1,4-dimethylcyclohexane (3-ax) and trans-1,4-dimethylcyclohexane (3’-ax), leading to the

same activation energy.

17

The stereochemistry of the methyl group to the reacting center has little impact on the

interaction energy, which is intrinsic whether an axial or equatorial hydrogen is abstracted. As in

Table 1.2, oxidation of cis-1,2-dimethylcyclohexane (4-ax) and trans-1,2-dimethylcyclohexane

(4’-ax) give virtually identical interaction energies. The distortion energy for 4’-ax is raised by

0.5 kcal/mol due to developing torsional strain in the transition state as the reacting site becomes

more sp2 in character. Overall the activation energy for 4’-ax is 0.6 kcal/mol higher than that of

4-ax.

Compared with tertiary C-H bonds, activations of secondary C-H bonds require higher

activation energies, which are around 21.5 kcal/mol in Fig. 1.3, due to the electrophilic nature of

the oxidizing reagent, compared with those activation energies for tertiary C-H bonds, which are

below 20.0 kcal/mol. This agrees with Curci’s experiments that tertiary C-H bonds are generally

more reactive than secondary C-H bonds, and the usual greater stability of tertiary radicals.17

The

differences in activation energies between secondary C-H oxidation (~18 kcal/mol) and tertiary

C-H oxidation (15-16 kcal/mol) is comparable to, or slightly higher, than the difference between

secondary and tertiary C-H bond dissociation enthalpies (BDEs) (98 and 96 kcal/mol,

respectively47

) because there is polarization and partial positive charge build up at the oxidized

carbon in the transition states.

18

1.4.4 Influence of axial methyl substitution on reactivity

Figure 1.6. Optimized reactant and transition state geometries for the reaction of DMDO with

substituted cyclohexanes. Energies are in kcal/mol. Individual components of distortion energies

(alkane distortion, dioxirane distortion) are in parenthesis.

To systematically investigate the effects of 1,3-diaxial methyl-methyl interactions,

reactivities of the tertiary equatorial C-H bonds reacting on axial methylcyclohexane (5),

diaxial 1,3-dimethylcyclohexane (6), and triaxial 1,3,5-trimethylcyclohexane (7) were studied.

The reactants and transition states are shown in Fig. 1.6. The methyl groups in 5, 6, 7 were axial

19

rather than in the more stable equatorial position, to serve as models for rigid terpenes where

the methyl groups are fixed to be axial, as in Baran’s selective oxidation of 1.24

The relative rate

for oxidation of cis-1,3-dimethylcyclohexane (6) is slightly higher than the

trans-1,3-dimethylcyclohexane, as strain release operates in the former isomer, in agreement

with a ratio of 1.4-1.5:1 by TFDO oxidation generated in situ.

As more axial methyl substituents were added (5 to 6 to 7), the nearest H-H distance gets

shorter, and more strain is imposed as the H-H closed shell repulsion increases. In the reactants,

there is no significant H-H repulsion in 5, there is one H-H interaction with a 2.12 Å distance in

6, and there are three pairs of CH-CH interactions of 2.11 Å in 7. In the transition state for

abstraction of equatorial hydrogens, the methyl bends away from the ring, releasing 1,3-diaxial

strain. In 5, the H-H distances are above 2.4 Å, and no strain is released in the TS. In 6, the

closest H-H bond distance increases from 2.12 Å in the reactant to 2.19 Å in TS, releasing strain.

In 7, the two closest H-H distances increase from 2.11 Å to 2.20 Å and 2.21 Å, and further strain

release is expected.

The trend towards lower activation energy was reflected in the decreasing distortion

energies, from 32.1 (5-TS) to 31.2 (6-TS) to 30.4 kcal/mol (7-TS). The interaction energies are

nearly the same, increasing by 0.2 kcal/mol (less negative) with each additional axial methyl

group, but this change is smaller than the increase in distortion energies, which dominates the

trend in reactivity. Most of the difference in distortion energy results from the hydrocarbon

reactant. The steric acceleration proposed to result from strain release by Eschenmoser,17

is

manifested in lower distortion energy in the TS. In other words, the reactants are pre-distorted

towards the TS geometry, and the activation energy is lowered. We have referred to this

20

phenomenon as distortion-acceleration.48

The higher reactivity along the series is also paralleled

by slightly earlier transition states along the series, and so the distortion energy of the dioxirane

decreases in the series.

1.4.5 Strain Release and Enhanced Reactivity towards Oxidations in Steroidal Systems

The strain release (or distortion-acceleration) effect has been taken advantage of by

Baran24

to generate selectivity in C-H oxidations, as in Scheme 1.2. A 1:1 mixture of 8’ and 9’

gave the corresponding oxidation products in a ratio of 3:1. The strain release effect was

proposed to account for the enhancement in reactivity of 8’ as compared with 9’. A similar trend

in reactivity was observed in oxidation by ozone, which produced 8’ and 9’ in a 4:1 ratio.

We modeled the reaction with methyl(trifluoromethyl)dioxirane (TFDO) and

dichloromethane as the solvent. The results are shown in Figure 1.7. A model study with DMDO

and these substrates in the gas phase is summarized in Table 1.2.

Scheme 1.2. Reaction products in a competition reaction

21

Figure 1.7. Optimized reactant and transition state geometries for the reaction of TFDO with (a)

8 and (b) 9. R = -C(O)NHCH2CF3. The iPr substituents on the decalin rings of 8’ and 9’ were

omitted for simplicity. Energies are in kcal/mol. Individual components of distortion energies

(alkane distortion, dioxirane distortion) are in parenthesis. Energies with dichloromethane

solvation correction are in square brackets.

Reactants 8 and 9, models for 8’ and 9’, and transition states for DMDO 8-TS and 9-TS

are shown in Fig. 1.7(a) and Fig. 1.7(b), respectively. In Fig. 1.7(a), the two axial methyl groups

in 8 lead to strong 1,3-diaxial strain, with the nearest H-H distance being 2.05 Å. This distance

becomes 2.08 Å in 8-TS, which reduces the strain. The remaining H-H distances stay rather

constant, contributing little to the activation energies due to strain release. The energy required to

22

distort 8 into 8-TS is 26.3 kcal/mol, mainly from the dioxirane distortion, due to the stiffness of

the ring.

The aforementioned methyl-methyl interaction in 8 is alleviated when the bridge methyl is

replaced by a hydrogen atom for 9 in Fig. 1.7(b). Compared with the case of 8, there is less strain

release. In other words, more energy was required to distort the reactant 9 into its transition states

9-TS, with distortion energy increased to 26.8 kcal/mol; the activation energy is 11.4 kcal/mol.

The 0.5 kcal/mol difference in distortion energy, although rather small, dominates over the

interaction energies, and lead to 0.3 kcal/mol lower activation energy for 8-TS than 9-TS. With

solvent correction, there is 0.6 kcal/mol lower in activation enthalpy for 8-TS, in agreement with

observed 3:1 experimental ratio. For model study with DMDO, predicted activation energy for

8-TS is 0.4 kcal/mol lower than 9-TS (16.4 kcal/mol and 16.8 kcal/mol, respectively). The axial

methyl group gives only 0.4 kcal/mol rate enhancement for the substituted decalin, compared to

0.7 kcal/mol with substituted cyclohexane, which is likely due to the rigidity of the ring, as

reflected in the distortion energies of the ring systems.

1.4.6 Selective Oxidations of Sclareolide

Sclareolide, 10, has five methylenes and ten distinct secondary hydrogens. Experimentally,

10 is oxidized at both the C2 and C3 positions (steroid numbering),49

and ratios of products are

given in Table 1.3.

Electronically, oxidation of C3 is favored because it is remote from the

electron-withdrawing lactone that deactivates the methylenes at C1, C5 and C6 and steric effects

23

favor oxidation at C2. For reagents that behave like small reagents, such as dioxiranes and

Fe(PDP) there is marginal selectivity between oxidation at the C2 and C3 positions. The use of a

manganese porphyrin results in considerably higher site selectivity.50

In the case of the

Rh-mediated amination employing a reagent that behaves like a large reagent, excellent C3

selectivity is observed.5n,24

Table 1.3 Experimental and calculated ratios for C2/C3 products from 10. Energies in kcal/mol.

oxidant C2:C3 ref Calc ΔE‡(C2) ΔE

‡(C3axial) ΔE

‡(C3equatorial)

DMDO 1.5:1a

49 ~1:1

b 22.2

b 22.2

b 22.7

b

TFDO 2.1:1~1:2.0c

49 2.7:1

d 10.2

d 12.1

d 10.9

d

tBu-TFDO 2.4:1e

49 8.3:1

d 11.9

d 14.3

d 13.1

d

Rh2(esp)2f Only C2

49

Mn(TPP)Clg

Fe(PDP)

7:1

1.4:1

50

21

a 1:1 acetone:DCM, <5% conversion

b UB3LYP/6-311++G(d,p), gas phase

c Ratios with a variety of solvents gave from 2.1:1 with hexafluoroisopropanol to 1:2.0 with

acetonitrile.

d UB3LYP/6-311++G(d,p), calculated ratio from activation energies in 298 K with

CPCM/UAHF in acetonitrile

e In acetonitrile solvent

f Amination products

g A ratio of 3:1 was observed for fluorination with Mn(TMP)Cl.

Table 1.4 Analysis of site selectivities on 10 for TFDO vs tBu-TFDO. Energies in kcal/mol.

oxidant TS ΔH‡(solv)

a ΔH

‡ ΔE

‡ ΔEd

‡(10) ΔEd

‡(TFDO) ΔEd

‡(10 ΔEi

24

+TFDO)

TFDO C2 6.7 12.2 15.7 8.7 20.2 28.9 -13.2

TFDO C3-ax 8.8 12.7 15.9 8.5 20.1 28.6 -12.7

TFDO C3-eq 7.5 12.8 16.1 9.2 20.3 29.5 -13.4

tBu-TFDO C2 8.3 14.2 17.8 9.9 20.8 30.7 -12.9

tBu-TFDO C3-ax 10.7 15.1 18.7 10.3 20.8 31.1 -12.4

tBu-TFDO C3-eq 9.5 15.0 18.7 11.1 20.9 32.1 -13.4

a Calculated solvent with CPCM/UAHF in acetonitrile

Table 1.4 shows details of the computed activation barriers with different oxidants. The

bulky tBu group of tBu-TFDO increases the activation energies by about 2 kcal/mol, reflected in

the distortion energies especially from the substrate 10. The interaction energies are essentially

the same, since the methyl group in TFDO and tBu in tBu-TFDO have similar electronic effects.

The reactants and transition states geometries are shown in Figure 1.8 for TFDO and Figure 1.9

for tBu-TFDO. With solvent correction (acetonitrile), the axial TS becomes disfavored probably

due to a higher energy cost for the rearrangement of solvent molecules. Also a higher C2:C3

selectivity is expected. A thorough study of solvent effects is underway.

Figure 1.8 gives the geometries of sclareolide 10 and three transition states for C-H

oxidations on the C2/C3 C-H bonds with TFDO. The methyl-hydrogen strain is released, as the

bond lengths for nearest H-H pairs, which are 2.10 Å and 2.12 Å in 10, increased to 2.11 Å and

2.18 Å in 10-eq-C2-TS, but stays the same on 10-ax-C3-TS and 10-eq-C3-TS. Not surprisingly,

the strain release effect is not as pronounced as is the case with the relief of a methyl-methyl 1,3

diaxial interaction. Accordingly, a mixture of products results.

25

Figure 1.8. Optimized reactant and transition state geometries for the reaction of TFDO with

sclareolide 10. Energies are in kcal/mol. Individual components of distortion energies (alkane

distortion, dioxirane distortion) are in parenthesis. Energies with acetonitrile solvation correction

are in square brackets.

Figure 1.9 gives the geometries of sclareolide 10 and three transition states for C-H

oxidations on the C2/C3 C-H bonds with the bulkier oxidant tBu-TFDO. The site selectivity for

26

oxidation on the C2 versus C3 position is predicted to be 0.9 kcal/mol. Experimentally a higher

C2/C3 ratio was observed for tBu-TFDO than for TFDO. The steric effect of the tert-butyl group

raises the activation barriers for tBu-TFDO oxidations compared with TFDO by about 2

kcal/mol. This energy difference is also consistent with the low yield and low levels of reactivity

observed with tBu-TFDO compared with TFDO. The trace conversion affected by tBu-TFDO

may also reflect the low concentration of this reagent in the reaction mixture, as it is formed in situ.

The increased steric effects also render the equatorial hydrogen on C3 as susceptible to oxidation

as the axial hydrogen at the same position.

27

Figure 1.9. Optimized reactant and transition state geometries for the reaction of tBu-TFDO

with sclareolide 10. Energies are in kcal/mol. Individual components of distortion energies

(alkane distortion, dioxirane distortion) are in parenthesis. The prime labels on TS distinguish the

transition states for tBu-TFDO oxidation with those with DMDO. Energies with acetonitrile

solvation correction are in square brackets.

28

1.5 Conclusions

The site selectivities and stereoselectivities of sp3 C-H activations in cyclohexane

derivatives by dioxiranes were investigated computationally. CASPT2 calculations established

the appropriate transition state for C-H oxidations by dioxiranes, which corresponds to that

located using unrestricted DFT. The full pathway for dioxirane oxidation was established. The

distortion/interaction model was used to understand the origins of selectivity in each case. In the

absence of axial substitution of cyclohexanes there is no innate preference for either axial or

equatorial C-H activation: eclipsing interactions slightly favor axial activation, but this is

counterbalanced by the greater hyperconjugative stabilization in equatorial activation. The axial

and equatorial C-H bonds of cyclohexane have similar reactivities.

However, with increasing axial substitution, computed activation barriers decrease, in

accord with experiment. The enhancement in equatorial reactivity by geminal and diaxial

substitution arises as a consequence of strain release in going to the transitions state.25

The

distortion/interaction model allows for a quantification of the acceleration that arises as a result

of strain release. The strain release effect is around 0.7 kcal/mol when the strain involving two

1,3-diaxial methyl groups is released. The eudesmane 8 is more reactive than 9 due to the strain

release effect, but the C2/C3 site selectivity for sclareolide is dominated by steric effects and can

be controlled by the bulkiness of the reagents.

29

1.6 Experimental Section

1.6.1 General Procedures

All reactions were carried out under an inert nitrogen atmosphere with dry solvents under

anhydrous conditions unless otherwise stated. Dry acetonitrile (MeCN) was obtained by passing

the previously degassed solvents through activated alumina columns. Product ratios were

determined by analysis of the crude 1H NMR spectra. Yields refer to chromatographically and

spectroscopically (1H-NMR) homogeneous materials, unless otherwise stated. Reagents were

purchased at the highest commercial quality and used without further purification, unless

otherwise stated. Reactions were monitored by thin-layer chromatography (TLC) carried out on

0.25 mm E. Merck silica gel plates (60F-254 or RP-18 F254s) using UV light as the visualizing

agent and an acidic solution of p-anisaldehyde, phosphomolybdic acid, cerric ammonium

molybdate, Seebach’s stain (PMA and CAM), ninhydrin, or potassium permanganate and heat as

developing agents. E. Merck silica gel (60, particle size 0.043-0.063 mm) was used for flash

column chromatography.

1.6.2 General Procedure for Oxidation using TFDO Generated in situ.

The substrate (0.2 mmol) was dissolved in MeCN (6.0 mL, 0.03M) by stirring and

sonicating. To this solution was added aqueous sodium phosphate buffer (6.0 mL, pH = 7.5, 25

mM) and aqueous Na2EDTA (1.3 mL, 40 mM). The solution was then cooled to 4 C (ambient

temperature) and cool (4 C) 1,1,1-trifluoroacetone (0.143 mL, 1.60 mmol, 8.0 equiv) was

added. Oxone

(KHSO5 0.5KHSO4 0.5K2SO4, 738 mg, 2.4 mmol, 12 equiv) was then added

30

in a single portion and the suspension was rapidly stirred for 36 hours. Then, the mixture was

warmed to room temperature and extracted with Et2O (3 15 mL). The combined organic layers

were washed with brine (50 mL), dried over MgSO4, filtered and concentrated.

Similar conditions for in situ dioxirane formation for C-H oxidation have been

developed.51

References 21, 52, and 53 include spectral data for the sclareolide oxidation

products. Reference 54 includes data for the products from oxidation of

1,3-dimethylcyclohexanes. tert-butyl trifluoromethylketone could be synthesized according to

ref. 55. Alternatively, CF3TMS could be added dropwise to a mixture of methyl pivalate and

excess CsF at 0 °C.

1.7 References

(1) Yamaguchi, J.; Yamaguchi, A. D.; Itami, K. Angew. Chem. Int. Ed. 2012, 51, 8960.

(2) Doyle, M. P.; Goldberg, K. I. Acc. Chem. Res. 2012, 45, 777.

(3) Crabtree, R. H. Chem. Rev. 2010, 110, 575.

(4) Brueckl, T.; Baxter, R. D.; Ishihara, Y.; Baran, P. S. Acc. Chem. Res. 2012, 45, 826.

(5) For recent reviews, see: a) Davies, H. M. L.; Beckwith, R. E. J.; Chem. Rev. 2003, 103, 2861;

b) Punniyamurthy, T.; Velusamy, S.;Iqbal, J. Chem. Rev. 2005 105, 2329; c) Godula, K.; Sames,

D. Science 2006, 312, 67; d) Dick, A. R.; Sanford, M. S. Tetrahedron 2006, 62, 2439; e) Davies,

H. M. L.; Manning, J. R. Nature 2008, 451, 417 f) Giri, R.; Shi, B.-F.; Engle, K. M.; Maugel, N.;

Yu, J.-Q. Chem. Soc. Rev. 2009, 38, 3242; g) Daugulis, O.; Do, H.-Q.; Shabashov, D. Acc. Chem.

Res. 2009, 42, 1074. h) Mkhalid, I.A.I.; Barnard, J.H.; Marder, T. B.; Murphy, J. M.; Hartwig, J.

F. Chem. Rev. 2010, 110, 890; i) Shul’pin, G. B. Org. Biomol. Chem. 2010, 8, 4217; j) Jazzar, R;

Hitce, J.; Renaudat, A.; Sofack-Kreutzer, J.; Baudoin, O. Chem. Eur. J. 2010, 16, 2654; k) Lyons,

T. W.; Sanford, M. S. Chem. Rev. 2010, 110, 1147; l) “C-H Activation”: Topics in Current

Chemistry (Eds.: J.-Q. Yu, Z. Shi), Springer, Berlin, 2010, 292. m) Doyle, M. P.; Duffy, R.;

31

Ratnikov, M.; Zhou, L. Chem. Rev. 2010, 110, 704; n) Roizen, J. L.; Harvey, M. E.; Du Bois, J.

Acc. Chem. Res. 2012, 45, 911.

(6) Murray, R. W.; Jeyaraman, R. J. Org. Chem. 1985, 50, 2847.

(7) Mello, R.; Fiorentino, M.; Sciacovelli, O.; Curci, R. J. Org. Chem. 1988, 53, 3890.

(8) Curci, R.; D'Accolti, L.; Fusco, C. Acc. Chem. Res. 2006, 39, 1.

(9) Curci, R.; Dinoi, A.; Rubino, M. F. Pure Appl. Chem. 1995, 67, 811.

(10) Bravo, A.; Bjorsvik, H. R.; Fontana, F.; Minisci, F.; Serri, A. J. Org. Chem. 1996, 61, 9409.

(11) Curci, R.; Dinoi, A.; Fusco, C.; Lillo, M. A. Tetrahedron Lett. 1996, 37, 249.

(12) Bravo, A.; Fontana, F.; Fronza, G.; Minisci, F.; Zhao, L. H. J. Org. Chem. 1998, 63, 254.

(13) Romney, D. K.; Miller, S. J. Org. Lett. 2012, 14, 1138.

(14) Newhouse, T.; Baran, P. S. Angew. Chem. Int. Ed. 2011, 50, 3362.

(15) Barton, D. H. R. J. Chem. Soc. 1953, 1027.

(16) Barton, D. H. R.; Beviere, S. D.; Chavasiri, W.; Csuhai, E.; Doller, D.; Liu, W. G. J. Am.

Chem. Soc. 1992, 114, 2147.

(17) Schreiber, J.; Eschenmoser, A. Helv. Chim. Acta. 1955, 38, 1529.

(18) D'Accolti, L.; Fiorentino, M.; Fusco, C.; Rosa, A. M.; Curci, R. Tetrahedron Lett. 1999, 40,

8023.

(19) Mello, R.; Fiorentino, M.; Fusco, C.; Curci, R. J. Am. Chem. Soc. 1989, 111, 6749.

(20) a) Gonzalez-Nunez, M. E.; Castellano, G.; Andreu, C.; Royo, J.; Baguena, M.; Mello, R.;

Asensio, G. J. Am. Chem. Soc. 2001, 123, 7487; b) Gonzalez-Nunez, M. E.; Royo, J.; Mello, R.;

Baguena, M.; Ferrer, J. M.; de Arellano, C. R.; Asensio, G.; Prakash, G. K. S. J. Org. Chem.

32

2005, 70, 7919.

(21) Chen, M. S.; White, M. C. Science 2010, 327, 566.

(22) Chen, M. S.; White, M. C. Science 2007, 318, 783.

(23) Ishihara, Y.; Baran, P. S. Synlett 2010, 1733.

(24) Chen, K.; Baran, P. S. Nature 2009, 459, 824.

(25) Chen, K.; Eschenmoser, A.; Baran, P. S. Angew. Chem. Int. Ed. 2009, 48, 9705.

(26) Du, X. H.; Houk, K. N. J. Org. Chem. 1998, 63, 6480.

(27) Glukhovtsev, M. N.; Canepa, C.; Bach, R. D. J. Am. Chem. Soc. 1998, 120, 10528.

(28) Paton, R. S.; Kim, S.; Ross, A. G.; Danishefsky, S. J.; Houk, K. N. Angew. Chem. Int. Ed.

2011, 50, 10366.

(29) Lan, Y.; Houk, K. N. J. Am. Chem. Soc. 2010, 132, 17921.

(30) Ess, D. H.; Houk, K. N. J. Am. Chem. Soc. 2008, 130, 10187.

(31) Ess, D. H.; Houk, K. N. J. Am. Chem. Soc. 2007, 129, 10646.

(32) Gaussian 09, Revision B.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, N. J.; 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.;

33

Cioslowski, J.; Fox, D. J. Gaussian, Inc., Wallingford CT, 2009.

(33) Aquilante, F.; De Vico, L.; Ferre, N.; Ghigo, G.; Malmqvist, P. A.; Neogrady, P.; Pedersen, T.

B.; Pitonak, M.; Reiher, M.; Roos, B. O.; Serrano-Andres, L.; Urban, M.; Veryazov, V.; Lindh, R.

J. Comput. Chem. 2010, 31, 224.

(34) Harding, L. B.; Goddard, W. A. J. Am. Chem. Soc. 1978, 100, 7180.

(35) Wadt, W. R.; Goddard, W. A. J. Am. Chem. Soc. 1975, 97, 3004.

(36) Bach, R. D.; Andres, J. L.; Owensby, A. L.; Schlegel, H. B.; Mcdouall, J. J. W. J. Am. Chem.

Soc. 1992, 114, 7207.

(37) Shustov, G. V.; Rauk, A. J. Org. Chem. 1998, 63, 5413.

(38) Freccero, M.; Gandolfi, R.; Sarzi-Amade, M.; Rastelli, A. Tetrahedron Lett. 2001, 42, 2739.

(39) Freccero, M.; Gandolfi, R.; Sarzi-Amade, M.; Rastelli, A. J. Org. Chem. 2003, 68, 811.

(40) Fokin, A. A.; Tkachenko, B. A.; Korshunov, O. I.; Gunchenko, P. A.; Schreiner, P. R. J. Am.

Chem. Soc. 2001, 123, 11248.

(41) Cremer, D.; Kraka, E.; Szalay, P. G. Chem. Phys. Lett. 1998, 292, 97.

(42) Ensing, B.; Buda, F.; Gribnau, M. C. M.; Baerends, E. J. J. Am. Chem. Soc. 2004, 126, 4355.

(43) Kumar, D.; de Visser, S. P.; Shaik, S. J. Am. Chem. Soc. 2003, 125, 13024.

(44) Wu, Y. D.; Houk, K. N. J. Am. Chem. Soc. 1987, 109, 908.

(45) Wu, Y. D.; Houk, K. N.; Trost, B. M. J. Am. Chem. Soc. 1987, 109, 5560.

(46) Damm, W.; Giese, B.; Hartung, J.; Hasskerl, T.; Houk, K. N.; Zipse, H. J. Am. Chem. Soc.

1992, 114, 4067.

(47) Luo, Y. R. CRC Handbook of Chemistry and Physics, 91st ed. (CRC Press, Boca Raton, FL,

2010)

(48) Gordon, C. G.; Mackey, J. L.; Jewett, J. C.; Sletten, E. M.; Houk, K. N.; Bertozzi, C. R. J.

34

Am. Chem. Soc. 2012, 134, 9199.

(49) Newhouse, T., PhD thesis with Baran, P. S.

(50) (a) Liu, W.; Groves, J. T. J. Am. Chem. Soc. 2010, 132, 12847; (b) Liu, W.; Huang, X.;

Cheng, M.-J.; Nielsen, R.J.; Goddard, W.A., III; Groves, J.T.; Science 2012, 337, 1322.

(51) Annese, C.; D’Accolti, L.; Fusco, C.; Curci, R. Org. Lett. 2011, 13, 2142.

(52) Cambie, R. C.; Joblin, K. N.; McCallum, N. K. Aust. J. Chem. 1970, 23, 1439.

(53) Choudhary, M. I.; Musharraf, S. G.; Sami, A.; Rahman, A.-u. Helv. Chim. Acta. 2004, 87,

2685.

(54) Senda, Y.; Ishiyama, J.; Imaizumi, S. Tetrahedron 1975, 31, 1601.

(55) Wiedemann, J.; Heiner, T.; Mloston, G.; Prakash, G. K. S.; Olah, G. A. Angew. Chem. Int.

Ed. 1998, 37, 820.

35

Chapter 2. Formation of Chelated Z-selective Ruthenium Metathesis Catalysts:

Computational Study on C–H Activation Mechanism, Reactivity, and Selectivity

2.1 Abstract

The mechanism and origins of formation of chelated Z-selective Ru catalysts were explored

using density functional theory. The key step involves an intramolecular C-H bond activation of

the NHC ligand by carboxylates introduced by anion exchange. The C-H activation follows a

concerted metalation-deprotonation mechanism, where carboxylate abstracts the hydrogen from

the “bottom” position, forming a six-membered ring with the metal. Torsional strains from

closed−shell repulsions between the chelating groups dominate over electronic effects for

determining the barriers for the C-H activation process.

2.2 Introduction

Olefin metathesis is a powerful methodology for the construction of new carbon-carbon

double bonds. Since the discovery of metathesis, nearly 400 ruthenium olefin metathesis

catalysts were reported by the end of 2009.1 Along with improved reactivity and selectivity, the

thermal stability, oxygen- and moisture-tolerance of the catalyst have been greatly improved

over the years. New ligands, especially the N-heterocyclic carbenes and benzylidene ether ligand

had special value.

36

For most olefin metathesis catalysts, E olefins were preferentially obtained due to their

greater thermodynamical stability. A highly Z-selective olefin metathesis catalyst was reported in

2011,2 followed by continuous improvements.

3-5 Z-selective catalysts have already been applied

in several synthetic routes.6 The Z-selective breakthrough was achieved by an unexpected C-H

activation of the previously known E-selective catalyst, 1.2,3

(Scheme 2.1) Intermediate 2 was

obtained by replacing the chloride ligands of 1 with pivalate, and 2 unexpectedly underwent

rapid C-H activation to give a C-chelated, Z-selective catalyst, 3. The Ru-carbon bond to the

mesityl group was formed via intramolecular C-H activation.

Scheme 2.1. Generation of the Z-selective catalyst 3 through C-H activation

An even more efficient Z-selective catalyst, 6, was obtained following the same

procedure, without the detection of anion-exchange intermediates.2-4

The Ru-adamantyl bond

was originally formed by ligand exchange promoted by silver pivalate, but the reaction also

occurs even with sodium pivalate (Scheme 2.2).5

37

Scheme 2.2. Generation of the Z-selective catalyst 6 through C-H activation

Computational studies on ruthenium C-H activation have been reported, but mostly on

(aromatic) sp2 C-H bonds. Cavallo studied computationally the C-H activation of a benzene sp

2

C-H bond as part of a study of the decomposition pathway of the second generation Grubbs

catalyst.7 Berlinguette studied C-H activation on the benzene ligand of bis-tridentate ruthenium

complexes.8 Ruthenium (II) catalyzed sp

2 C-H activation was reviewed by Dixneuf.

9 Few cases

of ruthenium catalyzed intramolecular sp3 C-H activation with carboxylic acid ligands have been

reported.10

Murai,11

Sames,12

and Maes employed the Ru3(CO)12 catalyst for the activation of sp3

C-H bonds adjacent to a nitrogen atom.13,14

Computational studies of both sp2 and sp

3 C-H

activation by various transition metals were recently reviewed by Eisenstein,15

and various

mechanisms in carboxylate-assisted transition-metal-catalyzed C-H bond functionalization is

reviewed by Ackermann.16

We have explored the intramolecular C-H activation that forms Z-selective olefin

metathesis catalyst, using density functional theory. We have found that the mechanism follows

a concerted metallation-deprotonation (CMD) process, which is similar to the palladium

catalyzed catalyzed intramolecular sp3 C-H activation with carboxylic acids studied earlier by

38

Fagnou’s group,17,18

and our group.19

We have investigated transition state (TS) geometries,

explored the origins of selectivities with different types of C-H bonds, and have investigated the

influence of both anionic and NHC ligands on the rates of intramolecular C-H activations.

2.3 Computational Details

All calculations were performed with Gaussian 09.20

Geometry optimizations were

performed with B3LYP and the LANL2D2 basis set for Ru and the 6-31G(d) basis set for other

atoms. Single point energies were calculated at the M06/SDD–6-311+G(d,p) level. Solvation

corrections for single point energies were calculated with the SMD solvation model using THF

as the solvent. The reported free energies and enthalpies include zero-point energies and thermal

corrections calculated at 298 K with B3LYP/SDD-6-31G(d).

2.4 Results and Discussions

2.4.1 Transition state geometries for the C-H activation

C-H activations on the metal centers were proposed to follow several classes of

mechanism, such as oxidative addition, sigma-bond metathesis, and electrophilic activation.21

Among these mechanism, carboxylate-assisted C-H bond functionalizations were mostly

proposed to follow the mechanism of base-assisted deprotonation,16

which was named as

concerted metallation-deprotonation (CMD) by Fagnou.17

39

Depending on the side of attack and the coordination atom of the carboxylate, there are

four different ways the dipivalate intermediate 5 can undergo C-H activation. The four transition

state geometries are shown in Figure 2.1. The carboxylic acid attack from the bottom (TS-A and

TS-C), breaking the Ru-O(benzylidene) bond, or the acid attack from the side (TS-B and TS-D).

In TS-A and TS-B, one of the oxygen of the pivalate carboxylic is coordinated to Ru, while the

other oxygen deprotonates the adamantyl C-H bond, thus forming a six-membered transition

state. In TS-C and TS-D, only one oxygen is involved, forming a four-membered transition state.

The most favored transition state, TS-A, has the forming and breaking bonds in a six-membered

ring with the pivalate oxygen deprotonating from the bottom.

Figure 2.1. Four possible transition state geometries for C-H activation of 4. The Ru-O bond

lengths and TS bond lengths are labeled in Å.

40

2.4.2 The free energy profile for the formation of the chelated ruthenium catalyst

Experimentally, the dicarboxylate intermediate 2 was observed as an intermediate in the

formation of catalyst 3 from 1, but intermediate 5 was not characterized in the formation of

catalyst 6 from 4. No chloride-containing C-H activated catalyst has been observed. We explored

whether a monochloro monopivalate might also give C-H activation. The computed free energy

diagram for C-H activation of 4 (1a) by a dipivalate or a monochlorate, mono pivalate is shown

in Figure 2.2.

Figure 2.2. The free energy profile of the C-H activation of 4(1a) to form the C-H activated

catalyst (1g).

The lower energy pathway (colored in blue) involves an anion exchange step, forming a

stable complex 1c with both Cls replaced with pivalates, which coordinates better with

ruthenium. Isomerization of 1c to a less stable complex 1d gives the geometry of CMD (see the

41

previous section). The rate- and stereo-determining C-H activation step (1e-ts) requires an

overall barrier of 23.5 kcal/mol (1d -> 1e-ts). The complex 1f is formed after the C-H activation

and finally the active catalyst 1g is obtained.

An alternative pathway of C-H activation with only one carboxylate and one chloride ligand

was also studied (colored in pink in Figure 2.2).The free energy barrier (1h-ts) is 28.4 kcal/mol,

around 5 kcal/mol less favored than the dicarboxylate pathway, as the other pivalate is bidentate,

which might give additional stabilization in the transition state.

2.4.3 Regio- and stereoselectivity in CH activations

There are three possible hydrogens which could be subject to replacement by Ru in C-H

activation of the admantyl/mesityl NHC. These are shown in Scheme 2.3 with their

corresponding transition states are shown in Figure 2.3.

42

Scheme 2.3. Three possible sites for C-H activation of 1a, leading to three possible chelated

catalysts 1g.

Figure 2.3. Substrate and TS geometries for C-H activation on three different sites of 1c.

The 1e-ts-A, with the adamantyl hydrogen being abstracted cis to the alkylidene gives the

lowest activation barriers of 23.5 kcal/mol, which leads to the Z-selective catalyst 3, as observed

in the experiment. C-H activation on the mesityl group gives the highest barrier of 27.7 kcal/mol.

The there is substantial distortion in 1e-ts-F due to the steric repulsion of the other ortho Me

group on mesityl and the NHC prevents the Me substituent on mesityl rotating to the proximity

of the Ru, as reflected in the short H-H distance (2.16Å) and increased C-H and H-H repulsion

on the adamantyl side as well. C-H activation on the adamantyl hydrogen trans to the alkylidene

is disfavored, probably due to the repulsion between the Ru=C electrons and C-H bonds. As in

Figure 2.4, the C-H bond is anti to the Ru=C bond in 1e-ts-A, taking a staggered conformation,

while in 1e-ts-E the C-H bond is syn to the Ru=C bond, taking an eclipsed conformation.

43

Figure 2.4. Steric effects on stereoselectivity for the cis- and trans-TS for C-H activation of 1c.

Atoms not directly not directed linked to the reactive center are omitted for clarity.

2.4.4 Effects of chelating group and N-substituents on reactivity

Table 2.1. Activation energies for C-H activation of various N-alkyl groups

Entry Alkyl-H G‡

H‡ BDE

1

23.5 23.8 98.0

44

2

20.0 20.5 96.4

3

16.2 16.8 98.1

4

15.2 16.4 100.7

5

25.6 26.0 87.4

The reactivity toward C-H activation of several different alkyl groups was studied and is

summarized in Table 2.1. The activation is facilitated with less steric demanding groups. The

activation of the secondary hydrogen on the cyclohexane ring (TS2) requires 20.0 kcal/mol in

free energy, 3.5 kcal/mol lower than TS1. The activation barrier for the primary hydrogen on the

methyl, TS3 and TS4, are significantly lower, which are only 16.2 and 15.2 kcal/mol,

respectively. Replacing the adamantyl group with mesityl increases the activation barrier to 25.6

kcal/mol, probably due to the closed shell repulsion between the methyl group on mesityl and the

NHC rings, as in the case of C-H activation on the mesityl C-H (1e-ts-F).

Table 2.2. Activation energies for C-H activation with different N-aryl groups

45

Entry Aryl group G‡

H‡

1

23.5 23.8

7

23.5 23.1

8

21.3 22.4

9

22.0 23.1

Substituent effects on the N-aryl groups were investigated, and the results are summarized in

Table 2.2. C-H activation of the new catalyst 7 has the same activation free energy as 1.

Enhanced activity towards C-H activation was observed when one or two of the ortho methyl

groups were replaced with fluorine. As in Figure 2.5, there is stabilizing effect of O-H interaction

in 1c, and destabilizing effect from C-H repulsion in 1e-ts-A, both lead to higher barriers. These

effects are absent when the ortho methyl groups are replaced with fluorines.

46

Figure 2.5. Substrate and TS geometries for C-H activation on three different sites of 5.

Table 2.3. Activation energies for C-H activation of 5- and 6-membered NHC rings

NHC G‡

H‡

1

23.5 23.8

6

17.8 17.9

We also studied the effect of NHC ring size on the C-H activation barrier, and the results

are in Table 2.3. The activation barrier decreases to 17.8 kcal/mol when the five-membered NHC

(1) is replaced with the six-membered NHC (6). The 5 kcal/mol lower barrier arises because

47

additional carbon in the ring brings the N-adamantyl group closer to the pivalate group, which

facilitates the CH activation. The C-H activation process with 6 is also more exergonic than with

the five-membered complex 1 (reaction energies are -12.6 and -3.0 kcal/mol for 6 and 1,

respectively).

Table 2.4. Activation energies for C-H activation with different carboxylate ligands

Entry Anion RCOO- G

‡ H

‡ pKaH

1 tBuCOO- 23.5 23.8 5.01

10 CH3COO- 24.0 24.4 4.75

11 PhCOO- 26.2 25.5 4.2

The anion effect on the barriers of C-H activation is shown in Table 2.4. Replacing the

pivalate with acetate increases the barrier by 0.5 kcal/mol, while with the benzoate the barrier is

2.7 kcal/mol higher. Notice the strength of the conjugate acid follows the order: benzoic acid

(pKa 4.2) > acetic acid (pKa 4.75) > pivalic acid (pKa 5.01). Since an acid is formed as the

byproduct, formation of a strong acid is less favored, leading to a higher barrier.

48

2.5 Conclusions

The mechanism of C-H activation of the Grubbs II catalyst upon converting halide to

carboxylate ligands has been elucidated. The C-H activation proceeds via a six-membered,

bottom-attack transition state with a high stereoselectivity. Reducing the steric demands of the

N-alkyl group leads to lower activation barriers. The steric effect on the ortho position of N-aryl

group has minor impact on the C-H activation barrier, while substitution of methyl with fluorine

decreases the barrier. Replacing the five membered NHC with six membered NHC also

facilitates the CH activation.

.

2.6 References

(1) Vougioukalakis, G. C.; Grubbs, R. H. Chem. Rev. 2010, 110, 1746.

(2) Endo, K.; Grubbs, R. H. J. Am. Chem. Soc. 2011, 133, 8525.

(3) Keitz, B. K.; Endo, K.; Herbert, M. B.; Grubbs, R. H. J. Am. Chem. Soc. 2011, 133, 9686.

(4) Keitz, B. K.; Endo, K.; Patel, P. R.; Herbert, M. B.; Grubbs, R. H. J. Am. Chem. Soc. 2012,

134, 693.

(5) Rosebrugh, L. E.; Herbert, M. B.; Marx, V. M.; Keitz, B. K.; Grubbs, R. H. J. Am. Chem.

Soc. 2013, 135, 1276.

(6) Herbert, M. B.; Marx, V. M.; Pederson, R. L.; Grubbs, R. H. Angew. Chem. Int. Ed. 2013,

52, 310.

(7) Poater, A.; Cavallo, L. J. Mol. Cat. A. – Chem. 2010, 324, 75.

(8) Muise, S. S. R.; Severin, H. A.; Koivisto, B. D.; Robson, K. C. D.; Schott, E.; Berlinguette,

C. P. Organometallics 2011, 30, 6628.

49

(9) Arockiam, P. B.; Bruneau, C.; Dixneuf, P. H. Chem. Rev. 2012, 112, 5879.

(10) Rousseau, G.; Breit, B. Angew. Chem. Int. Ed. 2011, 50, 2450.

(11) Chatani, N.; Asaumi, T.; Yorimitsu, S.; Ikeda, T.; Kakiuchi, F.; Murai, S. J. Am. Chem. Soc.

2001, 123, 10935.

(12) Pastine, S. J.; Gribkov, D. V.; Sames, D. J. Am. Chem. Soc. 2006, 128, 14220.

(13) Bergman, S. D.; Storr, T. E.; Prokopcova, H.; Aelvoet, K.; Diels, G.; Meerpoel, L.; Maes, B.

U. W. Chem.-Eur. J. 2012, 18, 10393.

(14) Peschiulli, A.; Smout, V.; Storr, T. E.; Mitchell, E. A.; Elias, Z.; Herrebout, W.; Berthelot,

D.; Meerpoel, L.; Maes, B. U. W. Chem.-Eur. J. 2013, 19, 10378.

(15) Balcells, D.; Clot, E.; Eisenstein, O. Chem. Rev. 2010, 110, 749.

(16) Ackermann, L. Chem. Rev. 2011, 111, 1315.

(17) Gorelsky, S. I.; Lapointe, D.; Fagnou, K. J. Am. Chem. Soc. 2008, 130, 10848.

(18) Gorelsky, S. I.; Lapointe, D.; Fagnou, K. J. Org. Chem. 2012, 77, 658.

(19) Giri, R.; Lan, Y.; Liu, P.; Houk, K. N.; Yu, J. Q. J. Am. Chem. Soc. 2012, 134, 14118.

(20) Gaussian 09, Revision B.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, N. J.; 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.

(21) Labinger, J. A.; Bercaw, J. E. Nature 2002, 417, 507.

50

Chapter 3. Mechanism and Enantioselectivity in Palladium-Catalyzed Conjugate Addition

of Arylboronic Acids to -Substituted Cyclic Enones: Insights from Computation and

Experiment

3.1 Abstract

Enantioselective conjugate additions of arylboronic acids to -substituted cyclic enones have

been reported previously from our laboratories. Air and moisture tolerant conditions were

achieved with a catalyst derived in situ from palladium(II) trifluoroacetate and the chiral ligand

(S)-t-BuPyOx. We now report a combined experimental and computational investigation on the

mechanism, the nature of the active catalyst, the origins of the enantioselectivity, and the

stereoelectronic effects of the ligand and the substrates of this transformation. Enantioselectivity

is controlled primarily by steric repulsions between the t-Bu group of the chiral ligand and the

-methylene hydrogens of the enone substrate in the enantiodetermining carbopalladation step.

Computations indicate that the reaction occurs via formation of a cationic arylpalladium(II)

species, and subsequent carbopalladation of the enone olefin forms the key carbon-carbon bond.

Studies of non-linear effects and stoichiometric and catalytic reactions of isolated (PyOx)Pd(Ph)I

complexes show that a monomeric arylpalladium-ligand complex is the active species in the

selectivity-determining step. The addition of water and ammonium hexafluorophosphate

synergistically increases the rate of the reaction, corroborating the hypothesis that a cationic

51

palladium species is involved in the reaction pathway. These additives also allow the reaction to

be performed at 40 °C and facilitate an expanded substrate scope.

3.2 Introduction

Asymmetric conjugate addition has become a familiar reaction manifold in the synthetic

chemists’ repertoire.1

Though seminal reports involved highly reactive organometallic

nucleophiles,2

systems were rapidly developed that involved functional-group-tolerant

organoboron nucleophiles. Namely, Hayashi pioneered the use of rhodium/BINAP catalysts for

the asymmetric conjugate addition of a number of boron-derived nucleophiles.3 As an

economical alternative to the rhodium systems, Miyaura pioneered the use of chiral

palladium-phosphine catalysts to address similar transformations4 and Minnaard reported a

palladium-catalyzed asymmetric conjugate addition using a catalyst formed in situ from

palladium trifluoroacetate and commercially available (S,S)-MeDuPhos.5

More recently, asymmetric conjugate addition has become a useful strategy for the challenge

of constructing asymmetric quaternary stereocenters.6 Again, many earlier developed methods

involved highly reactive diorganozinc,7

triorganoaluminum,8

and organomagnesium9

nucleophiles, however, more recently chiral rhodium/diene systems have been shown to

construct asymmetric quaternary stereocenters with functional group-tolerant organoboron

nucleophiles.10

While rhodium systems are highly developed and exhibit a wide substrate scope,

the high cost of the catalyst precursors and oxygen sensitivity of the reactions are undesirable.

Despite progress in palladium-catalyzed conjugate additions for the formation of tertiary

52

stereocenters,11

no conditions were amenable to the synthesis of even racemic quaternary

centers until Lu and coworkers disclosed a dicationic, dimeric palladium/bipyridine catalyst in

2010.12

However, it was not until our recent report that a palladium-derived catalyst was

employed to generate an asymmetric quaternary stereocenter via conjugate addition chemistry.13

Scheme 3.1. Asymmetric conjugate addition with (S)-t-BuPyOx ligand.

We employed a catalyst derived in situ from Pd(OCOCF3)2 and a chiral pyridinooxazoline

(PyOx) ligand,14

(S)-t-BuPyOx (ligand 1, Scheme 3.1). This catalyst facilitates the asymmetric

conjugate addition of arylboronic acids to -substituted enones in high yield and good

enantioselectivity. Importantly, this reaction is highly tolerant of air and moisture, and the chiral

ligand, while not yet commercially available, is easily prepared.15

Initial results with the

Pd/PyOx system were reported rapidly due to concerns over competition in the field. Indeed,

recent publications prove palladium-catalyzed conjugate addition to be a burgeoning field of

research.16

After the initial disclosure, we observed that in addition to catalyzing conjugate

additions to 5-, 6-, and 7-membered enones, the Pd/PyOx catalyst successfully reacted with

chromones and 4-quinolones.17

Intrigued by the broad substrate scope and operational simplicity

of this highly asymmetric process, we conducted a thorough study to optimize the reaction

conditions, including measures to reduce the catalyst loading, lower the reaction temperature,

53

and further generalize the substrate scope. We also performed mechanistic and computational

investigations toward elucidating the catalytic cycle, active catalyst species, and the

stereoelectronic effects on enantioselectivity of this reaction.

3.3 Results

3.3.1 Effects of Water on Catalyst Turnover

In our initial report,13

we were able to demonstrate that the addition of up to 10 equivalents

of water had no deleterious effect. Despite this, water was not considered as an important

additive in the initially reported conditions because the stoichiometric arylboronic acid was

believed to be a sufficient proton source to turn over the catalyst. In considering the overall

reaction scheme, a more precise analysis of the mass balance of the reaction led us to reconsider

the importance of water as a participant in the overall transformation (Scheme 3.2a).

These considerations proved to be essential during the scale up of the reaction. Attempts to

use the original conditions (with no water added) failed to convert enone 2 efficiently, generating

the desired ketone (3) only in moderate yield (Scheme 3.2b). We reasoned that when the reaction

is performed on a small scale under ambient atmosphere the moisture present in the air and on

the glassware could be sufficient to drive the reaction to completion. On a larger scale, however,

where a more significant quantity of water was necessary, this was no longer true. Gratifyingly,

upon the addition of as little as 1.5 equivalents of water to the reaction mixture, both reactivity

and the enantioselectivity were restored (Scheme 3.2c), affording ketone 3 in high yield and ee.

54

Scheme 3.2. (a) Examination of reaction mass balance. (b) Absence of water prohibits scale-up.

(c) Addition of water facilitates larger scale reactions.

We next sought to measure deuterium incorporation at the carbonyl -position as a method

to determine the source of the proton utilized in reaction turnover. Reactions were performed

substituting deuterium oxide for water and observed by 1H and

2H NMR analysis (Figure 3.1).

Using phenylboronic acid, the reaction afforded ketone 3 in similar yield and enantioselectivity

(Figure 3.1a). Likewise, substitution of phenylboroxine ((PhBO)3) for phenylboronic acid and

deuterium oxide for water (Figure 3.1b) resulted in identical yield, albeit with slightly depressed

ee observed in ketone 3. Analysis of ketone 3 by 1H NMR (Figure 3.1c) showed significant

deuterium incorporation at the -position of the carbonyl, even in the presence of phenylboronic

acid.18

As expected, a higher degree of deuterium incorporation was observed in the reaction

where phenyl boroxine was substituted for the boronic acid, however, the similar level of

incorporation in both experiments suggested that the deuterium oxide was the agent assisting

reaction turnover regardless of the use of protic or aprotic boron reagent.

O

P h

O

H

H O B (O H )2P h B (O H )2 H 2 O

OO (S ) - t -B u P y O x (6 m o l % )

P d (O C O C F 3 )2 (5 m o l % )

P h B (O H )2 (2 e q u iv )

C lC H 2 C H 2 C l, 6 0 °C , 1 2 h

5 e q u iv H 2 OP h

9 7 % y ie ld , 9 1 % e e

2 2 m m o l s c a le

OO (S ) - t -B u P y O x (6 m o l % )

P d (O C O C F 3 )2 (5 m o l % )

P h B (O H )2 (2 e q u iv )

C lC H 2 C H 2 C l, 6 0 °C , 1 2 h

N o A d d e d H 2 OP h

In c o m p le te C o n v e rs io n

o n 1 g s c a le (9 m m o l)

2

2

2 3

3

3

(a )

(b )

(c )

55

Figure 3.1. (a) deuterium incorporation using PhB(OH)2. (b) deuterium incorporation using

(PhBO)3. (c) 1H NMR data measuring deuterium incorporation by integral comparison of

-protons relative to H5, control: treatment of ketone 3 to deuterium incorporation conditions.

3.3.2 Effects of Salt Additives on Reaction Rate

Satisfied with our ability to perform the reaction on scale, we turned our attention toward

improving the catalyst activity. We observed that nearly all previous literature reports regarding

palladium-catalyzed conjugate addition utilized cationic precatalysts featuring

weakly-coordinative anions (PF6–, SbF6

–, BF4

– etc.). We reasoned that the substitution of the

trifluoroacetate counterion with a less coordinative species could lead to an increase in reaction

rate. With this goal in mind, we examined a series of salt additives containing weakly

coordinative counterions. We viewed the strategy for the in situ generation of the catalyst as the

more practical and operationally simple alternative to the design, synthesis, and isolation of a

new dicationic palladium precatalyst.

OO

(S ) - t -B u P y O x (6 m o l % )

P d (O C O C F 3 )2 (5 m o l % )

N H 4 P F 6 (3 0 m o l % )

C lC H 2 C H 2 C l, 6 0 °C , 1 2 h

D 2 O

P h

B (O H )2

OO (S ) - t -B u P y O x (6 m o l % )

P d (O C O C F 3 )2 (5 m o l % )

N H 4 P F 6 (3 0 m o l % )

C lC H 2 C H 2 C l, 6 0 °C , 1 2 h

D 2 O

P h

B

OB

O

BO P hP h

P h

(9 9 % y ie ld )

(9 6 % y ie ld )

P h

O

H 1

H 2

H 3

H 4

e n try

b o ro n ic a c id

b o ro x in e

0 .6 7 0 .5 2 0 .8 8 1 .0 0

0 .5 2 0 .4 1 0 .6 6 1 .0 0

H 2 H 3 + H 4 H 5H 1

in te g ra l re la t iv e to H 5

(a )

(b )

(c )

2

2

3

3

D

D

H 5

c o n tro l 0 .9 4 0 .9 0 1 .6 7 1 .0 0

56

We investigated a number of salt additives to test this mechanistic hypothesis (Table 3.1).

Coordinating counterions like chloride (entry 1) shut down reactivity. Pleasingly, as per our

hypothesis, weakly coordinating counterions with sodium cations (entries 2–4) facilitated swift

reaction, albeit with depressed ee. Tetrabutylammonium salts (entries 5–6) encountered slow

reaction times, but good enantioselectivity. Sodium tetraphenylborate (entry 7), however, failed

to deliver appreciable quantities of the quaternary ketone 3, as rapid formation of biphenyl was

observed. Ammonium salts (entires 8–9) provided the desired blend of reaction rate and

enentioselectivity. We concluded that the hexafluorophosphate anion (entry 9) gave the optimal

combination of short reaction time with minimized loss of enantioselectivity.

Table 3.1. Effect of salt additives on reaction rate.a

aConditions: phenylboronic acid (0.5 mmol), 3-methylcyclohexen-2-one (0.25 mmol), water (5

equiv), additive (30 mol %), Pd(OCOCF3)2 (5 mol %) and (S)-t-BuPyOX (6 mol %) in

ClCH2CH2Cl (1 mL) at 40 °C . bGC yield utilizng tridecane standard.

cee was determined by

chiral HPLC. dReaction checked at 83% conversion as determined by GC analysis.

e n trya

t im e (h ) y ie ld (% )b

e e (% )c

1

2

3

2 4 tra c e - - -

2 4 tra c e - - -

6 9 7 8 7

4

5

6

7

2 4 9 8 9 0

5 9 9 8 1

2 4 9 5 8 8

a d d it iv e

n - (B u )4 P F 6

8 8 1d 8 8

N a C l

N a B P h 4

N a P F 6

N a S b F 6

N a B F 4

n - (B u ) 4 B F 4

O O

P h

(S ) - t -B u P y O X , P h B (O H )2

P d (O C O C F 3 )2

a d d it iv e

C lC H 2 C H 2 C l, 4 0 °C

2 3

1 5 9 3 8 98 N H 4 B F 4

1 2 9 6 9 19 N H 4 P F 6

57

Based on our previous observations regarding the beneficial nature of water as an additive,

we next explored the combined effect of water and hexafluorophosphate counterions. We found

addition of both water and ammonium hexafluorophosphate were the most successful for

increasing reactivity (Table 3.2). Water alone is insufficient to alter reactivity (entry 1), though

the use of water with 30 mol % ammonium hexafluorophosphate greatly reduced the reaction

time (entry 2) to only 1.5 hours with minimal effect on yield or ee. Furthermore, this

combination of additives allowed the reaction to proceed at temperatures as low as 25 °C with 5

mol % palladium and 6 mol % ligand, and lowering of catalyst loadings to only 2.5 mol % of

palladium and 3 mol % ligand at 40 °C (entry 3). We determined that optimal conditions for

the reaction with lower catalyst loading to be 5 equivalents of water, 30 mol % ammonium

hexafluorophosphate at 40 °C (entry 4), conditions that reproduce the original result at milder

temperature and lower catalyst loadings. The reaction was extraordinarily tolerant of the amount

of water, with both 10 (entry 5) and 20 (entry 6) equivalents of water having minimal effect on

the yield or ee. Loadings of ammonium hexaflurophosphate can be as low as 5 mol % (entry 7)

or 10 mol % (entry 8) with reactions completed in 24 hours. Stoichiometric additive (entry 9)

gave no additional benefit (entry 4). Thus, we optimized the additive amounts to be 30 mol %

ammonium hexafluorophosphate and 5 equivalents of water.

Table 3.2. Effect of water and NH4PF6 on reaction rate.

58

aConditions: phenylboronic acid (1.0 mmol), 3-methylcyclohexen-2-one (0.5 mmol), water,

NH4PF6, Pd(OCOCF3)2 and (S)-t-BuPyOx in ClCH2CH2Cl (2 mL). bIsolated yield.

cee was

determined by chiral HPLC.

Though increased rates were observed at 60 °C, the newly-found ability to perform reactions

at 40 °C promoted superior reactivity of many substrates (Table 3.3). In fact, many substrates

that exhibited high enantioselectivities under the original 60 °C reaction conditions suffered from

poor yields. Reacting these substrates at 40 °C with the addition of ammonium

hexafluorophosphate and water promoted significantly higher isolated yields. Arylboronic acids

containing halides, such as m-chloro- (4a) and m-bromophenylboronic acid (4b) reacted with

good enantioselectivity, but each substrate was originally marred by low yield using our original

conditions. However, when reacted under the newly optimized reaction conditions, the isolated

yield for the addition of chlorophenylboronic acid increased from 55% to 96% and for

bromophenylboronic acid from 44% to 86%. Even m-nitroboronic acid (4c) reacted with higher

isolated yield. Notably, some ortho-substituted boronic acids, such as o-fluorophenylboronic

OO

e n trya

t im e (h ) y ie ld (% )b

e e (% )c

1

2

3

1 0

5

1 2 9 9 9 1

te m p

(°C )

6 0

1 .5 9 3 8 86 0

1 2 9 5 9 14 0

P d

(m o l % )

5

4

5

6

7

3 6 9 8 9 12 5

1 2 9 6 8 94 0

1 2 9 5 9 04 0

8

5 2 4 9 3 9 04 0

H 2 O

(e q u iv .)

N H 4 P F 6

(m o l % )

- - -

3 0

5 3 0

5 3 0

1 0 3 0

2 0 3 0

5

9

5 2 4 9 3 9 24 01 0

5 1 2 9 7 8 84 01 0 0

P h

lig a n d

(m o l % )

5

5

6

6

6

2 .5 3

2 .5 3

2 .5 3

2 .5 3

2 .5 3

2 .5 3

(S ) - t -B u P y O x , P h B (O H )2

P d (O C O C F 3 ) 2

H 2 O , N H 4 P F 6

C lC H 2 C H 2 C l

2 3

59

acid (4d), reacted more successfully under the milder reaction conditions, leading to increased

isolated yield of 70%.

Table 3.3. Increased reaction yields with different arylboronic acid substrates under new reaction

conditions.a

aBlue font: reported yield and ee of 5 in the absence of NH4PF6 and water with reactions

performed at 60 °C; red font: yield and ee of 5 with additives. Conditions: boronic acid (1.0

mmol), 3-methylcyclohexen-2-one (0.5 mmol), NH4PF6 (30 mol %), water (5 equiv.),

Pd(OCOCF3)2 (5 mol %) and (S)-t-BuPyOx (6 mol %) in ClCH2CH2Cl (2 mL) at 40 °C . Isolated

yield reported, ee was determined by chiral HPLC.

3.3.3 Non-Linear Effect Correlation of Catalyst and Product Enantioenrichment

Despite optimization of catalytic conditions for this highly enantioselective process, we were

unsure of the nature of the active catalyst. For example, some rhodium conjugate addition

systems have been shown to involve trimeric ligand/metal complexes.19

Furthermore, Lu and

coworkers reported the use of the palladium dimer [(bpy)Pd(OH)]2•2BF4 as a precatalyst for

conjugate addition.12

We aimed to rule out the in situ formation and kinetic relevance of such

dimers in our system. In seeking to support our hypothesized monomeric ligand-metal complex,

4 d4 c

B r N O 2

5 5 % Y ie ld 9 7 % e e

(H O )2 B (H O )2 B (H O )2 BC l (H O )2 B

9 6 % Y ie ld 9 6 % e e

4 4 % Y ie ld 8 6 % e e

8 4 % Y ie ld 8 4 % e e

4 0 % Y ie ld 9 2 % e e

8 1 % Y ie ld 9 1 % e e

3 2 % Y ie ld 7 7 % e e

7 0 % Y ie ld 7 7 % e e

F

4 a 4 b

OO(S ) - t -B u P y O x (6 m o l % )

P d (O C O C F 3 )3 (5 m o l % )

N H 4 P F 6 , H 2 O

C lC H 2 C H 2 C l

4 0 °C , 1 2 hB (O H ) 2

RR

2 4 5

60

we performed a non-linear effect study to determine the relationship between the ee of the ligand

and the ee of the generated product.20

The endeavor was to exclude dimeric (ML)2 complexes

from kinetic relevance, clarifying the monomeric nature of the active catalyst.21

Five reactions

were performed using a catalyst with different level of enantiopurity (racemic, 20%, 40%, 60%

and 80% ee), and the obtained enantioselectivities were plotted against ee of the catalyst mixture

(Figure 3.2). The obtained data clearly demonstrates that a non-linear effect is not present, and

this observation strongly supports the action of a single, monomeric (ML)-type Pd/PyOx catalyst

as the kinetically relevant species.21

While the precise nature of the active catalyst species is

unknown, isolated (PyOx)Pd(OCOCF3)2 serves as an identically useful precatylst, delivering

ketone 3 in 99% yield and 92% ee.

Figure 3.2. Determination of linearity between catalyst ee and product ee.

61

3.3.4 Computational Investigations of the Reaction Mechanism

Despite the results of the non-linear effect study agreeing with the proposed monomeric

Pd/PyOx catalyst, no formal exploration of the mechanism of this transformation has been

reported. Our initial hypothesis concerning the mechanism of the Pd/PyOx-catalyzed asymmetric

conjugate addition were well informed by the seminal work of Miyaura,22

however, the

heterogeneous nature of the reaction medium, undefined nature of the precise catalyst,23

and

complicating equilibrium of organoboron species make kinetic analysis and thorough

mechanistic study extremely challenging.24

Previously, we performed density functional theory (DFT) calculations to investigate the

mechanism of palladium-catalyzed conjugate addition of arylboronic acids to enones, explicitly

studying a catalytic palladium(II)/bipyridine system in MeOH solvent similar to that developed

by Lu.12,25,26

Calculations indicated that the mechanism involves three steps: transmetallation,

carbopalladation (i.e. alkene insertion), and protonation with MeOH. Monomeric cationic

palladium complexes are the active species in the catalytic cycle. The carbopalladation is

calculated to be the rate- and stereoselectivity- determining step (Scheme 3.3). Now, we have

performed DFT investigations on the catalytic cycle of reactions with the Pd/PyOx manifold and

the effects of substituents and ligand on reactivity and enantioselectivity. The calculations were

performed at the theoretical level found satisfactory in our previous study of the Pd/bipyridine

system. Geometries were optimized with BP8627

and a standard 6-31G(d) basis set (SDD basis

set for palladium). Solvent effects were calculated with single point calculations on the gas phase

geometries with the CPCM solvation model in dichloroethane. All calculations were performed

with Gaussian 03.28

62

Scheme 3.3. Enantioselectivity-determining step in asymmetric conjugate addition of

arylboronic acids to cyclic enones.

1O O

B (O H )2

N N

O

t -B u

P d (O C O C F 3 )2

C lC H 2 C H 2 C l, 6 0 °C

2

E n a n t io s e le c t iv ity -D e te rm in in g S te p

3

N

P d

N

O O

H

QuickTim e?and a

PNG decompressor

are needed t o see t his pict ure.

63

Figure 3.3. Computed potential energy surface of the catalytic cycle (shown in black), the

alternative direct nucleophilic addition pathway (via 9-ts, shown in blue), and the isomeric

carbopalladation pathway (via 16-ts, shown in green).

The computed potential energy surface for the catalytic cycle is shown in Figure 3.3. To

simplify the computations of the mechanisms, a model ligand, in which the t-Bu group on the

t-BuPyOx ligand was replaced by H, was used in the calculations of the mechanisms and the full

ligand was used in the calculations of enantioselectivities which will be discussed below.

Calculations on the reaction mechanism with the full ligand scaffold, however, generated a

similar reaction diagram, and the rate- and stereo- determining steps were unchanged (not

shown). The first step involves transmetallation of cationic Pd(II)-phenylborate complex 6 to

generate a phenyl palladium complex. Transmetallation requires a relatively low free energy

barrier of 15.6 kcal/mol (7-ts) with respect to complex 6 and leads to a phenyl palladium

64

complex (8). Complex 8 undergoes ligand exchange to form a more stable phenyl

palladium-enone complex 12, in which the palladium binds to the enone oxygen atom. Complex

12 isomerizes to a less stable complex 13 and then undergoes carbopalladation of the enone

(14-ts) to form the new carbon–carbon bond. The carbopalladation step requires an activation

free energy of 21.3 kcal/mol (12 14-ts), and is the stereoselectivity-determining step. The

regioisomeric carbopalladation transition state 16-ts requires 5.6 kcal/mol higher activation free

energy than 14-ts, indicating the formation of the -addition compound 17 is unlikely to occur.

Coordination of one water molecule to 15 leads to a water-palladium enolate complex 18, and

finally facile hydrolysis of 18 via 19-ts affords product complex 20. Liberation of the product 3

from 20 and coordination with another molecule of phenyl boronic acid regenerates complex 6 to

complete the catalytic cycle. The computed catalytic cycle demonstrates some similarities with

the Pd/bipyridine system in our previous computational investigation, which also involves

monomeric cationic palladium as the active species and a catalytic cycle of transmetallation,

carbopalladation, and protonation (with MeOH instead of H2O).

We also considered an alternative pathway involving direct nucleophilic attack of the phenyl

boronic acid at the enone while the Pd catalyst is acting as a Lewis acid to activate the enone and

directs the attack of the nucleophile (9-ts, Figure 3.3). This alternative pathway requires an

activation free energy of 58.9 kcal/mol, 43.3 kcal/mol higher than the transmetallation transition

state 7-ts. Thus, this alternative pathway was excluded by calculations.

65

3.3.5 Experimental and Computational Investigations of the Enantioselectivities

With the aforementioned optimized reaction conditions and computational elucidation of the

mechanism and stereoselectivity-determining transition states, we explored the effects of ligand

and substrate on enantioselectivities by both experiment and computations. The

enantioselectivity-determining alkene insertion step involves a four-membered cyclic transition

state, which adopts a square-planar geometry. When a chiral bidentate ligand, such as

(S)-t-BuPyOx, is employed, there are four possible isomeric alkene insertion transition states.

The 3D structures of the alkene insertion transition states in the reaction of

3-methyl-2-cyclohexenone with (S)–t-BuPyOx ligand are shown in Figure 3.4. In 1-TS-A and

1-TS-B, the phenyl group is trans to the chiral oxazoline on the ligand, and in 1-TS-C and

1-TS-D, the phenyl group is cis to the oxazoline. 1-TS-A, which leads to the predorminant

(R)-product, is the most stable as the t-Bu group is pointing away from other bulky groups.

1-TS-C leads to the same enantiomer, but with an activation enthalpy 2.6 kcal/mol higher than

1-TS-A. The difference is likely to result from steric effects between the t-Bu on the ligand and

the phenyl group, as indicated by the C-H and C-C distances labeled in Figure 3.4. 1-TS-B and

1-TS-D lead to the minor (S)-product, which are ~3 kcal/mol less stable than 1-TS-A as a result

of the repulsions between the t-Bu on the ligand and the phenyl group. In 1-TS-B, the

cyclohexenone ring is syn to the t-Bu group on the ligand. The shortest H–H distance between

the ligand and the enone is 2.30 Å, suggesting some steric repulsions. In contrast, no

ligand-substrate steric repulsions are observed in 1-TS-A, in which the cyclohexenone is anti to

the t-Bu. 1-TS-A is also stabilized by a weak hydrogen bond between the carbonyl oxygen and

the hydrogen geminal to the t-Bu group on the oxazoline. The O–H distance is 2.16Å. Therefore,

66

the enthalpy of 1-TS-B is 2.3 kcal/mol higher than that of 1-TS-A. This corresponds to an ee of

94%, which is very similar to the experimental observation (93%). Enantioselectivities were

computed from relative enthalpies of the transition states. The selectivities computed from Gibbs

free energies are very similar and are given in the SI.

We then investigated the effects of substituents on the ligand, in particular, at the 4 position

of the oxazoline. The activation enthalpies of four alkene insertion pathways and the computed

and experimental ee for the reaction of 3-methyl-2-cyclohexenone and phenyl boronic acid are

summarized in Table 3.4. The t-Bu substituted PyOx ligand is found to be the optimum ligand

experimentally (Table 3.4, entry 1). Replacing t-Bu with smaller groups, such as i-Pr, i-Bu, or

Ph, dramatically reduces the ee.

The bulky t-Bu substituent on the ligand is essential not only to discriminate the

diastereomeric transition states 1-TS-A and 1-TS-B, but also fix the orientation of the ligand to

point the chiral center cis to the cyclohexenone. The energy difference between 1-TS-C and

1-TS-D, in which the chiral center on the ligand is trans to the cyclohexenone, is diminished.

67

Figure 3.4. The optimized geometries of transition states in the enantioselectivity-determining

alkene insertion step of the reaction of 3-methyl-2-cyclohexenone and phenyl boronic acid with

(S)–t-BuPyOx ligand. Selected H–H, C–H, and O–H distances are labeled in Å.

When the (S)-i-PrPyOx ligand is used, the alkene insertion transition states with phenyl

trans to the oxazoline (2-TS-A and 2-TS-B) are also preferred. Thus, the enantioselectivity is

determined by the energy difference between 2-TS-A and 2-TS-B. The (R)-product (via 2-TS-A)

is favored with a computed ee of 67%, slightly higher than the experimental ee (40%). The

optimized geometries of 2-TS-A and 2-TS-B are shown in Figure 3.5 and the activation energies

of all four transition states are shown in Table 3.4, entry 2. The i-Pr substituted ligand manifests

via similar steric effects to (S)-t-BuPyOx, with, as expected, slightly weaker steric control. The

lower enantioselectivity is attributed to the weaker steric repulsions between the i-Pr and the

cyclohexenone in 2-TS-B than those with the t-Bu in 1-TS-B. The shortest distance between the

68

hydrogen atoms on the ligand and the cyclohexenone is 2.35 Å in 2-TS-B, slightly longer than

the H–H distance in 1-TS-B (2.30 Å). Less steric repulsions with the (S)-i-PrPyOx lead to 2.8

kcal/mol lower activation barriers for 2-TS-B compared to 1-TS-B. The ligand steric effects on

the activation energies of the major pathway TS-A are smaller; the i-Pr substituted 2-TS-A is

only 0.6 kcal/mol more stable than the t-Bu substituted 1-TS-A.

Similarly, when the (S)-i-BuPyOx or (S)-PhPyOx ligands are used, the enantioselectivity is

further decreased to 0.8 kcal/mol (52% ee) for (S)-i-BuPyOx and 1.0 kcal/mol (65% ee) for

(S)-PhPyOx. (Table 3.4, entries 3 and 4). These results agree well with the experimental trend.

Table 3.4. Activation energies and enantioselectivities of alkene insertion with (S)-t-BuPyOX,

(S)-i-PrPyOX, (S)-i-BuPyOX, and (S)-PhPyOx ligands.

TS R1 H

‡ a ee

b

TS-A TS-B TS-C TS-D

1 t-Bu 19.2 21.5 21.8 22.2 94%

[93%]

2 i-Pr 18.6 19.7 20.1 20.0 67%

[40%]

3 i-Bu 18.5 19.3 20.4 20.4 52%

[24%]

4 Ph 18.0 19.0 19.6 20.2 65%

[52%] aThe values are activation enthalpies in kcal/mol calculated at the BP86/6-31G(d)–SDD level

and the CPCM solvation model in dichloroethane. bExperimental ee were obtained under

standard conditions and are given in square brackets.

N

P d

N

OO

R 1

H N

P d

N

O

R 1

H

O

N

P d

N

O

O

R 1

HN

P d

N

O

R 1

H

O

n -T S -A n -T S -B n -T S -C n -T S -D

69

Electronically differentiated PyOx ligands were also studied, and the results are summarized

in Table 3.5. Electron-withdrawing or donating groups at the 4-position of the PyOx ligand

showed minimal effects on the activation barriers, and were calculated to have minimal effect on

product ee. With the electron-withdrawing CF3 and the electron-donating OCH3 on the

4-position of the ligand, the activation enthalpies of alkene insertion increase by only 0.3

kcal/mol and 0.1 kcal/mol, respectively. The calculated ee are essentially identical among these

three ligands. Experimentally, depressed ee was observed with both the 4-CF3 and the 4-OCH3

substituted ligands (entries 5 and 6). This confirms that the enantioselectivity is mainly attributed

to the ligand/substrate steric repulsions.

Table 3.5. Remote ligand substituent effects on activation energies and enantioselectivities of

alkene insertion.

TS R

2 H

‡ a ee

b

TS-A TS-B TS-C TS-D

1 H 19.2 21.5 21.8 22.2 94%

[93%]

5 CF3 19.5 21.7 21.8 22.3 93%

[81%]

6 OCH3 19.3 21.5 21.5 22.2 93%

[78%] aThe values are activation enthalpies in kcal/mol calculated at the BP86/6-31G(d)-SDD level

and the CPCM solvation model in dichloroethane. bExperimental ee are given in square brackets.

N

P d

N

OO

H N

P d

N

O

H

O

N

P d

N

O

OHN

P d

N

O

H

O

n -T S -A n -T S -B n -T S -C n -T S -D

R 2R 2

R 2R 2

70

2 - T S - A

H‡

= 1 8 . 6 k c a l/ m o l

2 - T S - B

H‡

= 1 9 . 7 k c a l/ m o l

7 - T S - A

H‡

= 1 6 . 7 k c a l/ m o l

7 - T S - B

H‡

= 1 8 . 6 k c a l/ m o l

(a )

(b )

Figure 3.5. The optimized geometries of transition states in the enantioselectivity-determining

alkene insertion step of the reaction of (a) 3-methyl-2-cyclohexenone and phenyl boronic acid

with (S)-i-PrPyOx ligand; (b) 3-methyl--2-pentenolide and phenyl boronic acid with

(S)-t-BuPyOx ligand. Selected H–H, C–H, and O–H distances are labeled in Å.

The transition state structures shown in Figure 3.5 indicate that the steric control mainly

arises from the repulsion of the C' hydrogens on the cyclohexenone with the ligand. We then

investigated the effects of substitution at the ' position and replacement of the CH2 group with

O. The reactivity and enantioselectivity of the reactions of lactone (Table 3.6, entry 7) and

','-dimethylcyclohexenone (Table 3.6, entry 8) with the (S)-t-BuPyOx ligand were computed.

The enantioselectivity of lactone is predicted to be lower than cyclohexenone. The enthalpy of

7-TS-B is 1.8 kcal/mol higher than that of 7-TS-A, corresponding to an ee of 88%.

Experimentally, the ee of the lactone product is 59%, also significantly lower than that with

cyclohexenone. The optimized geometries of 7-TS-A and 7-TS-B are shown in Figure 3.5b.

71

Replacing the CH2 group with O decreases the ligand–substrate steric repulsion in 7-TS-B is

smaller than that in 1-TS-B. This results in decreased enantioselectivity.

Methyl substitution at the ' position of cyclohexenone increases the steric repulsion with

the t-Bu group on the ligand. Computations predicted increased enenatioselectivity with

','-dimethylcyclohexenone (99% ee, Table 3.6, entry 8).29

However, experimentally, the ee is

comparable with the reaction of 3-methylcyclohexenone.

Table 3.6. Activation energies and enantioselectivities of alkene insertion with substrates varying

at the '-position.

TS X H‡ a

ee b

TS-A TS-B TS-C TS-D

1 CH2 19.2 21.5 21.8 22.2 94%

[93%]

7 O 16.7 18.6 19.4 19.6 88%

[57%]

8 C(CH3)2 17.9 22.0 20.0 20.8 99%

[90%] aThe values are activation enthalpies in kcal/mol calculated at the BP86/6-31G(d)-SDD level

and the CPCM solvation model in dichloroethane. bExperimental ee are given in square brackets.

Experimental isolated yields: TS-1: 99%, TS-7: 49%, TS-8: 9%.

We also considered the electronic effects of arylboronic acids on enantioselectivity (Table

3.7). Computations predicted that para-electron-withdrawing substituents lead to increases in the

activation barrier in alkene insertion, probably due to the electrophilicity of the -carbon of the

enone, and thus are predicted to afford slightly decreased enantioselectivities. Both

N

P d

N

OX O

H N

P d

N

O

H

X

O

N

P d

N

O

XOHN

P d

N

O

H

X

O

n -T S -A n -T S -B n -T S -C n -T S -D

72

para-acetylphenylboronic acid (9-TS-A) and para-trifluoromethylphenylboronic acid (10-TS-A)

are predicted to react with 92% ee. However, both excellent enantioselectivities (96% ee) and

excellent yields (99% isolated yield) are observed experimentally. Thus, the electronic effects

of phenyl substituents on enantioselectivities are minimal, though slightly increased

enantioselectivities are observed experimentally with the use of electron-withdrawing

substituents.

Table 3.7. Activation energies and enantioselectivities of alkene insertion with various boronic

acids.

TS R3 H

‡ a ee

b

TS-A TS-B TS-C TS-D

1 H 19.2 21.5 21.8 22.2 94%

[93%]

9 CH3CO 21.6 23.7 24.5 24.8 92%

[96%]

10 CF3 21.3 23.4 23.6 24.2 92%

[96%] aThe values are activation enthalpies in kcal/mol calculated at the BP86/6-31G(d)-SDD level

and the CPCM solvation model in dichloroethane. bExperimental ee are given in square brackets.

Experimental isolated yields: TS-1: 99%, TS-9: 1%, TS-10: 99%, TS-11: 99%.

Table 3.8. Enantioselectivity trends in pyridinooxazoline and related ligand frameworks.a

N

P d

N

OO

H N

P d

N

O

H

O

N

P d

N

O

OHN

P d

N

O

H

O

n -T S -A n -T S -B n -T S -C n -T S -D

R 3R 3

R 3

R 3

73

aConditions: 3-methylcyclohexen-2-one (0.25 mmol), phenylboronic acid (0.5 mmol),

ClCH2CH2Cl (1 mL). Yields are isolated yields, ee determined by chiral HPLC.

Permutations of the pyridinooxazoline ligand framework corroborate the calculated data and

suggest that a number of factors affect enantioselectivity. First, the steric demand of the chiral

group on the oxazoline greatly impacts the observed enantioselectivity in the reaction (Table

3.8). Only t-BuPyOx (1) yields synthetically tractable levels of enantioselectivity, while the less

sterically demanding i-PrPyOx (21), PhPyOx (22), and i-BuPyOx (23) all exhibit greatly

diminished selectivity. Oxazoline substitution patterns also affect enantioselectivity.

Substitution at the 4-position appears to be required for high selectivity, as substitution at the

5-position yields practically no enantioselectivity (ligand 24). Electronic variation in the PyOx

framework was observed to have a large effect on the rate of the reaction but, disappointingly,

N

N

O

i -P r

N

N

O

P h

N

N

O

P h

2 1

9 9 % y ie ld

4 0 % e e

2 2

9 9 % y ie ld

5 2 % e e

2 4

9 9 % y ie ld

2 % e e

N

N

O

t -B u

N

N

O

i-P r2 7

1 3 % y ie ld

2 0 % e e

2 8

1 4 % y ie ld

0 % e e

O

B (O H )2

O

lig a n d (6 m o l % )

P d (O C O C F 3 ) 2 (5 m o l % )

C lC H 2 C H 2 C l, 6 0 °C , 1 2 h

+

N

N

O

i-B u

2 3

9 5 % y ie ld

2 4 % e e

N

N

O

t -B u

1

9 9 % y ie ld

9 3 % e e

N

N

O

t -B u

2 5

9 9 % y ie ld

8 1 % e e

N

N

O

t -B u

2 6

9 6 % y ie ld

7 8 % e e

H C F 3 O M e

P h

2 3

74

led to depressed stereoselectivity. CF3-t-BuPyOx (25) afforded the conjugate addition product in

99% yield and 81% ee. Surprisingly, MeO-t-BuPyOx (26) afforded the product in similar yield

and only 78% ee. Finally, substitution at the 6-position of the pyridine (ligands 27 and 28)

greatly diminished both reactivity and selectivity, perhaps due to hindered ligand chelation to

palladium.

Thus, we have concluded that enantioselectivity is controlled by the steric repulsion between

the substituent on the chiral pyridinooxazoline ligand and the cyclohexyl ring. The bulkier t-Bu

substituent on the (S)-t-BuPyOx ligand leads to greater enantioselectivity than the reactions with

(S)-i-PrPyOx or (S)-PhPyOx. Similarly, substrates with less steric demand adjacent the carbonyl

exhibit lower enantioselectivities; for example the reaction of a lactone substrate (Table 3.6,

TS-7) yields lower enantioselectivity due to smaller repulsions between the lactone oxygen and

the t-Bu group.

3.3.6 Experimental Investigation of Arylpalladium(II) Intermediates and Formation of the

Key C–C Bond

Experiments aiming to corroborate the calculated mechanism have been performed. We

sought to observe the formation of the key C–C bond between an arylpalladium(II) species and

the enone substrate in the absence of exogenous phenylboronic acid. Complexes 29 were

synthesized as an intractable mixture of isomers, and were treated with AgPF6 in situ to generate

the [(PyOx)Pd(Ph)]+

cation. Gratifyingly, complexes 29 serve as a competent precatalyst at 5 mol

% loading, and affords ketone 3 in 96% yield and 90% ee (Table 3.9, entry 1). Varying the

amount of complex utilized in proportion with phenylboronic acid, however, leads to significant

75

production of biphenyl above 5 mol % (Table 3.9, entries 2–5). Utilizing even 25 mol % (entry

2) resulted in significant increase in biphenyl production (16% yield), and reduction in yield of

the desired ketone 3 to 79%. Increasing complex loadings to 45 and 65 mol % (entries 3 and 4)

leads to negligible production of ketone 3, and nearly quantitative formation of biphenyl relative

to catalyst loading. Furthermore, attempts to use the complex as a stoichiometric reagent in the

place of PhB(OH)2 lead to no observed product (entry 5), and exclusive formation of biphenyl.

We hypothesize that quantitative generation of the reactive arylpalladium cation intermediate in

high relative concentration leads quickly to disproportionation and formation of biphenyl and

palladium(0). Omission of the AgPF6 in favor of 30 mol % NH4PF6 leads to isolation of only

22% yield of the conjugate addition product (entry 6). Finally, a control experiment demonstrates

that AgPF6 is incapable of catalyzing the reaction itself (entry 7).30

This control further supports

the computational results, which indicate a transmetallation-based mechanism as opposed to a

Lewis acid-catalyzed pathway.

Table 3.9. Arylpalladium(II) catalyzed conjugate addition.a

OO

C o m p le x 2 9 , A g P F 6

H 2 O (5 e q u iv .)

C lC H 2 C H 2 C l, 4 0 °C , 1 2 h P h

B (O H ) 2

e n try c o m p le x 2 9 (m o l % ) A g P F 6 (m o l % ) P h B (O H )2 (m o l % )

y ie ld (% )

3 b ip h e n y l

1

2

3

4

5

5

2 5

4 5

6 5

1 0 5

1 0

5 0

9 0

1 3 0

2 1 0

7 0 2 0 2 0 0

0

4 0

6 0

8 0

1 0 0 9 6

7 9

1 1

5

0

3

1 6

5 3

5 8

8 7

0 0

2 3

6b

5 0 2 2 02 0 0

76

a Conditions: 3-methylcyclohexen-2-one (0.25 mmol), pheylboronic acid (equiv as stated),

ClCH2CH2Cl (1 mL). Yields are isolated yields, ee determined by chiral HPLC. bReaction

performed with 30 mol % NH4PF6

Concerned by our inability to observe consistent product formation at 40 °C, we sought

alternative experimental verification that a putative cationic arylpalladium(II) species is capable

of reacting to form conjugate addition products. Thus, we performed the stoichiometric reaction

of the isomeric phenyl palladium iodide complexes (29) with AgPF6 and 3-methylcyclohexenone

at cryogenic temperatures, allowing the mixture to warm slowly to room temperature before

quenching with trifluoroethanol.31

We observed both conjugate addition product and biphenyl,

with the desired adduct (3) isolated in 30% yield (Scheme 3.4a). Curiously, the conjugate

addition adduct was isolated in only 55% ee. We considered that the isomeric mixture of phenyl

palladium iodide isomers formed configurationally stable cationic species at cryogenic

temperature.32

Repeating the experiment, we substituted triphenylphosphine for

methylcyclohexenone and observed the reaction at low temperature utilizing 1H and

31P NMR

(Scheme 3.4b). Indeed, two 31

P signals corresponding to phosphine-bound palladium(II) species

(30) were observed at 28 and 34 ppm. No isomerization was observed upon warming the

isomeric mixture to room temperature.33

Indeed, we were able to isolate and characterize the

mixture of arylpalladiumphosphine cations. With evidence for the configurationally stable

arylpalladium cation, we rationalized the observed 55% ee for the direct reaction of mixture 29

with methylcyclohexenone corresponds directly to the isomeric ratio of the complex: a ratio of

1.3:1 represents a 56:44 ratio of isomers. Assuming that the major isomer reacts with 92% ee,

and the minor isomer reacts with no stereoselectivity to give racemic products,34

a net

stereoselectivity of 53.7% ee would be predicted for the product. Presuming configurational

77

stability of the arylpalladium(II) cation as suggested by the triphenylphosphine trapping

experiment, the diminished enantioselectivity observed in this result is unsurprising. Thus, we

have obtained experimental verification of the key C–C bond forming event of the Pd/PyOx

asymmetric conjugate addition occurring from a quantitative generated arylpalladium(II) cation

in the presence of enone substrate and absence of an exogenous arylboronic acid.

Scheme 3.4. (a) Direct formation of C–C bond from arylpalladium(II) cation. (b)

Triphenylphosphine trapping experiments demonstrates configurational stability of

arylpalladium cation.

3.4 Summary and Discussion

Several experimental results have been described that support the DFT-calculated

mechanism for the Pd/PyOx-catalyzed asymmetric conjugate addition of arylboronic acids to

N

P d

I

N

O

t -B u

1 . 3 -m e th y lc y c lo h e x e n o n e ,

A g P F 6 , C H 2 C l2 , – 7 8 °C

2 . C F 3 C H 2 O H , 2 0 °C

O

P h

3

3 0 % y ie ld

5 5 % e e

N

P d

I

N

O

t -B u

N

P d

P P h 3

N

O

t -B u

A g P F 6 , P P h 3

C D 2 C l2 , – 7 8 °C

+ P F 6–

(a )

(b )

2 9

m ix o f is o m e rs

3 0

T w o 3 1

P S ig n a ls O b s e rv e d

N

I

P d

N

O

t -B u

2 9

m ix o f is o m e rs

N

I

P d

N

O

t -B u

N

P h 3 P

P d

N

O

t -B u

+

P F 6–

78

cyclic enone (Figure 3.4); specifically, the role of the palladium catalyst has been addressed.

Calculations and previous experimental work by Miyaura on palladium-catalyzed conjugate

addition suggest that a transmetallation event occurs to transfer the aryl moiety from the boronic

acid species to the palladium catalyst.4

Our calculations indicated that the Pd/PyOx system

operates under a similar manifold, and demonstrated a significant energy difference

(transmetallation is favored by over 40 kcal/mol, Figure 3.4) between the potential roles of the

palladium catalyst, suggesting that the role of the palladium species is not simply that of a Lewis

acid. Furthermore, it is difficult to rationalize the high degree of enantioselectivity imparted by

the chiral palladium catalyst if it is assumed that the metal acts only as a Lewis acid and is not

directly mediating the key C–C bond-forming step.35

Finally, a number of Lewis- and acidic

metal salts were substituted for palladium with no product observed, further indicating that

palladium-catalyzed conjugate addition is likely not a Lewis acid-catalyzed process.36

While highly Lewis-acidic, the role of cationic palladium(II) is to provide a vacant

coordination site for the enone substrate to approach the catalyst. The presence of a cationic

intermediate is further supported by the observed rate acceleration of non-coordinating anionic

additives such as PF6– and BF4

– salts (Table 3.1). Conversely, the addition of coordinating

anions, such as chloride, inhibited the reaction (Table 3.1, entry 1). This counterion effect was

evident even from choice of palladium(II) precatalysts.13

For instance, Pd(MeCN)2Cl2 was not

a suitable precatalyst, nor were any palladium(II) halides. Additionally, Pd(OAc)2 only afforded

modest yield of conjugate adducts, whereas the less coordinating counterion present in

Pd(OCOCF3)2 afforded complete conversion.

Satisfied that the palladium catalyst was not acting as a Lewis acid, we next sought to

demonstrate the viability of the hypothesized arylpalladium(II) species as a catalyst (Table 3.9,

79

entries 1–5). While serving as a suitable precatalyst under reaction conditions, arylpalladium(II)

mixture 29 failed to facilitate the conjugate addition reaction when used in stoichiometric

quantities in the presence of AgPF6 (entry 5). We rationalize this outcome to be the result of the

highly reactive nature of the quantitatively generated arylpalladium(II) cation. Significant

biphenyl formation- occuring even under dilute conditions representative of the catalytic reaction

itself- suggest that alternative reaction pathways, such as disproportionation, readily out compete

the desired insertion reaction. This reactive nature of the arylpalladium(II) cation led us to

propose performing the stoichiometric reactions at low temperatures (Scheme 3.4). Successfully

demonstrating that the key C–C bond could be generated, albeit in modest yield, from

(PyOx)Pd(Ph)(I) (29) corroborates both the precise role of the arylpalladium(II) cation in the

calculated mechanism as well as the transition state put forth in the calculations. However, the

modest yield of this process and requisite cryogenic temperatures prompted, again, consideration

of the role of the arylboronic acid.37

Calculations suggest that the presence of boronic acids as

Lewis basic entities may serve to stabilize these reactive intermediates under the reaction

conditions (Figure 3.3, cationic arylpalladium 8).38

This suggestion is consistent with the

successful use of arylpalladium(II) mixture 29 as a precatalyst in the presence of arylboronic

acids (Table 3.9, entry 1).

Lastly, water (or another proton source) is required for efficient turnover of the reaction.

Considerations of reaction scale (Scheme 3.2) and deuterium incorporation experiments (Figure

3.1) suggest that water is the likely protonation agent, despite numerous other protic sources in

the heterogenous reaction mixture. Attempts to use other, miscible proton sources (MeOH,

phenol, t-BuOH, etc.) typically resulted in 10–15% less enantioselectivity observed.39

2,2,2-Trifluoroethanol can be successfully substituted for water with minimal loss of

80

enantioselectivity, however it affords no supplementary benefit and water is the preferred

additive for all reactions.24

Water in combination with NH4PF6 serves to facilitate milder

reaction conditions (Table 3.2), which in turn greatly increases the synthetic scope with respect

to challenging arylboronic acid nucleophiles (Table 3.3). Synthetic yields were observed to

double in many cases, greatly increasing the utility of these transformations. The functional

group tolerance of the Pd/PyOx system is unprecedented for asymmetric conjugate addition; it

encompasses a wide array of halides, carbonyl functional groups, protected phenols, acetamides,

free hydroxyl groups, and even nitro groups. Many of these groups are incompatible with

traditional rhodium- and copper-catalyzed conjugate additions due to the reactivity with the

nucleophiles used or the strong coordination of the functional group to the metal catalyst.

Figure 3.6. Proposed catalytic cycle for Pd/PyOx-catalyzed conjugate addition of arylboronic

acids to cyclic enones.

L

P d N

N

P d N

N

O

P hP h

OOP d

P d

L

NN

N

L

N

O

O

P hX L

P d

NN B (O H )2

t ra n s m e ta lla t io n

in s e r t io n

p ro to n a t io n

c o o rd in a t io n

+

+

+

H 2 O

3 1

3 2

3 3

3 4 3 5

3

2

+

+

81

The combined results described herein have allowed us to put forth the following catalytic

cycle (Figure 3.6). The cationic catalyst, represented as (PyOx)Pd(X)(L) (31), undergoes

transmetallation with an arylboronic acid to yield cationic (PyOx)Pd(Ar)(L) (32). Substrate

coordination forms cationic arylpalladium(II) 33, which undergoes rate and

enantioselectivity-determining insertion of the aryl moiet -system to afford

C-bound palladium enolate 34. Tautomerization to the O-bound palladium enolate (35), or direct

protonolysis of the C-bound enolate (34), liberates the conjugate addition product (3) and

regenerates a cationic palladium(II) species for reentry into the catalytic cycle.

3.5 Conclusions

In conclusion, we have reported experimental and computational results that corroborate a

single PyOx ligand/metal complex as the active catalytic species in the palladium-catalyzed

asymmetric conjugate addition of arylboronic acids to enones for the construction of quaternary

stereocenters. We have used computational models to rule out a suggested alternative mechanism

in which the palladium catalyst is acting as a Lewis acid to activate the enone. The preferred

mechanism involves formal transmetallation from boron to palladium, rate- and

enantioselectivity-determining carbopalladation of the enone olefin by a cationic palladium

species, and protonolysis of the resulting palladium-enolate. We have taken advantage of these

mechanistic insights to develop a modified reaction system whereby the addition of water and

ammonium hexafluorophophate increase reaction rates, and can facilitate lower catalyst loadings.

The modified conditions have opened the door to new substrate classes that were inaccessible by

the initially published reaction conditions. Furthermore, we have demonstrated that this

82

operationally simple reaction is tolerant of ambient atmosphere and capable of producing

enantioenriched, -quaternary ketones on multi-gram scale. The steric and electronic effects of

the boronic acid and enone substrates and the ligand on enantioselectivities were elucidated by a

combined experimental and computational investigation. The enantioselectivity is mainly

controlled by the steric repulsion of the t-Bu substituent of the oxazoline on the ligand and the

C' position hydrogens of the cyclohexenone substrate in the alkene insertion transition state.

Further investigations of both the scope of this transformation and its application toward

natural product synthesis are current underway in our laboratories.

3.6 Materials and Methods

Unless otherwise stated, reactions were performed with no extra precautions taken to

exclude air or moisture. Air and moisture sensitive reactions were carried out under an

atmosphere of argon using Schlenk and glovebox techniques. Diethyl ether, hexanes, benzene

and toluene were dried and deoxygenated by passing through columns packed with activated

alumina and Q5, respectively. Deionized water was purified with a Barnstead NANOpure

Infinity UV/UF system. Dichloroethane was dried by stirring over CaH2 and subsequent

distillation, followed by deoxygenation by three freeze–pump–thaw cycles. Deuterated solvents

were dried by distillation from CaH2 (CD2Cl2) and deoxygenated by three freeze-pump-thaw

cycles. Commercially available reagents were used as received from Sigma Aldrich unless

otherwise stated. Enone substrates were purchased from Sigma Aldrich (3-methylcyclohexenone,

2-cyclohexene-1-one, chromone) or prepared according to literature procedure.40

Pyridinooxazoline ligands were synthesized according to literature procedures.41

Iodobenzene

83

was obtained from Sigma Aldrich and purified by distillation and deoxygenated by three freeze–

pump–thaw cycles. AgPF6 was obtained from ABCR and used without further purification.

Reaction temperatures were controlled by an IKAmag temperature modulator. Thin-layer

chromatography (TLC) was performed using E. Merck silica gel 60 F254 precoated plates (250

nm) and visualized by UV fluorescence quenching, potassium permanganate, or p-anisaldehyde

staining. Silicycle SiliaFlash P60 Academic Silica gel (particle size 40-63 nm) was used for flash

chromatography. Analytical chiral HPLC was performed with an Agilent 1100 Series HPLC

utilizing a Chiralcel OJ column (4.6 mm x 25 cm) obtained from Daicel Chemical Industries, Ltd

with visualization at 254 nm and flow rate of 1 mL/min, unless otherwise stated. Analytical

chiral SFC was performed with a JASCO 2000 series instrument utilizing Chiralpak (AD-H or

AS-H) or Chiralcel (OD-H, OJ-H, or OB-H) columns (4.6 mm x 25 cm), or a Chiralpak IC

column (4.6 mm x 10 cm) obtained from Daicel Chemical Industries, Ltd with visualization at

210 or 254 nm. 1H and

13C NMR spectra were recorded on a Varian Inova 500 (500 MHz and

125 MHz, respectively) and a Varian Mercury 300 spectrometer (300 MHz and 75 MHz,

respectively). Data for 1H NMR spectra are reported as follows: chemical shift (δ ppm)

(multiplicity, coupling constant (Hz), integration). Data for 1H NMR spectra are referenced to the

centerline of CHCl3 (δ 7.26), CD2Cl2 (δ 5.32), or C2D4Cl2 (δ 3.72) as the internal standard and

were reported in terms of chemical shift relative to Me4Si (δ 0.00). Data for 13

C NMR spectra

were referenced to the centerline of CDCl3 (δ 77.0), CD2Cl2 (δ 53.8), or C2D4Cl2 (δ 43.6) and

were reported in terms of chemical shift relative to Me4Si (δ 0.00). Infrared spectra were

recorded on a Perkin Elmer Paragon 1000 Spectrometer and are reported in frequency of

absorption (cm-1

). High resolution mass spectra (HRMS) were obtained on an Agilent 6200

Series TOF with an Agilent G1978A MultiMode source in electrospray ionization (ESI),

84

atmospheric pressure chemical ionization (APCI) or mixed (MultiMode ESI/APCI) ionization

mode. Optical rotations were measured on a Jasco P-2000 polarimeter using a 100 mm

path-length cell at 589 nm.

3.7 References

1 (a) P. Perlmutter, in Conjugate Addition Reactions in Organic Synthesis, Tetrahedron Organic

Chemistry Series 9; Pergamon, Oxford, 1992; (b) K. Tomioka, Y. Nagaoka, in Comprehensive

Asymmetric Catalysis, Vol. 3 (Eds: E. N. Jacobsen, A. Pfaltz, H. Yamamoto, Springer-Verlag,

New York, 1999; Chapter 31; (c) Gini, F.; Hessen, B.; Feringa, B. L.; Minnaard, A. J. Chem.

Commun. 2007, 710–712.

2 For reviews, see: (a) Harutyunyan, S. R.; den Hartog, T.; Geurts, K.; Minnaard, A. J.; Feringa,

B. L. Chem. Rev. 2008, 108, 2824–2882. (b) Alexakis, A.; Backvall, J. E.; Krause, N.; Pamies,

O.; Dieguez, M. Chem. Rev. 2008, 108, 2796–2893. (c) Lopez, F.; Minnarard, A. J.; Feringa, B.

L. Acc. Chem. Res. 2007, 40, 179–188. (d) Hayashi, T.; Yamasaki, K. Chem. Rev. 2003, 103,

2829–2844. (e) Feringa, B. L. Acc. Chem. Res. 2000, 33, 346–353. (f) Rossiter, B. E.; Swingle,

N. M. Chem. Rev. 1992, 92, 771–806.

3 (a) Shintani, R.; Duan, W.-L.; Hayashi, T. J. Am. Chem. Soc. 2006, 128, 5628–5629. (b)

Hayashi, T.; Takahashi, M.; Takaya, Y.; Ogasawara, M.; J. Am. Chem. Soc. 2002, 124, 5052–

5058. (c) Takaya, Y.; Ogasawara, M.; Hayashi, T. J. Am. Chem. Soc. 1998, 120, 5579–5580.

4 (a) Nishikata, T.; Yamamoto, Y.; Miyaura, N. Angew. Chem., Int. Ed. 2003, 42, 2768–2770.

(b) Nishitaka, T.; Yamamoto, Y.; Miyaura, N. Chem. Lett. 2007, 36, 1442–1443. (c) Nishitaka,

T.; Kiyomura, S.; Yamamoto, Y.; Miyaura, N. Synlett 2008, 2487–2490.

5 Gini, F.; Hessen, B.; Minnaard, A. J. Org. Lett. 2005, 7, 5309–5312.

6 For reviews on the synthesis of quaternary stereocenters, see: (a) Denissova, I.; Barriault, L.

Tetrahedron 2003, 59, 10105–10146. (b) Douglas, C. J.; Overman, L. E. Proc. Natl. Acad. Sci.

U.S.A. 2004, 101, 5363–5267. (c) Christoffers, J.; Baro, A. Adv. Synth. Catal. 2005, 347, 1473–

1482. (d) Trost, B. M.; Jiang, C. Synthesis 2006, 369–396. (e) Mohr, J. T.; Stoltz, B. M. Chem.

Asian J. 2007, 21, 1476–1491. (f) Cozzi, P. G.; Hilgraf, R.; Zimmermann, N. Eur. J. Org. Chem.

2007, 5969–5994.

85

6 For an excellent comprehensive review, see: Hawner, C.; Alexakis, A. Chem. Commun. 2010,

46, 7295–7306.

7 (a) Wu, J.; Mampreian D. M.; Hoveyda, A. H. J. Am. Chem. Soc. 2005, 127, 4584–4585. (b)

Hird, A. W.; Hoveyda, A. H. J. Am. Chem. Soc. 2005, 127, 14988–14989. (c) Lee, K.-S.; Brown,

M. K.; Hird, A. W.; Hoveyda, A. H. J. Am. Chem. Soc. 2006, 128, 7182–7184. (d) Brown, M.

K.; May, T. L.; Baxter, C. A.; Hoveyda, A. H. Angew. Chem., Int. Ed. 2007, 46, 1097–1100. (e)

Fillion, E.; Wilsily, A. J. Am. Chem. Soc. 2006, 128, 2774–2775. (f) Wilsily, A.; Fillion, E. J.

Org. Chem. 2009, 74, 8583–8594. (g) Dumas, A. M.; Fillion, E. Acc. Chem. Res. 2010, 43, 440–

454. (h) Wilsily, A.; Fillion, E. Org. Lett. 2008, 10, 2801–2804.

8 (a) d’Augustin, M.; Palais, L.; Alexakis, A. Angew. Chem., Int. Ed. 2005, 44, 1376–1378. (b)

Vuagnoux-d’Augustin, M.; Alexakis, A. Chem.–Eur. J. 2007, 13, 9647–9662. (c) Palais, L.;

Alexakis, A. Chem.–Eur. J. 2009, 15, 10473–10485. (d) Fuchs, N.; d’Augustin, M.; Humam, M.;

Alexakis, A.; Taras, R.; Gladiali, S. Tetrahedron: Asymmetry 2005, 16, 3143–3146. (e)

Vuagnoux-d’Augustin, M.; Kehrli, S.; Alexakis, A. Synlett, 2007, 2057–2060. (f) May, T. L.;

Brown, M. K.; Hoveyda, A. H. Angew. Chem., Int. Ed. 2008, 47, 7358–7362. (g) Ladjel, C.;

Fuchs, N.; Zhao, J.; Bernardinelli, G. Alexakis, A. Eur. J. Org. Chem. 2009, 4949–4955. (h)

Palais, L.; Mikhel, I. S.; Bournaud, C.; Micouin, L.; Falciola, C. A.; Vuagnoux-d’Augustin, M.;

Rosset, S.; Bernardinelli, G.; Alexakis, A. Angew. Chem., Int. Ed. 2007, 46, 7462–7465. (i)

Hawner, C.; Li, K.; Cirriez, V.; Alexakis, A. Angew. Chem., Int. Ed. 2008, 47, 8211–8214. (j)

Müller, D.; Hawner, C.; Tissot, M.; Palais, L.; Alexakis, A. Synlett 2010, 1694–1698. (k)

Hawner, C.; Müller, D.; Gremaud, L.; Felouat, A.; Woodward, S.; Alexakis, A. Angew. Chem.,

Int. Ed. 2010, 49, 7769–7772.

9 (a) Martin, D.; Kehrli, S.; d’Augustin, M.; Clavier, H.; Mauduit, M.; Alexakis, A. J. Am.

Chem. Soc. 2006, 128, 8416–8417. (b) Kehrli, S.; Martin, D.; Rix, D.; Mauduit, M.; Alexakis,

A. Chem.–Eur. J. 2010, 16, 9890–9904. (c) Hénon, H.; Mauduit, M.; Alexakis, A. Angew.

Chem., Int. Ed. 2008, 47, 9122–9124. (d) Matsumoto, Y.; Yamada, K.-I.; Tomioka, K. J. Org.

Chem. 2008, 73, 4578–4581.

10 (a) Shintani, R.; Tsutsumi, Y.; Nagaosa, M.; Nishimura, T.; Hayashi, T. J. Am. Chem. Soc.

2009, 131, 13588–13589. (b) Shintani, R.; Takeda, M.; Nishimura, T.; Hayashi, T. Angew.

Chem., Int. Ed. 2010, 49, 3969–3971.

11 For excellent review articles, see: (a) Gutnov, A. Eur. J. Org. Chem. 2008, 4547–4554. (b)

Christoffers, J.; Koripelly, G.; Rosiak, A.; Rössle, M. Synthesis 2007, 1279–1300.

12 Lin, S.; Lu, X. Org. Lett. 2010, 12, 2536–2539.

13 Kikushima, K.; Holder, J. C.; Gatti, M.; Stoltz, B. M. J. Am. Chem. Soc. 2011, 133, 6902–

6905.

86

14

PyOx ligand scaffolds are increasingly common in asymmetric catalysis, see: (a) Podhajsky,

S. M.; Iwai, Y.; Cook-Sneathen, A.; Sigman, M. S. Tetrahedron 2011, 67, 4435–4441. (b)

Aranda, C.; Cornejo, A.; Fraile, J. M.; García-Verdugo, E.; Gil, M. J.; Luis, S. V.; Mayoral, J.

A.; Martínez-Merino, V.; Ochoa, Z. Green Chem. 2011, 13, 983–990. (d) Pathak, T. P.;

Gligorich, K. M.; Welm, B. E.; Sigman, M. S. J. Am. Chem. Soc. 2010, 132, 7870–7871. (d)

Jiang, F.; Wu, Z.; Zhang, W. Tetrahedron Lett. 2010, 51, 5124–5126. (e) Jensen, K. H.; Pathak,

T. P.; Zhang, Y.; Sigman, M. S. J. Am. Chem. Soc. 2009, 131, 17074–17075. (f) He, W.; Yip,

K.-T.; Zhu, N.-Y.; Yang, D. Org. Lett. 2009, 11, 5626–5628. (g) Dai, H.; Lu, X. Tetrahedron

Lett. 2009, 50, 3478–3481. (h) Linder, D.; Buron, F.; Constant, S.; Lacour, J. Eur. J. Org. Chem.

2008, 5778–5785. (i) Schiffner, J. A.; Machotta, A. B.; Oestreich, M. Synlett 2008, 2271–2274.

(j) Koskinen, A. M. P.; Oila, M. J.; Tois, J. E. Lett. Org. Chem. 2008, 5, 11–16. (k) Zhang, Y.;

Sigman, M. S. J. Am. Chem. Soc. 2007, 129, 3076–3077. (l) Yoo, K. S.; Park, C. P.; Yoon, C.

H.; Sakaguchi, S.; O’Neill, J.; Jung, K. W. Org. Lett. 2007, 9, 3933–3935. (m) Dhawan, R.;

Dghaym, R. D.; St. Cyr, D. J.; Arndtsen, B. A. Org. Lett. 2006, 8, 3927–3930. (n) Xu, W.; Kong,

A.; Lu, X. J. Org. Chem. 2006, 71, 3854–3858. (o) Malkov, A. V.; Stewart Liddon, A. J. P.;

Ramírez-López, P.; Bendová, L.; Haigh, D.; Kocovsky, P. Angew. Chem., Int. Ed. 2006, 45,

1432–1435. (p) Abrunhosa, I.; Delain-Bioton, L.; Gaumont, A.-C.; Gulea, M.; Masson, S.

Tetrahedron 2004, 60, 9263–9272. (q) Brunner, H.; Kagan, H. B.; Kreutzer, G. Tetrahedron:

Asymmetry 2003, 14, 2177–2187. (r) Cornejo, A.; Fraile, J. M.; García, J. I.; Gil, M. J.;

Herrerías, C. I.; Legarreta, G.; Martínez-Merino, V.; Mayoral, J. A. J. Mol. Catal. A: Chem.

2003, 196, 101–108. (s) Zhang, Q.; Lu, X.; Han, X. J. Org. Chem. 2001, 66, 7676–7684. (t)

Zhang, Q.; Lu, X. J. Am. Chem. Soc. 2000, 122, 7604–7605. (u) Perch, N. S.; Pei, T.;

Widenhoefer, R. A. J. Org. Chem. 2000, 65, 3836–3845. (v) Bremberg, U.; Rahm, F.; Moberg,

C. Tetrahedron: Asymmetry 1998, 9, 3437–3443. (w) Brunner, H.; Obermann, U.; Wimmer, P.

Organometallics 1989, 8, 821–826.

15 The ligand was prepared as described in the literature, see: Brunner, H.; Obermann, U. Chem.

Ber. 1989, 122, 499–507.

16 (a) Xu, Q.; Zhang, R.; Zhang, T.; Shi, M. J. Org. Chem. 2010, 75, 3935–3937; (b) Zhang,

T.; Shi, M. Chem.–Eur. J. 2008, 14, 3759–3764; (c) Gottummukkala, A. L.; Matcha, K.; Lutz,

M.; de Vries, J. G. ; Minnaard, A. J. Chem.–Eur. J. 2012, 18, 6907–6914.

17 Holder, J. C., Marziale, A. N., Gatti, M.; Mao, B.; Stoltz, B. M. Chem.–Eur. J. 2013, 19, 74–

77.

18 We believe the small amount of deuterium incorporation at the methylene adjacent the enone

occurs via substrate enolization during the extended reaction times under the mild reaction

conditions.

19 Duan, W.-L.; Iwamura, H.; Shintani, R.; Hayashi, T. J. Am. Chem. Soc. 2007, 129, 2130–

2138.

87

20

Guillaneux, D.; Zhao, S.-H.; Samuel, O.; Rainford, D.; Kagan, H. B. J. Am. Chem. Soc. 1994,

116, 9430–9439. Further evidence for the monomeric nature of the Pd/PyOx catalyst was

provided by comparison of diffusion rates of various PyOx/Pd complexes by diffusion oriented

spectroscopy (DOSY NMR). These studies suggest an upper bound for the molecular weight of

the solution species of the catalyst, demonstrating that dimeric (ML)2 complexes are not present.

21 (a) Inanaga, J.; Furuno, H.; Hayano, T. Chem. Rev. 2002, 102, 2211–2226. (b) Kagan, H. B.

Synlett 2001, 888–899. (c) Girard, C.; Kagan H. B. Angew. Chem., Int. Ed. 1998, 37, 2922–

2959.

22 Nishitaka, T.; Yamamoto, Y.; Miyaura, N. Organometallics 2004, 23, 4317–4324.

23 The catalyst itself is known to be soluble and not zero valent palladium nanoparticles due to

exclusion by a mercury drop test, see ref 17. For examples of the mercury drop test, see: (a) Ines,

B.; SanMartin, R.; Moure, M. J.; Dominguez, E. Adv. Synth. Catal. 2009, 351, 2124–2132. (b)

Ogo, S.; Takebe, Y., Uehara, K., Yamazaki, T., Nakai, H., Watanabe, Y., Fukuzumi, S,

Organometallics 2006, 25, 331–338.

24 Attempts to study the reaction by NMR initially saw no reaction progress due to the inability

to stir the immiscible reaction mixture in an NMR tube. Substitution of 2,2,2-trifluoroethanol

for water as the super stoichiometric proton source facilitated the reaction to proceed in the

absence of stirring with no detriment to the observed enantioselectivity, however the reaction

was necessarily performed at sufficiently elevated temperature that observation of the initial

rate was not tenable and, thus, kinetic study was abandoned.

25 Lan, Y.; Houk, K. N. J. Org. Chem. 2011, 76, 4905–4909.

26 For related computational studies on palladium-catalyzed conjugate additions of arylboronic

acids to enones: (a) Nishikata, T.; Yamamoto, Y.; Gridnev, I. D.; Miyaura, N. Organometallics

2005, 24, 5025–5032. (b) Sieber, J. D.; Liu, S.; Morken, J. P. J. Am. Chem. Soc. 2007, 129,

2214–2215. (c) Dang, L.; Lin, Z.; Marder, T. B. Organometallics 2008, 27, 4443–4454.

27 (a) Becke, A. D. J. Chem. Phys. 1993, 98, 1372–1377. (b) Becke, A. D. J. Chem. Phys. 1993,

98, 5648–5652. (c) Perdew, J. P.; Chevary, J. A.; Vosko, S. H.; Jackson, K. A.; Pederson, M. R.;

Singh, D. J.; Fiolhais, C. Phys. Rev. B 1992, 46, 6671–6687. (d) Perdew, J. P. Phys. Rev. B 1986,

33, 8822–8824.

28 Gaussian 03, Revision C.02, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A.

Robb, J. R. Cheeseman, J. A. Montgomery, Jr., T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam,

S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A.

Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T.

Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross,

V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R.

88

Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J.

Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick,

A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J.

Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J.

Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B.

Johnson, W. Chen, M. W. Wong, C. Gonzalez, and J. A. Pople, Gaussian, Inc., Wallingford CT,

2004.

29 Here, it is assumed the isomeric trans and cis phenylpalladium complexes cannot undergo

rapid isomerization before alkene insertion and thus the enantioselectivity is determined by the

energy difference between 8-TS-A and 8-TS-B, since the transmetallation leading to the trans

isomer is favored (see SI). If trans/cis isomerization is faster than alkene insertion, the

enantioselectivity will be determined by the energy difference between 8-TS-A and 8-TS-D (98%

ee).

30 Other Lewis acids also proved incapable of catalyzing the reaction, including AlCl3 and

Sn(OTf)2.

31 These experiments were followed by NMR (

1H,

13C ), however no discrete intermediates were

successfully characterized.

32 Attempts to separate the isomeric mixture of phenylpalladium iodide complexes failed by

both conventional silica gel flash chromatography and preparatory HPLC.

33 Optimization of the cis/trans isomerization transition state of the cationic phenyl

palladium(II) complex failed to locate a TS. Scan of the reaction coordinate indicates the barrier

of isomerization is higher than 10 kcal/mol with respect to the cationic phenyl palladium

complex 11. This suggests the cis/trans isomerization via the dissociative mechanism via

isomerization of the tri-coordinated complex 11 cannot occur.

34 Computations indicated that alkene insertion to the minor isomer of phenyl palladium

complex, in which the Ph is cis to the oxazoline, yields very low enantioselectivity. The chiral

oxazoline is now trans to the enone, and thus the stereocontrol is diminished (see 1-TS-C and

1-TS-D in Figure 3.5).

35 A similar argument about imparted enantioselectivity can be made to rule out the catalytic

activity of palladium(0) nanoparticles. Additionally, a mercury(0) poisoning test has ruled out

the activity of palladium(0) nanoparticles in the Pd/PyOx manifold. See ref. 17.

36 Brønsted acid catalysis was also ruled out, as the substitution of trifluoroacetic acid for

Pd(OCOCF3)2 proved unable to catalyze the reaction. Protic acids are not tenable catalysts in the

absence of palladium salts.

89

37

We have computed the effects of coordination with phenylboronic acid to activate the

carbonyl on the enone in the alkene insertion step. No acceleration on alkene insertion was

observed computationally with either Lewis acid or hydrogen bonding coordination. (Result not

included).

38 The suggestion of boronic acid stabilization of arylpalladium cationic intermediates is

consistent with the observation that boron species lacking hydroxyl groups serve as poor

substrates for the reaction. For example, greatly diminished yields (and high degrees of biphenyl

formation) are observed with the use of NaBPh4 or KF3BPh as the phenyl donor species. (Not

included)

39 See Supporting Information of ref. 13 for details on the sensitivity of enantioselectivity of the

Pd/PyOx system to polar, coordinating solvents. The addition of 5 equiv. of alcoholic co-solvent

as a proton source is generally detrimental to the enantioselectivity and, occasionally, to the yield

of the reaction.

40 Shintani, R.; Yamagami, T.; Kimura, T.; Hayashi, T. Org. Lett. 2005, 7, 5317–5319.

41 (a) i-PrPyOx: Frauenlob, R.; McCormack, M. M.; Walsh, C. M.; Bergin, E. Org. Biomol.

Chem. 2011, 9, 6934–6937. (b) PhPyOx: Malkov, A. V.; Stewart-Liddon, A. J. P.; McGeoch, G.

D.; Ramirez-Lopez, P.; Kocovsk, P. Org. Biomol. Chem. 2012, 10, 4864–4877. (c) i-BuPyOx:

Brunner, H.; Obermann, U. Chem. Ber. 1989, 122, 499–507. (d) 4-OMePyOx, Jensen, K. H.;

Webb, J. D.; Sigman, M. S. J. Am. Chem. Soc. 2010, 132, 17471–17482. (e) 4-CF3PyOx: Jensen,

K. H.; Webb, J. D.; Sigman, M. S. J. Am. Chem. Soc. 2010, 132, 17471–17482. (f) 5-PhPyOx:

Brunner, H.; Obermann, U. Chem. Ber. 1989, 122, 499–507. (g) t-BuQuinOx: He, W.; Yip, K-T.;

Zhu, N-Y.; Yang, D. Org. Lett. 2009, 11, 5626–5628. (h) i-PrQuinOx: Zhang, Y.; Sigman, M. S.

J. Am. Chem. Soc. 2007, 129, 3076–3077.

90

Chapter 4. Stereoselectivities of Diels-Alder Reactions of 5-Substituted Cyclopentadienes: Effects

of Hyperconjugative Aromaticity on Reactivity and Stereoselectivity

4.1 Abstract

In the Diels-Alder cycloaddition of 5-substituted cyclopentadienes, the dienophile could add

either anti or syn to the substituents, which is the π-facial stereoselectivity. A computational analysis

based on distortion/interaction theory is conducted with SCS-MP2, which was employed due to its

agreement with the high-level composite theory G4. Activation and reaction energies are affected

substantially by the 5-substituents and characterized by the bending of C5-X out-of-plane angles to the

C1-C4-C5 plane. Due to the bending of C5-X bonds in the reactant dienes, attack of the dienophiles syn

to a donor and anti to an acceptor are favored since less distortion is required to reach the

corresponding transition states. The bending of C5-X out-of-plane angles in each substituted

cyclopentadiene results from the interactions between the π orbitals of diene and the 5-substituent.

4.2 Introduction

The Diels-Alder cycloaddition is one of the most versatile tools in organic synthesis.1-3

The

generation of complexity with high stereoselectivity provides considerable utility in synthesis and its

atom efficiency stands out in the pursuit of green chemistry.4 In addition to the throughly-studied

exo/endo stereoselectivity, π- facial stereoselectivity provides a potent method of asymmetric induction.

One general selectivity of this type involves the direction of approach of a dienophile to a 5-substituted

91

cyclopentadiene. (Scheme 4.1) When a dienophile, YZ, reacts with a 5-substituted cyclopentadiene,

either syn or anti adducts can be obtained.

Scheme 4.1. The π- facial stereoselectivity

Our interest in the π- facial stereoselectivity of Diels-Alder cycloaddition of 5-substituted

cyclopentadienes was rekindled by the late David Gin,5 who visited UCLA shortly before his tragic

death. In Gin’s total synthesis of aconitine alkaloids, the π- facial stereoselectivity in Diels-Alder

cycloaddition was employed as a key factor to obtain the desired product (3a). Scheme 4.2 shows the

reaction of a cyclopropene dienophile, which adds to the methoxy group on the cyclopentadiene

derivative, with a syn:anti ratio of around 6:1.

Scheme 4.2. Utilization of the π-facial stereoselectivity in total synthesis by David Gin

Despite of its broad application, the origin of the π-facial stereoselectivity remains controversial.

Among the effort of various groups, Burnell and his group first identified the importance of

deformation energy, which we call distortion energy, on stereoselectivities of these reactions. This

92

paper constitutes a theoretical update and an elaboration of the elegant theoretical studies based on

Hartree-Fock theory by Burnell. We developed a straightforward explanation and predictive model

based upon distortion/interaction theory.

4.3 Background

In 1955, the surprisingly contrasteric cycloaddition of 5-acetoxycyclopentadiene to ethylene and

other dienophiles was reported by the Winstein and Woodward groups.6 Since then, a great volume of

examples were identified.

The experimental studies are led by the Burnell group, along with extensive efforts from other

groups as well. Fallis7,8

performed a study of oxygen, sulfur and nitrogen 5-substituents and found that

when the C5 substituent involves N or O bonding to C-5, the dienophile tends to add syn to the

substituent, while a SH substitution on C5 gives mixed products and a SMe on C5 gives a 10:1 ratio on

anti/syn products. Burnell carried out extensive studies on π- facial stereoselectivities of various

5-substituents, and found that steric effects dominate in the absence of heteroatoms for 5-alkyl

cyclopentadienes.9,10

Their investigations of 5-methylcyclopentadiene Diels-Alder reactions revealed

that the stereoselectivity is relatively insensitive to the nature of dienophiles.11

They also studied the

inverse electron demand 1,2,3,4,5-pentachloro-5-X-cyclopentadienes and showed that the π- facial

stereoselectivity follows the same pattern as in nonchlorinated cyclopentadienes.12

Despite numerous experimental investigations, the origin of the π- facial stereoselectivity

remains unsettled, even controversial. The first rationale, in 1973, was Anh's proposal that

stereoselectivities are determined by a stabilizing interaction between the heteroatom lone-pair and the

dienophile π*-LUMO.13

This explains the syn preference of the acetoxy group, but the repulsive

93

interaction between the heteroatom n-orbitals and the diene π-HOMO was neglected. Furthermore,

related heteroatoms, such as the stronger donor S in SMe, give an anti preference.

The Burnell group expanded their experimental studies with extensive calculations with the

Hartree-Fock theory. With bulky hydrocarbon substitutions on C5, the steric effect is dominant and

overrides other factors in π-facial stereoselectivity, as established by Houk and Burnell. For smaller and

heteroatom substituents, Burnell identified the importance of deformation energy on the reactivities and

stereoselectivities of reactions of 5-substituted cyclopentadienes. They showed good correlations

between the diene deformation energies and activation energies. They also showed that the differences

between the syn and anti deformation energies correlate with the differences in syn and anti activation

energies. While the distortion energies of anti additions to 5-substituted cyclopentadienes are in a small

range around the cyclopentadiene, the syn distortion energies are in a much wider range. The

dienophile approaches preferentially at the face with a smaller change of C1-C5-X angle. A modified

steric factor in the reactant diene describing the bond size of C5-X has been proposed. In spite of theses

correlations, the origin of the distortion from the reactant to the transition state (TS) is still unknown.

Various theories attempt to rationalize the π-facial stereoselectivity other than distortion.

Inagaki elaborated the Fukui orbital mixing scheme and argued that syn/anti facial selectivity comes

from the tilting of the orbitals of C1 and C4 in the diene π*-LUMO. This tilting is the consequence of

σ-p orbital mixing between the C1/C4 σ orbital and the π* orbital of the cyclopentadiene, induced by

the heteroatom lone pair p orbital. This argument applies as well for electronegative 5-substituents.14,15

Inagaki extended the arguments to both electronegative and less electronegative heteroatom

substituents.16

Cieplak rationalized the syn preference to the better donor ability of C5-X than C5-H.17

Cieplak’s theory always favors bond formation anti to the best donor, which would lead to a wider

94

range in anti activation energies. This seems quite contrary to the experiment results and to Burnell’s

calculations.

Electrostatic effects have been proposed as an alternative explanation. Kahn and Hehre argues

that the more nucleophilic face of diene is subject to attack by the electron-poor dienophile,18

but it

does not apply to inverse electron demand Diels-Alder additions.12

In summary, the origin of π- facial stereoselectivity is still uncertain. “The general applicability

and predictive value of the proposed explanations are not obvious at the present stage.”19

We have revisited this problem with several types of density functional and ab initio wave

function based methods, and have come up with a new general explanation of this phenomenon. We

report a theoretical exploration of these reactions using high-accuracy methods, both DFT and ab initio.

The distortion/interaction theory, combined with the effects of substituents on the structures of

5-substituted cyclopentadienes, provide a simple explanation on the origin of the π- facial

stereoselectivity.

4.4 Computational Details

All the calculations were performed with Gaussian 09.20

The B3LYP density functional21,22

was

employed with the 6-31G(d) basis set for geometry optimizations. Single point energies were obtained

with a larger 6-311++G(d,p) basis set. Various density functional methods (DFTs) as well as some

wave-functional-based methods (WFTs) and multi-determinant and high level composite method were

tested here. B3LYP has been widely used for geometry optimizations and predicting relative energies,

but to calculate absolute energies, it is subject to significant systematic errors, most relevantly to this

work, on the Diels-Alder reaction of cyclopentadiene and ethene.23

In addition to B3LYP, M06-2X24

95

was tested here since it has been shown to provide more accurate thermodynamics for main-group

elements. B2PLYP-D25,26

was also applied, which incorporates the dispersion correction in the

functional. Furthermore, Grimme’s WFT-based spin-component-scaled second-order Møller-Plesset

perturbation theory (SCS-MP2),27

which exhibits a significant improvement over the original MP2,28

was also used here. High-level composite methods, G3B3,29

CBS-QB3,30,31

and G432

were employed to

compare with the DFT results.33,34

The latest development in the high-level composite method series,

G4, is based on a sequence of single point energy calculations and give a total energy effectively at

CCSD(T,full)/G3LargeXP+HFlimt level. It is applicable for compounds containing first-, second-, and

thired-row main atoms, and exhibits the least mean absolute deviation from 454 experimental energies

in the G3/05 test set, as low as 0.83 kcal/mol, which is accurate enough to beyond the experimental

limit. Due to its accuracy, G4 energies are taken here as the reference values.

4.5 Results and Discussions

4.5.1 Comparison of various theoretical levels

We have investigated the substituent effects on the activation free energies in Scheme 4.3 with 17

5-substituted cyclopentadienes reacting with ethene. There are a total of 35 reactions since ethene

attacking in both syn and anti fashions have been considered except the parent reaction, which is

cyclopentadiene reacting with ethene. The substituents range from a potent π-donor like OMe, to a

strong π-acceptor like BH2. Activation energies in each reaction are calculated with the highest level

method G4 as a reference, along with other high-accuracy composite methods including G3B3 and

CBS-QB3, and the density functional theories including B3LYP, M06-2X, SCS-MP2, B3PLYP-D. For

comparison, free energies was also calculated from the HF/6-31G(d) single point and HF.6-31G(d)

96

geometry, as adopted in Burnell’s previous analysis. The predicted activation energies by various

theoretical methods are plotted against the G4 values. (Figure 4.1)

Scheme 4.3. Benchmarking reactions in anti/syn fashion

97

yHF = 0.82x + 29.4

R2 = 0.76

yB2PLYP-D = 0.81x + 14.9

R2 = 0.82

yB3LYP = 0.98x + 7.3

R2 = 0.95

yM06-2X = 1.10x - 3.8

R2 = 0.98

ySCS-MP2 = 0.96x + 1.1

R2 = 0.99

yG3B3 = 0.97x + 0.3

R2 = 1.00

yMP2 = 1.24x - 15.6

R2 = 0.96

yCBS-Q B3 = 0.99x - 1.6

R2 = 0.99

15.0

20.0

25.0

30.0

35.0

40.0

45.0

50.0

55.0

60.0

65.0

25.0 30.0 35.0 40.0

HF

B2PLYP-D

B3LYP

SCS-MP2

G3B3

M06-2X

CBS-QB3

MP2

98

Figure 4.1. Calculated activation free energies (ΔGX‡) in 35 reactions (kcal/mol). The regression

equation is ΔGX‡ (method) = slope * ΔGX

‡ (G4) + interception.

High-level composite theoretical methods G3B3, CBS-QB3 and G4 agree well with each other.

CBS-QB3 consistently underestimates the free energies by 2.1 kcal/mol. SCS-MP2, G3B3 and

M06-2X give excellent agreement with the G4 values, with 0.3 kcal/mol, 0.5 kcal/mol and 0.6 kcal/mol

MAD, respectively, while the computational cost of SCS-MP2 and M06-2X are much lower.

The HF, B2PLYP-D, and B3LYP barriers are overestimated by around 24 kcal/mol, 9 kcal/mol,

and 7 kcal/mol, respectively. The failure of HF absolute barriers is probably due to the neglect of

correlation energy, while MP2 does not provide an accurate prediction of correlation energies and the

barriers are underestimated by 8 kcal/mol. Therefore, these methods are not suitable for predicting

absolute barriers.

The slope is 0.99 for CBS-QB3, closest to 1, followed by B3LYP (0.98), G3B3 (0.97), and

SCS-MP2 (0.96). These methods are acceptable to predict relative barriers. On the other hand, M06-2X

has a higher slope of 1.10, while MP2 is the highest at 1.24. HF and B2PLYP-D give the smallest slope

of 0.82 and 0.81, respectively. The HF values are scattered around the linear regression line, with the

most significant deviation coming from a 2 kcal/mol underestimation at the lower end for F (syn) and

the higher end for BH2 (both syn and anti). B2PLYP fails in these cases as well, with a 1.5 kcal/mol

underestimation. Therefore the overall slope is decreased greatly for these two methods.

To further test the performance of the theories aforementioned on the π-facial

stereoselectivities, relative anti/syn free energies are compared with G4 values (Figure 4.2).

99

Figure 4.2. Comparison of various theories against G4 in predicting π-facial stereoselectivities ΔΔGX‡

(anti-syn, kcal/mol) The regression equation is ΔΔGX‡ (method) = slope * ΔΔGX

‡ (G4) + interception.

While various theories lead to large differences in predicting the free energies, their performances

on the anti/syn selectivity are much better, less than 1.0 kcal/mol in MAD from G4. The composite

methods, G3B3 and CBS-QB3, provide the least interception of 0.1 kcal/mol and -0.1 kcal/mol,

respectively. B3LYP and SCS-MP2 also exhibit minor deviation from G4 with an interception of -0.2

kcal/mol. MP2, B2PLYP-D are also within the range of ±0.5 kcal/mol, while the deviation for M06-2X

and HF are a little higher, at 0.6 kcal/mol and -0.8 kcal/mol, respectively. This means that M06-2X

yHF = 1.18x - 0.8

R2 = 0.98

yB2PLYP-D = 0.99x + 0.4

R2 = 0.96

yG3B3 = 1.01x + 0.1

R2 = 1.00

yB3LYP = 0.99x - 0.2

R2 = 0.97

ySCS-MP2 = 1.05x - 0.2

R2 = 0.99

yM06-2X = 1.11x + 0.6

R2 = 0.99

yCBS-Q B3 = 1.09x - 0.1

R2 = 0.99

yMP2 = 1.20x - 0.3

R2 = 0.96

-8.0

-6.0

-4.0

-2.0

0.0

2.0

4.0

6.0

8.0

-6.0 -4.0 -2.0 0.0 2.0 4.0 6.0

HF

B2PLYP-D

B3LYP

SCS-MP2

G3B3

M06-2X

CBS-QB3

MP2

100

gives systematic overestimation of ΔG‡

a-s favoring syn attack by around 0.6 kcal/mol, while HF

systematically gives 0.8 kcal/mol preference for anti attack. The slopes for G3B3, B3LYP and

B2PLYP-D are closest to 1, being 1.01, 0.99 and 0.99, respectively, followed by a slightly higher slope

of 1.05 with SCS-MP2. Considering both the interceptions and the slopes, G3B3 shows the best

agreement with G4, while B3LYP is the best choice between the non-composite methods, follows

closely by B2PLYP-D and SCS-MP2.

Our analysis with distortion/interaction theories requires an economic and accurate method that is

able to provide both quantitative estimation of absolute energies, and the anti/syn stereoselectivities.

SCS-MP2 provides lowest MAD and SD for absolute barriers, and performance the second best on the

anti/syn stereoselectivities. Due to its excellent performance both on absolute and relative activation

free energies, SCS-MP2 has been employed in the following studies in the present work. All the

energies were calculated by SCS-MP2/6-311++G(d,p)//B3LYP/6-31G(d), unless otherwise specified.

4.5.2 π-Facial stereoselectivity and geometrical distortions of reactants and transition states,

characterized by the C5-X out-of-plane angle

To trace the origin of differences in anti/syn activation barriers, we analyzed the TS geometries

with 5-substituted cyclopentadiene models, including cyclopentadiene (4), 5-methoxycyclopentadiene

(5), and 5-methborylcyclopentadiene (6). The TS structures for ethene with 4 (TS4), with 5 in anti/syn

fashion (TS5a, TS5s), and with 6 in anti/syn fashion (TS5a, TS5s), were plotted in Figure 4.3.

A parameter θ, characterizing the bending of the substituent groups X, is the C5-X out-of plane

angle with respect to the plane defined by C1, C4 and C5. (Scheme 4.4)

Scheme 4.4. The C5-X out-of-plane angle, θ

101

Figure 4.3. The Diels-Alder transition state geometries for dienophile ethene with cyclopentadiene

(TS4), anti/syn to 5-methoxycyclopentadiene (TS5a and TS5s), and 5-methborylcyclopentadiene

(TS6a and TS6s). The C5-X out-of-plane angle are in red, and the C5-H out-of-plane angles are in

black, except for TS4, where X=H.

As shown in Figure 4.3, in the transition state for cyclopentadiene reacting with ethene (TS4), the

out-of-plane angle is larger for syn C5-H (133°) than anti C5-H (119°). This broken symmetry result

from a preference for staggering of the vicinal bonds related to the bond to the atom addition.35-37

(Scheme 4.5)

Scheme 4.5. Staggered conformation in the transition state for the cyclopentadiene-ethene reaction

102

When one of the C5-H is substituted by a methoxy or methboryl group, the transition states have

subtle changes in θ on the anti/syn π-faces. Compared with the θ anti to ethene in TS4 (119°), in anti

TSs, it is slightly increased to 123° in TS5a, while it stays the same in TS6a. θ syn to ethene are

slightly decreased from 133° in TS4 to 129° in TS5a and 130 in TS6a. Overall, the changes in θ are

smaller then 5° in anti TSs with the representative 5-substituents, both a strong π-donor like MeO and a

strong π-acceptor like MeBH. On the other hand, the changes in θ with 5-substitution are greater for

syn TSs. The θ anti to ethene in TS4 decreases to 115° in TS5s and 114° in TS6s, while the θ syn to

ethene increases to 136° in TS5s and, most notable, 145° in TS6s.

Other than the change of C5-X and C5-H bending angles, the forming C-C bond distance increases

from 2.25 in TS4 to 2.28 in both TS5a and TS5s with the substitution of MeO, denoting earlier TS, and

decreases to 2.20 in TS6a and almost the same (2.24) in TS6s.

Overall the changes of TS geometries are small, even with the 5-substituents of vastly different

electron demand. The change of out-of-plane angles are within 5° except for the C5-B θ in syn TS

(TS6s), with the changes in anti TSs even smaller than changes in syn TSs. The forming C-C bond

length does change slightly, but it is almost the same for 5 in both anti/syn TS.

Since the activation barrier is the energetic differences between TS and reactant, and the TS

structures are similar, the π-facial stereoselectivity lies in the difference in the reactants. With the same

dienophile ethene, the difference in the diene was analyzed.

103

Figure 4.4. The geometries of reactant diene. Cyclopentadiene (4), 5-methoxycyclopentadiene (5). and

5-methborylcyclopentadiene (6). The C5-X out-of-plane angle are labeled in red. The C5-H

out-of-plane angles are in black, except for 4, where X=H.

Figure 4.4 shows the geometries of the cyclopentadienes 4, 5, and 6. While the C5-X out-of-plane

angles in the transition states are similar, they are quite different in the corresponding reactants. θ in the

symmetrical diene 4 is 127°, while that of 5 is asymmetrical, with the C5-O bond bending away from

the π system, increasing θ to 137°. As we will discuss later, the bending of θ results from a strong

interaction between the 5-substituent and the π system. Due to the bending of C5-O θ, the C5-H θ in 5

also decreases to 119°. On the other hand, C5-B out-of-plane angle θ in 6 has decreased substantially to

109°, while C5-H θ in 6 has increased to 143°.

Comparing with the anti and syn addition transition states in Figure 4.3, the C5-O θ of the reactant

5 (137°) resemble more like that in TS5s (136°) than TS5a (123°). Due to the geometrical similarity,

lower activation barrier for the former is expected. This is supported by our SCS-MP2 computation that

the activation energies are 12.4 kcal/mol (TS5s) and 17.4 kcal/mol (TS5a). On the other hand, the

C5-B θ in reactant 6 is drastically decreased, thus requiring substantial bending out for both anti and

syn TS, however, the difference in θ between 6 (109°) and TS6s (145°) is much greater than TS6a

(119°). Therefore lower barrier for the latter is expected. Indeed our calculation gives 22.4 kcal/mol

and 18.9 kcal/mol for TS6s and TS6a, respectively.

104

To check the relationship between θ and activation energies, the activation energies for the

seventeen substituents were listed in Table 4.1, along with the C5-X out-of-plane angle θ. θ0 are those

in the reactants, while θTS are those in the anti/syn transition states. Also listed are distortion/interaction

energies, which will be talked about in the next section.

Table 4.1. Distortion/Interaction energies for various substituents (kcal/mol), and out-of-plane angle θ*

X ΔEa‡ ΔEd

‡ ΔEi

‡ θ0 TSa/s θTS

MeO 17.4 21.7 -4.3 137 a 123

12.4 18.6 -6.3 s 136

OH 18.3 22.6 -4.3 135 a 122

13.4 19.1 -5.7 s 136

F 15.7 20.2 -4.4 132 a 122

11.2 16.6 -5.4 s 135

CF3 17.4 22.6 -5.2 131 a 126

19.4 25.1 -5.7 s 145

NH2 16.5 21.1 -4.6 134 a 126

14.7 20.7 -5.9 s 142

CN 16.1 21.9 -5.8 130 a 123

14.1 21.8 -7.7 s 137

Cl 15.7 20.6 -5.0 129 a 123

14.3 20.9 -6.6 s 139

SH 16.2 21.3 -5.1 132 a 127

16.7 23.9 -7.2 s 139

Me 17.3 21.8 -4.5 131 a 125

17.5 23.3 -5.8 s 142

H 16.5 21.8 -5.3 127 a* 119

s* 133

Br 15.5 20.8 -5.3 126 a 121

105

16.0 21.8 -5.9 s 138

NO 14.1 19.9 -5.8 116 a 115

14.5 21.2 -6.7 s 132

CH=CH2 16.2 21.3 -5.1 128 a 123

17.1 23.3 -6.2 s 141

CHO 15.3 20.6 -5.3 120 a 120

17.4 24.9 -7.5 s 140

C≡CH 16.9 22.2 -5.3 132 a 124

14.5 21.4 -7.0 s 139

MeBH 18.9 24.6 -5.7 109 a 118

22.4 29.7 -7.3 s 145

BH2 20.0 25.7 -5.7 110 a 116

23.9 30.7 -6.8 s 144

SiH3 17.7 23.4 -5.7 120 a 123

23.0 29.5 -6.5 s 144

* The hydrogen syn or anti to the approaching ethene.

C5-substitutions affect more on the reactant structures than in the TS structures, while the effect is

larger for syn than anti TS. θ0 covers a range of 28°, about twice as large as those of syn and anti θTS,

which are only 13° and 12°, respectively. Substituents involving boron, and the nitrosyl group,

significantly affect the angle and the values lie far from other substituents, which makes the general

trend more difficult to observe. Taking them out make a better statistics of θ0 range from 120°-137°,

while syn and anti θTS range from 135°-145° and 122°-127°, respectively. Therefore, the range of θ0,

θTS (syn) and θTS (anti) are 17:10:5. The anti θTS for the strongest π-donor MeO and strongest acceptor

SiH3 are both 123°, the syn θTS are close (142° and 144°, respectively), while the θ0 with MeO and SiH3

differ by 17° (137° and 120°, respectively).

106

The observation agrees with the activation energies in Table 4.1 that anti activation energies are

similar, while the syn activation energies span a much wider range. The anti activation energies for the

strongest π-donor MeO and strongest acceptor SiH3 are quite close, which are 17.4 and 17.7 kcal/mol,

while the syn activation energies for MeO and SiH3 differ by over 10 kcal/mol, which are 12.4 and 23.0

kcal/mol, respectively. Another example is the halogen series. Anti activation energies for F, Cl, Br are

almost identical, which are 15.7, 15.7 and 15.5 kcal/mol, respectively, while their corresponding syn

activation energies are quite different, which are 11.2, 14.3 and 16.0 kcal/mol. Therefore, the π-facial

stereoselectivity is determined mainly through the reactants rather than transition states.

4.5.3 Distortion/Interaction energies in π-facial stereoselectivity

To provide a quantitative analyze of the activation energy ΔEa‡, we break it into distortion and

interaction energies. The distortion energy ΔEd‡, is the energy required to distort the reactants into the

corresponding TS structure, and the interaction energy ΔEi‡ is the change in energy from the reactant

electron configuration to that of TS. To provide a quantitative account of the geometric and

electrostatic effects in the activation barriers, we applied the distortion/interaction theory to all the 35

reactions and the results are combined with activation energies in Table 4.1.

As demonstrated in Table 4.1, activation energies are dominated by distortion energies, since they

are normally four to five times larger in magnitude than the interaction energies. As illustrated in

Figure 4.5, ΔEd‡ and ΔEa

‡ exhibit excellent linear relationship, while 5-X substituents affects syn TS

more profoundly than the anti TS.

107

Figure 4.5. The correlation between activation energies and distortion energies

The distortion energy not only dominates in activation energies, it also determines the π- facial

stereoselectivities, as indicated in Figure 4.6.

Figure 4.6. The correlation between π- facial stereoselectivities (ΔEa‡(a) - ΔE a

‡(s)) with differences in

anti and syn distortion energies

The slope of 1.0 and the 0.96 regression constant both demonstrate a perfect correlation for the

anti and syn distortion energies with the π- facial stereoselectivities. The 1.3 kcal/mol intercept denotes

108

a consistent favoring for the syn TS. As from Table 4.1, ΔEi‡(anti) are always around 1.3 kcal/mol

higher than ΔEi‡(syn), the interaction energies provide this bias for syn TS. This universal preference

might result from the electrostatic effect as Hehre mentioned.18

The dominance of distortion energies in activation energies, as well as a larger range for ΔEd‡ in

syn than anti TS, both follow the trend in the change of C5-X out-of-plane angles. If θ0 is closer to θTS

(syn) than θTS (anti), less distortion energy is required. Since distortion energy determines the

activation energy, if less change in θ is required for reactants to distort into TS, lower activation

energies are expected, as in the case for TS5s and TS6a.

Figure 4.7 demonstrated the correlation between distortion energies and the reactant C5-X

out-of-plane angle. The linear relationship of distortion energy and θ0 is semi-quantitative, with the

regression coefficient R2 = 0.58. The largest deviation coming from bulky groups, like -SH, -SiH3 and

-CF3, in which cases the steric repulsion also contribute to the increase of θ.

Figure 4.7. The correlation between distortion energies and the reactant C5-X out-of-plane angle

109

Although the correlation is only semi-quantitative, a simple “127°” criteria works for most cases,

even including some of the bulky groups. The reference θ angle of cyclopentadiene is 127°. For

5-substituted cyclopentadienes, if θ0 of the reactant is above 127°, the reaction proceeds in the syn

fashion, while if θ0 of the reactant is below 127°, the reaction proceeds in the anti fashion. If the θ0 of

the reactant is close to 127°, mixed products is expected. In spite of its simplicity, the “127°” criterion

works for almost all the systems. X = -SH, –CF3 and -CH=CH2 seems to be the only exceptions. The

first two are bulky groups and the cyclopentadiene has to make more distortion in the syn transition

state. In these cases, θ0 gets larger due to the bulky substituent, as reflected in the increasing of θTS in

both syn and anti transition states. X = -CH=CH2 is another exception since the vinyl α-hydrogen is

pointing towards the approaching ethene.

In summary, substituents affect the C5-X out-of-plane angles on the reactants, with those with

increased θ resemble transition states in the syn fashion, and vice versa, leading to the π- facial

stereoselectivity, as in Scheme 4.6.

Scheme 4.6. Substituent effects on the reactant geometries leading to π- facial stereoselectivity

4.5.4 Reaction energies and the reaction pathways

On average, C5-X θTS in syn and anti transition states differs by around 15° (Table 4.1), the

differences in the corresponding products are much smaller. For example, θrxn is only slightly different

110

(1°) on anti/syn faces when X=H, while θTS in different faces differ by 14° when X=H. In addition,

while syn θTS span a much larger range than anti, anti and syn θrxn span similar ranges.

Table 4.2. Reaction energies and C5-X out-of-plane angles (θrxn) for various substituents (kcal/mol),

and the relative (anti - syn) reaction energies (ΔΔErxn) and activation energies (ΔΔEa‡)

X ΔErxn θrxn ΔΔErxn ΔΔEa‡

MeO a -35.8 128 3.5 5.0

s -39.3 126

OH a -37.8 131 0.7 4.9

s -38.5 126

F a -37.9 129 3.3 4.6

s -41.2 125

CF3 a -31.4 131 0.5 -2.0

s -32.0 135

NH2 a -35.6 132 0.8 1.8

s -36.4 126

CN a -35.2 129 1.1 2.0

s -36.3 129

Cl a -36.0 129 2.5 1.4

s -38.5 129

SH a -33.8 133 0.4 -0.4

s -34.2 130

Me a -32.0 131 0.1 -0.2

s -32.1 132

H a -32.2 125 0.0 0.0

s 126

Br a -34.7 128 2.1 -0.5

s -36.8 128

111

NO a -34.7 125 1.5 -0.5

s -36.2 124

CH=CH2 a -32.6 130 0.0 -0.9

s -32.6 131

CHO a -31.4 129 -0.2 -2.1

s -31.2 132

C≡CH a -34.1 130 1.6 2.4

s -35.7 130

MeBH a -24.9 132 -0.6 -3.6

s -24.3 138

BH2 a -23.7 130 -0.6 -3.9

s -23.0 139

SiH3 a -26.3 129 -0.8 -5.4

s -25.5 135

Table 4.2 listed θ values in syn and anti products. The order of the substituents followed the

decreasing syn preference in activation energies by G4, as we did in Table 4.1. For the top five

substituents, which are the strongest π-donors that have N, O or F connecting to C5, anti θrxn is around

4° larger than syn θrxn. On the other hand, the last three substituents in Table 4.2, which are the

strongest π-acceptors, give anti θrxn that are around 6° smaller than the corresponding syn θrxn. θ values

in syn and anti products are almost the same for other substituents.

The syn and anti reaction energies are generally within 1.0 kcal/mol to each other, with the

largest differences occur for MeO and halogen substituents. Although F, Cl and Br lead to very

different π-facial selectivities, as indicated by a range of 5.1 kcal/mol for there ΔΔEa‡(anti - syn), there

relative reaction energies are quite close, with a range of only 1.2 kcal/mol. The rather constant relative

reaction energies, ΔΔEa‡(anti - syn), concur with the trend in θrxn.

112

Figure 4.8. The correlation between activation energies (Y axis) and reaction energies (X axis)

For a specific substituent, the syn/anti reaction energies are closer to each other comparing with

the differences in activation energies, while reaction energies from different substituents span a larger

range then the activation energies, as shown in the different X/Y scales in Figure 4.8. For the syn

transition states, the activation energies correlate better with the reaction energies, while the correlation

for anti transition states is much poorer.

Electron donating groups on C5, such as MeO, increases the activation energy, or destabilize

the diene, while electron-accepting groups like BHMe stabilize the diene. The least reaction energies

for 5-X cyclopentadiene Diels-Alder cycloaddition are around -25 kcal/mol for substituents BHMe,

BH2 and SiH3, and the highest ones are around -40 kcal/mol for F, OMe and OH syn attack.

y = 0.68x + 39.5

R2 = 0.95

y = 0.21x + 23.7

R2 = 0.38

10.0

15.0

20.0

25.0

-45.0 -40.0 -35.0 -30.0 -25.0 -20.0

ΔEa‡(anti)

ΔEa‡(syn)

113

Figure 4.9. The reaction pathways for ethene with cyclopentadiene (4), methoxycyclopentadiene (5)

and methborylcyclopentadiene (6) The reactants has been set as 0 kcal/mol in each pathway.

With the calculated energies, the complete reaction pathways for ethene with the parent

cyclopentadiene (4), and substituted diene with a strong π-donor (5) or π-acceptor (6), were plotted

together in Figure 4.9. Although the electronic demand is very different, the barriers for the two anti

TS, TS5a and TS6a, are close to each other, while the syn activation energies differ by 10 kcal/mol for

TS5s and TS6s. The reaction energy decreases greatly for Prod5s, but less so for Prod5a, while the

reaction energy for Prod6a and Prod6s are almost the same, both around 7.5 kcal/mol the that of

Prod4.

114

4.5.5 The origin of substituent-induced bending in the reactants – second order orbital

interaction

We attribute the bending of out-of-plane angle to a strong second order orbital interaction

between the 5-substituent and the π system. When there are only filled orbitals, like the nonbonding

orbitals on electronegative species like N, O, F, or Cl, the interaction with the π-HOMO of diene raises

the energy, since the interaction of two symmetry matched filled orbitals is unfavored. To minimize the

interaction, C5-X tends to move away from the diene, therefore increasing θ0 of the reactant. In the

case of X=MeO, this interaction is eliminated in the syn product Prod5s but only partially in the anti

product Prod5a, with the the energy of Prod5s even lower than Prod5a. The interaction of the lone

pairs on X with the π*-LUMO of diene could be favored, but their symmetry does not match and thus

not considered.

On the other hand, if there is a vacant orbital close in energy on X and matches the symmetry of

π-HOMO, the interaction aforementioned becomes favorable and C5-X tend to move towards the diene

to maximize the interaction, therefore decreasing θ0 of the reactant. This interaction stabilizes the

reactant and therefore less reaction energies were released in Prod6a and Prod6s.

To test on the proposed rationale, the diene HOMO orbital graphs are plotted in Figure 4.10.

The cyclopentadiene serves as a reference, MeO as a typical π-electron donor since it has nonbonding

lone pair electrons, and MeBH as a typical π-electron acceptor since it has vacant p orbitals.

115

Figure 4.10. Orbital interactions on model systems (X= H, –OMe, –BHMe)

For methoxycyclopentadiene, the interaction between oxygen lone pair and HOMO of diene is

repulsive and C5-X bend away from the diene moiety to minimize orbital mixing. For

5-MeBH-cyclopentadiene, the interaction between boron vacant p orbital and π-HOMO of diene is

favored and C5-X bend towards the diene moiety for enhanced orbital mixing.

4.5.6 Diels-Alder reaction energies and hydrogenation energies

Since hydrogenation of the 5-X cyclopentadienes eliminates the aforementioned secondary

orbital interactions between the substituent and the π system, turning the two conjugated double bond

into one, just the Diels-Alder reaction with ethene, a comparison of the reaction energy of two reactions

is made, as in Scheme 4.7.

Scheme 4.7. Comparison of 5-substituted cyclopentadiene Diels-Alder reactions and hydrogenations

116

The Diels-Alder reaction energies (ΔErxn) correlates very well with hydrogenation energies

(ΔEHY), as in Figure 4.11, with a slope 1.04 and intercept 0.2 kcal/mol, and the correlation coefficient

is 0.92. Therefore, this strong substituent effect comes from the interaction between the diene π-HOMO

and the 5-X substituent, with is shown in both reactions.

Figure 4.11. Comparison of 5-substituted cyclopentadiene Diels-Alder reaction energies (ΔErxn) and

hydrogenation energies (ΔEHY)

4.5.7 Hammett plots on the substitution effects- induction and resonance effects

To measure the extent that the π-facial stereoselectivity comes from induction and resonance

effects, the Hammett plot is plotted with the Taft values,38

and the results are in Figure 4.12 and Figure

4.13.

The overall correlation for the stereoselectivity dependence on the inductive effect is rather poor,

with the substituent separated into two groups, although a good linear relationship is shown within each

group, as in Figure 4.12. The top group, colored in blue, is smaller, while the bottom group, colored in

yellow, is larger. The separation indicates that inductive effect alone does not determine the

stereoselectivity, and is largely affected by steric effects. On the other hand, some correlation with the

y = 1.04x + 0.2

R2 = 0.92

-45.0

-40.0

-35.0

-30.0

-25.0

-20.0

-40.0 -35.0 -30.0 -25.0 -20.0

ΔErxn(a)

ΔErxn(s)

117

resonance effect is shown with R2 = 0.42, is it rather significant since the substituent X is connected to

a saturated carbon and does not participate directly to the resonance of the diene π systems. The

correlation indicates a different mechanism for substituent X to participate the resonance of the diene π

systems. That substantiates a possible second-order orbital interaction as we demonstrated. Smaller

groups are above the line, while larger groups are below the line. Since when the substituent gets

larger, the steric in the syn TS increases, which raises its energy, making ΔΔEa‡(a-s) lower. The

correlation with resonance constant gives further support for the 5-substituent interaction with the

interaction with π systems.

Figure 4.12. Hammett plot of the selectivity dependence on the inductive effect

118

Figure 4.13. Hammett plot of the selectivity dependence on the resonance effect

4.5.8. Comparison with Torquoselectivity

The proposed secondary orbital interaction scheme in the π-facial stereoselectivity could be

compared with the torquoselectivity, which addresses the inward/outward rotation on the ring opening

reaction of 3-X cyclobutene.39,40

With electronegative groups, the secondary orbital interaction favors

syn for π-facial stereoselectivity, while it is repulsive and favors outward rotation for torquoselectivity.

Just as the Diels-Alder cycloaddition of 5-X cyclopentadiene and ethene favors ethene attack

syn to electronegative groups, or π-donor groups, the ring opening of 3-X cyclobutene favors that the

π-donor groups to rotate outward. As in Figure 4.14, a reasonable correlation of the two contrasteric

reactions is given, and torquoselectivity is more sensitive to substituent effects since the substituents

are linked directly to the π systems.

119

Figure 4.14. π-facial stereoselectivity Ea‡(a-s) vs. torquoselectivity Ea

‡(i-o)

4.6 Conclusions

Facial selectivity has long been discovered experimentally and through computational modeling,

which inspires numerous discussions about its origin. For bulky substituents like CF3 or those with

third row elements, except for highly electronegative –Cl, steric effects dominates. For smaller

substituent, electronic effect dominates and contrasteric products is obtained for π-electron

withdrawing groups.

However, regardless of the type of substituents, the activation energies are determined by

bending of C5-X out-of-plane angles, which originates from second order orbital interactions between

X and the π-HOMO. Therefore a quick qualitative reference for prediction of facial selectivities for

Diels-Alder reactions is given. If the C5-X out-of-plane angle in 5-substituted cyclopentadiene is

120

greater than 127°, the angle in cyclopentadiene, syn product is expected, while anti product is expected

when the C5-X out-of-plane angle is less than 127°, except for –SH, –CH=CH2 and –CF3. The

secondary orbital interaction scheme could be expanded to our previous torquoselectivity scheme.

4.7 References

(1) Woodward, R. B.; Katz, T. J. Tetrahedron Lett. 1959, 19.

(2) Woodward, R. B.; Katz, T. J. Tetrahedron 1959, 5, 70.

(3) Woodward, R. B. J. Am. Chem. Soc. 1942, 64, 3058.

(4) Cheong, P. H. Y.; Legault, C. Y.; Um, J. M.; Celebi-Olcum, N.; Houk, K. N. Chem. Rev. 2011,

111, 5042.

(5) Wilmot, J. T. Ph. D. thesis, University of Illinois at Urbana-Champaign, 2010.

(6) Winstein, S.; Shatavsky, M.; Norton, C.; Woodward, R. B. J. Am. Chem. Soc. 1955, 77, 4183.

(7) Macaulay, J. B.; Fallis, A. G. J. Am. Chem. Soc. 1990, 112, 1136.

(8) Macaulay, J. B.; Fallis, A. G. J. Am. Chem. Soc. 1988, 110, 4074.

(9) Xidos, J. D.; Poirier, R. A.; Pye, C. C.; Burnell, D. J. J. Org. Chem. 1998, 63, 105.

(10) Poirier, R. A.; Pye, C. C.; Xidos, J. D.; Burnell, D. J. J. Org. Chem. 1995, 60, 2328.

(11) Burnell, D. J.; Valenta, Z. J. Chem. Soc.-Chem. Commun. 1985, 1247.

(12) Burry, L. C.; Bridson, J. N.; Burnell, D. J. J. Org. Chem. 1995, 60, 5931.

(13) Anh, N. T. Tetrahedron 1973, 29, 3227.

(14) Ishida, M.; Inagaki, S. J. Synth. Org. Chem. Jpn. 1994, 52, 649.

(15) Ishida, M.; Beniya, Y.; Inagaki, S.; Kato, S. J. Am. Chem. Soc. 1990, 112, 8980.

(16) Ishida, M.; Inagaki, S. In Orbitals in Chemistry; Inagaki, S., Ed.; Springer-Verlag Berlin:

Berlin, 2009; Vol. 289, p 183.

(17) Cieplak, A. S. Chem. Rev. 1999, 99, 1265.

(18) Kahn, S. D.; Hehre, W. J. J. Am. Chem. Soc. 1987, 109, 663.

(19) Ohwada, T. Chem. Rev. 1999, 99, 1337.

121

(20) Frisch, M. J. T., 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, Jr., J. A.; 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, N. J.; 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. ; B.01 ed.; Gaussian, Inc.: Wallingford CT, 2009.

(21) Becke, A. D. Phys Rev A 1988, 38, 3098.

(22) Lee, C. T.; Yang, W. T.; Parr, R. G. Phys. Rev. B 1988, 37, 785.

(23) Guner, V.; Khuong, K. S.; Leach, A. G.; Lee, P. S.; Bartberger, M. D.; Houk, K. N. J. Phys.

Chem. A 2003, 107, 11445.

(24) Zhao, Y.; Truhlar, D. G. Theor Chem Acc 2008, 120, 215.

(25) Grimme, S. J. Chem. Phys. 2006, 124.

(26) Schwabe, T.; Grimme, S. Phys Chem Chem Phys 2007, 9, 3397.

(27) Grimme, S. J. Chem. Phys. 2003, 118, 9095.

(28) Moller, C.; Plesset, M. S. Phys Rev 1934, 46, 0618.

(29) Baboul, A. G.; Curtiss, L. A.; Redfern, P. C.; Raghavachari, K. J. Chem. Phys. 1999, 110,

7650.

(30) Montgomery, J. A.; Frisch, M. J.; Ochterski, J. W.; Petersson, G. A. J. Chem. Phys. 2000, 112,

6532.

(31) Montgomery, J. A.; Frisch, M. J.; Ochterski, J. W.; Petersson, G. A. J. Chem. Phys. 1999, 110,

2822.

(32) Curtiss, L. A.; Redfern, P. C.; Raghavachari, K. J. Chem. Phys. 2007, 126.

(33) Zou, L. F.; Shen, K.; Fu, Y.; Guo, Q. X. J Phys Org Chem 2007, 20, 754.

(34) Zou, L. F.; Fu, Y.; Shen, K.; Guo, Q. X. Theochem-J. Mol. Struct. 2007, 807, 87.

(35)Houk, K. N.; Paddon-Row, M. N.; Rondan, N. G.; Wu, Y. D.; Brown, F. K.; Spellmeyer, D. C.;

Metz, J. T.; Li, Y.; Loncharich, R. J. Science 1986, 231, 1108.

(36) Ando, K.; Green, N. S.; Li, Y.; Houk, K. N. J. Am. Chem. Soc. 1999, 121, 5334.

(37) Wu, Y. D.; Houk, K. N.; Paddon-Row, M. N. Angew. Chem. Int. Ed. 1992, 31, 1019.

(38) Hansch, C.; Leo, A.; Taft, R. W. Chem. Rev. 1991, 91, 165.

(39) Rondan, N. G.; Houk, K. N. J. Am. Chem. Soc. 1985, 107, 2099.

122

(40) Kirmse, W.; Rondan, N. G.; Houk, K. N. J. Am. Chem. Soc. 1984, 106, 7989.