zirconium and palladium catalyzed telescopic synthesis of ... · zirconium and palladium catalyzed...

119
Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes by Jordan Anthony Evans A thesis submitted in conformity with the requirements for the degree of Master of Science Department of Chemistry University of Toronto © Copyright by Jordan Anthony Evans 2014

Upload: others

Post on 22-Jan-2020

23 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes

by

Jordan Anthony Evans

A thesis submitted in conformity with the requirements for the degree of Master of Science

Department of Chemistry University of Toronto

© Copyright by Jordan Anthony Evans 2014

Page 2: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

ii

Zirconium and Palladium Catalyzed

Telescopic Synthesis of (E)-Alkenes

Jordan Anthony Evans

Master of Science

Department of Chemistry

University of Toronto

2014

Abstract

Alkenes are remarkably versatile motifs as they can be further functionalized by a vast array of

addition, reduction, and oxidation reactions. Thus their efficient synthesis is highly desired.

Over the past 35 years, the Suzuki-Miyaura cross-coupling reaction has emerged as a powerful

synthetic tool for the formation of carbon-carbon bonds. Herein is described the development of

a one-pot two-step protocol for the synthesis of (E)-alkenes comprising palladium-catalyzed

Suzuki-Miyaura cross-coupling of aryl or heteroaryl halides, including chlorides, with alkenyl

pinacolboronates, prepared in situ via solvent-free zirconium-catalyzed hydroboration of

terminal alkynes. Avoiding isolation of intermediates saves time and reduces waste. The regio-

and stereochemistry of the alkene is set by initial hydrozirconation of the alkyne. Addition of

water to the second step deactivates the zirconocene catalyst, which is otherwise deleterious to

cross-coupling. Thus this sequence exploits the water tolerance of the Suzuki-Miyaura reaction.

Page 3: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

iii

Acknowledgments

I would like to thank my supervisor, Mark Lautens, for giving me the tremendous opportunity to

study in his laboratory and for fostering such a stimulating and supportive environment. It was

truly a privilege to work among such thoughtful, talented, and self-directed people. Thank you

to my colleagues, especially Dave and Christine, for their friendship and willingness to share

their expertise. I could not have done this without you. Thank you to Mark Taylor for reviewing

this thesis. I would also like to thank my friends and family for their support.

Page 4: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

iv

Table of Contents

Acknowledgments .......................................................................................................................... iii

Table of Contents ........................................................................................................................... iv

List of Tables ................................................................................................................................. vi

List of Figures ............................................................................................................................... vii

List of Schemes ............................................................................................................................ viii

List of Abbreviations ..................................................................................................................... ix

1 Introduction ................................................................................................................................ 1

1.1 Domino Chemistry in Organic Synthesis ........................................................................... 1

1.1.1 Transition Metal-Catalysis in Domino Chemistry .................................................. 1

1.2 Suzuki-Miyaura Reaction in Organic Synthesis ................................................................. 5

1.2.1 Proposed Mechanism of the Suzuki-Miyaura Reaction .......................................... 5

1.2.2 Brief History of the Suzuki-Miyaura Reaction ....................................................... 8

1.2.3 Development and Activity of Dialkylbiaryl Phosphine Ligands for the Suzuki-

Miyaura Reaction .................................................................................................. 12

1.3 Preparation of Organoboron Coupling Partners for the Suzuki-Miyaura Reaction .......... 18

1.3.1 Rhodium-Catalyzed Hydroboration ...................................................................... 21

1.3.2 Zirconium-Catalyzed Hydroboration .................................................................... 25

1.4 Alkenes in Organic Synthesis ........................................................................................... 29

1.4.1 Bimetallic Catalyzed One-Pot Synthesis of Alkenes ............................................ 29

1.4.2 Palladium-Catalyzed Zirconium-Negishi Cross-Coupling Reaction .................... 30

1.4.3 Proposed Methodology ......................................................................................... 32

2 Results and Discussion ............................................................................................................. 33

2.1 Preparation of Alkenylboronates ...................................................................................... 33

2.1.1 Solvent-Free Zr-Catalyzed Hydroboration of Terminal Alkynes ......................... 33

2.1.2 Solvent-Free Zr-Catalyzed Hydroboration of Internal Alkynes ........................... 37

Page 5: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

v

2.1.3 Selecting a Solvent for Zr-Catalyzed Hydroboration ........................................... 38

2.2 Optimization of the Suzuki Cross-Coupling Reaction ...................................................... 40

2.2.1 Base and Solvent Screen ....................................................................................... 40

2.2.2 Further Optimization ............................................................................................. 41

2.3 Initial Conditions for One-Pot Two-Step Zr-Catalyzed Hydroboration/Pd-Catalyzed

Suzuki Coupling Sequence ............................................................................................... 43

2.3.1 Substrate Scope of Initial Conditions ................................................................... 43

2.4 Optimization of One-Pot Two-Step Zr-Catalyzed Hydroboration/Pd-Catalyzed Suzuki

Coupling Sequence ........................................................................................................... 46

2.4.1 Optimization with Aryl Bromides and Iodides ..................................................... 46

2.4.2 Extension to Aryl Chlorides .................................................................................. 47

2.5 Bimetallic Catalyzed Domino Attempt ............................................................................. 51

2.6 Optimized Conditions for One-Pot Two-Step Zr-Catalyzed Hydroboration/Pd-

Catalyzed Suzuki Coupling Sequence .............................................................................. 55

2.6.1 Substrate Scope of Optimized Conditions for Aryl Bromides .............................. 55

2.6.2 Substrate Scope of Optimized Conditions for Aryl Chlorides .............................. 57

2.6.3 Synthesis of (E)-2-Styrylbenzoxazole .................................................................. 59

2.7 Conclusions and Future Outlook ...................................................................................... 61

3 Experimental ............................................................................................................................ 63

3.1 General Considerations ..................................................................................................... 63

3.2 Synthesis of Starting Materials ......................................................................................... 65

3.3 Synthesis of Alkenylboronates ......................................................................................... 69

3.4 Synthesis of (E)-Alkenes .................................................................................................. 71

Appendix: Selected Spectra .......................................................................................................... 84

Page 6: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

vi

List of Tables

Table 1.2-1: Initial Results Reported by Suzuki and Miyaura ........................................................ 9

Table 1.2-2: Cross-Coupling of Heterocycles with Pd-XPhos ..................................................... 16

Table 1.3-1: Hydroboration of Alkynes with Pinacolborane ........................................................ 25

Table 1.3-2: Zirconium-Catalyzed Hydroboration of Alkynes with Pinacolborane ..................... 26

Table 2.1-1: Solvent-Free Zr-Catalyzed Hydroboration of Phenylacetylene ............................... 34

Table 2.1-2: Solvent-Free Zr-Catalyzed Hydroboration of N-Propargyl Indole .......................... 37

Table 2.1-3: Solvent-Free Zr-Catalyzed Hydroboration of Internal Alkynes ............................... 37

Table 2.1-4: Zr-Catalyzed Hydroboration of Phenylacetylene in Toluene ................................... 38

Table 2.1-5: Zr-Catalyzed Hydroboration of N-Propargyl Indole in Toluene .............................. 39

Table 2.2-1: Base and Solvent Screen for SMC ........................................................................... 40

Table 2.2-2: Further Optimization of SMC .................................................................................. 41

Table 2.3-1: Substrate Scope of Initial Conditions ....................................................................... 44

Table 2.4-1: Optimization with Aryl Bromides and Iodides ........................................................ 46

Table 2.4-2: Ligand Screen for Suzuki Coupling of Aryl Chlorides ............................................ 48

Table 2.4-3: Optimization with Aryl Chlorides using XPhos ....................................................... 48

Table 2.4-4: Optimization with Pd-XPhos-G2 ............................................................................. 50

Table 2.5-1: Effect of SMC Components on Zr-Catalyzed Hydroboration .................................. 52

Table 2.5-2: Effect of Cp2ZrHCl on Suzuki Coupling ................................................................. 54

Table 2.6-1: Substrate Scope of Optimized Conditions for Aryl Bromides ................................. 55

Table 2.6-2: Substrate Scope of Optimized Conditions for Aryl Chlorides ................................. 57

Table 2.6-3: Solvent and Base Screen for SMC with 2-Chlorobenzoxazole ................................ 59

Page 7: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

vii

List of Figures

Figure 1.1-1: Classification of Transition Metal Catalysis in Domino Reactions .......................... 2

Figure 1.2-1: Catalytic Cycle for the Suzuki-Miyaura Reaction .................................................... 6

Figure 1.2-2: Dialkylbiaryl Phosphine Ligands ............................................................................ 13

Figure 1.2-3: Stabilizing Interactions in Pd-SPhos Complexes .................................................... 15

Figure 1.2-4: Structural Features of Dialkylbiaryl Phosphine Ligand and Their Effect on the

Efficacy of Metal-Ligand Complexes in Catalysis ............................................................... 17

Figure 1.3-1: Organoborons Commonly Used in the Suzuki-Miyaura Reaction ......................... 18

Figure 1.3-2: Relationship Between Regioselectivity and Cyclic Transition State ...................... 20

Figure 1.3-3: Catalytic Cycle for the Rhodium-Catalyzed Hydroboration of Alkenes ................ 23

Figure 1.3-4: Catalytic Cycle for the Zirconium-Catalyzed Hydroboration of Alkynes .............. 27

Figure 1.3-5: Synthetic Applications of Zirconocene Hydrochloride (Schwartz’s Reagent) ....... 28

Figure 2.4-1: 2nd Generation XPhos Precatalyst .......................................................................... 49

Page 8: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

viii

List of Schemes

Scheme 1.1-1: Domino Formation of Dihydroquinolines .............................................................. 3

Scheme 1.1-2: Domino Formation of Aza-Dihydrodibenzoxepines .............................................. 4

Scheme 1.2-1: Possible Roles of Base in Transmetallation ............................................................ 7

Scheme 1.2-2: B-Alkyl Suzuki-Miyaura Reaction ....................................................................... 10

Scheme 1.2-3: Alkyl-Alkyl SMC of Alkyl Bromides that Possess β-Hydrogens ........................ 11

Scheme 1.2-4: Nickel-Catalyzed Alkyl-Alkyl SMC of Unactivated Secondary Bromides ......... 12

Scheme 1.2-5: Highly Fluoronated Aromatics via Pd-SPhos Catalyzed SMC ............................ 14

Scheme 1.2-6: Pd-SPhos Catalyzed SMC in Synthesis of Quinine and Quinidine ...................... 14

Scheme 1.3-1: Chemoselectivity of Uncatalyzed vs. Catalyzed Hydroboration .......................... 22

Scheme 1.3-2: Preparation of Chiral Alcohols ............................................................................. 24

Scheme 1.4-1: [Rh-Pd]-Catalyzed Hydrosilylation/Hiyama Cross-Coupling Sequence .............. 29

Scheme 1.4-2: Platinum-Catalyzed Diboration/Palladium-Catalyzed SMC Sequence ................ 29

Scheme 1.4-3: Hydrozirconation/Palladium-Catalyzed Cross-Coupling Sequence ..................... 30

Scheme 1.4-4: Pd-Catalyzed Zr-Negishi Coupling Toward Total Synthesis of (-)-Motuporin ... 31

Scheme 1.4-5: Coupling of Alkenylzirconocenes with Alkyl Halides ......................................... 31

Scheme 1.4-6: Proposed Methodology ......................................................................................... 32

Scheme 2.3-1: Initial Conditions for One-Pot Two-Step Sequence ............................................. 43

Scheme 2.5-1: Domino Attempt ................................................................................................... 51

Scheme 2.5-2: One-Pot Two-Step under Anhydrous Conditions ................................................. 53

Scheme 2.6-1: One-Pot Two-Step Synthesis of (E)-2-Styrylbenzoxazole ................................... 60

Page 9: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

ix

List of Abbreviations

Ac acetyl

acac acetylacetonate

Ar aryl

2-BuOH 2-butanol

9-BBN 9-borabicyclo[3.3.1]nonane

BINAP 2,2’-bis(diphenylphosphino)-1-1’-binapthyl

Bmim 1-butyl-3-methylimidazolium

BMS borane dimethylsulfide

Boc tert-butyloxycarbonyl

n-Bu n-butyl

t-Bu tert-butyl

n-BuLi n-butyllithium

n-BuOH n-butanol

cat catalyst or catalytic or catecholato

Cbz carbobenzyloxy

cod 1,5-cyclooctadiene

Cp cyclopentadienyl

Cy cyclohexyl

DART Direct Analysis in Real Time

dba dibenzylideneacetone

DCE dichloroethane

DCM dichloromethane

DI deionized

dioxane 1,4-dioxane

DME dimethoxyethane

DMF dimethylformamide

dppb 1,4-bis(diphenylphosphino)butane

dppf 1,1’-bis(diphenylphosphino)ferrocene

ee enantiomeric excess

ESI electrospray ionization

Page 10: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

x

equiv equivalent(s)

Et ethyl

EtOAc ethyl acetate

GC gas chromatography

h hour(s)

HBpin pinacolborane, 4,4,5,5-tetramethyl-1,3,2-dioxaborolane

HBcat catecholborane, 1,3,2-benzodioxaborole

HetAr heteroaryl

HRMS high resolution mass spectrometry

Hz hertz

IR infrared

L generic ligand

LDA lithium diisopropylamide

M generic metal or molar concentration

m meta

Me methyl

MHz megahertz

MIDA methyliminodicarbonic acid

min minute(s)

mol mole

MP melting point

Ms mesyl, methanesulfonyl

MW microwave irradiation

N/A not applicable

nm nanometres

NMP N-methyl-2-pyrrolidone

NMR nuclear magnetic resonance

o ortho

OTf triflate, trifluoromethanesulfonate

P product

p para

Ph phenyl

Page 11: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

xi

PhMe toluene

pin pinacolato

ppm parts per million

R generic group

QPhos 1,2,3,4,5-pentaphenyl-1’-(di-tert-butylphosphino)ferrocene

QUINAP 1-(2-diphenylphosphino-1-naphthyl)isoquinoline

rt room temperature

RuPhos 2-dicyclohexylphosphino-2’,6’-diisopropoxybiphenyl

SM starting material

SMC Suzuki-Miyaura cross-coupling

SPhos 2-dicyclohexylphosphino-2’,6’-dimethoxybiphenyl

TBAF tetra-n-butylammonium fluoride

Temp temperature

THF tetrahydrofuran

TLC thin-layer chromatography

TMB 1,3,5-trimethoxybenzene

Ts tosyl, p-toluenesulfonyl

UV ultraviolet

X generic halide

XPhos 2-dicyclohexylphosphino-2’,4’,6’-triisopropylbiphenyl

XPhos-Pd-G2 2nd Generation XPhos Precatalyst, Chloro(2-dicyclohexylphosphino-2′,4′,6′-

triisopropyl-1,1′-biphenyl)[2-(2′-amino-1,1′-biphenyl)]palladium(II)

Y generic group

Page 12: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

1

1 Introduction

1.1 Domino Chemistry in Organic Synthesis

That “it is better to prevent waste in the first place than to treat or clean it up afterwards” is one

of the basic tenets of green chemistry.1 To this end, domino chemistry is an ideal approach, as

multiple bond forming events occur under a single set of reaction conditions, rapidly generating

complex structures without the need to isolate intermediates. This saves time and reduces waste,

making it attractive to industry; thus a wide variety of domino processes have been developed. 2

1.1.1 Transition Metal-Catalysis in Domino Chemistry

Domino chemistry is a “chemical philosophy” that takes its cue from nature, wherein enzymes

catalyze the formation of multiple bonds in a single step within the highly complex environment

of a cell. Although not nearly as efficient, organic synthesis does have its advantages, such as

the ability to use reagents and catalysts that are unavailable to and/or incompatible with

biological systems. For example, transition metal catalysts can facilitate transformations that are

perhaps impossible in nature.

In 2001, Poli and co-workers3 introduced a classification system for transition metal-catalyzed

domino reactions (Figure 1.1-1), differentiating between “pure” and “pseudo” domino processes.

In a “pure” process, a single catalyst conducts a domino sequence within a single catalytic cycle;

whereas in a “pseudo” process, either a single (Type I) or multiple catalysts (Type II) conduct

two or more transformations in separate catalytic cycles with the formation of discrete

intermediates. Under the classification system put forth by MacMillan and co-workers,4 a

bimetallic “pseudo” domino process is equivalent to “cascade catalysis.” Such processes, in

which two transition metal catalysts operate independently in a domino fashion, are challenging

for a few reasons: chemoselectivity is often an issue as different catalysts can react differently

with different functional groups; additionally, redox reactions or the exchange of supporting

1 Anastas, P. T.; Warner, J. C. Green Chemistry: Theory and Practice; Oxford University Press: New York, 2000.

2 For reviews on domino chemistry, see: a) Tietze, L. F.; Beifuss, U. Angew. Chem. Int. Ed. 1993, 32, 131. b) Tietze,

L. F. Chem. Rev. 1996, 96, 115. c) Tietze, L. F.; Brasche, G.; Gericke, K. M. Domino Reactions in Organic

Synthesis; Wiley-VCH: Weinheim, 2006. 3 Poli, G.; Giambastiani, G. J. Org. Chem. 2002, 67, 9456.

4 Allen, A. E.; MacMillan, D. W. C. Chem. Sci. 2012, 3, 633.

Page 13: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

2

ligands between two metals can lead to deactivation of one or both catalysts. Nevertheless,

examples of bimetallic cascade catalysis have been reported,5 although one-pot two-step

protocols, in which the second catalyst is added upon completion of the first, are more common.

Figure 1.1-1: Classification of Transition Metal Catalysis in Domino Reactions

Our group has also contributed to the field of bimetallic cascade catalysis. Dihydroquinolines

were prepared via a rhodium-catalyzed hydroarylation/palladium-catalyzed C-N coupling

sequence employing a ([Rh(cod)OH]2/BINAP) / (Pd(OAc)2/XPhos) catalyst system (Scheme

1.1-1).6 Based on

31P NMR spectroscopy, it was revealed that while palladium could bind to both

phosphine ligands, rhodium did not bind XPhos to a measurable degree. This was crucial to the

5 For selected examples, see: a) Zimmermann, B.; Herwig, J.; Beller, M. Angew. Chem. Int. Ed. 1999, 38, 2372. b)

Crossy, J.; Bargiggia, F.; BouzBouz, S. Org. Lett. 2003, 5, 459. c) Ko, S.; Lee, C.; Choi, M. –G.; Na, Y.; Chang, S.

J. Org. Chem. 2003, 68, 1607. d) Kammerer, C.; Prestat, G.; Gaillard, T.; Madec, D.; Poli, G. Org. Lett. 2008, 10,

405. e) Cernak, T. A.; Lambert, T. H. J. Am. Chem. Soc. 2009, 131, 3124. f) Takahashi, K.; Yamashita, M.; Ichihara,

T.; Nakano, K.; Nozaki, K. Angew. Chem. Int. Ed. 2010, 49, 4488. 6 Panteleev, J.; Zhang, L.; Lautens, M. Angew. Chem. Int. Ed. 2011, 50, 9089.

Page 14: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

3

success of the domino reaction as Pd-BINAP was inactive in C-N coupling; therefore, any ligand

exchange between the two metals would have been deleterious. Indeed, when BINAP loading

exceeded 5.5 mol%, the reaction stalled at the arylated intermediate. This could not be remedied

by addition of excess XPhos as this also resulted in diminished yields, likely due to saturation of

the coordination sites on the palladium sphere. Thus a fine balancing of catalysts and ligands

was necessary to furnish the desired products in domino fashion.

Scheme 1.1-1: Domino Formation of Dihydroquinolines

More recently, our group reported the synthesis of aza-dihydrodibenzoxepines via a rhodium-

catalyzed hydroarylation/palladium-catalyzed C-O coupling sequence (Scheme 1.1-2).7

Arylation proceeded very rapidly, going to completion within a few minutes at room

temperature, whereas C-O coupling required higher temperatures and longer reaction times.

Domino reactivity was affected by pyridine electronics. While electron-poor pyridines reacted

smoothly in a domino fashion, electron-rich pyridines gave mixtures of product and arylated

intermediate. An asymmetric variant of this reaction was developed using β-substituted alkenyl

pyridines and a chiral diene ligand.

From this work, it is clear that thoughtful planning and careful optimization of reaction

parameters can lead to the development of domino reactions in which two metals operate without

interference. Given our group’s continued interest in this field, we endeavored to develop a one-

pot bimetallic catalyzed hydroboration/Suzuki-Miyaura cross-coupling sequence.

7 Friedman, A. A.; Panteleev, J.; Tsoung, J.; Huynh, V.; Lautens, M. Angew. Chem. Int. Ed. 2013, 52, 9755.

Page 15: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

4

Scheme 1.1-2: Domino Formation of Aza-Dihydrodibenzoxepines

Page 16: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

5

1.2 Suzuki-Miyaura Reaction in Organic Synthesis

Carbon-carbon bond formation is arguably the most fundamental transformation in organic

synthesis. Over the past four decades, cross-coupling reactions including the Mizoroki-Heck,8

Stille,9 Negishi coupling,

10 Suzuki-Miyaura,

11 and Hiyama coupling

12 reactions have emerged as

powerful tools for the chemo-, regio-, and stereoselective synthesis of carbon-carbon bonds.

Among these, the Suzuki-Miyaura cross-coupling (SMC) reaction, defined as the transition

metal-catalyzed coupling of an organoboron reagent with an organic halide, is arguably the most

useful for several reasons: relatively mild and versatile reaction conditions; low catalyst loading;

wide scope and functional group tolerance; commercial availability as well as ease of preparation

and handling of diverse boronic acid and ester coupling partners that are environmentally safer

than many other organometallic reagents; low toxicity and relative ease of removal of boron-

derived by-products; and tolerance of water. Given these advantages as well as its continuous

and extensive development, SMC has had a tremendous impact on organic synthesis.13

1.2.1 Proposed Mechanism of the Suzuki-Miyaura Reaction14

The proposed mechanism of the palladium(0)-catalyzed Suzuki cross-coupling of an aryl halide

with a generic aryl, alkenyl, or alkyl boronic acid or boronate in the presence of hydroxide as a

representative base is depicted below (Figure 1.2-1). As with other cross-coupling reactions, the

catalytic cycle is thought to begin with oxidative addition of the aryl halide to palladium(0),

which proceeds in a cis fashion, to form an arylpalladium(II) halide intermediate (A) that rapidly

8 For a recent review on the Mizoroki-Heck Reaction, see: de Meijere, A., Bräse, S. In Metal-Catalyzed Cross-

Coupling Reactions, Second Edition; Diederich, F., de Meijere, A., Eds.; Wiley-VCH: New York, 2008; Chapter 5. 9 For a recent review on the Stille reaction, see: Mitchell, T. In Metal-Catalyzed Cross-Coupling Reactions, Second

Edition; Diederich, F., de Meijere, A., Eds.; Wiley-VCH: New York, 2008; Chapter 3. 10

For a recent review on Negishi coupling, see: Negishi, E., Zeng, X., Tan, Z., Qian, M., Hu, Q., Huang, Z. In

Metal-Catalyzed Cross-Coupling Reactions, Second Edition; Diederich, F., de Meijere, A., Eds.; Wiley-VCH: New

York, 2008; Chapter 15. 11

For reviews on the Suzuki-Miyaura reaction, see: a) Miyaura, N.; Suzuki, A. Chem. Rev. 1995, 95, 2457. b)

Suzuki, A. J. Organomet. Chem. 1999, 576, 147. c) Bellina, F.; Carpita, A.; Rossi, R. Synthesis 2004, 15, 2419. d)

Miyaura, N. In Metal-Catalyzed Cross-Coupling Reactions, Second Edition; Diederich, F., de Meijere, A., Eds.;

Wiley-VCH: New York, 2008; Chapter 2. 12

For a recent review on Hiyama coupling, see: Denmark, S. E., Sweis, R. F. In Metal-Catalyzed Cross-Coupling

Reactions, Second Edition; Diederich, F., de Meijere, A., Eds.; Wiley-VCH: New York, 2008; Chapter 4. 13

a) Kotha, S.; Lahiri, K.; Kashinath, D. Tetrahedron 2002, 58, 9633. b) For a recent review on the application of

SMC to total synthesis, see: Heravi, M. M.; Hashemi, E. Tetrahedron 2012, 68, 9145. 14

a) Suzuki, A. Pure & Appl. Chem. 1985, 57, 1749. b) Miyarua, N. J. Organomet. Chem. 2002, 653, 54.

Page 17: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

6

isomerizes to a more stable trans configuration (B).15

Rate-determining16

transmetallation with

the boronic acid or ester followed by reductive elimination from the resulting di-

organopalladium(II) complex (C) gives the desired coupled product (D) and regenerates

palladium(0). Isomerization to the cis isomer is necessary before reductive elimination can

occur. If R = alkyl, then competing β-hydride elimination can occur to give an undesired

dehalogenation product.

Figure 1.2-1: Catalytic Cycle for the Suzuki-Miyaura Reaction

Transmetallation proceeds with retention of stereochemistry set by both the organoboron and

halide,17

indicating a four-centered hydroxo μ2-bridged transition state (Scheme 1.2-1, TS);

however, the mechanism by which this model arises (path A or B) is unresolved.18

Organoborons are highly inert to RPd(II)X and do not undergo transmetallation in the absence of

15 Casado, A. L.; Espinet, P. Organometallics 1998, 17, 954.

16 For a computational study of the transmetallation process in SMC, see: Braga, A. C. C.; Morgon, N. H.; Ujaque,

G.; Lledós, A.; Maseras, F. J. Organomet. Chem. 2006, 691, 4459. 17

Ridgway, B. H.; Woerpel, K. A. J. Org. Chem. 1998, 63, 458. 18

Matos, K.; Soderquist, J. A. J. Org. Chem. 1998, 63, 461.

Page 18: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

7

base; however, boron “ate” complexes (e.g. [ArBF3]K)19

readily undergo coupling with organic

electrophiles, indicating that quaternization of the boron centre enhances nucleophilicity of the

organic group. Therefore, R2B(OH)3

− may undergo transmetallation with R

1Pd(II)X via a

hydroxyborate-palladium complex (path A). Alternatively, a hydroxo-, alkoxo-, or

(acetoxo)Pd(II) complex may form via ligand exchange between R1Pd(II)X and base. Such

complexes readily undergo transmetallation with organoborons in the absence of base;20

therefore, the transition state may arise via coordination of the OH ligand to the boron sphere of

R2B(OH)2 (path B), a process which may be rate limiting.

21 The reactivity of R

1PdOH can be

attributed to the high basicity of such species22

as well as the high oxophilicity of the boron

centre. Soderquis and Matos18

postulated that palladium-catalyzed SMC of 9-alkyl-9-BBN with

iodobenzene in aqueous KOH proceeds via path A, whereas the less Lewis acidic 9-oxa-10-

borabicyclo[3.3.2]decane proceeds via path B, suggesting that the Lewis acidity of the

organoboron reagent may determine, at least in part, which path predominates.

Scheme 1.2-1: Possible Roles of Base in Transmetallation

More recently, Amatore et al.23

investigated the role of hydroxide in SMC and made several

interesting observations favouring path B: trans-[ArPdBr(PPh3)2] did not react with Ar’B(OH)2

in the absence of hydroxide, but trans-[ArPd(OH)(PPh3)2] did; furthermore, a large excess of

hydroxide inhibited the latter reaction, indicating that Ar’B(OH)3− does not undergo

19 Darses, S.; Genet, J. –P.; Brayer, J. –L.; Demoute, J. –P. Tetrahedron Lett. 1997, 38, 4393.

20 Miyaura, N.; Yamada, K.; Suginome, A.; Suzuki A. J. Am. Chem. Soc. 1985, 107, 972.

21 Moriya, T.; Miyaura, N.; Suzuki, A. Synlett 1994, 149.

22 Otsuka, S. J. Organomet. Chem. 1980, 191, 200.

23 Amatore, C.; Jutand, A.; Le Duc, G. Chem. Eur. J. 2011, 17, 2492.

Page 19: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

8

transmetallation with ArPd(OH)(PPh3)2. Additionally, PhB(OH)3− did not react with [(p-NC-

C6H4)PdBr(PPh3)2] (in the presence of excess bromide ions to quell the formation of the

hydroxopalladium species), disfavouring path A. In addition to its crucial role in the rate-

determining transmetallation process, the authors also observed that hydroxide increased the rate

of reductive elimination.

1.2.2 Brief History of the Suzuki-Miyaura Reaction

Akira Suzuki and Norio Miyaura introduced their eponymous cross-coupling reaction in 1979.24

In their seminal reports, the authors described the tetrakis(triphenylphosphine)palladium(0)-

catalyzed coupling of alkenylboronates with aryl, heteroaryl, and alkenyl halides in the presence

of a strong base to yield (E)-alkenes and conjugated alkadienes (Table 1.2-1). The authors noted

several features that have become characteristic of SMC; namely, that the reaction proceeded

with retention of configuration with respect to the alkenylboronate, the expected coupled product

was not obtained in the absence of base, Lewis bases such as triethylamine did not facilitate the

reaction, and electron poor aryl halides were more reactive than their electron rich analogues,

which is consistent with the effect of electron density on oxidative addition of aryl halides to

palladium(0).25

The authors also noted that aryl iodides were more reactive than aryl bromides

and that aryl chlorides, such as chlorobenzene (entry 6), were inert under their conditions. These

findings are in agreement with those of Fu and co-workers,26

who conducted chemoselective

Suzuki reactions with substrates bearing more than one halide/triflate and determined the order

of reactivity to be I > Br > OTf >> Cl under their catalyst system. Other reports have reversed

the reactivity of bromides and triflates.11

24 For pioneering work by Suzuki and Miyaura, see: a) Miyaura, N.; Yamada, K.; Suzuki, A. Tetrahedron Lett.

1979, 20, 3437. b) Miyaura, N.; Suzuki, A. J. Chem. Soc., Chem. Commun. 1979, 866. c) Miyaura, N.; Yano, T.;

Suzuki, A. Tetrahedron Lett. 1980, 21, 2865. d) Miyaura, N.; Yanagi, T.; Suzuki, A. Synth. Commun. 1981, 11, 513.

e) Miyaura, N.; Suzuki, A. J. Organomet. Chem. 1981, 213, C53. 25

Stille, J. K.; Lau, K. S. Y. Acc. Chem. Res. 1977, 10, 434. 26

Littke, A. F.; Dai, C.; Fu, G. C. J. Am. Chem. Soc. 2000, 122, 4020.

Page 20: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

9

Table 1.2-1: Initial Results Reported by Suzuki and Miyaura

Entry R-X mol% Pd Time (h) Yield (%)a

1 1 2 80

2

1 2 80

3

1 2 81

4 PhI 1 2 100

5 PhBr 1 4 98

6 PhCl 1 2 3

7 o-MePhBr 3 4 93

8 o-OMePhBr 3 4 81

9 p-ClPhBr 1 3 100

10

1 2 83

a Yield based on alkenyl or aryl halide and determined by GC.

Page 21: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

10

In the decade following their initial report of sp2-sp

2 cross-coupling, Suzuki and co-workers

27

reported the first palladium(0)-catalyzed inter- and intramolecular coupling of an sp2 hybridized

iodide or bromide with an sp3 hybridized primary or secondary organoborane, generated in situ,

and coined it the B-alkyl Suzuki-Miyaura cross-coupling reaction (Scheme 1.2-2).28

Scheme 1.2-2: B-Alkyl Suzuki-Miyaura Reaction

The key to the success of these sp2-sp

3 cross-couplings was careful choice of the catalyst,

dichloro[1,1’-bis(diphenylphosphino)ferrocene]palladium(II) [PdCl2(dppf)].29

As mentioned

above (1.2.1), the problem with coupling organometallic reagents that possess β-hydrogens is the

tendency of alkylpalladium complexes to undergo β-hydride elimination instead of reductive

elimination, generating a hydridopalladium species that reduces the halide. The [PdCl2(dppf)]

catalyst suppresses this undesired side-reaction as the large “bite angle” of the bidentate dppf

ligand forces the alkyl group and its coupling partner closer together about the square planar

Pd(II) complex, enforcing a cis geometry and promoting reductive elimination.30

Suzuki and co-

workers reported that secondary alkylboranes were significantly less reactive than primary

alkylboranes under their conditions, likely due to the slow rate of transmetallation of the former.

In 2008, Crudden and co-workers31

reported that secondary alkylboronate pinacol esters could be

27 a) Miyaura, N.; Ishiyama, T.; Ishikawa, M.; Suzuki, A. Tetrahedron Lett. 1986, 27, 6369. b) Miyarua, N.;

Ishiyama, T.; Sasaki, H.; Ishikawa, M.; Satoh, M.; Suzuki, A. J. Am. Soc. Chem. 1989, 111, 314. b) Sato, M.;

Suzuki, A. Chem. Lett. 1989, 1405. 28

For a review on B-alkyl SMC and its application in the total synthesis of natural products, see: Chemler, S. R.;

Trauner, D.; Danishefsky, S. J. Angew. Chem. Int. Ed. 2001, 40, 4544; and related references therein. 29

Hayahsi, T.; Konishi, M.; Kobori, Y.; Kumada, M.; Higuchi, T.; Hirotsu, K. J. Am. Chem. Soc. 1984, 106, 158. 30

Brown, J. M.; Guiry, P. J. Inorg. Chim. Acta. 1994, 220, 249. 31

Imao, D.; Glasspoole, B. W.; Laberge, V. S.; Crudden, C. M. J. Am. Chem. Soc. 2009, 131, 5024.

Page 22: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

11

successfully coupled with aryl iodides and bromides by employing silver oxide (Ag2O) as base,

which is believed to activate organoboron species toward transmetallation.32

In 2001, Fu and co-workers33

reported the first Suzuki coupling of primary alkyl bromides that

possess β-hydrogens by employing the bulky and electron rich tricyclohexylphosphine ligand

(PCy3) (Scheme 1.2-3). Prior to this, sp3 hybridized halides had proven to be challenging

substrates for SMC due to their slow rate of oxidative addition to palladium(0) and (if the alkyl

palladium(II)halide complex formed at all) the proclivity of β-hydride elimination to outcompete

cross-coupling. The authors reported that the reactions proceed under mild conditions at room

temperature. The following year, the Fu group extended their methodology to alkyl chlorides

using the same catalyst system, although higher reaction temperatures (90 °C) were required.34

Scheme 1.2-3: Alkyl-Alkyl SMC of Alkyl Bromides that Possess β-Hydrogens

Over the past several years, organic chemists have become increasingly interested in the use of

nickel catalysts for the Suzuki reaction. The allure of nickel is owed to its cheapness and earth-

abundance relative to palladium, as well as the broad range of traditionally challenging organic

electrophiles that can be successfully coupled in the presence of the appropriate nickel complex.

For example, Fu and co-workers have developed effective nickel-catalyzed methods for coupling

unactivated alkyl bromides and iodides with aryl35

and even alkylboranes (Scheme 1.2-4).36

Despite these advantages, the practicality of nickel-catalyzed SMC on an industrial scale is

presently limited as relatively high catalyst loadings and excess of external supporting ligands

are often required; furthermore, many nickel catalysts are air and moisture sensitive.37

32 a) Uenishi, J.; Beau, J. M.; Armstrong, R. W.; Kishi, Y. J. Am. Chem. Soc. 1987, 109, 4756. b) Hirabayashi, K.;

Kawashima, J.; Nishihara, Y.; Mori, A.; Hiyama, T. Org. Lett. 1999, 1, 299. 33

Netherton, M. R.; Dai, C.; Neuschütz, Fu, G. C. J. Am. Chem. Soc. 2001, 123, 10099. 34

Kirchhoff, J. H.; Dai, C.; Fu, G. C. Angew Chem. Int. Ed. 2002, 41, 1945. 35

a) Zhou, J.; Fu, G. C. J. Am. Chem. Soc. 2004, 126, 1340. b) González-Bobes, F.; Fu, G. C. J. Am. Chem. Soc.

2006, 128, 5360. 36

Saito, B.; Fu, G. C. J. Am. Chem. Soc. 2007, 129, 9602. 37

For a recent review on Ni-catalyzed SMC, see: Han, F. –S. Chem. Soc. Rev. 2013, 42, 5270.

Page 23: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

12

Scheme 1.2-4: Nickel-Catalyzed Alkyl-Alkyl SMC of Unactivated Secondary Bromides

1.2.3 Development and Activity of Dialkylbiaryl Phosphine Ligands for the Suzuki-Miyaura Reaction38

Prior to 1998, effective Suzuki coupling of electron-neutral and -rich aryl chlorides had not been

achieved.39

This was viewed as a serious limitation as aryl chlorides are the cheapest and most

readily available aryl halides, making them especially attractive to industry.40

The recalcitrance

of aryl chlorides can be attributed to the strength of the C-Cl bond (bond dissociation energies

for Ph-X at 298 K are: Cl: 96 kcal/mol; Br: 81 kcal/mol; I: 65 kcal/mol).41

Traditionally

employed palladium(0) catalysts supported by triarylphosphine ligands (e.g. Pd(PPh3)4) are not

sufficiently electron-rich to promote oxidative addition of aryl chlorides.42

Thus there was an

impetus to develop catalysts for the Suzuki reaction that could efficiently couple aryl chlorides.

In 1998, Buchwald and co-workers, through the course of developing dialkylbiaryl phosphine

ligands for C-N coupling,43

found that DavePhos (Figure 1.2-2, L1), in addition to being

effective for aminations, was also an effective supporting ligand for palladium-catalyzed Suzuki

couplings of aryl chlorides, including unactivated aryl chlorides.44

In the same year, Fu and co-

38 For a comprehensive overview of dialkylbiaryl phosphine ligands, see: Martin, R.; Buchwald, S. L. Acc. Chem.

Res. 2008, 41, 1461. 39

Activated, electron-poor aryl chlorides, particularly heteroaryl chlorides, have long been known to be suitable for

SMC. For a review on Pd-catalyzed reactions of heterocycles, see: Kalinin, V. N. Synthesis 1992, 413. 40

For a review on Pd-catalyzed coupling reactions of aryl chlorides, see: Littke, A. F.; Fu, G. C. Angew. Chem. Int.

Ed. 2002, 41, 4176; and related references therein. 41

Grushin, V.; Alper, H. Chem Rev. 1994, 94, 1047 42

It is well known that oxidative addition is faster with the use of electron-rich phosphine ligands. For a report, see:

Portnoy, M.; Milstein, D. Organometallics 1993, 12, 1665. 43

For recent examples of Pd-catalyzed amidations, see: a) Ikawa, T.; Barder, T. E.; Biscoe, M. R.; Buchwald, S. L.

J. Am. Chem. Soc. 2007, 129, 13001. b) Biscoe, M. R.; Barder, T. E.; Buchwald, S. L. Angew. Chem. Int. Ed. 2007,

46, 7232. 44

Old, D. W.; Wolfe, J. P.; Buchwald, S. L. J. Am. Chem. Soc. 1998, 120, 9722.

Page 24: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

13

workers reported efficient Suzuki coupling of both unactivated and hindered aryl chlorides using

the sterically encumbered and electron-rich Pd(PtBu3)2 catalyst.45

Figure 1.2-2: Dialkylbiaryl Phosphine Ligands

Buchwald and co-workers46

found that palladium catalysts with JohnPhos (L2) were significantly

more reactive than those with DavePhos (L1) for Suzuki couplings of both aryl bromides and

chlorides, indicating the dimethylamino group of DavePhos (L2) was not necessary for effective

catalysis. While CyJohnPhos (L3) provided a very active catalyst system for the coupling of

hindered substrates, JohnPhos (L2), its di-tert-butyl analogue, gave better results at room

temperature.47

The superior activity of catalysts supported by sterically demanding JohnPhos

(L2) was attributed to increased concentrations of monoligated complexes, L1Pd(0) and

L1Pd(II)(Ar)X, throughout the catalytic cycle (relative to the corresponding diligated species), as

oxidative addition of aryl halides to monoligated palladium(0) complexes is known to be much

faster than it is to more coordinatively crowded species.48

The reason for this reactivity is

intuitive: L1Pd(0) is smaller than L2Pd(0) and can therefore get closer to the aryl halide, which is

45 Littke, A. F.; Fu, G. C. Angew. Chem. 1998, 110, 3586; Angew. Chem. Int. Ed. 1998, 37, 3387.

46 Wolfe, J. P.; Buchwald, S. L. Angew. Chem. Int. Ed. 1999, 38, 2413.

47 Wolfe, J. P.; Singer, R. A.; Yang, B. H.; Buchwald, S. L. J. Am. Chem. Soc. 1999, 121, 9550.

48 a) Hartwig, J. F.; Paul, F. J. Am. Chem. Soc. 1995, 117, 5373. b) Barrios-Landeros, F.; Hartwig, J. F. J. Am.

Chem. Soc. 2005, 127, 6944.

Page 25: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

14

crucial for oxidatively adding aryl chlorides to palladium centers. The authors proposed that

transmetallation was also faster with monoligated complexes for the same reason.

At this point in the development of dialkylbiaryl ligands, a catalyst system for the effective

coupling of hindered arenes remained elusive. A further tuning of steric and electronic

properties led to the development of SPhos (L4), which proved to be the most universal ligand

developed to date. Palladium catalysts derived from SPhos (L4) exhibited unprecedented

activity, efficiently coupling ortho, ortho’-substituted aryl halides with ortho-substituted boronic

acids at exceedingly low catalyst loadings.49

Electron-poor boronic acids, which are traditionally

reluctant coupling partners due to their low nucleophilicity50

and susceptibility to metal-

catalyzed protodeboronation,51

were also effectively coupled (Scheme 1.2-5).52

Scheme 1.2-5: Highly Fluoronated Aromatics via Pd-SPhos Catalyzed SMC

SPhos (L4) has also seen extensive application in the synthesis of natural products.38

Shortly

after its development, SPhos (L4) was used by Jacobsen in the first catalytic asymmetric total

syntheses of the anitpyretic quinine and its enantiomer, quinidine (Scheme 1.2-6).53

Scheme 1.2-6: Pd-SPhos Catalyzed SMC in Synthesis of Quinine and Quinidine

49 Walker, S. D.; Barder, T. E.; Martinelli, J. R.; Buchwald, S. L. Angew. Chem. Int. Ed. 2004, 43, 1871. b)

50 Wong, M. S.; Zhang, X. L. Tetrahedron Lett. 2001, 42, 4087.

51 Kuivila, H. G.; Reuwer, J. F.; Mangravite, J. A. J. Am. Chem. Soc. 1964, 86, 2666.

52 Barder, T. E.; Walker, S. D.; Martinelli, J. R.; Buchwald, S. L. J. Am. Chem. Soc. 2005, 127, 4685.

53 Raheem, I. T.; Goodman, S. N.; Jacobsen, E. N. J. Am. Chem. Soc. 2004, 126, 706.

Page 26: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

15

Buchwald and co-workers attributed the longevity of catalysts supported by SPhos (L4) to the

stabilization of palladium intermediates via favourable interactions between the aryl ring and the

metal sphere (Figure 1.2-3). X-ray crystallography revealed that the SPhos (L4)/Pd(0) complex

possessed a Pd(0) η1-arene interaction with the ipso carbon of the bottom ring (A),

49 which may

confer stability.54

Based on computational studies,55

the most energetically favourable structures

for oxidative addition intermediates possessed either a Pd(II)-ipso carbon interaction (B) or a

Pd(II)-O interaction with either of the o-methoxy groups on the bottom ring (C). Compared to

other dialkylbiaryl phosphine ligands, this Pd-O interaction may further stabilize intermediates

throughout the catalytic cycle. In addition to stability, these Pd-arene interactions may also

improve reactivity by increasing steric bulk. Buchwald and co-workers also suggested that

monoligation may be directed by Pd-ipso carbon interactions, as they believed diligated species

would be too large to accommodate these favourable interactions.

Figure 1.2-3: Stabilizing Interactions in Pd-SPhos Complexes

The efficiency of Pd-SPhos (L4) in the coupling of aryl chlorides is the result of a fine balance

struck between ligand size and ability to enforce monoligation throughout the catalytic cycle.

Ligand L5 is substantially less electron-rich than SPhos (L4), but is nearly as effective as a

supporting ligand in the coupling of hindrered substrates at low temperatures. This implies that

while the electron-donating capacity of the phosphorus centre is important, it is secondary to size

for this ligand class.52

While the ortho, ortho-substituents on the bottom ring further increase the

size of the ligand, they also dramatically improve stability by preventing cyclometallation.56

54 Kočovosky, P.; Vyskočil, S.; Císařová, I.; Sejbal, J.; Tišlerová, I.; Smrčina, M.; Lloyd-Jones, G. C.; Stephen, S.

C.; Butts, C. P.; Murray, M.; Langer, V. J. Am. Chem. Soc. 1999, 121, 7714. 55

Barder, T. E.; Biscoe, M. R.; Buchwald, S. L. Organometallics 2007, 26, 2183. 56

Streiter, E. R.; Buchwald, S. L. Angew. Chem. Int. Ed. 2006, 45, 925.

Page 27: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

16

Nitrogen heterocycles are ubiquitous motifs in biologically and medicinally active compounds

and their efficient synthesis is highly desired; however, heteroaryl boronic acids have proven to

be difficult substrates in cross-coupling reactions, limiting their application in drug

development.57

While Pd-SPhos (L4) effectively coupled 3-thiophene boronic acid with

heteroaryl bromides, low yields were obtained with heteroaryl chlorides. The slow rate of

oxidative addition of heteroaryl chlorides combined with the tendency of 3-thiophene boronic

acids to decompose over time via protodeboronation makes such couplings particularly

challenging; however, Pd-XPhos (L6) proved effective.58

Pd-XPhos (L6) further proved itself to

be meritorious by efficiently coupling thiophene and pyridylboronic acids with a myriad of

heteroaryl chlorides, including conventionally challenging chloroaminopyridines (Table 1.2-2),

providing a fairly general method for the synthesis of heterobiaryls and increasing the scope of

the Suzuki-Miyaura reaction.

Table 1.2-2: Cross-Coupling of Heterocycles with Pd-XPhos

57 For a review on the synthesis and application of heterocyclic boronic acids, see: Tyrell, E.; Brookes, P. Synthesis

2003, 4, 469. 58

Billingsley, K.; Buchwald, S. L. J. Am. Chem. Soc. 2007, 129, 3358.

Page 28: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

17

Figure 1.2-4: Structural Features of Dialkylbiaryl Phosphine Ligand and Their Effect on

the Efficacy of Metal-Ligand Complexes in Catalysis

Page 29: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

18

1.3 Preparation of Organoboron Coupling Partners for the Suzuki-Miyaura Reaction

A variety of organoboron compounds are used as coupling partners for organic halides in the

Suzuki-Miyaura cross-coupling reaction (Figure 1.3-1), including boranes (I, II), boronic acids

(III), trifluoroborate salts (IV), and boronate esters (V – VIII).

Figure 1.3-1: Organoborons Commonly Used in the Suzuki-Miyaura Reaction

Boronic acids (III) are attractive coupling partners in Suzuki reactions as they are air-stable,

virtually non-toxic, and easy to handle crystalline solids; furthermore, a plethora of boronic acids

are commercially available. They do, however, have drawbacks. Under anhydrous conditions,

boronic acids form their cyclotrimeric anhydrides, boroxines, although this is inconsequential in

Suzuki reactions as they proceed regardless of hydrated state. More severe, many small boronic

acids are amphiphilic, which can complicate purification efforts. This problem can be remedied

by converting boronic acids to their corresponding esters, significantly reducing polarity.59

One of the cheapest and oldest methods of preparing boronic acids is the trapping of an

organometallic intermediate (Li, Mg) with a borate ester [B(OR)3] followed by an aqueous acidic

workup;60

however this method is limited by functional group compatibility. In the 1990s,

Miyaura and co-workers introduced a more tolerant method of preparing boronic acids and esters

that involved the palladium-catalyzed coupling of an aryl or alkenyl halide or triflate with a

59 For a comprehensive overview on boronic acids, see: Boronic Acids: Preparation and Applications in Organic

Synthesis, Medicines and Materials; Hall, D. G., Ed.; Wiley-VCH: Weinheim, 2011; pp 1 – 123. 60

Gilman, H.; Moore, L. O. J. Am. Chem. Soc. 1958, 80, 3609.

Page 30: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

19

diboryl ester such as bis(pinacolato)diboron.61

This method, known as “Miyaura borylation,”

was recently used in a one-pot two-step synthesis of unsymmetrical biaryls and heterobiaryls

derived from aryl and heteroaryl chlorides via a Pd-SPhos catalyzed Miyaura borylation/Suzuki

cross-coupling sequence.62

Organotrifluoroborate salts (IV), which can be prepared from reacting the appropriate boronic

acid or ester with potassium bifluoride,63

are monomeric, air-stable, easy to handle crystalline

solids that can be used in many of the reactions that use free boronic acids, including SMC.64

MIDA (methyliminodicarbonic acid) protected boronic acids (VIII) are monomeric, bench-top

and chromatography stable crystalline solids. Unlike polyenylboronic acids, polyenyl MIDA

boronates are stable; therefore, iterative Suzuki couplings of MIDA-protected

haloalkenylboronic acid building blocks are possible.65

Since its discovery by Herbert C. Brown in 1956,66

hydroboration of alkenes and alkynes with

various hydroborating reagents has been the most widely used approach to the preparation of

organoboron compounds. The addition of H-B across a C-C bond occurs in a cis fashion with

boron adding preferentially to the less substituted carbon; a trend that was exploited by Brown

shortly after his initial discovery in order to effectively hydrate alkenes in an anti-Markovnikov

fashion in his renowned hydroboration-oxidation sequence.67

Regioselectivity for the least

sterically hindered carbon is not absolute, however, and is in fact determined by a combination

of steric and electronic effects. In the hydroboration of alkenes,68

electrophilic borane is attacked

by the π electrons of the alkene. Since the C-B bond forms slightly faster than the C-H bond, a

partial negative charge builds on boron in the transition state (Figure 1.3-2). If boron adds to the

terminal carbon, then the more substituted carbon is electron deficient in the transition state (A);

conversely, if boron adds to the internal carbon, then the less substituted carbon is electron

deficient (B). Since electron donating alkyl groups have a stabilizing effect, transition state A is

61 Ishiyama, T.; Murata, M.; Miyarua, N. J. Org. Chem. 1995, 60, 7508.

62 Billingsley, K.; Barder, T. E.; Buchwald, S. L. Angew. Chem. Int. Ed. 2007, 46, 5359.

63 Vedejs, E.; Chapman, R. W.; Fields, S. C.; Lin, S.; Schrimpf, M. R. J. Org. Chem. 1995, 60, 3020.

64 For a review on organotrifluoroboronates, see: Darses, S.; Genêt. J. –P. Eur. J. Org. Chem. 2003, 4313.

65 Lee, S. J.; Gray, K. C.; Paek, J. S.; Burke, M. D. J. Am. Chem. Soc. 2008, 130, 466.

66 Brown, H. C.; Subba Rao, B. C. J. Am. Chem. Soc. 1956, 78, 5694.

67 Brown, H. C.; Zweifel, G. J. Am. Chem. Soc. 1959, 81, 247.

68 a) Brown, H.C.; Zweifel, G. J. Am. Chem. Soc. 1960, 82, 4708. b) Brown, H. C.; Sharp, R. L. J. Am. Chem. Soc.

1966, 88, 5851.

Page 31: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

20

lower in energy that of B; therefore hydroboration proceeds in an anti-Markovnikov fashion. To

complicate matters, if R is an aromatic group, then substituents on the ring have a directing effect

on hydroboration. In the case of styrenes, electron donating para-substituents can stabilize

partial positive charges at the benzylic position, improving regioselectivity for borylation at the

β-carbon (relative to styrene), whereas electron withdrawing substituents have a destabilizing

effect, resulting in an undesirable mixture of regioisomers.

Figure 1.3-2: Relationship Between Regioselectivity and Cyclic Transition State

Hydroboration of alkynes proceeds in a similar fashion.69

While hydroboration of terminal

alkynes is highly regioselective for borylation at the terminal position, a mixture of regioisomers

is usually obtained from internal alkynes. Whereas small hydroborating reagents can

dihydroborate alkynes, sterically hindered boranes (such as 9-BBN) stop after one pass to yield

the monohydroborated product.70

Among hydroborating reagents, the bulky dialkylborane 9-borobicyclo-[3.3.1]nonane (9-BBN),

which can be prepared from the reaction of 1,5-cyclooctadiene with borane, is the most

regioselective (>99.9% borylation at terminal carbon of 1-hexene, 98.5% borylation at β-carbon

of styrene);71

thus the resultant trialkylboranes (I) are popular for Suzuki couplings.

Dialkoxyboranes such as 4,4,6-trimethyl-1,3,2-dioxaborinane,72

1,3,2-benzodioxaborole

(catecholborane, HBcat),73

and 4,4,5,5-tetramethyl-1-3,2-dioxaborolane (pinacolborane,

HBpin)101

may also be used as hydroborating reagents to give organoboron compounds V, VI,

69 Brown, H. C.; Zweifel, G. J. Am. Chem. Soc. 1961, 83, 3834.

70 Brown, H. C.; Scouten, C. G.; Liotta, R. J. Am. Chem. Soc. 1979, 101, 96.

71 Brown, H. C.; Knights, E. F.; Scouten, C. G. J. Am. Chem. Soc. 1974, 96, 7765.

72 Woods, W. G.; Strong, P. L. J. Am. Chem. Soc. 1966, 88, 4667.

73 a) Brown, H. C.; Gupta, S. K. J. Am. Chem. Soc. 1971, 93, 1816. b) Brown, H. C.; Gupta, S. K. J. Am. Chem. Soc.

1972, 94, 4370. c) Brown, H. C.; Gupta, S. K. J. Am. Chem. Soc. 1975, 97, 5249.

Page 32: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

21

and VII respectively. The disadvantage of using these reagents is that they are less reactive than

dialkylboranes. Whereas hydroboration with 9-BBN typically proceeds at room temperature,

hydroboration of alkynes and alkenes with HBcat requires temperatures of 70 and 100 °C

respectively and reaction times of one to 24 hours.73c

However, there are advantages since

trialkoxyboranes are sensitive to air and moisture and thus must be prepared in situ, whereas

cyclic aliphatic boronate esters prepared from pinacol (VII) are bench-top and silica gel

chromatography stable, protected from the approach of water by a phalanx of methyl groups.

The hydroboration of alkenes and alkynes with less reactive dialkoxyboranes can be catalyzed by

transition metals.74

To date, a variety of transition metals have been used with varying degrees

of success, including nickel,75

iridium,76

rhodium,77

ruthenium,78

titanium,79

zirconium,103,105

samarium,80

and, recently, iron.81

The palladium-catalyzed hydroboration of enynes, generating

allenylboronates, has also been reported.82

Rhodium, the most commonly used, and zirconium

will be discussed in detail.

1.3.1 Rhodium-Catalyzed Hydroboration

In 1975, Kono and Ito reported that catecholborane can oxidatively add to Wilkinson’s catalyst

[Rh(PPh3)3Cl],83

but it was not until a decade later that a rhodium-catalyzed hydroboration

process was envisioned by Männig and Nöth, who reported that Wilkinson’s catalyst facilitates

the addition of catecholborane across alkenes and alkynes.84

While uncatalyzed hydroboration

with catecholborane requires heating at elevated temperatures for several hours,73c

Männig and

74 For reviews on transition metal-catalyzed hydroboration, see: a) Burgess, K.; Ohlmeyer, M. J. Chem. Rev. 1991,

91, 1179. b) Beletskaya, I.; Pelter, A. Tetrahedron 1997, 53, 4957. 75

a) Gridnev, I. D.; Suzuki, A. Organometallics 1993, 12, 589. b) Kabalka, G. W.; Narayana, C.; Reddy, N. K.

Synth. Commun. 1994, 24, 1019. 76

a) Evans, D. A.; Fu, G. C.; Hoveyda, A. H. J. Am. Chem. Soc. 1993, 58, 5307. b) Ohmura, T.; Yamamoto, Y.;

Miyaura, N. J. Am. Chem. Soc. 2000, 122, 4990. c) Yamamoto, Y.; Fujikawa, R.; Umemoto, T.; Miyaura, N.

Tetrahedron 2004, 60, 10695. 77

For selected examples, see: a) Evans, D. A.; Fu, G. C.; Hoveyda, A. H. J. Am. Chem. Soc. 1988, 110, 6917. b)

Hayashi, T.; Matsumoto, Y.; Ito, Y. J. Am. Chem. Soc. 1989, 111, 3426. c) Endo, K.; Hirokama, M.; Takeuchi, K.;

Shibata, T. Synlett 2008, 3231. 78

Burgess, K.; Jaspars, M. Organometallics 1993, 12, 4197. 79

He, X.; Hartwig, J. F. J. Am. Chem. Soc. 1996, 118, 1696. 80

Harrison, K. N.; Marks, T. J. J. Am. Chem. Soc. 1992, 114, 9220. 81

Zhang, L.; Peng, D.; Leng, X.; Huang, Z. Angew. Chem. Int. Ed. 2013, 52, 3676. 82

a) Satoh, M.; Nomoto, Y.; Miyaura, N.; Suzuki, A. Tetrahedron Lett. 1989, 30, 3789. b) Matsumoto, Y.; Naito,

M.; Hayashi, T. Organometallics 1992, 11, 2732. 83

Kono, H.; Ito, K.; Nagai, Y. Chem. Lett. 1975, 1095. 84

Männig, D.; Nöth, H. Angew. Chem. Int. Ed. 1985, 24, 878.

Page 33: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

22

Nöth observed that the catalyzed variant went to completion within minutes at room temperature

in the presence of Wilkinson’s catalyst (0.05 mol%, 45 mmol scale). Following distillative

workup, good yields were obtained from cyclopentene (83%) and 1-octene (77%) and a fair yield

from 1-hexyne (53%). Additionally, the authors noted that rhodium-catalyzed hydroboration

possessed markedly different chemoselectivity to that of the uncatalyzed variant, which was later

harnessed in the total synthesis of (+)-ptilocaulin (Scheme 1.3-1).85

In the presence of

Wilkinson’s catalyst, the terminal alkene was preferentially hydroborated over the ketone to give

desired product B following oxidation; whereas, in the absence of catalyst, catecholborane did

not touch the alkene, instead reducing the ketone to give A.

Scheme 1.3-1: Chemoselectivity of Uncatalyzed vs. Catalyzed Hydroboration

The mechanism of rhodium-catalyzed hydroboration of an alkene with catecholborane proposed

by Burgess and co-workers (Figure 1.3-3)86

begins with the dissociation of a triphenylphosphine

ligand from Wilkinson’s catalyst87

to generate a catalytically active 14-electron rhodium(I)

species (A). Oxidative addition of the B-H bond of catecholborane to A gives a 16-electron

boryl rhodium(III)hydride complex (B). The triisopropyl analogue of B has been isolated and

characterized.88

Coordination of the alkene with B trans to the chloride89

gives C, in which the

boryl and hydride ligands are also trans.90

Subsequent migratory insertion of the alkene into the

85 Cossy, J.; BouzBouz, S. Tetrahedron Lett. 1996, 37, 5091.

86 Burgess, K.; van der Donk, W. A.; Westcott, S. A.; Marder, T. B.; Baker, R. T.; Calabrese, J. C. J. Am. Chem.

Soc. 1992, 114, 9350. 87

Mechanism is general, can be applied to a variety hydroborating reagents and Rh sources. 88

Westcott, S. A.; Taylor, N. J.; Marder, T. B.; Baker, R. T.; Jones, N. L.; Calabrese, J. C. J. Chem. Soc. Chem.

Commun. 1991, 304. 89

Musaev, D. G.; Mebel, A. M.; Morokuma, K. J. Am. Chem. Soc. 1994, 116, 10693. 90

Widauer, C.; Grützmacher, H.; Zeigler, T. Organometallics 2000, 19, 2097.

Page 34: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

23

rhodium-hydride bond gives regioisomeric alkyl boryl rhodium(III) complexes D and E, from

which rate-limiting89,90

reductive elimination gives boronates F and G, respectively, and

regenerates rhodium(I) species A. Originally, Evans and Fu91

proposed a mechanism in which

coordination of the alkene to complex B proceeds with concurrent dissociation of a phosphine

ligand. Both mechanisms have been supported by subsequent studies; the debate continues.92

Figure 1.3-3: Catalytic Cycle for the Rhodium-Catalyzed Hydroboration of Alkenes

While the uncatalyzed hydroboration of alkenes gives the anti-Markovnikov boronate as the

major product (Figure 1.3-2, A), the catalyzed variant can give either the anti-Markovnikov or

the Markovnikov boronate as the major product depending on the ligands on the catalyst as well

as the steric and electronic properties of the reacting alkene.93

The anti-Markovnikov boronate is

the major product of rhodium-catalyzed hydroboration of aliphatic alkenes; whereas the

regioselectivity is usually reversed in the case of vinylarenes. Zhang et al. reported that the

hydroboration of styrene with catecholborane in the presence of Wilkinson’s catalyst gave the

Markovnikov or α-boronate as the major product (α:β was 94:6).94

To account for this, Hayashi

91 a) Evans, D. A.; Fu, G. C. J. Am. Chem. Soc. 1990, 55, 2280. b) Evans, D. A.; Fu, G. C.; Anderson, B. A. J. Am.

Chem. Soc. 1992, 114, 6679. 92

For a review, see: Huang, X., Lin, Z. Y. In Computational Modelling of Homogeneous Catalysts; Maseras, F.,

Lin, Z. Y., Eds; Kluwer: Dordrecht, 2002. 93

Smith III, M. R. Prog. Inorg. Chem. 1999, 48, 505. 94

Zhang, J.; Lou, B.; Guo, G.; Dai, L. J. Org. Chem. 1991, 56, 1670.

Page 35: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

24

proposed a modified mechanism for the rhodium-catalyzed hydroboration of vinylarenes that

proceeds through an η3-benzylrhodium complex (Figure 1.3-3, H).

95 Reductive elimination from

this key intermediate gives the α-boronate regioselectively.

An asymmetric variant of rhodium-catalyzed hydroboration has been developed using chiral

supporting ligands,96

particularly the axially chiral diphosphine ligands BINAP97

and QUINAP

(Scheme 1.3-2).98

The C-B bond of the chiral organoboron can be converted to a C-C,31,99

C-

N,100

or C-O95

bond with retention of stereochemistry, making them useful synthons for a variety

of functional groups.

Scheme 1.3-2: Preparation of Chiral Alcohols

95 Hayashi, T.; Matsumoto, Y.; Ito, Y. Tetrahedron: Asymmetry 1991, 2, 601.

96 For a review on enantioselective Rh-catalyzed hydroboration of alkenes. see: Carroll, A. –M.; O’Sullivan, T. P.;

Guiry, P. J. Adv. Synth. Catal. 2005, 347, 609. 97

a) Burgess, K.; Ohlmeyer, M. J. J. Org. Chem. 1988, 53, 5178. b) Sato, M.; Miyaura, N.; Suzuki, A. Tetrahedron

Lett. 1990, 31, 231. 98

a) Alcock, N. W.; Brown, J. M.; Hulmes, D. I. Tetrahedron: Asymmetry 1993, 4, 743. b) Alcock, N. W.; Brown, J.

M.; Hulmes, D. I. J. Chem. Soc. Chem. Commum. 1995, 395. 99

Chen, A.; Ren, L.; Crudden, C. M. J. Org. Chem. 1999, 64, 9704. 100

Fernandez, E.; Maeda, K.; Hooper, M. W.; Brown, J. M. Chem. Eur. J. 2000, 6, 1840.

Page 36: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

25

1.3.2 Zirconium-Catalyzed Hydroboration

In 1992, Knochel introduced pinacolborane (HBPin) or 4,4,5,5-tetramethyl-1-3,2-dioxaborolane

as an effective hydroborating reagent for alkynes and alkenes.101

Prepared from the addition of

borane dimethylsulfide (BMS) to a solution of pinacol in dry DMC, the resulting pinacolborane

solution was treated directly with alkyne or alkene (0.5 equiv) to yield the corresponding pinacol

boronate after several hours stirring at room temperature (Table 1.3-1). The use of less than 2

equivalents pinacolborane relative to alkyne or alkene led to incomplete conversion. The authors

noted that while pinacolborane could be distilled prior to use, it was unnecessary.

Table 1.3-1: Hydroboration of Alkynes with Pinacolborane

Entry R1 R

2 A : B : C

a Yield (%)

b

1 n-hexyl H 98 : 1 : 1 88

2 Ph H 96 : 4 : 0 64

3 Ph Me

85 : 0 : 15

(73 : 0 : 27)c

69

4 cyclohexene N/A 73

a Product ratio determined by GC.

b Isolated yield.

c Ratio of isomers obtained using catecholborane.

Knochel and co-workers highlighted several advantages afforded by the use of pinacolborane as

a hydroborating reagent instead of catecholborane; namely, milder reaction conditions, increased

101 Tucker, C. E.; Davidson, J.; Knochel, P. J. Org. Chem. 1992, 57, 3482.

Page 37: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

26

thermal stability, superior regioselectivity in the case of internal alkynes (Table 1.3-1, entry 3),

and stability of the resulting boronate esters to air, aqueous work-up, and silica gel

chromatography. In contrast, catechol boronates are sensitive to hydrolysis due to their Lewis

acidity and distillation yields unstable “glassy” solids.102

Table 1.3-2: Zirconium-Catalyzed Hydroboration of Alkynes with Pinacolborane

Entry R1

R2

A : B : C : Da

Yield (%)b

1 n-hexyl H 98 : 2 : 0 : 0 93

2 Ph H 97.2 : 0.8 : 0.7 : 0.9 75

3 i-propyl Me 96.9 : 2.2 : 0 0.9 94

4 t-butyl Me 100 : 0 : 0 : 0 91.5

5 Et Et 100 : 0 : 0 : 0 93

6 (EtO)2CH- H 81.9 : 10.8 : 7.3 : 0 82

a Product ratio determined by GC.

b Isolated yield.

In 1995, Srebnik and Pereira103

reported that zirconocene hydrochloride (Schwartz’s reagent)104

effectively catalyzes the hydroboration of alkynes at room temperature with stoichiometric

102 Zaidlewicz, M.; Meller, J. Collect. Czech. Chem. Commun. 1999, 64, 1049.

103 Pereira, S.; Srebnik, M. Organometallics 1995, 14, 3127.

104 Prepared via reduction of Cp2ZrCl2 with LiAlH4, see: Wailes, P. C.; Weigold, H. J. Organomet. Chem. 1970, 24,

405. For use in organic synthesis, see: a) Hart, D. W.; Schwartz, J. J. Am. Chem. Soc. 1974, 96, 8115. b) Schwartz,

J.; Labinger, J. A. Angew. Chem. Int. Ed. 1976, 15, 333.

Page 38: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

27

quantities of pinacolborane to give the corresponding pinacol boronates in excellent yields and

with excellent regio- and stereoselectivities (Table 1.3-2). Regioisomers were negligible with

the exception of 3,3-diethoxyprop-1-yne (entry 6), which the authors attributed to possible

coordination of pinacolborane with the oxygen atoms of the acetal. This was the first report of

transition metal-catalyzed hydroboration using pinacolborane.

Srebnik and Pereira proposed the following mechanism for hydroboration of alkynes catalyzed

by Schwartz’s reagent (Figure 1.3-4): hydrozirconation of the alkyne in a cis fashion, placing Zr

on the least sterically hindered carbon, followed by vinyl-hydride exchange between the resultant

organozirconocene chloride and pinacolborane to give the corresponding (E)-alkenylboronate

and regenerate the catalyst. It is interesting to note that the hydride reducing the alkyne initially

comes from the catalyst rather than pinacolborane.

Figure 1.3-4: Catalytic Cycle for the Zirconium-Catalyzed Hydroboration of Alkynes

The following year, Srebnik and Pereira reported that Schwartz’s reagent also catalyzes the

hydroboration of alkenes with pinacolborane.105,106

In the case of trans-4-octene, the terminal

105 Pereira, S.; Srebnik, M. J. Am. Chem. Soc. 1996, 118, 909.

106 Pereira, S.; Srebnik, M. Tetrahedron Lett. 1996, 37, 3283.

Page 39: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

28

alkyl boronate was obtained. This was expected as hydrozirconation of internal alkenes proceeds

to place Zr on the least hindered carbon via a series of additions/β-hydride eliminations.104b

In addition to catalyzing hydroboration, Schwartz’s reagent, named after its earliest champion,

Jeffery Schwartz at Princeton, has many applications in organic synthesis (Figure 1.3-5)107

and

has greatly expanded the field of organozirconium chemistry.108

Figure 1.3-5: Synthetic Applications of Zirconocene Hydrochloride (Schwartz’s Reagent)

107 For a review on the synthetic applications of organozirconocene complexes, see: Wipf, P.; Jahn, H. Tetrahedron

2004, 52, 12853; and related references therein. 108

For a review on organozirconium chemistry, see: Negishi, E. Dalton Trans. 2005, 827.

Page 40: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

29

1.4 Alkenes in Organic Synthesis

Alkenes are highly versatile motifs in organic synthesis as they can be further functionalized by a

vast array of addition, reduction, and oxidation reactions. Their fundamental nature is evidenced

by the fact that many of the first reactions taught to fledgling chemists are those of the carbon-

carbon double bond. Thus the efficient synthesis of alkenes is highly desired.

1.4.1 Bimetallic Catalyzed One-Pot Synthesis of Alkenes

Given the importance of alkenes as well as the numerous benefits afforded by cascade catalysis

(outlined at the beginning of this chapter), organic chemists have sought to develop efficient one-

pot syntheses of alkenes, including bimetallic catalyzed protocols. Selected recent examples

from the literature are depicted below, including a [Rh-Pd]-catalyzed hydrosilylation/Hiyama

cross-coupling sequence (Scheme 1.4-1),109

a microwave-assisted platinum-catalyzed

diboration/palladium-catalyzed Suzuki cross-coupling sequence (Scheme 1.4-2),110

and a

hydrozirconation/palladium-catalyzed cross-coupling sequence (Scheme 1.4-3).111

Scheme 1.4-1: [Rh-Pd]-Catalyzed Hydrosilylation/Hiyama Cross-Coupling Sequence

Scheme 1.4-2: Platinum-Catalyzed Diboration/Palladium-Catalyzed SMC Sequence

109 Thiot, C.; Schmutz, M.; Wagner, A.; Mioskowski, C. Chem. Eur. J. 2007, 13, 8971.

110 Prokopcová, H.; Ramírez, J.; Fernández, E.; Kappe, C. O. Tetrahedron Lett. 2008, 49, 4831.

111 Huang, B.; Wang, P.; Hao, W.; Cai, M. – Z. J. Organomet. Chem. 2011, 696, 2685.

Page 41: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

30

Scheme 1.4-3: Hydrozirconation/Palladium-Catalyzed Cross-Coupling Sequence

It should be noted that zirconium was not catalytic in the last example (Scheme 1.4-3); in fact, a

10% excess of zirconocene hydrochloride (Schwartz’s reagent) was required in the first step.

1.4.2 Palladium-Catalyzed Zirconium-Negishi Cross-Coupling Reaction

At a glance, hydrozirconation of alkynes and alkenes is an attractive process as the resultant

alkenyl- and alkylzirconocene chlorides are obtained with high regio- and stereoselectivities

under mild conditions; however, these organozirconium derivatives have proven to be poor

nucleophiles in comparison to organolithium and Grignard reagents, typically only reacting with

small electrophiles. For example, carbon monoxide readily inserts into the C-Zr bond to give the

corresponding aldehyde following an aqueous acidic work-up (Figure 1.3-5, A).112

Intrigued by the synthetic potential of organozirconium species, Negishi and co-workers turned

from conventional polar reactions to transmetallation. In the late 1970s, the group reported the

first nickel- and palladium-catalyzed cross-couplings of alkenylzirconocene chlorides with aryl

and alkenyl halides (Figure 1.3-5, G).113

The mechanism is believed to be analogous to the

Suzuki coupling of organoboron species with organic halides (Figure 1.2-1).114

Soon after, Negishi and co-workers discovered that the addition of zinc chloride significantly

improved the cross-coupling of sterically hindered alkenylzirconocenes and alkenyl halides.115

Presumeably, a transmetallation of the organic group from bulky zirconocene to the less

sterically demanding zinc salt precedeed the usual catalytic cycle. This Zr-to-Zn-to-Pd sequence

112 Bertelo, C. A.; Schwartz, J. J. Am. Chem. Soc. 1975, 97, 228.

113 a) Negishi, E.; Van Horn, D. E. J. Am. Chem. Soc. 1977, 99, 3168. b) Okukado, N.; Van Horn, D. E.; Klima, W.

L.; Negishi, E. Tetrahedron Lett. 1978, 1027. 114

Negishi, E.; Takahashi, T.; Baba, S.; Van Horn, D. E.; Okukado, N. J. Am. Chem. Soc. 1987, 109, 2393. 115

Negishi, E.; Okukado, N.; King, A. O.; Van Horn, D. E.; Spiegel, B. I. J. Am. Chem. Soc. 1978, 100, 2254.

Page 42: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

31

has been employed in key steps in a number of syntheses of complex natural products, including

motuporin, isolated from a marine sponge (Scheme 1.4-4).116

Scheme 1.4-4: Pd-Catalyzed Zr-Negishi Coupling Toward Total Synthesis of (-)-Motuporin

Initally limited to aryl and alkenyl halides, the palladium-catalyzed cross-coupling of

alkenylzirconocenes with alkyl halides was reported by the Fu group in 2004, the key feature

being the absence of a specialized ligand (Scheme 1.4-5).117

Scheme 1.4-5: Coupling of Alkenylzirconocenes with Alkyl Halides

Negishi recognized that using a stoichiometric amount of relatively expensive zirconocene was a

limitation and its use as a catalyst was highly desired. In 1978, Negishi and co-workers reported

the first successful zirconocene dichloride (Cp2ZrCl2)-catalyzed methylalumination of alkynes

with trimethylaluminum (Me3Al).118

The resulting alkenylalanes can be applied to a number of

116 Ho, T.; Panek, J. S. J. Org. Chem. 1999, 64, 3000.

117 Wiskur, S. L.; Korte, A.; Fu, G. C. J. Am. Chem. Soc. 2004, 126, 82.

118 a) Van Horn, D. E.; Negishi, E. J. Am. Chem. Soc. 1978, 100, 2252. b) Matsushita, H.; Negishi, E. J. Am. Chem.

Soc. 1981, 103, 2882.

Page 43: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

32

conventional polar reactions as well as nickel- and palladium-catalyzed cross-coupling reactions

as either the alkenylaluminum derivative or the corresponding iodide to which it can be readily

converted. The utility of zirconium-catalyzed carboalumination is evidenced by its application

in the total synthesis of more than 100 natural products.119

1.4.3 Proposed Methodology

Hydrozirconation of alkynes is an attractive approach to synthesizing alkenes given its mild

conditions as well as its excellent regio- and stereoselectivity; however, cost is a serious concern,

particularly in large scale reactions. While zirconium may be an inexpensive metal,

zirconocenes, although less expensive than many transition metal catalysts, may not be

considered inexpensive chemicals; thus their catalytic use in the synthesis of alkenes is highly

desirable. While zirconium-catalyzed carboalumination has proven incredibly useful,

trialkylaluminum reagents are highly pyrophoric, thus a milder approach would be welcomed.

Given our group’s history with bimetallic catalyzed domino reactions, we endeavored to develop

a one-pot two-step protocol for the efficient synthesis of alkenes with defined regio- and

stereoselectivity via a zirconium-catalyzed hydroboration/palladium-catalyzed Suzuki cross-

coupling sequence (Scheme 1.4-6) with the possibility of domino catalysis in mind.

Scheme 1.4-6: Proposed Methodology

119 For a review on Zr-catalyzed carboalumination and its application to total synthesis, see: Negishi, E.; Tan, Z.

Top. Organomet. Chem. 2004, 8, 139.

Page 44: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

33

2 Results and Discussion

2.1 Preparation of Alkenylboronates

We began developing a one-pot two-step Zr-catalyzed hydroboration/Pd-catalyzed Suzuki cross-

coupling protocol by examining the hydroboration of terminal alkynes with pinacolborane

(HBpin) catalyzed by zirconocene hydrochloride (Schwartz’s reagent). While the initial

conditions reported by Srebnik and Pereira employed DCM as a solvent,103

Wang et al. reported

in 2005 that hydrocarbon terminal alkynes readily underwent hydroboration with pinacolborane

at ambient temperature in the presence of catalytic amounts of Schwartz’s reagent (10 mol%) in

the absence of solvent, but did not elaborate further.120

In addition to being a greener approach,

the use of solvent-free or neat conditions was attractive as it eliminated the possibility of solvent

incompatibility with a sequential Suzuki cross-coupling. Thus the utility of solvent-free Zr-

catalyzed hydroboration of alkynes was determined.

2.1.1 Solvent-Free Zr-Catalyzed Hydroboration of Terminal Alkynes

We investigated the effects of reaction time, catalyst loading, temperature, and scale on the Zr-

catalyzed hydroboration of phenylacetylene with HBpin, as well as the sensitivity of the reaction

to both light and water (Table 2.1-1). All reactions employed a 10% excess of HBpin relative to

phenylacetylene and were carried out in 2 dram vials sealed under argon. In all cases,

appropriate care was taken to exclude both air and moisture from the reaction (except when

water was added deliberately).121

Styrene, presumably formed via protodemetallation of the

alkenylzirconocene intermediate, was detected in all cases, but in negligible amounts.

Hydroboration of phenylacetylene with HBpin at ambient temperature in the presence of 10

mol% Schwartz’s reagent gave the desired trans-β-styryl pinacolboronate (3.1) in excellent yield

after 24 hours (Entry 3). Neither regio- nor stereoisomers of the product were detected using 1H

NMR analysis of the crude mixture, which was quite clean. Although Schwartz’s reagent is light

120 Wang, Y. D.; Kimball, G.; Prashad, A. S.; Wang, Y. Tetrahedron Lett. 2005, 46, 8777.

121 Although Schwartz’s reagent was measured out in air, exposure was minimized as much as possible and the

reagent was stored under argon in a dessicator when not in use. Pinacolborane, sensitive to both air and moisture

according to its MSDS, was transferred to the reaction vessel under argon via syringe and stored under argon in a

refrigerator (4 °C) when not in use.

Page 45: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

34

sensitive according to its MSDS, covering the reaction vial with aluminum foil to reduce

exposure to light provided only a slight boost to the already excellent yield (Entry 4).

Regardless, reaction vials were covered in subsequent trials in an effort to maximize consistency.

Table 2.1-1: Solvent-Free Zr-Catalyzed Hydroboration of Phenylacetylene

Entry Scale

(mmol)

mol%

Zr cat

Temp

(°C)

Time

(h)

Add.

(equiv)

Yield A

(%)b

Yield B

(%)b

1 0.5 none rt 24 -- 0 0

2 0.5 none 110 30 -- 33 2

3 1.1 10 rt 24 -- 93 2

4a

1.1 10 rt 24 -- 95 2

5a

0.5 10 rt 24 -- 92 2

6a

0.2 10 rt 24 -- 65 3

7a

1.1 10 rt 16 -- 96(81) 2

8a

1.1 10 rt 12 -- 88 3

9a

1.1 5 rt 16 -- 85 2

10 1.1 5 45 16 -- 88 2

11 1.0 10 rt 16 H2O (1.0) 0 14

a Reaction vial covered with aluminum foil to reduce exposure to light.

b Yield determined using

1H NMR analysis

of crude with TMB as internal standard. Isolated yield following chromatography shown in parenthesis.

Decreasing the scale of the reaction from 1.1 to 0.5 mmol had no effect on yield (Entry 5);

however, the yield dropped significantly when the scale was further decreased to 0.2 mmol

(Entry 6). The cause was likely mechanical in nature: there was simply not enough material in

Page 46: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

35

the 2 dram reaction vial for the reagents to mix sufficiently. Use of a smaller reaction vessel

would presumably remove this limitation.

Reducing the reaction time from 24 to 16 hours did not affect the yield (Entry 7) and 81% of the

desired styrylboronate (3.1) was isolated following chromatography.122

Further reducing the

reaction time to 12 hours (Entry 8) resulted in a slight but noticeable drop in yield as well as the

detection of starting material in the crude reaction mixture. Thus a reaction time of 16 hours

appeared to be ideal for these conditions.

Reducing the catalyst loading from 10 to 5 mol% resulted in a roughly ten percent decrease in

yield (Entry 7 versus 9 respectively), which was somewhat restored by heating the latter at 45 °C

(Entry 10). Based on these data, two sets of reaction conditions of comparable efficacy emerged

for the Zr-catalyzed hydroboration of phenylacetylene with HBpin: 16 hours at ambient

temperature in the presence of 10 mol% Schwartz’s reagent (Entry 7), or 16 hours at 45 °C in the

presence of 5 mol% Schwartz’s reagent (Entry 10).

No desired product was observed when water was deliberately added to the reaction (Entry 11)

either due to deactivation of the catalyst or degradation of pinacolborane or both.

Knochel and co-workers101

reported that alkynes, including phenylacetylene, readily underwent

uncatalyzed hydroboration with HBpin, prepared in situ, at ambient temperature (Table 1.3-1);

however, we observed no reactivity in the absence of catalyst at ambient temperature with

commercially obtained HBpin (Entry 1). Some product was obtained after heating the same

reaction at 110 °C for 30 hours, but the yield was poor (Entry 2). Our conditions differed to

those of Knochel and co-workers, who employed 2.0 equivalents of HBpin (as opposed to our

1.1 equiv) and conducted their hydroborations in DCM; however, when we attempted to

reproduce Knochel’s conditions with phenylacetylene and commercially obtained HBpin (2.0

equiv) in distilled DCM, no product was detected using 1H NMR analysis, even after 30 hours.

Our results somewhat echo those of Srebnik and Pereira,103

who reported that hydroboration of

1-octyne with HBpin (1.05 equiv) in DCM in the absence of Schwartz’s reagent afforded only a

122 Product tended to trail on silica gel column, the majority eluting over several fractions and the remainder over

many more. Switching to an alumina column and/or increasing the polarity of the solvent system once the desired

product began eluting did little to remedy this issue.

Page 47: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

36

2% yield of the corresponding octenylboronate when HBpin was prepared in situ (using

Knochel’s procedure) or a 20% yield when HBpin was distilled prior to use. Srebnik and Pereira

attributed these differences to the fact that Knochel used 2.0 equivalents of HBpin, whereas they

used only 1.05 equivalents.

Fürstner and co-workers123

prepared styryl pinacolboronate (3.1) as a reagent for the total

synthesis of marine oxylipins on a 110 mmol scale in 89% yield under solvent-free conditions in

the absence of catalyst; however, the reaction was heated at 140 °C for 5 days. Their source of

HBpin was not specified, but it was presumably commercial. Whereas HBpin prepared in situ

from the reaction of pinacol with borane dimethylsulfide (BMS) in DCM and treated directly

with an alkyne gives the corresponding alkenylboronate in good yields at ambient temperature in

the absence of catalyst, it appears that commercially obtained HBpin does not.

Knochel and co-workers initially reported that Wilkinson’s catalyst does not catalyze the

hydroboration of alkynes or alkenes with HBpin,101

but Srebnik and Pereira later reported that it

is an excellent catalyst for such reactions.105

It is conceivable that the lack of reactivity observed

by Knochel was due to the presence of unreacted BMS and/or by-product(s) of HBpin

preparation. It would be useful to prepare HBpin in situ according to Knochel’s procedure and

compare its reactivity to that of HBpin obtained from a commercial source as well as determine

the effect of BMS on hydroboration with HBpin.

We applied our conditions for the Zr-catalyzed hydroboration of phenylacetylene with HBpin to

a non-hydrocarbon substrate in order to test the scope of the reaction. N-Propargyl indole (1.2)

proved suitable for our conditions (Table 2.1-2). The reaction proceeded smoothly at ambient

temperature to give boronate 3.2 in a good yield (Entry 1). Heating the reaction at 45 °C

improved the yield and decreased the formation of the N-allyl side-product (Entry 2); however, a

further increase in temperature afforded no additional benefit (Entry 3). Reducing the catalyst

loading to 5 mol% with compensatory heating at 45 °C gave a yield that was comparable to that

of the reaction conducted at ambient temperature in the presence of 10 mol% Schwartz’s reagent

(Entry 4 versus 1 respectively).

123 Hickman, V.; Kondoh, A.; Gabor, B.; Alcarazo, M.; Fürstner, A. J. Am. Chem. Soc. 2011, 133, 13471.

Page 48: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

37

Table 2.1-2: Solvent-Free Zr-Catalyzed Hydroboration of N-Propargyl Indole

Entry mol% Zr cat Temp (°C) Yield A (%)b

Yield B (%)b

1 10 rt 76 9

2 10 45 87 1

3 10 60 87 1

4 5 45 79(65) 5

a Reactions carried out on 0.50 mmol scale.

b Yield determined using

1H NMR analysis of crude with TMB as

internal standard. Isolated yield following chromatography shown in parenthesis.

2.1.2 Solvent-Free Zr-Catalyzed Hydroboration of Internal Alkynes

Hydroboration of internal alkynes proved sluggish under our conditions and heating at 60 °C for

60 hours was required to obtain moderate yields of the corresponding alkenylboronates (Table

2.1-3). A mixture of regioisomers was obtained from 1-phenyl-1-propyne (Entries 4 – 6) with

borylation occurring predominately at the least sterically hindered carbon, in conformity with the

trend of hydrozirconation of unsymmetrical internal alkynes.104

Regioselectivity decreased with

increased heating. Knochel and co-workers101

reported that uncatalyzed hydroboration of 1-

phenyl-1-propyne with HBpin (2.0 equiv), prepared in situ, provided the corresponding boronate

in a moderate yield after only several hours at ambient temperature in DCM (Table 1.3-1, Entry

3); however, the ratio of regioisomers they obtained was identical to the ratio we obtained from

heating at 60 °C for 60 hours (Entry 6).

Table 2.1-3: Solvent-Free Zr-Catalyzed Hydroboration of Internal Alkynes

Page 49: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

38

Entry R1

R2

Temp (°C) Yield A (%)b

A : Bc

1 Et Et rt 13 N/A

2 Et Et 45 41 N/A

3 Et Et 60 67 N/A

4 Ph Me rt 25 92 : 8

5 Ph Me 45 46 86 : 14

6 Ph Me 60 63 85 : 15

a Reactions carried out on 1.1 mmol scale.

b Yield determined using

1H NMR analysis of crude with TMB as

internal standard. c Regioisomeric ratios determined using

1H NMR analysis of crude.

2.1.3 Selecting a Solvent for Zr-Catalyzed Hydroboration

Solvent-free hydroboration is amenable to a one-pot two-step protocol in which a suitable

solvent for SMC is added upon completion of the first step; however, a domino reaction would

require a solvent that was suitable for both steps. We screened several solvents for the Zr-

catalyzed hydroboration of phenylacetylene with HBpin and found the reaction worked

moderately well in toluene at 60 °C (Table 2.1-4, Entry 3). This held promise for domino

reactivity as Suzuki reactions are often conducted in toluene.11

Ethereal solvents THF and

dioxane, also common reaction media for SMC, proved to be incompatible.

Table 2.1-4: Zr-Catalyzed Hydroboration of Phenylacetylene in Toluene

Entry mol% Zr cat Temp (°C) Time (h) Yield A (%)b

Yield B (%)b

1 10 rt 16 33 6

2 10 45 16 58 4

Page 50: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

39

3 10 80 24 67 2

4 none 80 24 11 0

a Reactions carried out on 0.22 mmol scale. Concentration: 0.55 M.

b Yield determined using

1H NMR analysis of

crude with TMB as internal standard.

Toluene proved to be an even better solvent for the Zr-catalyzed hydroboration of N-propargyl

indole (1.2) (Table 2.1-5). A good yield of boronate 3.2 was obtained after 16 hours at 45 °C in

the presence of 5 mol% Schwartz’s reagent (Entry 2).

Table 2.1-5: Zr-Catalyzed Hydroboration of N-Propargyl Indole in Toluene

Entry [1.2] (M) mol% Zr cat Temp (°C) Yield A (%)b

Yield B (%)b

1 1.0 10 rt 38 16

2 2.0 5 45 70 8

a Reactions carried out on 0.50 mmol scale.

b Yield determined using

1H NMR analysis of crude with TMB as

internal standard.

N-Propargyl isatin (1.4) proved to be an unsuitable substrate for Zr-catalyzed hydroboration.

The desired boronate was not detected using 1H NMR analysis of the crude, only a complex

mixture of decomposed starting material. This result was not surprising as ketones are known to

be reduced competitively with hydrozirconation of alkynes and alkenes in the presence of

Schwartz’s reagent.107

Page 51: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

40

2.2 Optimization of the Suzuki Cross-Coupling Reaction

For our optimization of the Suzuki coupling step we used trans-styryl pinacolboronate (3.1) as

the organoboron species and 4’-iodoacetophenone as its coupling partner. The latter was chosen

primarily for diagnostic purposes. In addition to being an ideal organic halide for Suzuki

couplings given the ease of oxidative addition of Ar-I bonds to Pd(0) – enhanced by the

electrophilic character imparted by the electron-withdrawing p-substituent – 4’-

iodoacetophenone was an easy-to-handle solid, and the acetyl group gave a distinct singlet in a

1H NMR spectrum that was sufficiently shifted in the spectrum of the coupled trans-stilbene

product to enable calculation of NMR yields using an internal standard. Lastly, stilbenes, true to

their name derived from the Greek word Stilbos, meaning “shining,” fluoresce under UV light,

making such products easy to detect using TLC analysis. 4-Acetyl-trans-stilbene (4.1) was also

visible on a silica gel column under UV light, making for easy purification by chromatography.

Tetrakis(triphenylphosphine)palladium(0) was chosen as the catalyst and, given its performance

in the Zr-catalyzed hydroboration step, toluene was chosen as the solvent.

2.2.1 Base and Solvent Screen

Among bases screened, Cs2CO3 was found to be the most effective in toluene (Table 2.2-1, Entry

3), significantly outperforming K2CO3 (Entry 2), implying the importance of the countercation.

In all cases, a 10:1 toluene/water mixture proved to be a more effective solvent system than dry

toluene (Entries 4 – 6), presumably due to increased solubility of the base and/or rate

enhancement of transmetallation by hydroxide.

Table 2.2-1: Base and Solvent Screen for SMC

Entry Base (equiv) Solvent Yield (%)b

1 K3PO4 (2.0) PhMe 46

Page 52: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

41

2 K2CO3 (2.0) PhMe 19

3 Cs2CO3 (2.0) PhMe 87

4 K3PO4 (2.0) PhMe/H2O (10:1) 90

5 K2CO3 (2.0) PhMe/H2O (10:1) 36

6 Cs2CO3 (2.0) PhMe/H2O (10:1) 96

a Reactions carried out on 0.20 mmol scale. Concentration: 0.2 M (aryl iodide in toluene).

b Yield determined by

1H

NMR analysis of crude with TMB as internal standard.

2.2.2 Further Optimization

The reaction was optimized for efficiency using toluene/water (10:1) as the solvent and Cs2CO3

as the base. The best yield of coupled product 4.1 was obtained using 1.2 equivalents of

boronate 3.1 and 5 mol% Pd catalyst (Entry 1); however, a satisfactory yield was obtained using

1.1 equivalents of 3.1 and only 1 mol% Pd catalyst (Entry 5). Although the yield of the latter

reaction was lower, we thought it was an acceptable result given the catalyst loading was reduced

by a factor of five, thus these conditions were chosen for the SMC step.

Table 2.2-2: Further Optimization of SMC

Entry equiv Cs2CO3 equiv 3.1 mol% Pd cat Yield (%)b

1 2.0 1.2 5.0 99(86)

2 2.0 1.1 5.0 99

3 2.0 1.0 5.0 84

4 2.0 1.1 2.5 89

Page 53: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

42

5 2.0 1.1 1.0 89(70)

6 1.2 1.1 1.0 66

a Reactions carried out on 0.20 mmol scale. Concentration: 0.2 M (aryl iodide in toluene).

b Yield determined using

1H NMR analysis of crude with TMB as internal standard. Isolated yield following chromatography shown in

parenthesis.

Page 54: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

43

2.3 Initial Conditions for One-Pot Two-Step Zr-Catalyzed Hydroboration/Pd-Catalyzed Suzuki Coupling Sequence

Our next objective was to combine the individual steps into a one-pot two-step bimetallic

catalyzed hydroboration/SMC sequence. To this end, Zr-catalyzed hydroboration of

phenylacetylene (1.1 equiv) with HBpin was carried out following our previously established

solvent-free procedure (Table 2.1-1, Entry 4). After stirring at ambient temperature for 24 hours,

the reaction was stopped and the crude boronate intermediate (3.1) was dissolved in toluene. To

this mixture, water, 4’-iodoacetophenone (1.0 equiv), Cs2CO3, and Pd(PPh3)4 (1 mol%) were

added in accordance with our previously established procedure for SMC (Table 2.2-2, Entry 5).

The reaction vial was sealed under argon and stirred at 100 °C. After 16 hours, the reaction was

stopped and 55% of the desired trans-stilbene (4.1) was isolated following work-up and

purification by chromatography (Scheme 2.3-1). Neither regio- nor stereoisomers of the product

were detected using 1H NMR analysis of the crude reaction mixture, although some of the

boronate intermediate (3.1) was present.

Scheme 2.3-1: Initial Conditions for One-Pot Two-Step Sequence

2.3.1 Substrate Scope of Initial Conditions

We were eager to attempt to apply the conditions for the one-pot two-step synthesis of 4-acetyl-

trans-stilbene (4.1, Scheme 2.3-1) to the synthesis of various (E)-alkenes, thus a substrate scope

was conducted (Table 2.3-1).

Page 55: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

44

Table 2.3-1: Substrate Scope of Initial Conditions

Entry Alkyne Aryl Halide Product Yield (%)b

1

58

2

48

3

50

4

39

5

62

6

48

7

63

Page 56: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

45

8

73

9

55

10

>10c

a Reactions carried out on 1.0 mmol scale in 3 dram vials. Concentration of coupling reaction: 0.25 M (aryl halide

in toluene). b Isolated yield following chromatography based on amount of aryl halide used.

c Based on conversion

of aryl chloride determined using 1H NMR analysis of crude.

In general, modest yields of the expected (E)-alkene were obtained over two steps. In addition to

aryl alkynes (Entries 1 and 2), cyclic enynes proved to be suitable substrates for these conditions

(Entry 5). Aliphatic alkynes were also suitable (Entry 6), although purification was somewhat

challenging as the product tended to co-elute with unreacted bromide 2.1, resulting in a

decreased isolated yield of 4.9. Yields obtained from the reactions of N-propargyl indole (1.2)

(Entries 7 – 9) were among the best; however, the first step required heating at 60 °C. When

conducted at ambient temperature, the solvent-free hydroboration of 1.2 was problematic as the

mixture remained solid and would not stir. The yield obtained from the reaction of 3-ethynyl

thiophene (Entry 3) was satisfactory given that thiophenes are conventionally challenging

substrates for SMC.58

The reaction with 3-ethynyl pyridine gave the poorest yield (Entry 4),

presumably due to the combination of an electron-poor organoboronate and an electron-rich aryl

halide. Whereas aryl and heteroaryl bromides were generally suitable coupling partners, aryl

chlorides were unreactive under our conditions (Entry 11). This was expected given Pd(0)

catalysts supported by triarylphosphine ligands are not sufficiently electron-rich to oxidatively

add aryl chlorides.42

Page 57: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

46

2.4 Optimization of One-Pot Two-Step Zr-Catalyzed Hydroboration/Pd-Catalyzed Suzuki Coupling Sequence

The yields of (E)-alkenes obtained using our initial conditions were not satisfactory, and the high

temperatures and long reaction times required to achieve them were synthetically impractical. In

addition, the incompatibility of aryl chlorides led us to try further experiments to improve the

effectiveness, efficiency, and scope of our one-pot two-step protocol.

2.4.1 Optimization with Aryl Bromides and Iodides

Doubling the concentration of the Suzuki coupling step from 0.25 to 0.5 M and switching from 3

to 2 dram vials resulted in the most dramatic improvement in yield, presumably promoting a

more thorough mixing of the biphasic reaction. Whereas our initial protocol yielded 55% of

coupled product 4.1 (Scheme 2.3-1) and left some of intermediate 3.1 unreacted after 16 hours,

simply doubling the concentration resulted in a 71% yield (NMR) of 4.1 and full conversion of

3.1 after 16 hours (Table 2.4-1, Entry 1). As an added benefit, reducing the amount of solvent

used by half makes this procedure more attractive from a green chemistry perspective.

Table 2.4-1: Optimization with Aryl Bromides and Iodides

Entry X mol%

Zr cat

Temp 1

(°C)

Time 1

(h)

mol%

Pd cat

Temp 2

(°C)

Time 2

(h)

Yield

(%)b

1 I 11 rt 24 1.0 100 16 71

2 I 11 rt 24 2.0 100 16 72

3 Br 11 rt 24 1.0 100 16 73

4 Br 11 rt 24 1.0 80 16 76(70)

5 Br 11 rt 24 1.0 60 16 63

Page 58: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

47

6 Br 11 rt 16 1.0 80 8 89

7 Br 5.5 rt 16 1.0 80 8 70

8 Br 5.5 45 16 1.0 80 8 89(74)

a Reactions carried out on 1.0 mmol scale in 2 dram vials. Concentration of coupling reaction: 0.5 M (aryl halide in

toluene). b Yield determined using

1H NMR analysis of crude with TMB as internal standard. Isolated yield

following chromatography shown in parenthesis.

Aryl bromides were as reactive as aryl iodides (Entry 3 versus 1 respectively), which was not

surprising in light of our recent substrate scope (Table 2.3-1). Continuing with aryl bromides,

we found that decreasing the temperature from 100 to 80 °C in the SMC step resulted in an

improved yield (Entry 4). Decreasing the temperature further, however, resulted in a drop in

yield (Entry 5), suggesting that 80 °C was an ideal temperature for this step. Halving the

reaction time of the SMC step to 8 hours resulted in a marked improvement in yield (Entry 6),

suggesting that the coupled product had been decomposing beyond this time.

Reducing the loading of Schwartz’s reagent from 10 to 5 mol% (relative to the alkyne) resulted

in a drop in yield and a messier reaction, complicating purification (Entry 7); however, heating

the hydroboration step to 45 °C afforded a clean reaction and excellent yield over two steps with

only 5 mol% Schwartz’s reagent (Entry 8). Accepting the trade-off between conducting the first

step at ambient temperature and halving the use of expensive zirconocene catalyst, we decided

the latter conditions were optimal for this reaction.

2.4.2 Extension to Aryl Chlorides

Aryl chlorides are the cheapest and most readily available aryl halides, albeit less reactive than

the more expensive bromides and iodides, thus we were eager to modify our one-pot two-step

protocol in order to be able to use them as substrates. To this end, we screened several

specialized ligands for the Suzuki coupling of alkenylboronate 3.1 with an aryl chloride,

including dialkylbiaryl phosphine ligands developed by Buchwald and co-workers, which were

discussed above (1.2.3). While SPhos (Figure 1.2-2, L4) and RuPhos (L8) were effective as

supporting ligands for the Pd-catalyzed cross-coupling of 3.1 with 4’-chloroacetophenone,

XPhos (L6) furnished the highest yields, particularly when the concentration of the reaction was

increased from 0.2 to 0.5 M (Table 2.4-2, Entry 4 and 5 respectively).

Page 59: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

48

Table 2.4-2: Ligand Screen for Suzuki Coupling of Aryl Chlorides

Entry Pd catalyst (mol%)/

Ligand (mol%) [ArCl] (M)

b Base (equiv)

Temp

(°C) Yield (%)

c

1 Pd2dba3 (2.0)/SPhos (4.0) 0.2 Cs2CO3 (2.0) 100 52

2 Pd2dba3 (2.0)/RuPhos (4.0) 0.2 Cs2CO3 (2.0) 100 66

3 Pd2dba3 (2.0)/tBuXPhos (4.0) 0.2 Cs2CO3 (2.0) 100 8

4 Pd2dba3 (2.0)/XPhos (4.0) 0.2 Cs2CO3 (2.0) 100 76

5 Pd2dba3 (2.0)/XPhos (4.0) 0.5 Cs2CO3 (2.0) 100 87

6 Pd(crotyl)QPhosCl (1.0) 0.8 KOtBu (1.2) 80 28

a Reactions carried out on 0.20 mmol scale.

b Concentration in toluene.

c Yield determined using

1H NMR analysis

of crude with TMB as internal standard.

We employed Pd-XPhos in our one-pot two-step sequence and found, after a short optimization,

that the SMC step went to completion within 4 hours at 80 °C in the presence of 1 mol% Pd(0)

and 2 mol% XPhos ligand (Table 2.4-3, Entry 5).

Table 2.4-3: Optimization with Aryl Chlorides using XPhos

Entry mol%

Zr cat

Time 1

(h)

mol%

Pd cat/L

Temp 2

(°C)

Time 2

(h)

Yield

(%)b

1 11 24 2.0/4.0 100 16 66

Page 60: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

49

2 11 24 2.0/4.0 80 16 80(73)

3 11 24 1.0/2.0 80 16 77

4 11 24 1.0/2.0 60 16 63

5 11 24 1.0/2.0 80 4 74

6 5.5 16 1.0/2.0 80 4 68

a Reactions carried out on 1.0 mmol scale in 2 dram vials. Concentration of coupling reaction: 0.5 M (aryl chloride

in toluene). b Yield determined using

1H NMR analysis of crude with TMB as internal standard. Isolated yield

following chromatography shown in parenthesis.

The commercially available 2nd

generation XPhos precatalyst (Figure 2.4-1) proved to be an even

more effective catalyst for the coupling of aryl chlorides (Table 2.4-4). The SMC step

completed within 2 hours at 80 °C in the presence of 1 mol% precatalyst (Entry 1). Reducing the

loading of Schwartz’s reagent to 5 mol% (relative to alkyne) in the hydroboration step resulted in

a lower yield overall (Entry 3); however, as with aryl bromides, reduced catalyst loading in the

first step could be compensated by heating at 45 °C (Entry 4). Again, preferring increased

heating to increased catalyst loading, we decided the latter conditions were optimal for a one-pot

two-step bimetallic catalyzed hydroboration/SMC sequence using aryl chlorides as substrates.

Apart from being operationally simpler, an added benefit of employing the XPhos palladacycle

(as opposed to adding Pd and XPhos ligand individually) was that only one equivalent of ligand

was used relative to Pd, improving efficiency.

Figure 2.4-1: 2nd Generation XPhos Precatalyst

Page 61: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

50

Table 2.4-4: Optimization with Pd-XPhos-G2

Entry mol% Zr cat Temp 1 (°C) mol%

Pd-XPhos-G2 Yield (%)

b

1 11 rt 1.0 83

2 5.5 rt 1.0 73

3 5.5 rt 2.0 73

4 5.5 45 1.0 80

a Reactions carried out on 1.0 mmol scale in 2 dram vials. Concentration of coupling reaction: 0.5 M (aryl chloride

in toluene). b Yield determined using

1H NMR analysis of crude with TMB as internal standard.

Page 62: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

51

2.5 Bimetallic Catalyzed Domino Attempt

Following optimization of two sets of conditions for the one-pot two-step synthesis of (E)-

alkenes from terminal alkynes and aryl halides via a Zr-catalyzed hydroboration/Pd-catalyzed

SMC sequence, a bimetallic catalyzed domino reaction was attempted. It was known from our

study of Zr-catalyzed hydroboration with HBpin that water must be excluded from the domino

reaction (Table 2.1-1, Entry 11); however, it was also known that dry toluene was a suitable

solvent for both Zr-catalyzed hydroboration (Table 2.1-4, Entry 3) and Pd-catalyzed SMC (Table

2.2-1, Entry 3), thus combining the two steps in domino fashion seemed plausible.

A solution of phenylacetylene and HBpin in dry toluene was added to a mixture of Schwartz’s

reagent, 4’-bromoacetophenone, Pd(PPh3)4, and Cs2CO3, and the resulting mixture was sealed

under argon and stirred at 80 °C for 16 hours (Scheme 2.5-1). Unfortunately, neither the desired

coupled product (4.1) nor the boronate intermediate (3.1) were detected using TLC or 1H NMR

analysis of the crude reaction mixture.

Scheme 2.5-1: Domino Attempt

The total absence of expected boronate intermediate 3.1 in the crude reaction mixture of the

domino attempt strongly suggested that one or more of the components of the SMC step impeded

Zr-catalyzed hydroboration. To determine which reagent(s) were deleterious, we carried out

several Zr-catalyzed hydroborations of phenylacetylene with HBpin in dry toluene, each in the

presence of one or more of the reagents used in the SMC step and in their relative amounts

(Table 2.5-1). In addition to water (Entry 2), which was already known to preclude Zr-catalyzed

hydroboration, Cs2CO3 inhibited the reaction (Entry 3), perhaps due to deprotonation of the

terminal alkyne and/or formation of an oxoborohydride species. 4’-Bromoacetophenone was

equally poor (Entry 4), likely due to the electrophilic carbonyl carbon. Proton NMR analysis of

Page 63: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

52

the crude reaction, while messy, suggested the presence of the alcohol; however, further

experimentation would be necessary to determine the reducing capability of HBpin and

Schwartz’s reagent under these conditions. A catalytic amount of Pd(PPh3)4 did not hinder

hydroboration (Entry 5), but only trace 3.1 was obtained in the presence of both the Pd catalyst

and 4’-bromoacetophenone (Entry 6). In the latter instance, a trace amount of coupled product

4.1 was detected in the crude reaction mixture, presumably arising from transmetallation

between the arylpalladium(II)bromide complex and either alkenylboronate 3.1 or, more directly,

its alkenylzirconocene precursor. Unlike 4’-bromoacetophenone, bromobenzene had no

detectable influence on hydroboration (Entry 7), and the combination of bromobenzene and

catalytic Pd(PPh3)4 decreased the expected yield of 3.1 by only a third (Entry 8). These data

suggest that use of an aryl halide lacking competing functional group(s) may enable domino

reactivity. Regardless, the base, which is necessary for the SMC step, remains a serious

impediment to the hydroboration step. Finding a base that facilitates Pd-catalyzed SMC without

hindering Zr-catalyzed hydroboration, should it exist, may solve this problem.

Table 2.5-1: Effect of SMC Components on Zr-Catalyzed Hydroboration

Entry Additive(s) (equiv)b

Yield A (%)c

Yield B (%)c

1 none 67 2

2 H2O (10) 0 0

3 Cs2CO3 (1.8) 8 3

4 4’-BrPhAc (0.91) 8 2

5 Pd(PPh3)4 (0.0091) 69 5

6

4’-BrPhAc (0.91) &

Pd(PPh3)4 (0.0091)

trace 3

Page 64: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

53

7 PhBr (0.91) 68 >1

8

PhBr (0.91) &

Pd(PPh3)4 (0.0091)

40 >1

a Reactions carried out on 0.22 mmol scale. Concentration: 0.55 M.

b Added directly before Cp2ZrHCl.

c Yield

determined using 1H NMR analysis of crude with TMB as internal standard.

Throughout the development of our one-pot two-step protocol, we had observed that the addition

of water to the crude product of the Zr-catalyzed hydroboration step (dissolved in toluene)

caused the mixture to lightly effervesce for a few minutes, although it did not warm to the touch.

While water was initially included in the SMC step simply because a toluene/water (10:1)

mixture provided better yields than dry toluene during optimization of the Suzuki reaction (Table

2.2-1), we now wondered if it served a dual purpose. Knowing that both Schwartz’s reagent and

HBpin are water sensitive, we hypothesized that the addition of water to the second step

deactivated the zirconocene catalyst and/or degraded any excess HBpin, either or both of which

may be deleterious to Pd-catalyzed cross-coupling. To test this, we conducted both one-pot two-

step reactions (the aryl bromide and aryl chloride variant) under anhydrous conditions (Scheme

2.5-2) and found that the addition of water to the SMC step was crucial. Whereas the optimized

one-pot two-step syntheses of 4.1 gave yields of 89% (NMR) from 4’-bromoacetophenone

(Table 2.4-1, Entry 8) and 80% (NMR) from 4’-chloroacetophenone (Table 2.4-4, Entry 4), only

10 and 17% yields, respectively, of 4.1 were obtained when water was omitted in the SMC step.

In both cases, unreacted aryl halide was present in the crude reaction mixture (85% 4’-bromo-

and 62% 4’-chloroacetophenone).

Scheme 2.5-2: One-Pot Two-Step under Anhydrous Conditions

Page 65: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

54

To test the effect of Schwartz’s reagent on SMC, we attempted to couple 4’-bromoacetophenone

with boronate 3.1 in dry toluene with and without the zirconocene present (Table 2.5-2).

Whereas 64% of coupled product 4.1 was obtained in the absence of Schwartz’s reagent (Entry

1), only 16% of 4.1 was obtained when 5.5 mol% of the zirconocene was present (Entry 2)

(amount employed in the hydroboration step of a typical one-pot two-step sequence).

Table 2.5-2: Effect of Cp2ZrHCl on Suzuki Coupling

Entry Additive (mol %) Yield (%)b

1 -- 64

2 Cp2ZrHCl (5.5) 16

a Reactions carried out on 0.20 mmol scale. Concentration: 0.5 M (aryl bromide in toluene).

b Yield determined

using 1H NMR analysis of crude with TMB as internal standard.

Based on these findings, the one-pot two-step Zr-catalyzed hydroboration/Pd-catalyzed SMC

sequence works for two reasons: 1) while the components of the second step severely hinder the

first, the components of the first step do not appear to hinder the second provided the second step

is conducted in the presence of water; and 2) in relation to the first point, Suzuki reactions are

water tolerant.

Page 66: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

55

2.6 Optimized Conditions for One-Pot Two-Step Zr-Catalyzed Hydroboration/Pd-Catalyzed Suzuki Coupling Sequence

Although unsuccessful in developing a bimetallic catalyzed domino reaction, we still had two

sets of optimized conditions for the efficient one-pot synthesis of (E)-alkenes from terminal

alkynes and aryl bromides or chlorides, thus we conducted a second substrate scope.

2.6.1 Substrate Scope of Optimized Conditions for Aryl Bromides

Following the optimized conditions for the one-pot two-step synthesis of 4-acetyl-trans-stilbene

(4.1) from phenylacetylene and 4’-bromoacetophenone (Table 2.4-1, Entry 8), we prepared a

variety of (E)-alkenes from various terminal alkynes and aryl- or heteroaryl bromides (Table

2.6-1). In some cases, the reaction temperature in the SMC step was increased from 80 to 100

°C and/or the Pd catalyst loading was increased from 1 to 2.5 mol% to achieve full conversion.

Reaction times were also increased in certain cases (see Experimental section).

Table 2.6-1: Substrate Scope of Optimized Conditions for Aryl Bromides

Entry Alkyne Aryl Bromide Product Yield (%)b

1

61

2

67

Page 67: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

56

3

58

4

59

5

89

6

78

a Reactions carried out on 1.0 mmol scale in 2 dram vials. Concentration of coupling reaction: 0.5 M (bromide in

toluene). b Isolated yield following chromatography based on amount of aryl bromide used.

Compared to the scope of our initial conditions (Table 2.3-1), the optimized conditions for our

one-pot two-step protocol gave superior yields of the desired (E)-alkenes in all cases where the

same substrates were reacted (Table 2.6-1) and generally under milder conditions. The yield of

coupled product 4.3 increased from 48% in our initial study to 61% (Entry 1), 4.6 increased from

62 to 67% (Entry 2), 4.14 from 73 to 89% (Entry 5), and 4.15 from 55 to 78% (Entry 6).

N-Propargyl indole (1.2) proved to be an excellent substrate for these conditions, the SMC step

going to completion in 8 to 10 hours at 80 °C in the presence of 1 mol% Pd catalyst to give

coupled products 4.14 (Entry 5) and 4.15 (Entry 6) in great yields over two steps. The protected

propargyl amine was not as reactive, and the SMC step required heating at 100 °C for 10 hours

in the presence of 2.5 mol% Pd catalyst to give 4.12 in a modest yield (Entry 4).

As expected, (PPh3)4Pd-catalyzed SMC was chemoselective for bromides over chlorides in the

presence of aryl halides bearing both, and bromides with para-chloro substituents were more

reactive substrates than those bearing ortho-chloro substituents. TLC analysis indicated 1-

bromo-4-chlorobenzene was consumed after 8 hours at 80 °C in the presence of 1 mol% Pd

catalyst (Entry 2). In contrast, SMC of 1-bromo-2-chlorobenzene required heating at 100 °C for

12 hours in the presence of 2.5 mol% Pd catalyst to achieve conversion, albeit in a lower yield

Page 68: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

57

(Entry 3). When the latter was subjected to the conditions of the former, only 22% of coupled

product 4.7 was obtained. Curiously, while 4.7 was a liquid at room temperature, its

constitutional isomer 4.6 was a solid with a melting point of 78 – 82 °C.

2.6.2 Substrate Scope of Optimized Conditions for Aryl Chlorides

Following the optimized conditions for the one-pot two-step synthesis of 4-acetyl-trans-stilbene

(4.1) from phenylacetylene and 4’-chloroacetophenone (Table 2.4-4, Entry 4), we prepared a

variety of (E)-alkenes from various aryl, heteroaryl, and aliphatic terminal alkynes and aryl- or

heteroarylchlorides (Table 2.6-1). In some cases, the reaction temperature in the SMC step was

increased from 80 to 100 °C and/or the Pd-XPhos precatalyst loading was increased from 1 to

2.5 mol% to achieve full conversion. Reaction times also varied (see Experimental section).

Table 2.6-2: Substrate Scope of Optimized Conditions for Aryl Chlorides

Entry Alkyne Aryl Chloride Product Yield (%)b

1

79

2

71

3

73

4

77

Page 69: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

58

5

86

6

71

7c

41

8

decomp.

a Reactions carried out on 1.0 mmol scale in 2 dram vials. Concentration of coupling reaction: 0.5 M (chloride in

toluene). b Isolated yield following chromatography based on amount of aryl chloride used.

c Hydroboration carried

out in toluene.

The optimized conditions for our one-pot two-step protocol incorporating aryl and heteroaryl

chlorides as substrates afforded good yields of the corresponding (E)-alkenes over two steps

(Table 2.6-2), with a couple of exceptions.

N-propargyl indole proved to be an excellent substrate for these conditions, providing the

corresponding (E)-alkenes in generally high yields over two steps (Entries 5 and 6); however, 1-

propargyl triazole proved to be more challenging (Entry 7). Unlike N-propargyl indole, which

readily underwent Zr-catalyzed hydroboration with HBpin in the absence of solvent at 45 °C, 1-

propargyl triazole required the use of toluene as a solvent for this step. When hydroboration of

1-propargyl triazole was attempted under solvent-free conditions, the mixture remained solid and

would not stir, even when the temperature was increased to 60 °C. Furthermore, SMC of the

corresponding boronate of 1-propargyl triazole required heating at 100 °C for 24 hours in the

presence of 2.5 mol% Pd catalyst to achieve conversion, whereas coupling of the corresponding

boronate (3.2) of N-propargyl indole went to completion within only 2 hours at 80 °C in the

presence of 1 mol% Pd catalyst and gave much higher yields of the resulting (E)-alkenes.

Page 70: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

59

2-Chloroquinoline proved to be a suitable coupling partner, furnishing good yields of the

corresponding 2-alkenylquinolines (Entries 3 and 4), but 2-chlorobenzoxazole decomposed

within 2 hours under these conditions and no desired product was detected using 1H NMR

analysis of the crude reaction mixture (Entry 8).

2.6.3 Synthesis of (E)-2-Styrylbenzoxazole

In 2007, Buchwald and co-workers58

reported the cross-coupling of aryl boronic acids with a

myriad of conventionally challenging heteroaryl chlorides employing Pd-XPhos as the catalyst,

n-butanol as the solvent, and K3PO4 as the base. Although it was not among the heteroaryl

halides coupled, we applied Buchwald’s conditions to the SMC of 2-chlorobenzoxazole with 3.1

as the coupling partner (Table 2.6-3).

Table 2.6-3: Solvent and Base Screen for SMC with 2-Chlorobenzoxazole

Entry Solventb mol%

Pd-XPhos-G2 Base Temp (°C) Yield (%)

c

1 PhMe (distilled) 2.5 Cs2CO3 100 decomp.

2 n-BuOH (reagent A.C.S.) 2.5 Cs2CO3 100 26

3 n-BuOH (reagent A.C.S.) 2.5 K3PO4 100 47

4 2-BuOH (anhydrous) 2.5 K3PO4 100 58

5 2-BuOH (anhydrous) 2.5 K3PO4 80 55

6 2-BuOH (anhydrous) 5.0 K3PO4 100 48

a Reactions carried out on 0.20 mmol scale.

b Concentration: 0.5 M.

c Yield determined by

1H NMR analysis of

crude with TMB as internal standard.

The combination of n-butanol and K3PO4 gave 4.18 in a modest yield (Entry 3), which was

improved by switching to 2-butanol (Entry 4). The superior performance of 2-butanol was

Page 71: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

60

presumably due to its higher grade as opposed to some special property conferred by it being a

branched rather than a linear alcohol. Distillation of n-butanol would presumably improve its

utility as a solvent.

Although some product formed, 2-chlorobenzoxazole still decomposed in butanol and was not

detected in the reaction mixture after 4 hours. Increasing the catalyst loading from 2.5 to 5.0

mol% had a negative impact on yield (Entry 6), suggesting the oxidative addition product was

unstable. For this substrate, it appeared cross-coupling was in competition with decomposition.

The rate of transmetallation was presumably enhanced in the alcohol relative to toluene, perhaps

through formation of a butoxypalladium(II) or butoxyborate species, thus cross-coupling was fast

enough to generate some desired product even in the presence of a decomposition pathway.

We applied the modified conditions for SMC of 2-chlorobenzoxazole with 3.1 to our one-pot

two-step protocol and isolated coupled product 4.18 in a 49% yield (Scheme 2.6-1), which was

deemed satisfactory given the challenging nature of this substrate.

Scheme 2.6-1: One-Pot Two-Step Synthesis of (E)-2-Styrylbenzoxazole

Page 72: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

61

2.7 Conclusions and Future Outlook

Two sets of conditions have been described for the one-pot two-step or telescopic synthesis of

(E)-alkenes. In the first step, zirconium-catalyzed hydroboration of a terminal alkyne with

pinacolborane generates the corresponding alkenyl pinacolboronate intermediate, which can then

undergo palladium-catalyzed cross-coupling with an aryl or heteroaryl halide in the same

reaction vessel. A variety of terminal alkynes can be used including aryl and aliphatic alkynes,

cyclic enynes, and propargylic indoles, benzotriazoles, and amines. Initial hydrozirconation of

the alkyne sets the regio- and stereochemistry of the resulting alkenylboronate, which is retained

in the cross-coupling step. Aryl and heteroaryl chlorides, which are unreactive under traditional

Suzuki conditions employing palladium catalysts derived from triarylphosphine ligands, can be

used as coupling partners when bulky, electron-rich, dialkylbiaryl phosphine ligands are

employed. Among catalysts screened, the preformed XPhos palladacycle proved to be the most

efficient under these conditions.

This bimetallic catalyzed sequence has several key features: relative ease of set-up; mild

conditions for the hydroboration step; lack of need to isolate boronate intermediate, which saves

time, reduces waste, and likely improves overall yield; and low catalyst loading in the cross-

coupling step. Furthermore, zirconium is much less expensive and more earth-abundant than

many other transition metals employed in catalyzed hydroboration, particularly rhodium. This

sequence was not amenable to domino reactivity as the components of the cross-coupling step

hinder hydroboration. In kind, the components of the hydroboration step hinder cross-coupling;

however, the addition of water to the second step prevents this from happening by deactivating

the otherwise deleterious zirconocene catalyst and/or degrading excess pinacolborane. Thus this

sequence succeeds by exploiting the water tolerance of the Suzuki-Miyaura reaction.

In general, the desired (E)-alkenes were isolated in good yields over two steps on a 1.0 mmol

scale. N-Propargyl indole proved to be a particularly effective substrate for these conditions,

thus an efficient method for the one-pot synthesis of N-cinnamyl indole derivatives has been

provided. These conditions are also applicable to the efficient synthesis of 2-alkenylquinolines.

Page 73: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

62

In 2010, Lam and co-workers124

reported the asymmetric rhodium-catalyzed addition of boronic

acids to alkenylheteroarenes using chiral diene ligands and 2-hexenylquinoline (4.10) as the

model substrate. In light of Lam’s work, and the fact that rhodium does not bind XPhos as

previous work by our group has demonstrated,6 we propose that our conditions may be amenable

to a sequential rhodium-catalyzed hydroarylation step.

124 Pattison, G.; Piraux, G.; Lam, H. W. J. Am. Chem. Soc. 2010, 132, 14373.

Page 74: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

63

3 Experimental

3.1 General Considerations

All metal-catalyzed reactions were carried out under an inert atmosphere of dry argon or dry

nitrogen using glassware that was oven- or flame-dried and cooled under nitrogen. Work-up and

isolation of products of all reactions was conducted on the bench-top using standard techniques.

Organic solutions were concentrated under reduced pressure on a Büchi rotary evaporator and a

high vaccum pump in tandem. Unless otherwise stated, yields quoted are isolated yields.

Commercial reagents were purchased from Sigma-Aldrich, Strem Chemicals, Alfa Aesar,

Combi-Blocks, or BDH Chemicals and were used without further purification. Zirconocene

hydrochloride (Schwartz’s reagent) was purchased from Alfa Aesar and palladium catalysts from

Strem Chemicals. Pinacolborane was purchased from Sigma-Aldrich and stored under argon in a

refrigerator (4°C). Anhydrous cesium carbonate was ground and stored under argon in a

dessicator. Toluene and tetrahydrofuran were distilled over sodium; dichloromethane was

distilled over calcium hydride; 2-butanol was purchased from Alfa Aesar and N,N-

dimethylformamide from Fisher Scientific and both were used as received.

Reactions were monitored by thin layer chromatography (TLC) on EMD Silica Gel 60 F254

plates. Visualization of the developed plates was enabled with UV light (254 or 365 nm) and/or

by staining with either KMnO4 or p-anisaldehyde. Silica gel flash column chromatography was

performed on Silicycle 230 – 400 mesh silica gel.

NMR characterization data were collected on a Varian Mercury 300 or a Bruker Avance III

spectrometer operating at 300 or 400 MHz for 1H NMR, 75 or 100 MHz for

13C NMR, 128 MHz

for 11

B NMR, and 377 MHz for 19

F NMR. 1H NMR spectra were internally referenced to the

residual solvent signal (CDCl3 = 7.26 ppm) or TMS. 13

C NMR spectra were internally

referenced to the residual solvent signal (CDCl3 = 77.0 ppm). Chemical shifts (δ) are reported in

ppm and are given to the nearest 0.01 ppm. Coupling constants (J) are reported in Hz and are

given to the nearest 0.1 Hz. Signal multiplicities are as follows: s = singlet, d = doublet, t =

triplet, q = quartet, p = pentet, m = mulitplet, br = broad. Regioisomeric ratios reported were

obtained by 1H NMR analysis of crude reaction mixtures with a 10 second relaxation delay.

Page 75: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

64

NMR yields were obtained by 1H NMR analysis of crude reaction mixtures with a 10 second

relaxation delay and 1,3,5-trimethoxybenzene (TMB) as the internal standard.

Melting points reported were determined using a Fisher-Johns melting point apparatus and are

uncorrected. Infrared (IR) spectra were obtained using a Perkin-Elmer Spectrum 1000 FT-IR

spectrometer as thin films from dichloromethane or chloroform. High resolution mass spectra

(HRMS) were obtained using a JEOL AccuTOF JMS-T1000LC equipped with an IONICS®

Direct Analysis in Real Time (DART) ion source or an ABI/Sciex Qstar mass spectrometer (ESI)

at Advanced Instrumentation for Molecular Structure (AIMS) in the Department of Chemistry at

the University of Toronto.

Page 76: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

65

3.2 Synthesis of Starting Materials

1,1-dibromo-4-phenyl-1-butene – Synthesized by adaptation of the

procedure reported by Gibtner et al.125

To a solution of CBr4 (2.0 equiv, 8.0

mmol, 2.65 g) and PPh3 (4.0 equiv, 16.0 mmol, 4.2 g) in DCM (10 mL) was added

hydrocinnamaldehyde (1.0 equiv, 4.0 mmol, 537 mg) at 0 °C. The mixture was stirred at this

temperature for 10 minutes then allowed to warm to room temperature. After 15 minutes at this

temperature, the mixture was concentrated in vacuo to give an intractable purple gum. Filtration

through a pad of silica gel with hexanes:EtOAc (50:50 v/v) (300 mL) as the eluent gave the

crude product as a pale yellow oil. Purification by silica gel flash column chromatography using

hexanes:EtOAc (49:1 v/v) as the eluent gave the gem-dibromoolefin as a clear colourless oil (630

mg, 54%). 1H NMR (400 MHz, CDCl3) δ 7.36 – 7.27 (m, 2H), 7.25 – 7.17 (m, 3H), 6.43 (t, J =

7.2 Hz, 1H), 2.75 (t, J = 8.2 Hz, 2H), 2.43 (dt, J = 8.2, 7.2 Hz, 2H). 13

C NMR (100 MHz,

CDCl3) δ 140.49, 137.58, 128.49, 128.34, 126.23, 89.46, 34.63, 33.84. The spectroscopic data

are in agreement with literature values.126

4-phenyl-1-butyne (1.1) – Synthesized by adaptation of the procedure reported

by Gibtner et al.125

To a flame-dried 100 mL round-bottom flask under an

atmosphere of argon was added (4,4-dibromobut-3-en-1-yl)benzene (1.0 equiv, 2.76 mmol, 800

mg) and dissolved in dry THF (10 mL). LDA (3.0 equiv, 0.5 M in dry THF, 16.6 mL) was

added dropwise with stirring at -78 °C, and the mixture was stirred at this temperature for 30

minutes then allowed to warm to room temperature and stirred for an additional 1 hour. At this

time, TLC analysis of the reaction mixture indicated incomplete conversion of starting material.

The mixture was re-cooled to -78 °C and n-BuLi (1.0 equiv, 2.34 M in hexanes, 1.18 mL) was

added dropwise. The mixture was stirred at this temperature for 15 minutes, after which TLC

125 Gibtner, T.; Hampel, F.; Gisselbrecht, J. –P.; Hirsh, A. Chem. Eur. J. 2008, 9, 408.

126 Yokota, M.; Fujita, D.; Ichikawa, J. Org. Lett. 2007, 9, 4639.

Page 77: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

66

analysis indicated complete conversion of the gem-dibromoolefin. The reaction was quenched

with 2M HCl solution (14 mL) followed by dilution with hexanes (30 mL). The aqueous phase

was extracted with hexanes (3 x 20 mL), and the combined organic layers were dried over

Na2SO4, filtered, and concentrated in vacuo to afford the crude product an an orange oil.

Purification by silica gel flash column chromatography using hexanes:EtOAc (24:1 v/v) as the

eluent gave the alkyne as a clear colourless oil (144 mg, 40%). 1H NMR (300 MHz, CDCl3) δ

7.34 – 7.24 (m, 2H), 7.24 – 7.15 (m, 3H), 2.83 (t, J = 7.6 Hz, 2H), 2.46 (td, J = 7.6, 2.6 Hz, 2H),

1.95 (t, J = 2.6 Hz, 1H). 13

C NMR (75 MHz, CDCl3) δ 140.31, 128.32, 128.31, 126.27, 83.69,

68.88, 34.76, 20.49. The spectroscopic data are in agreement with literature values.127

1-(prop-2-yn-1-yl)-1H-indole (1.2) – Synthesized according to the procedure

reported by Haider & Käferböck.128

No measures were taken to exclude air or

moisture. To a solution of indole (1.0 equiv, 10.0 mmol, 1.17 g) and propargyl bromide (80% in

toluene, 1.5 equiv, 15.0 mmol, 1.67 mL) in toluene (30 mL) were added tetrabutylammonium

bromide (0.050 equiv, 0.50 mmol, 161 mg) and 50% aqueous NaOH (6 mL), and the two-phase

mixture was stirred vigorously at room temperature. After 1 hour, the reaction mixture was

diluted with water (20 mL), the layers were separated, and the aqueous phase was extraced with

Et2O (20 mL). The combined organic layers were washed with water (3 x 20 mL), dried over

Na2SO4, filtered, and concentrated in vacuo. Purification by silica gel flash column

chromatography using hexanes:EtOAc (19:1 v/v) as the eluent gave the propargylated indole as a

yellow oil that eventually solidified to a beige solid (1.13g, 71%, MP = 64 °C (lit. 65 – 69)128

).

1H NMR (400 MHz, CDCl3) δ 7.65 (apparent dt, 1H), 7.42 (apparent dq, 1H), 7.29 – 7.23 (m,

1H), 7.22 (d, J = 3.2 Hz, 1H), 7.15 (ddd, J = 8.0, 7.1, 1.0 Hz, 1H), 6.54 (dd, J = 3.3, 0.9 Hz, 1H),

4.89 (d, J = 2.6 Hz, 2H), 2.40 (t, J = 2.6 Hz, 1H). 13

C NMR (100 MHz, CDCl3) δ 135.76,

128.87, 127.19, 121.86, 121.10, 119.85, 109.27, 102.08, 77.70, 73.48, 35.77. The spectroscopic

data are in agreement with literature values.128

127 Beshai, M.; Dhudshia, B.; Mills, R.; Thadani, A.N. Tetrahedron Letters 2008, 49, 6794.

128 Haider, N.; Käferböck, J. Tetrahedron 2004, 60, 6495.

Page 78: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

67

1-(prop-2-yn-1-yl)-1H-benzo[d][1,2,3]triazole (1.3) – Synthesized according

to the procedure reported by Katritzky et al.129

To a flame-dried 25 mL round-

bottom flask under an atmosphere of argon was added NaH (60 wt% in mineral

oil, 9.5 mmol, 380 mg) and suspended in dry DMF (3.5 mL). To the suspension was added a

solution of benzotriazole (1.0 equiv, 10.0 mmol, 1.19 g) in dry DMF (1.5 mL) at 0 °C, and the

mixture was stirred at this temperature for 20 minutes then allowed to warm to room temperature

and stirred for an additional 20 minutes. The mixture was cooled to 0 °C and propargyl bromide

(80% in toluene, 1.1 equiv, 11.0 mmol, 1.23 mL) was added dropwise, after which the mixture

was allowed to warm to room temperature. After stirring at this temperature for 2 hours, the

reaction was quenched with saturated NaHCO3 (10 mL), and the aqueous layer was extracted

with Et2O (3 x 10 mL). The combined organic layers were washed once with saturated NaHCO3

(20 mL) followed by water (3 x 20 mL), dried over MgSO4, filtered, and concentrated in vacuo.

Purification by silica gel flash column chromatography using hexanes:EtOAc (4:1 v/v) as the

eluent gave the propargylated benzotriazole as a yellow oil that eventually solidified to a light

yellow solid (846 mg, 54%, MP = 82 – 83 °C (lit. 57)130

). 1H NMR (400 MHz, CDCl3) δ 8.08

(dt, J = 8.4, 0.9 Hz, 1H), 7.72 (dt, J = 8.3, 1.0 Hz, 1H), 7.54 (ddd, J = 8.2, 7.1, 1.0 Hz, 1H), 7.41

(ddd, J = 8.1, 7.0, 1.0 Hz, 1H), 5.47 (d, J = 2.6 Hz, 2H), 2.50 (t, J = 2.6 Hz, 1H). 13

C NMR (100

MHz, CDCl3) δ 146.26, 132.42, 127.66, 124.16, 120.14, 109.74, 75.21, 75.04, 37.99. The

spectroscopic data are in agreement with literature values.130

1-(prop-2-yn-1-yl)indoline-2,3-dione (1.4) – Synthesized according to the

procedure reported by Leugn, C.; Tomaszewski, M.; Woo, S. New

Compounds. US Patent 185179 A1, February 5, 2007. To a solution of isatin

(1.0 equiv, 10.0 mmol, 1.47 g) in dry DMF (40 mL) was added Cs2CO3, and the mixutre was

stirred at room temperature. After 1.5 hours, propargyl bromide (80% in toluene, 1.2 equiv, 12.0

mmol, 1.34 mL) was added dropwise, and the mixture was stirred at room temperature overnight.

The reaction was quenched with saturated NaHCO3 (30 mL) and extracted into EtOAc (3 x 20

mL). The combined organic layers were washed with copious water (6 x 30 mL), dried over

MgSO4, filtered, and concentrated in vacuo. The residue was filtered through a pad of silica gel

with EtOAc as the eluent. Further purifiction was not necessary, and the propargylated isatin

129 Katritzky, A.R.; Button, M.A.C.; Denisenko, S.N. J. Heterocyclic Chem. 2000, 37, 1505.

130 Katritzky, A.R.; Oniciu, D.C.; Ghiviriga, I. Synth. Commun. 1997, 27, 1613.

Page 79: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

68

was obtained as a bright orange solid (1.62 g, 88%, MP = 154 – 159 °C (lit. 156 – 158)131

). 1H

NMR (400 MHz, CDCl3) δ 7.68 – 7.61 (m, 2H), 7.18 (td, J = 7.5, 0.8 Hz, 1H), 7.15 – 7.10 (m,

1H), 4.54 (d, J = 2.6 Hz, 2H), 2.31 (t, J = 2.5 Hz, 1H). 13

C NMR (100 MHz, CDCl3) δ 182.48,

157.13, 149.60, 138.37, 125.46, 124.18, 117.70, 111.05, 75.66, 73.31, 29.44. The spectroscopic

data are in agreement with literature values.131

5-bromo-1-tosyl-1H-indole (2.1) – Synthesized by adaptation of the procedure

reported by Wang & Liu.132

No measures were taken to exclude ambient air or

moisture. To a solution of 5-bromoindole (1.0 equiv, 10.2 mmol, 2.0 g) in toluene (20 mL) were

added tetrabutylammonium hydrogensulfate (0.070 equiv, 0.714 mmol, 242 mg) and 50%

aqueous KOH (12 mL) followed by p-toluenesulfonyl chloride (1.2 equiv, 12.2 mmol, 2.33 g) in

toluene (8 mL), and the resultant mixture was stirred at room temperature. After 2 hours, the

reaction was diluted with water (30 mL) and extracted into EtOAc (3 x 20 mL). The combined

organic layers were washed with water (2 x 50 mL) followed by brine (30 mL), dried over

MgSO4, filtered, and concentrated in vacuo. The crude product was filtered through a pad of

silica gel with EtOAc as the eluent. Further purification was not necessary, and the tosylated

indole was obtained as an off-white solid (3.33 g, 93%, MP = 162 °C (lit. 136 – 137)133

). 1H

NMR (400 MHz, CDCl3) δ 7.90 – 7.82 (m, 1H), 7.78 – 7.70 (m, 2H), 7.66 (d, J = 1.9 Hz, 1H),

7.56 (d, J = 3.7 Hz, 1H), 7.39 (dd, J = 8.8, 2.0 Hz, 1H), 7.26 – 7.20 (m, 2H), 6.59 (dd, J = 3.7,

0.8 Hz, 1H), 2.35 (s, 3H). 13

C NMR (100 MHz, CDCl3) δ 145.23, 134.99, 133.51, 132.44,

129.95, 127.54, 127.42, 126.76, 124.01, 116.74, 114.92, 108.24, 21.56. The spectroscopic data

are in agreement with literature values.133

131 Bouhfid, R.; Joly, N.; Essassi, El M.; Lequart, V.; Massoui, M.; Martin, P. Synth. Commun. 2011, 41, 2096.

132 Wang, K.; Liu, Z. Synth. Commun. 2010, 40, 144.

133 Prieto, M.; Zurita, E.; Rosa, E.; Muñoz, L.; Lloyd-Williams, P.; Giralt, E. J. Org. Chem. 2004, 69, 6812.

Page 80: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

69

3.3 Synthesis of Alkenylboronates

(E)-4,4,5,5-tetramethyl-2-styryl-1,3,2-dioxaborolane (3.1) – Synthesized

by adaptation of the procedure reported by Wang et al.120

Phenylacetylene

(1.0 equiv, 1.10 mmol, 113 mg) and Cp2ZrHCl (10 mol%, 28.4 mg) were

weighed into an oven-dried 2 dram vial and purged with nitrogen for one minute. To this

mixture, pinacolborane (1.1 equiv, 1.21 mmol, 155 mg) was added dropwise. The reaction vial

was sealed under argon, covered with aluminum foil to reduce exposure to light, and the mixture

was stirred at room temperature for 16 hours. The initially pale yellow solution became bright

orange over time, presumably indicating progression. After 16 hours, the mixture was filtered

through a thin pad of silica gel with Et2O (20 mL) as the eluent and concentrated in vacuo.

Purification by silica gel flash column chromatography using hexanes:Et2O (9:1 v/v initially then

4:1 v/v once product began eluting) as the eluent gave the styrylboronate as a clear colourless oil,

which crystallized upon storage in freezer (204 mg, 81%). 1H NMR (400 MHz, CDCl3) δ 7.52 –

7.45 (m, 2H), 7.40 (d, J = 18.4 Hz, 1H), 7.36 – 7.24 (m, 3H), 6.17 (d, J = 18.4 Hz, 1H), 1.30 (s,

12H). 13

C NMR (100 MHz, CDCl3) δ 149.43, 137.40, 128.81, 128.48, 126.97, 83.23, 24.74

(vinyl boron C absent). 11

B NMR (128 MHz, CDCl3) δ 30.35. The spectroscopic data are in

agreement with literature values.123

(E)-1-(3-(4,4,5,5-tetramethyl-1,3,2-dioxaborolan-2-yl)allyl)-1H-

indole (3.2) – Synthesized by adaptation of the procedure reported by

Wang et al.120

1-Propargyl indole (1.0 equiv, 0.50 mmol, 77.6 mg)

and Cp2ZrHCl (5 mol%, 6.5 mg) were weighed into an oven-dried 2 dram vial and purged with

nitrogen for one minute. To this mixture, pinacolborane (1.1 equiv, 0.55 mmol, 70.4 mg) was

added dropwise. The reaction vial was sealed under argon, and the mixture was stirred at 45 °C.

After 16 hours, the mixture was filtered through a thin pad of silica gel with Et2O (20 mL) as the

eluent and concentrated in vacuo. Purification by silica gel flash column chromatography using

hexanes:Et2O (9:1 v/v) as the eluent gave the title compound as an orange oil (91 mg, 65%). 1H

NMR (400 MHz, CDCl3) δ 7.65 (dt, J = 7.8, 1.0 Hz, 1H), 7.30 (dt, J = 8.3, 1.0 Hz, 1H), 7.21

(ddd, J = 8.2, 7.0, 1.2 Hz, 1H), 7.13 (ddd, J = 7.9, 7.1, 1.1 Hz, 1H), 7.08 (d, J = 3.2 Hz, 1H), 6.74

(dt, J = 18.0, 4.7 Hz, 1H), 6.54 (dd, J = 3.2, 0.9 Hz, 1H), 5.38 (dt, J = 17.9, 1.8 Hz, 1H), 4.82

(dd, J = 4.7, 1.9 Hz, 2H), 1.26 (s, 12H). 13

C NMR (400 MHz, CDCl3) δ 147.35, 136.06, 128.55,

Page 81: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

70

127.97, 121.50, 120.85, 119.33, 109.47, 101.50, 83.34, 49.82, 24.72 (vinyl boron C absent). 11

B

NMR (128 MHz, CDCl3) δ 29.83. The spectroscopic data are in agreement with literature

values.134

(Z)-2-(hex-3-en-3-yl)-4,4,5,5-tetramethyl-1,3,2-dioxaborolane (3.3) –

Synthesized by adaptation of the procedure reported by Wang et al.120

3-

Hexyne (1.0 equiv, 1.10 mmol, 90.4 mg) and Cp2ZrHCl (10 mol%, 28.4 mg)

were weighed into an oven-dried 2 dram vial and purged with nitrogen for one minute. To this

mixture, pinacolborane (1.1 equiv, 1.21 mmol, 155 mg) was added dropwise. The reaction vial

was sealed under argon and stirred at 60 °C. After 60 hours, the mixture was filtered through a

thin pad of silica gel with Et2O (20 mL) as the eluent and concentrated in vacuo. The yield was

calculated to be 67% using 1H NMR analysis of the crude reaction mixture.

103

(Z)-4,4,5,5-tetramethyl-2-(1-phenylprop-1-en-2-yl)-1,3,2-dioxaborolane

(3.4) – Synthesized by adaptation of the procedure reported by Wang et

al.120

1-Phenyl-1-propyne (1.0 equiv, 1.10 mmol, 128 mg) and Cp2ZrHCl

(10 mol%, 28.4 mg) were weighed into an oven-dried 2 dram vial and purged with nitrogen for

one minute. To this mixture, pinacolborane (1.1 equiv, 1.21 mmol, 1545 mg) was added

dropwise. The reaction vial was sealed under argon and stirred at 60 °C. After 60 hours, the

mixture was filtered through a thin pad of silica gel with Et2O (20 mL) as the eluent and

concentrated in vacuo. The yield was calculated to be 74% (mixture of regioisomers) using 1H

NMR analysis of the crude reaction mixture.135

The ratio of β- to α-borylation was 85:15.

134 For NMR spectra, see: Shade, R.E.; Hyde, A.M.; Olsen, J. –C.; Merlic, C.A. J. Am. Chem. Soc. 2010, 132, 1202.

135 For NMR spectra, see: Kim, H.R.; Jung, I.G.; Yoo, K.; Jang, K.; Lee, E.S.; Yun, J.; Son, S.U. Chem. Commun.

2010, 46, 758.

Page 82: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

71

3.4 Synthesis of (E)-Alkenes

General Procedure A: Initial One-Pot Two-Step Zr-Pd Catalyzed Synthesis of Disubstituted

(E)-Alkenes from Terminal Alkynes and Aryl Bromides and Iodides

The alkyne (1.1 equiv) and Cp2ZrHCl (11 mol%) were weighed into an oven-dried 3 dram vial

and purged with nitrogen for one minute. To this mixture, pinacolborane (1.21 equiv) was added

dropwise. The reaction vial was sealed and stirred at the indicated temperature. After 24 hours,

the crude mixture containing the alkenylboronate intermediate was dissolved in toluene (4.0

mL). To the resultant solution were added water (400 µL), aryl halide (1.0 equiv), Cs2CO3 (2.0

equiv), and Pd(PPh3)4 (1.0 mol %) in succession, purging the reaction vial with nitrogen

following each addition. Upon the addition of water, the mixture effervesced for a few minutes

but did not warm. The reaction vial was sealed under argon with a Teflon cap and stirred

vigorously at 100 °C. After 16 hours, the reaction was allowed to cool to room temperature,

filtered through a thin pad of silica gel with Et2O (20 mL) as the eluent, and concentrated in

vacuo. The crude product was purified by silica gel flash column chromatography using the

indicated mobile phase.

General Procedure B: Improved One-Pot Two-Step Zr-Pd Catalyzed Synthesis of Disubstituted

(E)-Alkenes from Terminal Alkynes and Aryl Bromides and Iodides

The alkyne (1.1 equiv) and Cp2ZrHCl (5.5 mol%) were weighed into an oven-dried 2 dram vial

and purged with nitrogen for one minute. To this mixture, pinacolborane (1.21 equiv) was added

dropwise. The reaction vial was sealed under argon with a Teflon cap and stirred at 45 °C. After

16 hours, the reaction was allowed to cool to room temperature and the crude mixture containing

the alkenylboronate intermediate was dissolved in toluene (2.0 mL). To the resultant solution

Page 83: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

72

were added water (200 µL), aryl halide (1.0 equiv), Cs2CO3 (2.0 equiv), and the indicated

amount of Pd(PPh3)4 in succession, purging the reaction vial with nitrogen following each

addition. If the aryl halide was an oil, it was added as a solution in toluene (2.0 mL). Upon the

addition of water, the mixture effervesced for a few minutes but did not warm. The reaction vial

was sealed under argon with a Teflon cap and stirred vigorously at the indicated temperature.

After the indicated time, the reaction was allowed to cool to room temperature, transferred to a

seperatory funnel with Et2O (20 mL), and quenched with saturated NH4Cl solution (20 mL). The

aqueous layer was extracted with Et2O (3 x 20 mL), and the combined organic layers were dried

over MgSO4, filtered, and concentrated in vacuo. The crude product was purified by silica gel

flash column chromatography using the indicated mobile phase.

General Procedure C: One-Pot Two-Step Zr-Pd Catalyzed Synthesis of (E)-Alkenes from

Terminal Alkynes and Aryl Chlorides

The alkyne (1.1 equiv) and Cp2ZrHCl (5.5 mol%) were weighed into an oven-dried 2 dram vial

and purged with nitrogen for one minute. To this mixture, pinacolborane (1.21 equiv) was added

dropwise. The reaction vial was sealed under argon with a Teflon cap and stirred at 45 °C. After

16 hours, the reaction was allowed to cool to room temperature and the crude mixture containing

the alkenylboronate intermediate was dissolved in toluene (2.0 mL). To the resultant solution

were added water (200 µL), aryl chloride (1.0 equiv), Cs2CO3 (2.0 equiv), and the indicated

amount of XPhos-Pd-G2 in succession, purging the reaction vial with nitrogen following each

addition. If the aryl chloride was an oil, it was added as a solution in toluene (2.0 mL). Upon

the addition of water, the mixture effervesced for a few minutes but did not warm. The reaction

vial was sealed under argon with a Teflon cap and stirred vigorously at the indicated

temperature. After the indicated time, the reaction was allowed to cool to room temperature,

Page 84: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

73

transferred to a seperatory funnel with Et2O (20 mL), and quenched with saturated NH4Cl

solution (20 mL). The aqueous layer was extracted with Et2O (3 x 20 mL), and the combined

organic layers were dried over MgSO4, filtered, and concentrated in vacuo. The crude product

was purified by silica gel flash column chromatography using the indicated mobile phase.

Reaction Scope:

(E)-1-(4-styrylphenyl)ethanone (4.1) – 1) To a 2 dram vial were added

(E)-4,4,5,5-tetramethyl-2-styryl-1,3,2-dioxaborolane (1.2 equiv, 0.24

mmol, 55.2 mg), toluene (1.0 mL), water (100 µL), 4’-iodoacetophenone

(1.0 equiv, 0.20 mmol, 246.05 mg), Cs2CO3 (2.0 equiv, 0.40 mmol, 130 mg), and Pd(PPh3)4 (5.0

mol%, 0.010 mmol, 11.6 mg) in succession, purging with nitrogen following each addition. The

reaction vial was sealed under argon with a Teflon cap and stirred at 100 °C. After 19 hours, the

mixture was filtered through a thin pad of silica gel with Et2O (10 mL) as the eluent and

concentrated in vacuo. Purification by silica gel flash column chromatography using

hexanes:EtOAc (9:1 v/v) as the eluent gave the title compound as an off-white solid (38 mg,

86%), which fluoresces under UV light (365 nm). The spectroscopic data are in agreement with

those reported below.

2) Synthesized according to general procedure A from phenylacetylene and 4’-iodoacetophenone

on a 1.0 mmol scale. Hydroboration run at room temperature. Purification by silica gel flash

column chromatography using hexanes:EtOAc (4:1 v/v) as the eluent gave the title compound as

a light brown solid (123 mg, 55%), which fluoresces under UV light (365 nm). The

spectroscopic data are in agreement with those reported below.

3) Synthesized according to general procedure B from phenylacetylene and 4’-

bromoacetophenone on a 1.0 mmol scale. Cross-coupling reaction employed 1.0 mol%

Pd(PPh3)4 and was run at 80 °C for 8 hours. The crude product was purified by silica gel flash

column chromatography using hexanes:EtOAc (9:1) as the eluent, co-eluting with a yellow trace

impurity. Subsequent trituration with hexanes gave the title compound as an off-white solid (164

mg, 74%, MP = 164 – 166 °C (lit. 148 – 150)136

), which fluoresces under UV light (365 nm). 1H

136 Iyer, S.; Kulkami, G.M.; Ramesh, C. Tetrahedron 2004, 60, 2163.

Page 85: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

74

NMR (400 MHz, CDCl3) δ 7.96 (d, J = 8.4 Hz, 1H), 7.59 (d, J = 8.4 Hz, 1H), 7.57 – 7.52 (m,

2H), 7.42 – 7.35 (m, 2H), 7.34 – 7.28 (m, 1H), 7.23 (d, J = 16.4 Hz, 1H), 7.13 (d, J = 16.4 Hz,

1H), 2.61 (s, 3H). 13

C NMR (100 MHz, CDCl3) δ 197.41, 141.98, 136.67, 135.94, 131.44,

128.84, 128.77, 128.29, 127.42, 126.79, 126.47, 26.56. The spectroscopic data are in agreement

with literature values. Only spectra for title compound synthesized by general procedure B are

shown below.137

(E)-1-(4-(4-methylstyryl)phenyl)ethanone (4.2) – Synthesized

according to general procedure A from 4-ethynyltoluene and 4’-

bromoacetophenone on a 1.0 mmol scale. Hydroboration run at room

temperature. Purification by silica gel flash column chromatography using hexanes:EtOAc (4:1

v/v) as the eluent gave the title compound as a beige solid (138 mg, 58%, MP = 167 – 170 °C),

which fluroesces under UV light (365 nm). 1H NMR (400 MHz, CDCl3) δ 7.95 (d, J = 8.4 Hz,

2H), 7.57 (d, J = 8.4 Hz, 2H), 7.44 (d, J = 8.1 Hz, 2H), 7.19 (d, J = 8.1 2H), 7.21 (d, J = 16.4 Hz,

1H), 7.09 (d, J = 16.4 Hz, 1H), 2.61 (s, 3H), 2.38 (s, 3H). 13

C NMR (100 MHz, CDCl3) δ

197.43, 142.22, 138.35, 135.75, 133.91, 131.41, 129.50, 128.84, 126.73, 126.42, 126.33, 26.55,

21.30. The spectroscopic data are in agreement with literature values.138

(E)-3-(3,5-difluorostyryl)thiophene (4.3) – 1) Synthesized according to

general procedure A from 1-ethynyl-3,5-difluorobenzene and 3-

bromothiophene on a 1.0 mmol scale. Hydroboration run at room

temperature. Purification by silica gel flash column chromatography using

hexanes as the eluent gave the title compound as a white solid (107 mg, 48%). The

spectroscopic data are in agreement with those reported below.

2) Synthesized according to general procedure B from 1-ethynyl-3,5-difluorobenzene and 3-

bromothiophene on a 1.0 mmol scale. Cross-coupling reaction employed 2.5 mol% Pd(PPh3)4

and was run at 100 °C for 8 hours. Purification by silica gel flash column chromatography using

hexanes as the eluent gave the title compound as a white solid (136 mg, 61%, MP = 92 – 93 °C).

1H NMR (400 MHz, CDCl3) δ 7.33 (tq, J = 4.1, 2.2 Hz, 3H), 7.12 (d, J = 16.2 Hz, 1H), 6.97 (dt,

137 Littke, A.F.; Fu, G.C. J. Org. Chem. 1999, 64, 10.

138 Sore, H.F.; Boehner, C.M.; MacDonald, S.J.F.; Norton, D.; Fox, D.J.; Spring, D.R. Org. Biomol. Chem. 2009, 7,

1068.

Page 86: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

75

J = 7.1, 2.1 Hz, 2H), 6.85 (d, J = 16.2 Hz, 1H), 6.69 (tt, J = 8.8, 2.3 Hz, 1H). 13

C NMR (100

MHz, CDCl3) δ 163.28 (dd, J(C-F) = 247.5, 13.2 Hz), 140.85 (t, J(C-F) = 9.6 Hz), 139.15,

126.53, 126.43 (t, J(C-F) = 3.0 Hz), 125.32, 124.78, 123.76, 108.87 (d, J(C-F) = 6.9 Hz), 108.68

(d, J(C-F) = 6.9 Hz), 102.52 (t, J(C-F) = 25.7 Hz). 19

F NMR (377 MHz, CDCl3) δ -110.31 (t, J

= 8.1 Hz). IR (cm-1

, thin film) 1117, 1247, 1337, 1414, 1449, 1589, 1616, 2366, 3050, 3097.

HRMS (DART+) Exact mass calc’d for C12H9F2S [M + H]+: 223.03930, found: 223.03855.

Only spectra for title compound synthesized by general procedure B are shown below.

(E)-3-(4-(trifluoromethyl)styryl)thiophene (4.4) – 1) Synthesized

according to general procedure A from 3-ethynylthiophene and 4-

bromobenzotrifluoride on a 1.0 mmol scale. Hydroboration run at room

temperature. Purification by silica gel flash column chromatography using hexanes as the eluent

gave the title compound as a snow-white solid (126 mg, 50%). The spectroscopic data are in

agreement with those reported below.

2) Synthesized according to general procedure C from 3-ethynylthiophene and 4-

chlorobenzotrifluoride on a 1.0 mmol scale. Cross-coupling reaction employed 1.0 mol%

XPhos-Pd-G2 and was run at 80 °C for 2 hours. Purification by silica gel flash column

chromatography using hexanes as the eluent gave the title compound as a snow-white solid (201

mg, 79%, MP = 136 – 140 °C). 1H NMR (400 MHz, CDCl3) δ 7.60 (d, J = 8.4 Hz, 2H), 7.56 (d,

J = 8.4 Hz, 2H), 7.39 – 7.31 (m, 3H), 7.21 (d, J = 16.3 Hz, 1H), 6.96 (d, J = 16.3 Hz, 1H). 13

C

NMR (100 MHz, CDCl3) δ 140.88 (q, J(C-F) = 1.5 Hz), 139.49, 129.07 (q, J(C-F) = 32.4 Hz),

127.03, 126.48, 126.30, 125.59 (q, J(C-F) = 3.8 Hz), 125.24, 124.81, 124.22 (q, J(C-F) = 271.5

Hz), 123.56. 19

F NMR (377 MHz, CDCl3) δ -62.41. IR (cm-1

, thin film) 1073, 1109, 1129,

1164, 1337, 1413, 2360, 3097. HRMS (DART+) Exact mass calc’d for C13H10F3S [M + H]+:

255.04553, found: 255.04458. Only spectra for title compound synthesized by general procedure

B are shown below.

(E)-3-(4-methylstyryl)pyridine (4.5) – Synthesized according to general

procedure A from 3-ethynylpyridine and 4-bromotoluene on a 1.0 mmol

scale. Hydroboration run at room temperature. Purification by silica gel

flash column chromatography using hexanes:EtOAc (4:1 v/v then 3:7 v/v once product began

eluting) as the eluent was attempted, but the product co-eluted with a trace impurity.

Page 87: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

76

Recrystallization from chloroform:hexanes furnished the title compound as an off-white solid

(76 mg, 39%, MP = 130 – 132 °C), which fluroesces under UV light (365 nm). 1H NMR (400

MHz, CDCl3) δ 8.70 (d, J = 2.3 Hz, 1H), 8.47 (dd, J = 4.8, 1.6 Hz, 1H), 7.79 (dt, J = 8.0, 2.0 Hz,

1H), 7.46 – 7.38 (m, 2H), 7.29 – 7.22 (m, 1H), 7.18 (d, J = 7.9 Hz, 2H), 7.13 (d, J = 16.4 Hz,

1H), 7.00 (d, J = 16.4 Hz, 1H), 2.36 (s, 3H). 13

C NMR (100 MHz, CDCl3) δ 148.38, 148.25,

138.11, 133.79, 133.07, 132.40, 130.64, 129.40, 126.50, 123.75, 123.40, 21.20. The

spectroscopic data are in agreement with literature values.139

(E)-1-chloro-4-(2-(cyclohex-1-en-1-yl)vinyl)benzene (4.6) – 1)

Synthesized according to general procedure A from 1-ethynylcyclohexene

and 1-bromo-4-chlorobenzene on a 1.0 mmol scale. Hydroboration run at

room temperature. Purification by silica gel flash column chromatography using hexanes as the

eluent gave the title compound as a white solid (136 mg, 62%). The spectroscopic data are in

agreement with those reported below.

2) Synthesized according to general procedure B from 1-ethynylcyclohexene and 1-bromo-4-

chlorobenzene on a 1.0 mmol scale. Cross-coupling reaction employed 1.0 mol% Pd(PPh3)4 and

was run at 80 °C for 8 hours. Purification by silica gel flash column chromatography using

hexanes as the eluent gave the title compound as a white solid (147 mg, 67%, MP = 78 – 82 °C).

1H NMR (400 MHz, CDCl3) δ 7.34 – 7.29 (m, 2H), 7.29 – 7.23 (m, 2H), 6.73 (d, J = 16.2 Hz,

1H), 6.38 (d, J = 16.2 Hz, 1H), 5.91 (t, J = 4.3 Hz, 1H), 2.32 – 2.13 (m, 4H), 1.79 – 1.59 (m,

4H). 13

C NMR (100 MHz, CDCl3) δ 136.55, 135.66, 133.19, 132.25, 131.49, 128.64, 127.26,

123.31, 26.16, 24.50, 22.49, 22.45. IR (cm-1

, thin film) 1013, 1091, 1171, 1268, 1403, 1490,

1694, 2861, 2935, 3029. HRMS (DART+) Exact mass calc’d for C14H16Cl [M + H]+:

219.09405, found: 219.09381. Only spectra for title compound synthesized by general procedure

B are shown below.

(E)-1-chloro-2-(2-(cyclohex-1-en-1-yl)vinyl)benzene (4.7) – Synthesized

according to general procedure B from 1-ethynylcyclohexene and 1-bromo-2-

chlorobenzene on a 1.0 mmol scale. Cross-coupling reaction employed 2.5

mol% Pd(PPh3)4 and was run at 100 °C for 12 hours. Purification by silica gel flash column

139 Kantam, M.L.; Reddy, P.V.; Srinivas, P.; Bhargava, S. Tetrahedron Lett. 2011, 52, 4490.

Page 88: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

77

chromatography using hexanes as the eluent gave the title compound as a pale yellow oil (126

mg, 58%). 1H NMR (400 MHz, CDCl3) δ 7.58 (dd, J = 7.8, 1.7 Hz, 1H), 7.35 (dd, J = 7.9, 1.4

Hz, 1H), 7.21 (td, J = 7.6, 1.4 Hz, 1H), 7.13 (td, J = 7.6, 1.7 Hz, 1H), 6.84 (d, J = 16.1 Hz, 1H),

6.75 (d, J = 16.1 Hz, 1H), 5.96 (tt, J = 4.0, 1.6 Hz, 1H), 2.33 (tq, J = 6.2, 2.1 Hz, 2H), 2.22 (tq, J

= 5.4, 2.5 Hz, 2H), 1.81 – 1.62 (m, 4H). 13

C NMR (100 MHz, CDCl3) δ 136.06, 136.03, 135.05,

133.02, 132.00, 129.63, 127.68, 126.71, 126.09, 120.68, 26.19, 24.54, 22.48, 22.43. IR (cm-1

,

thin film) 1035, 1052, 1124, 1210, 1256, 1350, 1440, 1470, 1588, 1614, 1630, 2366, 2829, 2859,

2928, 3058. HRMS (DART+) Exact mass calc’d for C14H16Cl [M + H]+: 219.09405, found:

219.09374.

(E)-5-(2-(cyclohex-1-en-1-yl)vinyl)benzo[d][1,3]dioxole (4.8) –

Synthesized according to general procedure C from 1-ethynylcyclohexene

and 5-chloro-1,3-benzodioxole on a 1.0 mmol scale. Cross-coupling

reaction employed 1.0 mol% XPhos-Pd-G2 and was run at 80 °C for 2 hours. Purification by

silica gel flash column chromatography using hexanes:EtOAc (49:1 v/v) as the eluent gave the

title compound as a yellow tacky solid (162 mg, 71%, MP = 50 – 54 °C). 1H NMR (400 MHz,

CDCl3) δ 6.96 (d, J = 1.7 Hz, 1H), 6.83 (dd, J = 8.1, 1.7 Hz, 1H), 6.75 (d, J = 8.0 Hz, 1H), 6.62

(d, J = 16.1 Hz, 1H), 6.37 (d, J = 16.1 Hz, 1H), 5.94 (s, 2H), 5.86 (t, J = 4.2 Hz, 1H), 2.30 – 2.14

(m, 4H), 1.78 – 1.59 (m, 4H). 13

C NMR (100 MHz, CDCl3) δ 147.98, 146.62, 135.71, 132.58,

131.04, 130.09, 124.26, 120.73, 108.26, 105.29, 100.91, 26.09, 24.55, 22.55, 22.51. IR (cm-1

,

thin film) 1040, 1195, 1233, 1251, 1354, 1447, 1488, 1504, 1690, 1722, 2836, 2860, 2929, 3027.

HRMS (DART+) Exact mass calc’d for C15H17O2 [M + H]+: 229.12285, found: 229.12372.

(E)-5-(hex-1-en-1-yl)-1-tosyl-1H-indole (4.9) – Synthesized according to

general procedure A from 1-hexyne and 5-bromo-1-tosyl-1H-indole on a

1.0 mmol scale. Hydroboration run at room temperature. Purification by silica gel flash column

chromatography using hexanes:EtOAc (4:1 v/v) as the eluent was attempted, but separation

proved difficult and title compound co-eluted with unreacted 5-bromo-1-tosyl-1H-indole.

Calculated yield based on isolated product (44 mg) and mixture of product and bromide was 231

mg, 65%. An additional two successive purifications by flash chromatography employing same

eluent gave the title compound as a viscous, clear, colourless oil (169 mg, 48%) along with a

mixture of product and bromide. 1H NMR (400 MHz, CDCl3) δ 7.91 (d, J = 8.6 Hz, 1H), 7.78 –

7.71 (m, 2H), 7.52 (d, J = 3.7 Hz, 1H), 7.45 (d, J = 1.6 Hz, 1H), 7.34 (dd, J = 8.7, 1.7 Hz, 1H),

Page 89: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

78

7.19 (d, J = 8.2 Hz, 2H), 6.61 (dd, J = 3.7, 0.8 Hz, 1H), 6.43 (dt, J = 15.7, 1.5 Hz, 1H), 6.20 (dt,

J = 15.7, 6.9 Hz, 1H), 2.31 (s, 3H), 2.22 (qd, J = 7.1, 1.4 Hz, 2H), 1.51 – 1.32 (m, 4H), 0.93 (t, J

= 7.2 Hz, 3H). 13

C NMR (100 MHz, CDCl3) δ 144.81, 135.22, 133.88, 133.52, 131.11, 130.61,

129.76, 129.47, 126.69, 126.66, 122.69, 118.55, 113.48, 109.21, 32.68, 31.55, 22.21, 21.46,

13.91. IR (cm-1

, thin film) 1093, 1127, 1145, 1174, 1188, 1223, 1274, 1287, 1306, 1372, 1439,

1456, 1597, 2857, 2871, 2927, 2956. HRMS (DART+) Exact mass calc’d for C21H24NO2S [M +

H]+: 354.15277, found: 354.15305.

(E)-2-(hex-1-en-1-yl)quinoline (4.10) – Synthesized according to general

procedure C from 1-hexyne and 2-chloroquinoline on a 1.0 mmol scale.

Cross-coupling reaction employed 2.5 mol% XPhos-Pd-G2 and was run at 80 °C for 10 hours.

Purification by silica gel flash column chromatography using hexanes:EtOAc (49:1 v/v) as the

eluent gave the title compound as a pale orange oil (155 mg, 73%). 1H NMR (400 MHz, CDCl3)

δ 8.03 (dd, J = 8.5, 1.1 Hz, 1H), 8.00 (d, J = 8.6 Hz, 1H), 7.70 (dd, J = 8.0, 1.4 Hz, 1H), 7.64

(ddd, J = 8.4, 6.8, 1.5 Hz, 1H), 7.47 (dd, J = 8.5, 0.9 Hz, 1H), 7.42 (ddt, J = 7.9, 6.9, 0.9 Hz, 1H),

6.81 (dt, J = 15.9, 6.6 Hz, 1H), 6.70 (dt, J = 15.9, 1.2 Hz, 1H), 2.43 – 2.22 (m, 2H), 1.59 – 1.47

(m, 2H), 1.47 – 1.34 (m, 2H), 0.93 (t, J = 7.2 Hz, 3H). 13

C NMR (100 MHz, CDCl3) δ 156.38,

147.99, 137.81, 135.95, 130.95, 129.34, 129.02, 127.28, 127.01, 125.65, 118.57, 32.61, 30.93,

22.22, 13.84. The spectroscopic data are in agreement with literature values.140

(E)-2-(4-phenylbut-1-en-1-yl)quinoline (4.11) – Synthesized

according to general procedure C from but-3-yn-1-ylbenzene and 2-

chloroquinoline on a 1.0 mmol scale. Cross-coupling reaction

employed 2.5 mol% XPhos-Pd-G2 and was run at 80 °C for 10 hours. Purification by silica gel

flash column chromatography using hexanes:EtOAc (9:1 v/v) as the eluent gave the title

compound as a bright orange oil (201 mg, 77%). 1H NMR (400 MHz, CDCl3) δ 8.03 (dd, J =

8.4, 1.2 Hz, 1H), 7.99 (d, J = 8.6 Hz, 1H), 7.74 – 7.67 (m, 1H), 7.64 (ddd, J = 8.4, 6.9, 1.5 Hz,

1H), 7.44 (d, J = 8.6 Hz, 1H), 7.44 – 7.40 (m, 1H), 7.33 – 7.25 (m, 2H), 7.25 – 7.15 (m, 3H),

6.84 (dt, J = 15.9, 6.4 Hz, 1H), 6.75 (d, J = 15.9 Hz, 1H), 2.85 (dd, J = 9.1, 6.6 Hz, 2H), 2.70 –

2.56 (m, 2H). 13

C NMR (100 MHz, CDCl3) δ 156.15, 147.98, 141.38, 136.54, 136.04, 131.51,

129.42, 129.06, 128.32 (2 signals overlapping), 127.32, 127.06, 125.89, 125.78, 118.61, 35.23,

140 Fakhfakh, M.A.; Franck, X.; Fournet, A.; Hocquemiller, R.; Figadère, B. Synth. Commun. 2002, 32, 2863.

Page 90: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

79

34.71. IR (cm-1

, thin film) 1118, 1314, 1427, 1453, 1497, 1503, 1556, 1597, 1615, 1653, 2854,

2936, 3026, 3059. HRMS (DART+) Exact mass calc’d for C19H18N [M + H]+: 260.14392,

found: 260.14382.

(E)-N,N-bis-Boc-3-(1-tosyl-1H-indol-5-yl) prop-2-en-1-amine (4.12)

– Synthesized according to general procedure B from N,N-bis-Boc-

propargyl amine141

and 5-bromo-1-tosyl-1H-indole on a 1.0 mmol scale. Cross-coupling

reaction employed 2.5 mol% Pd(PPh3)4 and was run at 100 °C for 10 hours. Purification by

silica gel flash column chromatography using hexanes:EtOAc (4:1) as the eluent gave the title

compound as an off-white tacky solid (309 mg, 59%, MP = 46 – 50 °C). 1H NMR (400 MHz,

CDCl3) δ 7.91 (d, J = 8.6 Hz, 1H), 7.78 – 7.71 (m, 2H), 7.53 (d, J = 3.6 Hz, 1H), 7.50 – 7.43 (m,

1H), 7.34 (dd, J = 8.6, 1.7 Hz, 1H), 7.21 (d, J = 8.1 Hz, 2H), 6.61 (d, J = 3.6 Hz, 1H), 6.57 (d, J

= 15.8 Hz, 1H), 6.19 (dt, J = 15.8, 6.3 Hz, 1H), 4.32 (dd, J = 6.3, 1.4 Hz, 2H), 2.33 (s, 3H), 1.51

(s, 18H). 13

C NMR (100 MHz, CDCl3) δ 152.37, 144.93, 135.23, 134.31, 132.35, 132.26,

131.07, 128.84, 126.83, 126.75, 124.50, 122.99, 119.31, 113.55, 109.11, 82.36, 48.17, 28.10,

21.52. IR (cm-1

, thin film) 1122, 1145, 1174, 1226, 1369, 1457, 1597, 1695, 1745, 2360, 2933,

2980. HRMS (ESI+) Exact mass calc’d for C28H34N2O6NaS [M + Na]+: 549.2029, found:

549.2027.

(E)-1-(3-(4-(trifluoromethyl)phenyl)allyl)-1H-indole (4.13) – 1)

Synthesized according to general procedure A from 1-(prop-2-yn-1-

yl)-1H-indole and 4-bromobenzotrifluoride on a 1.0 mmol scale.

Hydroboration run at 60 °C. Purification by silica gel flash column chromatography using

hexanes:EtOAc (9:1 v/v then 4:1 v/v once product began eluting) as the eluent gave the title

compound as a yellow solid (190 mg, 63%). The spectroscopic data are in agreement with those

reported below.

2) Synthesized according to general procedure C from 1-(prop-2-yn-1-yl)-1H-indole and 4’-

chloroacetophenone on a 1.0 mmol scale. Cross-coupling reaction employed 1.0 mol% XPhos-

Pd-G2 and was run at 80 °C for 2 hours. Purification by silica gel flash column chromatography

using hexanes:EtOAc (19:1 v/v) as the eluent gave the title compound as a yellow solid (260 mg,

141 Ko, E.; Burgess, K. Org. Lett. 2011, 13, 980.

Page 91: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

80

86%, MP = 110 – 112 °C). 1H NMR (400 MHz, CDCl3) δ 7.68 (dt, J = 7.9, 1.0 Hz, 1H), 7.53

(d, J = 8.1 Hz, 2H), 7.42 – 7.33 (m, 3H), 7.23 (td, J = 8.0, 7.5, 1.2 Hz, 1H), 7.19 – 7.11 (m, 2H),

6.57 (dd, J = 3.1, 0.9 Hz, 1H), 6.48 – 6.37 (m, 2H, olefinic), 4.90 (dd, J = 2.6, 1.1 Hz, 2H). 13

C

NMR (100 MHz, CDCl3) δ 139.66 (q, J(C-F) = 1.4 Hz), 136.08, 130.65, 129.60 (q, J(C-F) =

32.5 Hz), 128.75, 127.74 (2 signals overlapping), 126.61, 125.50 (q, J(C-F) = 3.8 Hz), 124.10 (q,

J(C-F) = 272.2 Hz), 121.74, 121.06, 119.60, 109.47, 101.80, 48.11 19

F NMR (377 MHz,

CDCl3) δ -62.49. IR (cm-1

, thin film) 1067, 1119, 1165, 1326, 1463, 1614, 2912, 3053. HRMS

(DART+) Exact mass calc’d for C18H15F3N [M + H]+: 302.11566, found: 302.11544. Only

spectra for title compound synthesized by general procedure C are shown below.

(E)-1-(4-(3-(1H-indol-1-yl)prop-1-en-1-yl)phenyl)ethanone (4.14)

– 1) Synthesized according to general procedure A from 1-(prop-2-

yn-1-yl)-1H-indole and 4’-bromoacetophenone on a 1.0 mmol scale.

Hydroboration run at 60 °C. Purification by silica gel flash column chromatography using

hexanes:EtOAc (4:1 v/v) as the eluent gave the title compound as a yellow solid (200 mg, 73%).

The spectroscopic data are in agreement with those reported below.

2) Synthesized according to general procedure B from 1-(prop-2-yn-1-yl)-1H-indole and 4’-

bromoacetophenone on a 1.0 mmol scale. Cross-coupling reaction employed 1.0 mol%

Pd(PPh3)4 and was run at 80 °C for 8 hours. Purification by silica gel flash column

chromatography using hexanes:EtOAc (4:1 v/v) as the eluent gave the title compound as a

yellow solid (244 mg, 89%, MP = 99 – 100 °C). 1H NMR (400 MHz, CDCl3) δ 7.93 – 7.85 (m,

2H), 7.72 – 7.65 (m, 1H), 7.43 – 7.35 (m, 3H), 7.23 (ddd, J = 8.3, 7.1, 1.2 Hz, 1H), 7.19 – 7.11

(m, 2H), 6.58 (dd, J = 3.1, 0.9 Hz, 1H), 6.54 – 6.41 (m, 2H, olefinic), 4.93 (d, J = 3.9 Hz, 2H),

2.58 (s, 3H). 13

C NMR (100 MHz, CDCl3) δ 197.39, 140.78, 136.22, 136.06, 130.96, 128.70,

128.69, 127.96, 127.74, 126.51, 121.71, 121.03, 119.56, 109.46, 101.77, 48.18, 26.53. IR (cm-1

,

thin film) 1182, 1267, 1314, 1357, 1410, 1463, 1484, 1511, 1602, 1678, 2918, 3052. HRMS

(DART+) Exact mass calc’d for C19H18NO [M + H]+: 276.13884, found: 276.13911. Only

spectra for title compound synthesized by general procedure B are shown below.

(E)-5-(3-(1H-indol-1-yl)prop-1-en-1-yl)-1-tosyl-1H-indole (4.15)

– 1) Synthesized according to general procedure A from 1-(prop-2-

yn-1-yl)-1H-indole and 5-bromo-1-tosyl-1H-indole on a 1.0 mmol

Page 92: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

81

scale. Hydroboration run at 60 °C. Purification by silica gel flash column chromatography using

hexanes:EtOAc (4:1 v/v) as the eluent gave the title compound as an off-white solid (233 mg,

55%). The spectroscopic data are in agreement with those reported below.

2) Synthesized according to general procedure B from 1-(prop-2-yn-1-yl)-1H-indole and 5-

bromo-1-tosyl-1H-indole on a 1.0 mmol scale. Cross-coupling reaction employed 1.0 mol%

Pd(PPh3)4 and was run at 80 °C for 10 hours. Purification by silica gel flash column

chromatography using hexanes:EtOAc (9:1 v/v) as the eluent gave the title compound as an off-

white solid (331 mg, 78%, MP = 163 – 166 °C). 1H NMR (400 MHz, CDCl3) δ 7.92 (d, J = 8.6

Hz, 1H), 7.80 – 7.71 (m, 2H), 7.71 – 7.65 (m, 1H), 7.54 (d, J = 3.7 Hz, 1H), 7.44 (d, J = 1.6 Hz,

1H), 7.40 (dd, J = 8.2, 1.1 Hz, 1H), 7.32 (dd, J = 8.7, 1.7 Hz, 1H), 7.25 – 7.11 (m, 5H), 6.58

(ddd, J = 11.5, 3.4, 0.9 Hz, 2H), 6.44 (dt, J = 15.8, 1.6 Hz, 1H), 6.23 (dt, J = 15.8, 5.8 Hz, 1H),

4.80 (dd, J = 5.8, 1.6 Hz, 2H) 2.32 (s, 3H). 13

C NMR (100 MHz, CDCl3) δ 144.94, 136.06,

135.10, 134.39, 132.17, 131.73, 131.05, 129.81, 128.71, 127.69, 126.92, 126.68, 124.23, 122.97,

121.53, 120.92, 119.49, 119.40, 113.57, 109.55, 109.10, 101.46, 48.36, 21.47. IR (cm-1

, thin

film) 1127, 1145, 1174, 1188, 1215, 1276, 1286, 1307, 1316, 1336, 1370, 1398, 1439, 1455,

1462, 1511, 1596, 2921, 3052, 3141. HRMS (DART+) Exact mass calc’d for C26H23N2O2S [M

+ H]+: 427.14802, found: 427.14757. Only spectra for title compound synthesized by general

procedure B are shown below.

NOTE: Compound appears to be light sensitive. Upon sitting on bench-top for several days,

compound became salmon-coloured. Minor decomposition observed by TLC analysis.

Filtration through silica gel with chloroform as the eluent recovered pure compound.

(E)-1-(3-(anthracen-9-yl)allyl)-1H-indole (4.16) – Synthesized

according to general procedure C from 1-(prop-2-yn-1-yl)-1H-indole

and 9-chloroanthracene on a 1.0 mmol scale. Cross-coupling reaction

employed 1.0 mol% XPhos-Pd-G2 and was run at 80 °C for 4 hours.

Purification by silica gel flash column chromatography using hexanes:EtOAc (19:1 v/v) as the

eluent gave the title compound as a yellow solid (236 mg, 71%, MP = 134 – 136 °C), which

fluoresces under UV light (365 nm). 1H NMR (400 MHz, CDCl3) δ 8.37 (s, 1H), 8.21 – 8.10

(m, 2H), 8.04 – 7.94 (m, 2H), 7.76 (dt, J = 7.9, 1.0 Hz, 1H), 7.61 (dd, J = 8.1, 1.1 Hz, 1H), 7.53 –

7.40 (m, 4H), 7.40 – 7.32 (m, 2H), 7.23 (ddd, J = 8.0, 7.1, 1.0 Hz, 1H), 7.10 (dq, J = 16.1, 1.5

Page 93: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

82

Hz, 1H), 6.67 (dd, J = 3.1, 0.9 Hz, 1H), 6.22 (dt, J = 16.1, 5.3 Hz, 1H), 5.15 (dd, J = 5.4, 1.8 Hz,

2H). 13

C NMR (100 MHz, CDCl3) δ 136.26, 133.23, 131.55, 131.28, 129.33, 128.84, 128.60,

128.29, 127.88, 126.50, 125.61, 125.47, 125.05, 121.75, 121.12, 119.60, 109.68, 101.85, 48.55.

IR (cm-1

, thin film) 1316, 1335, 1347, 1442, 1463, 1511, 1663, 3027, 3050. HRMS (DART+)

Exact mass calc’d for C25H20N [M + H]+: 334.15957, found: 334.15944.

(E)-1-(4-(3-(1H-benzo[d][1,2,3]triazol-1-yl) prop-1-en-1-

yl)phenyl)ethanone (4.17) – Synthesized by adaptation of general

procedure C from 1-(prop-2-yn-1-yl)-1H-benzo[d][1,2,3]triazole and

4’-chloroacetophenone on a 1.0 mmol scale. Modifications: hydroboration carried out in dry

toluene (0.5 mL), added directly before pinacolborane; 1.5 mL toluene added in coupling step.

Cross-coupling reaction employed 2.5 mol% XPhos-Pd-G2 and was run at 100 °C for 24 hours.

Purification by silica gel flash column chromatography using hexanes:EtOAc:DCM (5:4:1 v/v/v)

as the eluent gave the title compound as a white solid solid (115 mg, 41%, MP = 151 – 153 °C).

1H NMR (400 MHz, CDCl3) δ 8.09 (d, J = 8.4 Hz, 1H), 7.89 (d, J = 8.4 Hz, 2H), 7.55 (d, J = 8.3

Hz, 1H), 7.48 (ddd, J = 8.2, 6.9, 1.0 Hz, 1H), 7.42 (d, J = 8.4 Hz, 2H), 7.40 – 7.36 (m, 1H), 6.65

(dt, J = 15.9, 1.4 Hz, 1H), 6.53 (dt, J = 15.9, 6.0 Hz, 1H), 5.47 (dd, J = 6.0, 1.4 Hz, 2H), 2.57 (s,

3H). 13

C NMR (100 MHz, CDCl3) δ 197.31, 146.22, 140.04, 136.62, 133.07, 132.86, 128.74,

127.50, 126.70, 125.09, 124.01, 120.15, 109.44, 50.16, 26.55. IR (cm-1

, thin film) 1166, 1228,

1268, 1359, 1603, 1674, 2925, 3042. HRMS (ESI+) Exact mass calc’d for C17H16N3O [M +

H]+: 278.1287, found: 278.1280.

(E)-2-styrylbenzo[d]oxazole (4.18) – Reaction was carried out on a 1.0

mmol scale. Phenylacetylene (1.1 equiv) and Cp2ZrHCl (5.5 mol%) were

weighed into an oven-dried 2 dram vial and purged with nitrogen for one

minute. To this, pinacolborane (1.21 equiv) was added dropwise. The reaction vial was sealed

under argon with a Teflon cap and stirred at 45 °C. After 16 hours, the reaction was allowed to

cool to room temperature. To the crude mixture containing the alkenylboronate intermediate

were added 2-chlorobenzoxazole (1.0 equiv) in anhydrous 2-butanol (2.0 mL), K3PO4 (2.0 equiv)

and XPhos-Pd-G2 (2.5 mol%) in succession, purging the reaction vial with nitrogen following

each addition. The reaction vial was sealed under argon with a Teflon cap and stirred vigorously

at 100 °C. After 4 hours, the reaction was allowed to cool to room temperature, filtered through

a thin pad of silica gel with EtOAc (20 mL) as the eluent, and concentrated in vacuo.

Page 94: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

83

Purification by silica gel flash column chromatography using hexanes:EtOAc (9:1 v/v) as the

eluent gave the title compound as a yellow solid (109 mg, 49%, MP = 105 – 107 °C (lit. 79 –

80)). 1H NMR (400 MHz, CDCl3) δ 7.80 (d, J = 16.4 Hz, 1H), 7.76 – 7.69 (m, 1H), 7.65 – 7.57

(m, 2H), 7.53 (dt, J = 5.6, 3.4 Hz, 1H), 7.47 – 7.29 (m, 5H), 7.09 (d, J = 16.4 Hz, 1H). 13

C

NMR (100 MHz, CDCl3) δ 162.76, 150.39, 142.18, 139.42, 135.13, 129.74, 128.94, 127.53,

125.17, 124.47, 119.85, 113.94, 110.29. The spectroscopic data are in agreement with literature

values.142

142 Evindar, G.; Batey, R.A. J. Org. Chem. 2006, 71, 1802.

Page 95: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

84

Appendix: Selected Spectra

(Z)-4,4,5,5-tetramethyl-2-(1-phenylprop-1-en-2-yl)-1,3,2-dioxaborolane (3.4)

Page 96: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

85

(E)-4,4,5,5-tetramethyl-2-styryl-1,3,2-dioxaborolane (3.1)

Page 97: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

86

Page 98: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

87

(E)-1-(3-(4,4,5,5-tetramethyl-1,3,2-dioxaborolan-2-yl)allyl)-1H-indole (3.2)

Page 99: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

88

Page 100: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

89

(E)-1-(4-styrylphenyl)ethanone (4.1)

Page 101: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

90

(E)-1-(4-(4-methylstyryl)phenyl)ethanone (4.2)

Page 102: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

91

(E)-3-(3,5-difluorostyryl)thiophene (4.3)

Page 103: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

92

Page 104: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

93

(E)-3-(4-(trifluoromethyl)styryl)thiophene (4.4)

Page 105: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

94

Page 106: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

95

(E)-1-chloro-4-(2-(cyclohex-1-en-1-yl)vinyl)benzene (4.6)

Page 107: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

96

(E)-1-chloro-2-(2-(cyclohex-1-en-1-yl)vinyl)benzene (4.7)

Page 108: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

97

(E)-5-(2-(cyclohex-1-en-1-yl)vinyl)benzo[d][1,3]dioxole (4.8)

Page 109: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

98

(E)-5-(hex-1-en-1-yl)-1-tosyl-1H-indole (4.9)

Page 110: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

99

(E)-2-(hex-1-en-1-yl)quinoline (4.10)

Page 111: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

100

(E)-2-(4-phenylbut-1-en-1-yl)quinoline (4.11)

Page 112: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

101

(E)-N,N-bis-Boc-3-(1-tosyl-1H-indol-5-yl)prop-2-en-1-amine (4.12)

Page 113: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

102

(E)-1-(3-(4-(trifluoromethyl)phenyl)allyl)-1H-indole (4.13)

Page 114: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

103

Page 115: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

104

(E)-1-(4-(3-(1H-indol-1-yl)prop-1-en-1-yl)phenyl)ethanone (4.14)

Page 116: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

105

(E)-5-(3-(1H-indol-1-yl)prop-1-en-1-yl)-1-tosyl-1H-indole (4.15)

Page 117: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

106

(E)-1-(3-(anthracen-9-yl)allyl)-1H-indole (4.16)

Page 118: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

107

(E)-1-(4-(3-(1H-benzo[d][1,2,3]triazol-1-yl) prop-1-en-1-yl)phenyl)ethanone (4.17)

Page 119: Zirconium and Palladium Catalyzed Telescopic Synthesis of ... · Zirconium and Palladium Catalyzed Telescopic Synthesis of (E)-Alkenes Jordan Anthony Evans Master of Science Department

108

(E)-2-styrylbenzo[d]oxazole (4.18)