<<

Digital Comprehensive Summaries of Uppsala Dissertations from the Faculty of Pharmacy 221

Palladium(II)-Catalysed Heck and Addition Reactions

Exploring Decarboxylative and Desulfitative Processes

BOBO SKILLINGHAUG

ACTA UNIVERSITATIS UPSALIENSIS ISSN 1651-6192 ISBN 978-91-554-9717-0 UPPSALA urn:nbn:se:uu:diva-304746 2016 Dissertation presented at Uppsala University to be publicly examined in B21, BMC, Husargatan 3, Uppsala, Friday, 25 November 2016 at 09:15 for the degree of Doctor of Philosophy (Faculty of Pharmacy). The examination will be conducted in English. Faculty examiner: Professor Antonio de la Hoz Ayuso (Universidad de Castilla-La Mancha).

Abstract Skillinghaug, B. 2016. (II)-Catalysed Heck and Addition Reactions. Exploring Decarboxylative and Desulfitative Processes. Digital Comprehensive Summaries of Uppsala Dissertations from the Faculty of Pharmacy 221. 100 pp. Uppsala: Acta Universitatis Upsaliensis. ISBN 978-91-554-9717-0.

Palladium complexes have the ability to catalyse cross-coupling of two organic moieties through the formation of transient metal- bonds, thus bringing them closer to each other to facilitate the formation of a new bond. Palladium-catalysed coupling reactions are one of the most important carbon-carbon forming reactions available to organic chemists and many of these reactions rely on the reactivity of aryl-palladium complexes. The investigation of new aryl-palladium precursors is thus of great interest, especially as more sustainable and economic methods can be developed. This thesis describes the use of carboxylic acids and sodium arylsulfinates as such new arylating agents. Protocols for microwave-assisted palladium(II)-catalysed decarboxylative synthesis of electron-rich and 1,1-diarylethenes were developed. However, these transformations had very limited substrate scopes which prompted the investigation of sodium arylsulfinates as alternative arylating agents. These substrates were employed in the microwave- assisted palladium(II)-catalysed desulfitative addition to nitriles, and the substrate scope was demonstrated by combining a wide array of sodium arylsulfinates and nitriles to yield the corresponding aryl ketones. The application of the desulfitative reaction in a continuous flow setup was demonstrated, and aluminium oxide was identified as safe alternative to borosilicate glass as a reactor material. The mechanisms of the decarboxylative and desulfitative transformations were investigated by density functional theory (DFT) calculations. The desulfitative reaction was also investigated by direct electrospray ionization mass spectrometry (ESI-MS), providing further mechanistic insight. Finally, a protocol for the safe and convenient synthesis of a wide range of sodium arylsulfinates was developed.

Keywords: Palladium, , palladium(II) catalysis, synthesis, Heck, carboxylic acid, sulfinic acid, sodium sulfinate, nitrile, , ketone, aryl ketone, electrospray ionization mass spectrometry, density functional theory, microwave heating, continuous flow

Bobo Skillinghaug, Department of Medicinal Chemistry, Organic Pharmaceutical Chemistry, Box 574, Uppsala University, SE-75123 Uppsala, Sweden.

© Bobo Skillinghaug 2016

ISSN 1651-6192 ISBN 978-91-554-9717-0 urn:nbn:se:uu:diva-304746 (http://urn.kb.se/resolve?urn=urn:nbn:se:uu:diva-304746)

May you live in interesting times

List of Papers

This thesis is based on the following papers, which are referred to in the text by their Roman numerals.

I Fardost, A.,† Skillinghaug, B.,† Svensson, F.,† Wakchaure, P., Wejdemar, M., Larhed, M., Sköld. C. (2016) Mechanistic In- vestigation of Palladium(II)-Catalyzed Decarboxylative Synthe- sis of Electron-Rich Styrenes and 1,1-Diarylethenes. Manu- script. II Skillinghaug, B., Sköld., C., Rydfjord, J., Svensson, F., Beh- rends, M., Sävmarker, J., Sjöberg, P. J. R., Larhed, M. (2014) Palladium(II)-Catalyzed Desulfitative Synthesis of Aryl Ke- tones from Sodium Arylsulfinates and Nitriles: Scope, Limita- tions, and Mechanistic Studies. J. Org. Chem., 79, 12018−12032. III Skillinghaug, B., Rydfjord, J., Sävmarker, J., Larhed, M. (2016) Microwave Heated Continuous Flow Palladium(II)- Catalyzed Desulfitative Synthesis of Aryl Ketones. Org. Pro- cess Res. Dev., Under revision. IV Skillinghaug, B., Rydfjord, J., Odell, L. (2016) Synthesis of Sodium Aryl Sulfinates from Aryl Bromides Employing 1,4- Diazabicyclo[2.2.2]octane Bis(sulfur dioxide) Adduct (DAB- SO) as a Bench-stable, Gas-free Alternative to SO2. Tetrahe- dron Lett., 57, 533-536.

†These authors contributed equally.

Reprints were made with permission from the respective publishers.

Author Contribution Statement

The following contributions to the papers included in this thesis were made by the author:

I Performed preparative work and characterisation of compounds, performed single point energy calculations, collated experi- mental data and contributed significantly during the manuscript preparation. II Performed the majority of the preparative work and characteri- sation of compounds, aided in the design of and participated in the electrospray ionisation mass spectrometry (ESI-MS) study, collated experimental data and drafted the manuscript. III Performed the optimisation work and synthesised the majority of the compounds, collated experimental data and contributed significantly during the manuscript preparation. IV Performed the optimization work, the majority of the prepara- tive work and characterisation of compounds, collated experi- mental data and drafted the manuscript.

Contents

Introduction ...... 11 Synthetic Organic Chemistry ...... 11 Catalysis ...... 13 Palladium Catalysis ...... 14 Palladium-Catalysed Cross-Coupling Reactions ...... 16 The Mizoroki- ...... 19 Palladium(II) Catalysis ...... 21 The Oxidative Heck Reaction ...... 22 Palladium-Catalysed Addition Reactions ...... 25 Metal-Catalysed Decarboxylative Reactions ...... 26 Sulfinic Acids and Metal-Catalysed Desulfitative Reactions ...... 28 Density Functional Theory ...... 31 Microwave-Assisted Organic Synthesis ...... 32 Continuous Flow Chemistry ...... 34 Electrospray Ionisation Mass Spectrometry ...... 35 Research Aims ...... 37 Palladium(II)-Catalysed Decarboxylative Synthesis of Electron-Rich Styrenes and 1,1-Diarylethenes (Paper I) ...... 38 Background ...... 38 Optimisation of Reaction Conditions for the Synthesis of Styrenes ...... 39 Investigation of the Scope of the Palladium(II)-Catalysed Decarboxylative Synthesis of Styrenes ...... 44 Investigation of the Reaction Mechanism using Density Functional Theory ...... 45 Palladium(II)-Catalysed Decarboxylative Synthesis of 1,1- Diarylethenes ...... 48 Palladium(II)-Catalysed Desulfitative Addition of Sodium Arylsulfinates to Nitriles (Paper II-III) ...... 51 Background ...... 51 Initial Investigation ...... 51 Optimisation of Reaction Conditions for the Synthesis of Aryl Ketones ...... 53 Investigation of the Reaction Mechanism using Electrospray Ionisation Mass Spectrometry ...... 55

Investigation of the Reaction Mechanism using Density Functional Theory ...... 56 Investigation of the Scope of the Microwave-Assisted Organic Synthesis of Aryl Ketones ...... 60 Continuous Flow Microwave-Assisted Organic Synthesis of Aryl Ketones ...... 65 Synthesis of Sodium Arylsulfinates Using a Solid Source of Sulphur Dioxide (Paper IV) ...... 72 Background ...... 72 Development of the Protocol ...... 73 Investigation of the Scope of the Synthesis of Sodium Arylsulfinates .... 73 Conclusions ...... 77 Acknowledgements ...... 79 References ...... 82

Abbreviations

TFA Trifluoroacetic Acid NMR Nuclear Magnetic Resonance GC-MS Gas Chromatography Mass Spectrometry ESI-MS Electrospray Ionisation Mass Spectrometry DBA Dibenzylideneacetone THF Tetrahydrofuran DMF Dimethyl formamide NMP N-methyl pyrrolidone DMSO Dimethyl sulfoxide DFT Density functional theory Ar Aryl CF Continuous flow CF-MAOS Continuous flow microwave-assisted organic synthesis TS Transition state p-BQ Para-benzoquinone TEMPO 2,2,6,6-tetramethylpiperidine-1-oxyl HPLC High Performance Liquid Chromatography or High Pressure Liquid Chromatography EDG Electron donating group LC Liquid chromatography UV Ultra violet

Introduction

Synthetic Organic Chemistry Organic chemicals were once believed to possess a vital force, which could only be imbued by living organisms. This mainstream theory, called vital- ism, was effectively disproved in 1828 when Wöhler chemically synthesised urea without the need of a living kidney.1 Since then, the field of organic chemistry has flourished, leading to numerous important discoveries of nov- el compounds and materials prepared through abiotic chemical reactions. Today organic chemistry has been re-defined as the science regarding all carbon based molecules (without the influence of vitalism).2 Synthetic organic chemistry is the art of constructing such carbon-based organic molecules from smaller building blocks. These new organic mole- cules are formed through chemical transformations that are often performed in sequence and in a specific order. A schematic synthetic route from simple building blocks a more complex structure is illustrated in Figure 1. Reactions that form new carbon-carbon bonds are of particular interest as tools in syn- thetic organic chemistry and ever since the first carbon–carbon bond for- mation performed by Kolbe in 1845 in the synthesis of , carbon– carbon bond-forming reactions have had a profound role in the evolution of chemical synthesis.3 Years of development have now provided organic chemists with a variety of reactions, enabling the construction of complex structures from relatively simple building blocks. The development of new such tools is often referred to as method development, and involves the ex- ploration of a certain transformation, especially with regards to investigation of the synthetic scope and limitations. A significant part of method devel- opment is the optimisation of the conditions for a transformation with re- gards to the selectivity and yield of the desired product.

Figure 1. Schematic synthesis of complex structures from simple building blocks.

11 Synthetic organic chemistry has come to include parts of , which is the study of compounds that include carbon-metal bonds. In the synthesis of organic molecules these carbon-metal bonds can often be transient, yet responsible for the desired chemical transformations. An important subset of organometallic chemistry is metal-mediated synthetic organic chemistry. Metal-mediated synthetic organic chemistry has become very important in this context through the development of transition metal- mediated cross-coupling reactions. A schematic representation of a metal- mediated cross- is shown in Figure 2.

Figure 2. Schematic representation of a metal-mediated cross-coupling reaction, a process where a metal successively forms transient metal-carbon bonds with two organic moieties, thus bringing them in close proximity to each other to facilitate the formation of a new bond.

The first metal-mediated coupling reactions ever reported were performed by Glaser in 1869 and involved the - and silver-mediated homocoupling of acetylenes.4 Half a century later, Ullmann developed homocoupling of ortho-halo-nitrobenzenes mediated by super-stoichiometric amounts of cop- per.5 In 1855 Wurtz6,7 demonstrated the sodium-mediated homodimerisation of alkyl halides, later extended in 1862 by Fittig8 to include aryl halides. However the high reactivity of sodium limited its synthetic utility, leading to the investigation and development of the less nucleophilic magnesium-based Grignard reagents.9 Early metal-promoted coupling reactions were non- catalytic, poorly selective and limited to homocoupling. Arguably, the Gri- gnard reaction remains one of the most important carbon-carbon forming tools at the chemist’s disposal to make new organic compounds. The reactiv- ity of Grignard reagents stems from the highly polarised magnesium-carbon bond that allows them to react readily with a wide range of electrophiles. During the early days of the last century the use of organolithium reagents also became an essential part of the organic chemist’s toolbox. Undoubtedly, these reagents are of great importance in synthetic organic chemistry and have been exploited in this thesis.

12 The focus of this thesis is on synthetic organic chemistry and method devel- opment in the context of transition metal catalysis, and more specifically palladium(II) catalysis, to expand its applicability and scope. Over the past few decades, three major waves of synthetic development have occurred in the field of transition metal catalysed cross-coupling reactions. The first wave of discovery involved the use of different metals such as copper, silver, nickel and palladium as the catalyst. In the following wave of discoveries various coupling partners were investigated, and lastly, the scope of such reactions was progressively widened.10 Some of these important discoveries will be discussed herein.

Catalysis The term catalysis was introduced by Berzelius in 183511,12 and describes the phenomenon where a chemical entity (catalyst) lowers the energy barrier for a reaction and thus increases the rate of a reaction without being consumed itself. All life on earth is dependent on catalysis as the chemical reactions that take place in biological systems are often transformations catalysed by enzymes. Catalytic processes can either occur at the interface between two phases (heterogeneous) or in one phase, most often a solution (homogenous). A schematic mechanism of a catalytic process is illustrated in Figure 3 and the energy diagram is presented in Figure 4.

Figure 3. Schematic representation of a catalytic process, more specifically a metal- catalysed cross-coupling reaction.

One of the most recognised examples of catalysts are the catalytic converters used in the automobile industry to convert harmful pollutants into less toxic substances. These are heterogeneous catalysts (e.g. solid platinum or a com- bination of palladium and rhodium) and function by lowering the activation energies for the reduction of nitrous oxides, oxidation of carbon monoxide to carbon dioxide and combustion of unburnt hydrocarbons.13 Similarly, the same principle can be used to lower the activation energies for different transformations in synthetic organic chemistry.

13 Figure 4. A non-catalysed and a catalysed process, leading to the same product. The catalyst (metal) lowers the energy barrier for the reaction and thus increases the rate of the reaction without being consumed itself.

The Haber-Bosch process uses a heterogeneous catalyst to enable the reac- tion of atmospheric with hydrogen to produce ammonia.14 This chemical transformation allows artificial fixation of nitrogen and represents the most important source of fertiliser world-wide. The drastic increase in the world population during the last 50 years has in part been accredited to the Haber-Bosch process, due to its impact on agriculture.15,16 In fact, it has been estimated that 80% of the nitrogen in the proteins of an average human originates from this process.17

Palladium Catalysis The focus of this thesis is on homogeneous palladium(II)-catalysed Heck and addition reactions but a brief introduction to certain palladium(0)- catalysed reactions that are of particular relevance to the work described herein will be given. The historical, mechanistic, theoretical, and practical aspects of palladium-catalysed carbon-carbon bond forming reactions have been immensely discussed in the literature, and the reader is referred to the books18–21 and reviews3,10,22,23 covering this well-established field for a com- prehensive overview. Palladium is a late transition metal first discovered in 1803 by Wollaston as an impurity in platinum.24 It took more than 150 years before its useful- ness began to be realised when PdCl2 was exploited in the catalytic trans- formation of ethene to ethanal in a reaction now known as the Wacker pro- cess.25 One of the properties that makes palladium such a useful element in

14 catalysis is that it can relatively easily switch between the oxidation states 0 and +II. Palladium has a relatively high electronegativity (2.2 on the Pauling scale) and forms a relatively non-polar palladium-carbon bond. Also, palla- dium is relatively soft due to shielding of the nuclear charge by the 10 or 8 d-electrons (for palladium(0) or palladium(II), respectively). In addition, the relatively large van der Waals radius promotes formation of bonds with soft π- and σ-donors. Due to these properties, and to the closely lying highest occupied molecular orbital (HOMO) and lowest unoccupied molecular or- bital energies (LUMO), palladium is predisposed to undergo concerted reac- tions and displays lower reactivity than many other organometallic species toward polar functional groups such as ketones. Palladium prefers the for- mation of tetrahedral d10 and square planar d8 complexes.26 Importantly, palladium catalysis has provided chemists with the ability to forge carbon-carbon bonds between or within functionalised and sensitive substrates, leading to new opportunities in many fields of research. In mod- ern days many important industrial processes rely on heterogenous palladi- um-based catalysts, while homogenous catalysts have become exceedingly important in chemical research for the formation of new carbon-carbon bonds, especially in drug discovery.23 Indeed, palladium has a wide array of applications - ranging from the use as a building block in optoelectronics,27 nanoelectronics,28,29 fuel cells,30 solar cells,31 cancer treatment,32 photogra- phy,33 to the use as a catalyst in the preparation of fine chemicals23,34 and pharmaceuticals23,35. Palladium-catalysed cross-couplings have a wide scope with regards to forming carbon-carbon bonds with exceedingly high selectivity, and this has been exploited in the total synthesis of complicated molecules.3 Since the early development of cross-coupling reactions, the use of palla- dium as a catalyst has competed with more cost-effective metals like copper and iron. However, due to the properties described above, palladium cataly- sis retains certain advantages such as enabling the conversion of less reactive substrates, better performance at relatively low temperatures and high cata- lyst turn-over numbers. The reactivity of the palladium catalyst can also be modulated by since the polarity of the carbon-palladium bond can be tuned by coordinating species. The importance of palladium catalysis as a synthetic tool to make new compounds is illustrated by the fact that Richard Heck, and Ei- ichi Negishi were awarded the Nobel prize in 2010 for their contribution to the development of palladium catalysis.10

15 Palladium-Catalysed Cross-Coupling Reactions After being instructed at the Hercules Powder Company to “do something with transition metals”,36 Heck commenced his pioneering work and report- ed seven back-to-back articles on palladium-mediated reactions in 1968,37–43 including one of the first palladium-catalysed arylations of olefins (see Scheme 1).38 This reaction employed toxic arylmercury chlorides as ary- lating agents and used CuCl2 as a catalytic reoxidant together with oxygen as the terminal reoxidant. Fujiwara independently converted the previously stoichiometric arylation of olefins into a catalytic process, using silver ace- tate as the reoxidant.44 This type of palladium(II)-catalysed reaction is often referred to as the oxidative Heck reaction and is discussed in more detail in later chapters.

Scheme 1. Catalytic arylation of methyl acrylate reported by Heck.

The arylation of olefins was further developed independently by Mizoroki45 and Heck46 in 1971-1972 into a protocol that employed aryl iodides instead of the toxic arylmercury chlorides. Furthermore, the use of a to regen- erate the palladium(0) catalyst obviated the need for an oxidant. During the same period Kumada47 and Corriu48 independently discovered that aryl Gri- gnard reagents could be used in the nickel-catalysed cross-coupling of ole- finic halides, a reaction which was later translated into the corresponding palladium-catalysed process.49,50 The cross-coupling with nucleophiles was further developed by Negishi who was able to employ less reactive zinc rea- gents.51 In 1975 Sonogashira and Hagihara reported an elegant protocol for the direct use of via the corresponding cuprates formed in situ in the palladium-catalysed coupling with aryl halides.52 During the following dec- ade, several variations of the palladium-catalysed cross-coupling reaction using different coupling reagents such as stannanes (Stille-Migita),53,54 bo- ronic acids (Suzuki-Miyaura)55,56 and silanes (Hiyama-Hatanaka)57 were developed. An overview of the most important palladium-catalysed cross- coupling reactions is shown in Scheme 2.

16

Scheme 2. Selected examples of palladium-catalysed cross-coupling reactions.

Although the Stille-Migita reaction has a very wide substrate scope and is probably the most powerful cross-coupling tool available, its usefulness is reduced by the need for toxic stannanes that also complicate purification. Conversely, the Suzuki-Miyaura reaction uses relatively environmentally benign boron compounds as substrates and has emerged as one of the most important tools in both synthetic organic chemistry and medicinal chemis- try.58 Though the details of the mechanisms of the palladium-catalysed cross- coupling reactions in Scheme 2 vary in many different aspects, these cou- pling reactions share certain fundamental processes. The steps involved in the palladium cross-coupling reaction are illustrated in the general catalytic cycle in Scheme 3, and involves three major steps: , transmetalation and . Many palladium(0)-catalysed reactions are performed using palladium(II) salts that are reduced in situ to generate a complex where the palladium atom has the oxidation state 0. The palladium centre in these complexes prefers coordination to four ligands that each contribute two electrons to form a stable tetrahedral 18 electron com- plex. In solution, the loss of two neutral ligands (e.g. phosphines) leads to the formation of a catalytically active low coordinate d10 14 electron palla- dium complex to which an electrophile can undergo addition.59 Each catalyt- ic cycle is initiated by the oxidative addition of an organohalide (I, Br, Cl) or pseudohalide (e.g. triflate, diazonium salt, tosylate), through the cleavage of a covalent C-X bond by a palladium(0) species. This generates an organo– palladium species and as implied by the name of the transformation, the oxidation state of palladium is in turn increased to +II. Bulky ligands disso- ciate more readily and thus promote oxidative addition due to the increased

17 concentration of the low coordinated palladium complex.60 It should be not- ed that alkyl halides that contain -hydrogens are often incompatible sub- strates due to rapid -hydrogen elimination and most cross-coupling reac- tions have been limited to aryl or vinyl halides (although examples of alkyl- alkyl cross-coupling reactions do exist).61–65

0 Pd L4 R X 0 Pd L2

Reductive Oxidative Elimination Addition

II II L2RPd L2XPd Trans- metalation

MX M R Scheme 3. General mechanism for palladium-catalysed cross-coupling reactions.

Following oxidative addition, a second organyl group is introduced to the palladium catalyst via transmetallation from an organometallic reagent to give the corresponding diaryl, aryl-alkenyl or aryl-alkynyl palladium(II) complex. The driving force for the transmetallation is the formation of a more stable and generally less polarised carbon-metal bond. Thus, transmetalation occurs more readily for organometals or organometalloids (Mg, Cu, Zn, Sn, B or Si) that contain elements with low electronegativity relative to palladium.66,67 In the last step of the catalytic cycle, the organyl groups must be oriented in a cis fashion and rearrangement occurs through a migration of one of the organyl groups. Through reductive elimination a new -bond is formed be- tween the two organic palladium(II)-coordinating groups and the catalyst is reduced back to the original active palladium(0) complex.68

18 In traditional cross-coupling reactions, a nucleophile (e.g. an organome- tallic reagent) is coupled with an electrophile (e.g. an aryl halide) which acts as an organic oxidant that is included in the product. In oxidative coupling reactions, such as the oxidative Heck reaction, two nucleophiles are coupled and the reaction usually requires a terminal oxidant, which is not included in the final product. There is also an emerging field of reductive cross-coupling reactions, where the coupling between two electrophiles is facilitated by the use of a terminal reductant (see Scheme 4).

Scheme 4. Principal difference between traditional, oxidative and reductive cross- coupling reactions.69

The Mizoroki-Heck reaction Like the palladium(0)-catalysed cross-coupling reactions described above, the Mizoroki-Heck reaction45,46 proceeds via an initial oxidative addition. However, an organometallic reagent is not required to introduce the second organyl group as the Mizoroki-Heck reaction does not involve a transmetal- lation step.45,46 Instead, an olefin coordinates to the aryl-palladium interme- diate to give a π-complex, followed by carbopalladation of the olefin (see Scheme 5). The palladium intermediate adds to the double bond through a concerted syn-insertion process (migratory insertion) forming a new carbon- carbon bond and a -complex is generated. Migratory insertion can either occur at the terminal or internal carbon of substituted olefins leading to the formation of the linear or branched product.

19

Scheme 5. Mechanism of the neutral Mizoroki-Heck reaction resulting in internal or terminal arylation.

Supposing that there is a vacant coordination site on the metal centre, an internal rotation of the Cα–Cβ bond with respect to palladium positions a - hydrogen syn to the metal centre. Subsequently, elimination of the - hydrogen generates a palladium hydride complex coordinated to the prod- uct.22,70–72 The product is released by reversible dissociation from the palla- dium hydride complex. Analogously to the cross-coupling reactions, the catalytically active palladium(0) complex is regenerated in the last step. However, this is achieved through a base-mediated removal of a proton from the palladium hydride complex instead of a reductive elimination. The palla- dium hydride complex needs to be quickly scavenged by a base after release of the product or addition to the product or the starting olefin can lead to double bond migration or double bond isomerisation (see Scheme 6). 22,70–73

20

Scheme 6. Double bond migration by re-addition and elimination of palladium hy- dride.

As mentioned above, the Mizoroki-Heck reaction can either result in termi- nal or internal arylation of the olefin, depending on the position in which insertion occurs. The regioselectivity of the insertion is dependent on the electronic and steric properties of the π-complex74,75 and the outcome of the reaction depends on the ,76 the solvent77 and the substrates78 used. Electron-poor olefins favour terminal arylation since the internal carbon is more electron-rich and coordinates the palladium centre more strongly. In addition, the -complex leading to terminal arylation is also sterically fa- voured.73,74,79,80 Electron rich substrates generally result in mixtures of the internal and terminal arylation products due to competing electronic and steric factors. In this case, migratory insertion to form the internal - complex is electronically favoured but sterically disfavoured.

Palladium(II) Catalysis Several important processes that rely on palladium(II)-catalysed transfor- mations such as the synthesis of vinyl acetate from and acetic acid,81 the previously mentioned Wacker process82 and the oxidative Heck reaction37 were discovered over 50 years ago. However, the requirement of a stoichiometric amount of palladium or a palladium reoxidant in many palla- dium(II)-mediated reactions hampered further development of these process- es.83 Instead, research was focused on palladium(0) catalysis which has en- joyed considerable development during the last part of the 20th century. Alt- hough palladium(II)-catalysed reactions have received relatively little atten- tion in the academic literature until recently, there has now been a revival in the field of palladium(II)-catalysis fueled by the demand for more environ- mentally benign and sustainable methods as more benign substrates can po- tentially be employed. Palladium(II) catalysis is distinguished from palladium(0)-catalysed cross-coupling reactions in the way that the transition metal enters the cata- lytic cycle. Whereas the first step in the palladium(0) catalytic cycle is an oxidative addition of an aryl halide or pseudohalide, the first step in the cor- responding palladium(II) catalytic cycle is an electrophilic palladation,44,84–88

21 a decarboxylation,89,90 desulfination91–93 or an organyl transmetalation of an arylmetal or arylmetalloid such as mercury,37 boron,94 tin,95 silicon,96 anti- mony97 or phosphorous98. Recently, considerable efforts to obtain the key species by direct C-H activation has been demonstrated. Although this strategy is desirable from an economic and sustainability per- spective, cross-coupling reactions generally display higher regiocontrol. The remaining steps are analogous to palladium(0)-catalysed reactions except the regeneration of the active catalyst where many palladium(II)- catalysed reactions require stoichiometric amounts of a reoxidant. Many 99–101 102–104 105 different oxidants such oxygen, para-benzoquinone, MnO2, desyl chloride,106 TEMPO,101 copper82 and silver89 salts, and peroxides101 can be used. However, the use of copper or silver salts is expensive, TEMPO and desyl chloride are also relatively expensive and para-benzoquinone is toxic. Additionally, the use of reoxidants produces stoichiometric amounts of waste which contributes to poor sustainability and can often complicate puri- fication. However, there are palladium(II)-catalysed cross-coupling reactions acting through different mechanisms where the product releasing step does not produce metallic palladium or palladium hydride.107–109 Therefore, these reactions do not require an oxidant making them highly desirable. The use of more effective ligand and reoxidation systems has addressed the major problems associated with palladium(II) catalysis, allowing the use of catalytic amounts of palladium.100,110–113 The most common arylating agents used in palladium(II) catalysis are ar- yl boronic acids since these are widely commercially available, tolerate a broad variety of functional groups and are relatively environmentally benign. However, the shelf life of aryl boronic acids is limited due to protodeborona- tion and to the formation of trimeric cyclic anhydrides. Further, the handling of certain boronic acids is not convenient due to their waxy appearance.114 Consequently, there is great interest in developing new protocols employing alternative aryl-palladium precursors.

The Oxidative Heck Reaction As described above, the pioneering work of Nobel Prize laureate Richard Heck in the field of palladium-promoted carbon-carbon coupling reactions was initially focused on the use of palladium(II) transformations.38 Inde- pendently of Heck,37 Fujiwara and Moritani84–86 reported on the palladi- um(II)-catalysed direct arylation of olefins with benzene using silver acetate as an oxidant in 1968.44 The palladium(II)-catalysed reactions were developed over time and Heck demonstrated that organoboron substrates could be used in the palladium(II) mediated reaction with , although stoichiometric amounts of 94 Pd(OAc)2 were required. Uemura later developed the first catalytic exam- ple of the oxidative Heck reaction using boronic acids for the arylation of

22 alkenes, although the mechanism originally suggested included the oxidative addition of palladium(0).115 Similar oxidative Heck reactions of organoboron reagents with alkenes and alkynes could be performed using Cu(OAc)2 as the oxidant.116 This development to render the reaction catalytic by inclusion of an oxidant suggested that the organopalladium species at the beginning of the catalytic cycle is formed through transmetallation. Subsequently, several different oxidant systems and arylating agents have been investigated.89,98,117 Notably, in 2003 the use of oxygen as the sole oxidant in the oxidative Heck reaction was developed which represents a more sustainable protocol than previous examples.99,100

Scheme 7. Mechanism of the cationic oxidative Heck reaction resulting in internal or terminal arylation.

The mechanism of the cationic oxidative Heck reaction is presented in Scheme 7. As in other palladium(II)-catalysed reactions the first step is an electrophilic palladation, a decarboxylation/desulfination or an organyl transmetalation. The subsequent steps are the same insertion and -hydride elimination processes as in the Mizoroki-Heck reaction. In contrast to the Mizoroki-Heck reaction, the active catalyst is generally regenerated by reox- idation of the palladium hydride intermediate.

The Mizoroki-Heck and the oxidative Heck reactions can proceed through either neutral (Scheme 5) or cationic (Scheme 7) pathways.80,118 The path-

23 way influences the regiochemical outcome of the reaction since the insertion step may be affected and result in different -complexes. The effect on regi- oselectivity is the same for the Mizoroki-Heck and oxidative Heck reaction since only the first step is different. Coordination of the olefin must be pre- ceded by the dissociation of a ligand. If a neutral ligand is dissociated the resulting -complex is neutral. However, if an anionic X ligand dissociated instead, the resulting -complex is cationic. Bidentate ligands are less likely to dissociate and thus favour a cationic reaction pathway. Weakly associated - - - anionic ligands such as OTf, O2CCF3 or OAc and polar solvents that assist in the exchange of the anionic ligands also favour a cationic mechanism. Analogously to the Mizoroki-Heck reaction, palladium insertion at the in- ternal carbon is electronically favoured for electron-poor substrates but is less accessible due to unfavourable sterics. Conversely, palladium insertion at the terminal carbon is electronically favoured for electron-rich substrates and is more easily accessible. When electron-rich substrates are used, these factors compete and a mixture of the internal and terminal arylation products is often obtained. When the reaction operates through a cationic mechanism the electronic factor becomes more important and electron-rich substrates generally give predominantly the internal arylation product. The migratory insertion in the neutral and the cationic pathway is highlighted in Scheme 8.

Steric clash R R Ar Ar L Pd Ar L Pd X R X L R Neutral pathway Mixture of L Pd Ar R regioisomers X R Ar Ar L Pd Ar L Pd X R L X L L Pd Ar X

R L L L Ar L Pd Ar L Pd Ar L Pd X R Cationic pathway Ar R Steric R clash Scheme 8. Regioselectivity for electron-rich substrates in the oxidative Heck reac- tion acting through a neutral or cationic pathway.

-elimination from the palladium centre can also occur with heteroatom groups such as -OAc, -OH, -Br and -Cl. However, whereas -hydride elimi- nation occurs in a syn fashion, -elimination of heteroatom groups occurs through a trans mechanism with the following relative rates: halide > -OAc > -OR > -OH. -elimination of a heteroatom group directly regenerates the

24 active palladium(II) catalyst which obviates the need for a stoichiometric amount of oxidant.107–109

Palladium-Catalysed Addition Reactions Despite the relatively non-polar metal-carbon bond, palladium shows some reactivity towards polar functional groups such as aldehydes and nitriles. The palladium-catalysed direct addition of boronic acids to aldehydes119 and conjugate addition to ,-unsaturated ketones120 has been reported. Formal addition-elimination reactions catalysed by palladium have been exploited in the synthesis of aryl ketones from in situ generated anhydrides from carboxylic acids and arylboronic acids.121–124 The palladium(II)- catalysed addition of boron compounds to alkynes and subsequent intramo- lecular cyclisation with esters has been reported.125 Interestingly, the direct palladium(II)-catalysed addition and cyclisation of alkynes containing an aldehyde, ketone, or nitrile functional group has also been demonstrated.126 Since the first report by Garves on the stoichiometric palladium-mediated addition of sodium sulfinates to nitriles in 1970, a considerable number of reports concerning the palladium-catalysed addition to nitriles has appeared in the literature.127 Larock has been especially active in the field, and has demonstrated the palladium-catalysed addition of aryl iodides to alkynes and subsequent cyclisation with a nitrile functionality to yield naphthylamines.128,129 This strategy was further expanded to give 2,3- diarylindenones by changing the substrate.130 Further, the synthesis of aryl ketones by palladium-catalysed addition of arylboronic acids131 or simple arenes132,133 to nitriles to give an intermediate ketimine134 and subsequent hydrolysis has been reported.132,135 Palladium-catalysed addition reactions of aryltrifluoroborates136 or arylcarboxylic acids136 to cyanamides results in the formation of amidines, which are less prone to hydrolysis and can be isolated in good yield. The palladium-catalysed addition of arylboronic acids to ni- triles has also been exploited for subsequent formation of benzofurans.133 An overview of palladium-catalysed addition reactions is presented in Scheme 9.

25

Scheme 9. Principal palladium-catalysed addition reactions.

Metal-Catalysed Decarboxylative Reactions Highly activated carboxylic acids such as -oxo-acids, diphenylacetic acids, certain heterocyclic carboxylic acids, propiolic carboxylic acids and di-ortho substituted benzoic acids may protodecarboxylate spontaneously at moderate temperature or upon treatment with an acid. The decarboxylation of -oxo- acids has been exploited extensively as a textbook strategy for the synthesis of various organic compounds, especially via -alkylation of malonic esters (see Scheme 10).

Scheme 10. -Alkylation of a malonic ester and subsequent decarboxylation.

Although the ability of enzymes to catalyse the decarboxylation of a wide range of carboxylic acids has been used to some extent for the production of commodity chemicals,137 the use of metal catalysts is more convenient for the organic chemist. The first metal-mediated protodecarboxylation reactions were reported in 1901 by Pesci138 who employed mercury to effect the trans- formation, and later Shepard139 demonstrated the use of copper bronze or nickel in 1930. Subsequently, the corresponding silver-mediated decarboxy- lative process was reported.140,141 If protonolysis of the intermediate aryl- metal species is avoided, it can be coupled with a carbon electrophile instead (see Scheme 11).

26

Scheme 11. Schematic metal-mediated decarboxylation and subsequent reactions.

In 1966, Nilsson developed a crossed Ullman type coupling reaction be- tween aryl-copper reagents derived from ortho-nitrobenzoic acid and aryl iodides142 but the first highly useful protocol was achieved when the copper- mediated decarboxylation was combined with a palladium catalyst in the cross-coupling of benzoic acids and aryl bromides.143,144 However, stoichio- metric amounts of Cu(II) salts were required until a catalytic Cu(I)-mediated system was developed. As described in previous chapters, there is great interest in developing new protocols employing alternative aryl palladium precursors. Decarboxy- lative metalation is a desirable strategy for the generation of aryl palladium intermediates in cross-coupling reactions since the only waste produced is carbon dioxide. Although decarboxylative processes are well-known, it was not until Myers developed the palladium-catalysed decarboxylative Heck reaction that it gained significant attention (Scheme 12).89,90 This reaction required high catalyst loading and used three equivalents of expensive silver salts as the reoxidant but was later developed to use para-benzoquinone as the palladi- um reoxidant.145

Scheme 12. The decarboxylative Heck reaction reported by Myers.

Since these early discoveries, different mono- and bimetallic palladium(II)- catalysed cross-coupling reactions of carboxylic acids have been developed.89,146,147 Some examples include non-oxidative Suzuki-Myaura and Sonogashira-Hagihara type cross-couplings,143,148,149 oxidative Suzuki- Myaura and Sonogashira-Hagihara type cross-couplings151,152 and oxidative Heck reactions.90,145,150–153 The scope of palladium-catalysed decarboxylative reactions is limited to ortho substituted substrates.154 Theoretical investiga-

27 tions suggest that the carboxylic acid functionality must be orthogonal to the plane of the aromatic system in order for decarboxylative metalation to occur and the ortho substituents help force the carboxyl group into this confor- mation.155,156 A simplified representation of the palladium-catalysed decar- boxylation mechanism is presented in Scheme 13.

Scheme 13. Simplified palladium-catalysed decarboxylation mechanism.

Bimetallic catalyst systems that exploit the decarboxylative activity of cop- per143,147 or silver157,158 and subsequent transmetalation of palladium to initate palladium-catalysed cross-coupling have been used to expand the scope of decarboxylative reactions.146,159–161 However, high loadings of copper or silver salts are generally required in these systems which detracts significant- ly from the main advantage of using carboxylic acids as arylating agents, namely the formation of CO2 as the only by-product.

Sulfinic Acids and Metal-Catalysed Desulfitative Reactions The properties and reactivity of sulfinic acids were first reported in the be- ginning of the 1900s.162–164 Sulfinic acids are also naturally occurring in bio- logical systems as cysteine is reversibly oxidised to the corresponding sulfin- ic acid.165 Sulfinic acids are sensitive to oxidation to the corresponding sul- phonic acids and disproportionation to form sulfinyl sulfones, and are there- fore most commonly handled as their sodium salt, since these are generally less prone to undergo these unwanted reactions.166 Several synthetic approaches have been suggested for the production of sulfinic acids or sulfinate salts. The most common strategy is the reduction of sulfonyl chlorides. This can be achieved either in the presence of sodium sulphite to give the sodium sulfinate167 or by direct reaction of sulfonyl chlo- rides and zinc to give zinc bis(sulfinates)168,169. Other examples include pal- ladium-catalysed coupling of diazonium tetrafluoroborates with sulphur dioxide and hydrogen gas,170 Friedel–Crafts sulfination with sulfonyl chlo- rides171 and the reaction between Grignard172,173 or organolithium reagents174 and sulphur dioxide. From a retrosynthetic perspective, these latter examples are the most straightforward approach to sulfinate salts but rely on the use of toxic and corrosive sulphur dioxide gas. Sulfinates are valuable precursors in organic synthesis and act as either electrophiles or nucleophiles through either a sulfonylative or desulfitative

28 process depending on the reaction conditions.175 An overview of selected examples of the reactivity of sulfinic acids and sulfinates is presented in Scheme 14. Sulfinates can form sulfonyl thioesters in the presence of thiols,176 sulfonyl cyanides in the presence of cyanogen chloride177 and they can react with dinitrogentetraoxide to form sulfonyl nitrites178. Sulfinic acids are readily reduced to disulfides,179,180 which is a key process in the sulfenyl- ation of indoles by sulfinic acids in the presence of toluenesulfonic acid and 181 n-Bu4NI . Sulfinates can react directly with sulphur to form sodium me- thanethiosulfonate182 or can act as electrophiles in electrophilic aromatic substitution reactions to give the corresponding aryl sulfoxides upon treat- ment with trifluoromethanesulfonic acid.183 Desulfitative gem- difluoroolefination of aldehydes and ketones has been demonstrated via in situ generated sodium sulfinates.184 Sulfinates can also be exploited as a source of sulfonyl radicals. Sulfinates can be oxidised directly by TBHP185 or CAN186 and also react with iodine to form sulfonyl iodides that can be homolytically cleaved to form sulfonyl radicals187. Sulfinates spontaneously form radicals in DMSO which has been exploited in the metal-free decar- boxylative coupling with cinnamic acids to form vinyl sulfones.188 Sulfinic acids react via a postulated radical mechanism with NBS to form sulfonyl bromides, which were exploited in the synthesis of β-bromo sulfones from sulfinic acids and styrenes.189

Scheme 14. Selected examples of reactions illustrating the various transformation that sulfinic acids and sulfinate salts may undergo.

Sulfinic acids or sulfinate salts can also undergo desulfitative metalation and subsequent coupling reactions. As described in the previous chapter, palladi- um-catalysed decarboxylative processes require the use of additional metal salts or ortho-substituted very electron-rich aromatic substrates. Sulfinates have emerged as a promising analogous alternative to these arylating agents

29 without the requirement for ortho substitution. The first desulfitative palla- dium-mediated reaction was reported in 1970127 but it was not until 2011 that the first catalytic examples appeared in the literature when Miao,91 Deng92 and Larhed93 independently reported on the palladium-catalysed addition of sulfinic acids or sodium sulfinates to nitriles. The latter report was a com- munication of the preliminary findings in the project discussed in Paper II. Since 2011, the use of desulfitative metalation has been extended to So- nogashira cross-coupling,190 Heck-type reactions,191–194 homocoupling,195 palladium catalysed phosphonation196 and conjugate 1,4-addition197,198. Se- lected examples of reported palladium-catalysed desulfitative reactions are shown in Scheme 15. Several examples of palladium-catalysed desulfitative cross-coupling reactions by C-H activation of azoles,199,200 indoles,201,202 coumarins194 and polyfluoroarenes203 have been reported. Though C-H acti- vation is a desirable strategy in synthesis for economic and environmental reasons, cross-coupling reactions generally display superior regioselectivity.

Scheme 15. Selected examples of reported palladium-catalysed desulfitative reac- tions.

The simplified proposed mechanism for desulfination is analogous to the mechanism suggested for palladium(II)-catalysed decarboxylation and the acid functionality is expected to be orthogonal to the plane of the ring in the transition state (Scheme 16).204 The mechanisms of the desulfination and decarboxylation processes are investigated in detail in Paper II.

30

Scheme 16. Simplified desulfination mechanism.

The reader is referred to the comprehensive reviews that have been pub- lished recently for a full account of the applicability and synthesis of sul- finate salts.175,205,206

Density Functional Theory

Quantum mechanical calculations enable the study of the energies associated with chemical transformations of molecules by studying their electronic properties. Arguably, one of the most fundamental developments in compu- tational chemistry was the development of the Hartree-Fock method that can be employed to iteratively solve the time-independent Schrödinger equa- tion.207 Ab initio quantum mechanical computational methods, such as the Hartree-Fock method, are very time consuming as they are dependent on 3N spatial coordinates and N spin coordinates, where N is the number of elec- trons.

General time-independent Schrödinger equation:208 HΨ = EΨ

Instead of solving the Schrödinger equation, density functional theory (DFT) calculations aim to solve a corresponding equation for the electron density. Becke developed a hybrid theory by merging Hartree-Fock and DFT which has been instrumental in the development of DFT as the standard method for calculations (especially the B3LYP hybrid functional) used today.209–211 DFT calculations rely on basis sets that are mathematical functions typically rep- resenting atomic orbitals that are linearly combined to represent the molecu- lar orbitals. Different numbers of basis functions for the core and valence electrons can be used in conjunction with additional functions to account for polarisation. Also, additional corrections for long range electron-electron interactions (dispersion) have been developed, providing better accuracy of the calculations.212 The computational details will not be covered herein, instead a brief description of the use of these methods as tools in synthetic organic chemistry will follow.

31

Figure 5. An energy minimum (right) and a saddle point or transition state (left) on potential energy surfaces.

The energy of a molecule can be represented by a potential energy surface which is the energy as a function of the geometry of the molecule. The bond lengths and angles between the atoms in a molecule are the dimensions and the geometry is a point in these dimensions. If only two bond lengths are considered, an analogy can be made with a landscape where directions (east- west and north-south) represent the bond lengths and the height corresponds to the energy. To map out the reaction mechanism the geometries of hypoth- esised intermediates are optimised and the energies are calculated. The opti- misation of the geometry finds the closest energy minimum and all isomeric forms and conformations must be investigated in order to find the global energy minimum. The transition state in a reaction is the point with the high- est energy along the reaction coordinate, but is an energy minimum in all other dimensions. In mathematical terms, the transition state is a saddle point on the potential energy surface (see Figure 5). Due to the rapid development of computers and processing speed, the study of more complex systems by computational methods has become pos- sible, and these methods constitute important tools for theoretical investiga- tions of reaction mechanisms.

Microwave-Assisted Organic Synthesis For centuries, chemists and alchemists alike relied on open fires as a source of heat for their chemical transformations. The Bunsen burner was invented in 1855213 and improved the temperature control and probably reduced the number of accidents in chemical laboratories. Most modern laboratories now use even safer sources of heating such as oil baths, metal heating blocks or electric mantles. However, the principle of all these methods is the same type of thermal heating where the inner wall of the reaction vessel heats the reaction mixture by conduction. This results in a temperature gradient where the reactor walls have a significantly higher temperature than the average temperature in the reaction mixture (although the gradient is often reduced

32 by stirring the reaction mixture). The different temperature profiles for heat- ing a reaction vial with conventional heating and microwave heating is pre- sented in Figure 6. The first reports on the use of microwaves in organic synthesis appeared in the 1980s and these reactions were performed in domestic microwave ovens.214,215 Since then dedicated microwave instruments have become commonplace in organic chemistry labs since they enable safe processing at elevated temperatures and pressures, and because of the generally reduced reaction times.216–220 An additional benefit of using microwaves as a source of heating is often increased yields.221–224 The higher selectivity that is some- times observed in microwave-heated reactions can probably be attributed to more homogenous heating and avoiding the wall effects associated with conventional heating. Reactions can typically be performed at higher tem- peratures (above the boiling point of the solvent) and pressures when using dedicated microwave instruments since the reaction vessels are sealed sys- tems. In the event of an explosion there is also a safety benefit of these sys- tems as the temperature and pressure is monitored on-line and the reaction is performed in a purpose-built steel cavity.

Figure 6. Temperature profiles for heating a reaction vial with conventional heating (left) and microwave heating (right).222

Microwaves are electromagnetic radiation with frequencies of 0.3 to 300 GHz and the frequency used for heating is typically 2.45 GHz. This frequen- cy has been assigned to avoid interference with radar and telecommunica- tions.223,225 Microwave radiation can transfer energy to molecules as heat through ionic conduction or dipolar polarisation. When an electromagnetic field is applied, ions start to oscillate through the solution and the kinetic energy of the ions is transformed into heat in the solution (see Figure 7). This phenomenon is referred to as ionic conduction. Dipoles also respond to electromagnetic fields and will rotate in order to adjust to the alternating electric field. If a suitable frequency is applied there

33 will be a phase difference between the dipoles and the electromagnetic field which causes constant rotation and conversion of kinetic energy into heat by friction and collision between molecules. This phenomenon is referred to as dipolar polarisation.225

+ - + -

Figure 7. Microwaves are electromagnetic radiation (left) that transfers energy to molecules as heat through ionic conduction (middle) or dipolar polarization (right).

How a given molecule is affected by the electromagnetic field depends on the dielectric constant (’), and the dielectric loss constant (’’) is a measure of how efficiently the microwave energy is transformed into heat for a given substance. The loss tangent (tan  = ’’/’) is dependent on these constants and is often used to describe the properties of materials in microwave fields and represents the ability of a specific substance to be heated by microwaves at a certain frequency and temperature.226,227

Continuous Flow Chemistry Continuous flow processes have been considered an emerging technology for many years, but have not been widely adopted by organic chemistry la- boratories. One reason for this is likely the technological hurdle associated with using complicated flow chemistry setups. As commercial systems be- come more user-friendly and affordable they will likely become yet another routine tool for the synthesis of compounds in both academia and industry, especially since there are certain benefits over conventional batch synthesis. Better selectivity and higher yields compared to batch can sometimes be achieved when reactions are performed in continuous flow systems.228,229 This can be explained by the constant reaction environment as the product and any side products are instantly removed from the reactor. Also, the safety associated with continuous flow chemistry processes is an important benefit. Highly toxic or explosive reagents can be used or highly exothermic reactions can be performed at relatively low risk, as only a small amount of the reaction mixture volume is heated.228 Also, instead of per- forming the reaction on a larger scale, the reaction can be continuously per- formed for a longer time (scaled out).

34 Flow processes were early on adopted for industrial applications but it was not until the 1970s230 that examples of flow chemistry appeared in the academic literature. However, the field did not receive much interest except for the development of continuous flow peptide synthesis.231 Continuous flow systems currently used in academia are typically based on microfluidics and microreactors but some examples of flow systems that use larger reactors exist. Most commercial continuous flow systems that are available today rely on conductive heating via oil,232 air,233 electric re- sistance234 or induction235,236. Microwave technology enables rapid heating and has been applied in both stop-flow and continuous-flow setups.237–244 The use of microwave technology in flow chemistry was first reported by Strauss in 1994.245 These early continuous flow setups used modified batch microwave systems. Dedicated continuous flow microwave systems have since been developed that allow for efficient direct heating of the reaction mixture. High performance liquid chromatography (HPLC) pumps or syringe pumps are commonly used to move the reaction mixture through the reactor. Syringe pumps tolerate a higher amount of particles and the use of harsh reagents while HPLC pumps tolerate higher pressures. Peristaltic pumps are another alternative that even allow the use of slurries. The main drawback of using continuous flow methodologies is that homogenous reaction setups are often needed, especially when micro-reactors are used since the capillaries are easily blocked by particles.246

Electrospray Ionisation Mass Spectrometry Mass spectrometry (MS) is arguably one of the most important analytical technique available to chemists. Coupled with electrospray ionisation (ESI), which is a soft ionisation technique that only gives few fragmentation prod- ucts, large and relatively sensitive molecules and complexes can be studied. Examples include biomolecules247,248 and organometallic species. ESI-MS is an especially useful tool to directly detect and study charged reaction inter- mediates.249–251 The ionisation method also allows for the study of species that are not inherently charged in the sample. Certain elements have characteristic natural isotopic distributions that can be exploited in the analysis of the mass spectra to identify ions of interest. Bromine and chlorine are typical examples of such elements that show char- acteristic patterns. Analogously, palladium also naturally exists as several different isotopes which can be exploited in the analysis of mass spectra of palladium complexes (see Figure 8). By systematic replacements of palladi- um ligands and substrates, the identity of the ions can be elucidated. The mechanisms of e.g. Mizoroki-Heck251–254 and Suzuki-Miyaura255,256 reactions

35 have been investigated by electrospray ionisation mass spectrometry (ESI- MS).

100 102 104 106 108 110 112 m/z Figure 8. Natural isotopic distribution of palladium.

Although ESI-MS enables the detection of short-lived and unstable reac- tion intermediates at very low concentrations, only ions formed in the ion source are observed and complexes that remain neutral cannot be detected. Intensities correlate with abundance in the mass spectrometer, not in the original sample and care should be taken during the interpretation of the spectra.

36 Research Aims

In diverse research fields, ranging from biology to the development of new display materials, there is an increasing demand for custom molecules. This is especially true in contemporary drug discovery and the search for novel molecular features as potential new pharmacophores. The high level of com- plexity of these organic compounds has fuelled immense effort in the devel- opment of novel synthetic methods to carry out specific transformations, including carbon-carbon bond forming reactions. The formation of new car- bon-carbon bonds allows extension and decoration of a carbon framework, providing a robust and versatile synthetic platform in organic and medicinal chemistry. In the last few decades, tremendous advancements in carbon-carbon bond formations have been observed. The palladium-catalysed coupling reactions that were developed during the end of the last century are among the most important carbon-carbon forming reactions. Many of these reactions rely on the reactivity of aryl-palladium complexes and the investigation of new aryl- palladium precursors is thus of great interest.

The aims of this thesis were: to explore the use of carboxylic acids and sodium arylsulfinates as such new arylating agents, to develop new useful tools for the organic chemist that are convenient and robust, and to provide experimental and theoretical insight into the mechanisms of these reactions.

37 Palladium(II)-Catalysed Decarboxylative Synthesis of Electron-Rich Styrenes and 1,1- Diarylethenes (Paper I)

Background As described in previous chapters, decarboxylative metalation has emerged as an attractive alternative to boronic acids to generate the key aryl- palladium complex in palladium-catalysed reactions, and Myers developed the first decarboxylative Heck reaction in 2002 that used substituted benzoic acids as the arylating agent.89 The use of arylcarboxylic acids is desirable as the only byproduct from the aryl palladium precursor is CO2. In addition, carboxylic acids are less sensitive to air and moisture compared to many commonly used organometallic reagents. Unfortunately the scope for palla- dium-catalysed decarboxylative reactions is currently limited to ortho- substituted arylcarboxylic acids.154,155,257 Styrene is traditionally produced by dehydrogenation of ethyl benzene, catalysed by potassium-promoted iron oxide.258 Styrenes are mainly used for the production of polystyrene materials but are also useful substrates in Diels-Alder,259 oxidation260 and metathesis261 reactions. Due to the useful- ness of styrenes, several palladium(0)-catalysed cross-couplings such as Hiyama-Hatanaku,262 Stille-Migita,263–265 Mizoroki-Heck76,266–270 and Suzuki- Miyaura271,272 reactions for the synthesis of styrenes have been developed. However, these approaches have some drawbacks – the Hiyama-Hatanaku coupling relies on expensive vinyl silanes, the stannanes used in the Stille- Migita reaction are toxic, the Mizoroki-Heck reaction uses flammable eth- ylene and the Suzuki-Miyaura reaction either relies on unstable vinyl iodide, vinyl tosylates that are not commercially available, or expensive trivinylcy- clotriboroxane. To address these economic, safety and environmental issues, an adapted decarboxylative protocol was envisioned using vinyl acetate, an inexpensive vinyl source that only generates acetate as the byproduct, based on a previ- ously reported synthesis of styrenes using boron compounds and vinyl ace- tate.273

38 Thus, starting from the reaction conditions for a similar decarboxylative protocol for electron-rich olefins,274 a model reaction between 2,6- dimethoxybenzoic acid 1 and vinyl acetate 2 was conducted employing Pd(O2CCF3)2 and 6-methyl-2,2’-bipyridyl (8f) as the catalyst system. Analy- sis of the product mixture by GC-MS showed formation of the desired sty- rene product 3 but also revealed the formation of several side products (see Scheme 17).

Scheme 17. Products observed by GC-MS during the initial reaction.

Optimisation of Reaction Conditions for the Synthesis of Styrenes An optimisation study was undertaken to investigate the reaction conditions to improve the selectivity and productivity of the desired styrene product 3. The reaction outcome was monitored by GC-MS of the product and side products in the reaction mixture. The ratio between the integral for products 3a-7a and the internal standard (2,3-dimethylnaphthalene) was used to gauge the conversion. In conjunction, the conversion of the arylcarboxylic acid starting material 1a was monitored by LC-UV (254 nm) by comparing the integral to that of the internal standard. All optimisation reactions were car- ried out using Pd(O2CCF3)2 as the pre-catalyst and 6-methyl-2,2’- dimethylbipyridine (8f) as the bidentate ligand unless otherwise specified. An extensive solvent survey was carried out and the stoichiometry of 2 proved to be important for the generation of 3a. The optimal solvent system included 2 in a large excess as part of the solvent system (see Figure 9). Se- lected reactions performed in solvent systems that provided promising selec- tivity were repeated without an internal standard and the product was isolat- ed by column chromatography. Encouragingly, the reaction performed in vinyl acetate (2):DMF 3:1 gave 68% isolated yield of the desired styrene product 3a.

39

3a 1a 4a+5a 6a 7a 1,2

1

65%a 0,8 68%a

0,6

0,4 Compound/IS ratio Compound/IS 0,2

0 O 2 H NMP THF DMF Neat 2a (75:25) 2a/DMF Dioxane DMF/DMSO No (95:5), 8f Figure 9. Selected results from the screening of solvent systems for the palladi- um(II)-catalysed decarboxylative synthesis of electron-rich styrenes. The propor- tions of the solvent systems are based on the volume. aIsolated yield.

The impact of the addition of an acid or the corresponding base was investi- gated. The addition of trifluoroacetic acid (TFA) or sodium trifluoroacetate (NaO2CCF3) in the reaction mixture did not significantly improve the reac- tion outcome. However, the use of 10% sodium trifluoroacetate led to in- creased inhibition of side product formation and slightly higher selectivity for 3a was observed. Thus, 10% sodium trifluoroacetate was included in subsequent reactions. Next, twelve different palladium salts and complexes were investigated as the palladium source for the reaction together with 8f as the ligand (see Fig- ure 10). Although no specific trend regarding the choice of pre-catalyst con- cerning the reaction outcomes could be discerned, the use of Pd(O2CCF3)2, Pd(OAc)2 and tetrakis(MeCN)Pd(BF4)2 provided the highest amounts of 3a as determined by GC-MS.

40 0,9 3a 4a + 5a 6a 7a 52%a 0,8 70%a a 0,7 68% 0,6 0,5 0,4 0,3 0,2 Compound/IS ratio 0,1 2

0 2 2 2 2 2 2 3 ) ) 3 3 Cl - Cl 2 2 2 2 ) ) 2 2 5 CH 3 ) PdCl 2 2 4 (dba) CCF H 2 Cl 2 3 2 PdCl 2 Pd(OAc) Pd(acac) (1-naph- Pd -C 2 3 Pd CCH 2 PdCl] Pd(BF Pd(O ( η Pd(PPh Pd(MeCN) thyl)] [(Cinnamyl)- [PPh Tetrakis(MeCN) Pd(O Figure 10. Screening of pre-catalysts for the palladium(II)-catalysed decarboxyla- tive synthesis of electron-rich styrenes. aIsolated yield.

These reactions were repeated without internal standard and the styrene product was isolated. Pd(O2CCF3)2 and tetrakis(MeCN)Pd(BF4)2 have weak- ly coordinating counterions and gave 68% and 70% isolated yields of 3a, while Pd(OAc)2 gave lower yield. Using Pd(O2CCF3)2 as the pre-catalyst, a ligand survey was carried out in which a variety of bidentate nitrogen and phosphorous containing ligands were investigated. The structures of the ligands used in the screening are depicted in Scheme 18 and the reaction outcomes as determined by GC-MS in Figure 11. The nitrogen containing ligands could be divided into two dif- ferent groups with different reaction outcomes. Ligands bearing a substituent next to one or both of the (such as ligand 8f) favoured the for- mation of 3a, while all other ligands favoured formation of 7. The reactions with ligands 8f, 8m, 8n and 8p which produced the highest ratios of 3a/IS were repeated without internal standard and the styrene product was isolated. Ligands 8f and 8m gave comparable isolated yields, and 8m was selected as the ligand of choice due to its lower price and higher market availability.

41

Scheme 18. The bidentate ligands screened for the palladium(II)-catalysed decar- boxylative synthesis of electron-rich styrenes.

1 3a 4a + 5a 6a 7a

58% 0,8 53% 68% 69%

0,6

0,4

Compound/IS ratio 0,2

0 8a 8b 8c 8d 8e 8f 8g 8h 8i 8j 8k 8l 8m 8n 8o 8p 8q 8r 8s

Figure 11. Screening of various bidentate ligands for the palladium(II)-catalysed decarboxylative synthesis of 3a.

Due to their equal catalytic performance, the use of Pd(O2CCF3)2 and tetrakis(MeCN)Pd(BF4)2 was investigated further in the evaluation of reac- tion time and temperature (see Figure 12). In general, lowering the reaction temperature (140 °C < 120 °C < 100 °C) seemed to be beneficial in the reac- tion using Pd(O2CCF3)2, and increasing the reaction time from 60 min to 120 min improved the yield of 3a. The reactions performed at 100 °C and 120 °C for 120 min were equally productive and both reactions were repeat- ed without internal standard and 3a was isolated in 74% and 78% yield, re- spectively.

42 Conversely, the use of tetrakis(MeCN)Pd(BF4)2 was most productive at 100 °C and shorter reaction time (60 min), in which 3 was obtained at 71% yield. Lower conversion was observed at 80 °C and a drastic decrease in conversion was observed at 140 °C, possibly due to slower reaction rate and decomposition, respectively. Contrary to the observation seen in the use of Pd(O2CCF3)2, the conversions did not improve when the reaction time was increased to 120 min.

1,2 3a 1a 4a + 5a 6a 7a

1 71%a 0,8 74%a 78%a 0,6

0,4

Compound/IS ratio 0,2

0 Temp. (°C): 80 100 120 140 100 120 140 100 120 140 100 120 140 Time (min): 60 60 60 60 120 120 120 60 60 60 120 120 120

Tetrakis(MeCN)Pd(BF ) Pd(O2CCF3)2 4 2 Figure 12. Screening of time and temperature for the palladium(II)-catalysed decar- boxylative synthesis of 3a. aIsolated yield.

Thus, Pd(O2CCF3)2 was identified as the catalyst of choice and subsequently an optimisation of the catalyst loading was carried out. The most productive loading was determined as 5 mol%, with the use of 8m as the bidentate lig- and in 10% excess. Reducing the catalyst loading to 2.5 mol% led to a slight decrease in reaction yield, while increasing the catalyst loading did not im- prove the yield.

OMe O 5 mol% Pd(O2CCF3)2 OMe 5.5 mol% 8m OH 10 mol% Na(O2CCF3) OMe 1.5 mL 2/DMF (3:1) OMe 120 CMW,2h 1a 3a Isolated yield: 78% Scheme 19. Optimised conditions for the palladium(II)-catalysed decarboxylative synthesis of 3a.

After optimisation of the reaction conditions the reaction between 1a and 2 provided the corresponding styrene 3a in a good isolated yield of 78% (see Scheme 19).

43 Investigation of the Scope of the Palladium(II)- Catalysed Decarboxylative Synthesis of Styrenes Using the optimised reaction conditions, a wide range of ortho-substituted arylcarboxylic acids were employed as substrates in the palladium(II)- catalysed decarboxylative Heck-type reaction (see Table 1). The alkoxy- substituted arylcarboxylic acids 1a and 1d gave good yields of the corre- sponding styrene products 3a and 3d, but disappointingly the use of all other arylcarboxylic acids resulted in poor yields. Arylcarboxylic acids 1c, 1g, 1i and 1j gave predominantly the corresponding transvinylation side products 7 (as determined by GC-MS).

Table 1. Scope of arylcarboxylic acids in the palladium(II)-catalysed decarboxyla- tive synthesis of styrenes.

Isolated yields.

Arylcarboxylic acid 1f did not result in full conversion and the purification of the crude product was troublesome. No activation of the 3-bromo func- tionality during the reaction of arylcarboxylic acid 1b was observed by GC- MS, and the low yield compared to 1a is likely due to the lower electron density. The low yields obtained for most of the arylcarboxylic acids above can be explained by the formation of side products. Thus, to gain further insight into the reaction pathways leading to the formation of the desired product and side products a computational study was initiated.

44 Investigation of the Reaction Mechanism using Density Functional Theory To investigate the reaction mechanism and side reactions theoretical studies by DFT calculations were performed and focused particularly on the for- mation of 5. A schematic proposed catalytic cycle for the decarboxylative synthesis of styrenes is shown in Scheme 20. The carboxylic acid coordi- nates the palladium centre in a bidentate fashion and forms a cationic com- plex. Following decarboxylation, vinyl acetate is coordinated. Migratory insertion leaves one coordination site empty which is then occupied by a carboxylate ion. Finally, -acetate elimination forms the product and the palladium(II) catalyst is regenerated.

Scheme 20. Schematic proposed catalytic cycle the palladium(II)-catalysed decar- boxylative synthesis of styrenes.

The detailed reaction mechanism involves both cationic and neutral com- plexes and calculations are associated with a high degree of uncertainty when comparing charged and neutral species. Previously this problem has been partially mitigated by calculating the final electronic energies at the B3LYP-LACV3P** level, after performing geometry optimisation and vi-

45 brational analysis using the LACVP* basis set.275 Inspired by the success in this study, this strategy was used for the calculations described herein. Though insight regarding the pathways to all the different side products are of value, the internal arylation of 3 to give electron-rich and sterically congested 1,1-diarylethenes 5 is especially interesting as these compounds have not previously been synthesised through Heck type reactions. The ary- lation of 3 may give both stilbene 4 and 1,1-diarylethene 5, and these were indeed among the observed side products in the initial synthesis of styrene 3. Arylation reactions of electron-rich 4-substituted styrenes acting through a cationic mechanism have previously been shown to favour internal aryla- tion.276 A computational investigation to compare the competing arylation of 2 (producing 3a) and subsequent terminal or internal arylation of 3a (produc- ing 4a or 5a) was performed. The calculated free energy profiles are shown in Figure 13.

100 L Ar L N L TS-IIb L Pd L = (87) L Pd Ar N Ar OAc L OMe L TS-I L Pd L Pd Ar (65) H Ar = Ar Ar Ar OMe TS-III L TS-IIa (52) L Pd Ar 50 (60) L L H Ar L L Pd L Pd V L L Pd Ar Ar (35) L Pd O Ar H Ar OAc Ar H Ar III O I (23) IV (19) (20) VI (17) ) -1 II 0 (0) (kJ mol

rel L

G L Pd Ar Ar

L L MeO Pd OMe OMe MeO OMe OMe -50 O O OMe OMe OMe OMe VII (-68) 3a 4a 5a Figure 13. Free energy profile for arylation of vinyl acetate 2 (green) and terminal (blue) or internal (purple) arylation of 3a.

Complex I was chosen as the starting point for this investigation since the decarboxylation step and formation of the aryl–palladium complex is identi- cal for the investigated arylation pathways (the study of the decarboxylation step by DFT calculations is discussed in Paper II). Thus, starting from I three

46 migratory insertion transition states (TS-I, TS-IIa, and TS-IIb) were inves- tigated. Proceeding from complex I to TS-I to eventually form the product 3a is associated with a free energy requirement of 46 kJ mol-1. During the reaction the concentration of 3a is increased and thereby also the concentration of complex II, enabling the competing pathway to side products 4a and 5a. As complex II is actually 19 kJ mol-1 lower in energy than I, the calculations suggest that the product 3a impedes the migratory insertion step towards the formation of itself. Comparing TS-IIa and TS-IIb, which lead to the branched product 5a and the linear product 4a, respectively, shows that the internal arylation (TS- IIa) is favored with an energy difference of 27 kJ mol-1. This observation is in agreement with the expectations for an oxidative Heck reaction of an elec- tron-rich olefin acting through a cationic pathway. According to the compu- tational results the internal second arylation via TS-IIa is predicted to be preferred by 5 kJ mol-1 even compared to TS-I. However, the small energy difference between the transition states for the formation of 3a and 5a make conclusions sensitive for errors in the calculation method. Also, the large excess of 2 used in the reaction protocol optimised for the synthesis of 3a could influence the preferred reaction pathway. Complex III is formed through the arylation via TS-IIa and internal rota- tion results in IV where a -hydrogen is positioned for a syn-elimination which occurs via TS-III to form complex V. Ligand substitution with depro- tonated 1a gives complex VI and the product 5a is released. The catalyst has to be regenerated from the palladium–hydride complex VI in order for the next catalytic cycle to begin. The only reoxidant included in the experimental protocol optimised for the synthesis of 3 was the atmos- pheric oxygen from air in the reaction vial. Para-benzoquinone was compu- tationally evaluated as a reoxidant to investigate the viability of a protocol for the selective synthesis of 5a. The formation of complex VII with para- benzoquinone associated to palladium(0) shows an exergonic pathway that should provide a thermodynamic driving force for the reaction towards 5a. Regeneration of palladium(II) has previously been investigated in a related system that indicates a viable low-energy pathway.277 Therefore, the regen- eration of palladium(II) from complex VII was not included in the computa- tional investigation presented herein.

47 Palladium(II)-Catalysed Decarboxylative Synthesis of 1,1-Diarylethenes The computational investigation described above indicated that there is a low energy difference between the formation of the styrene product 3 and a second internal arylation to produce 1,1-diarylethene 5a. To favour the dia- rylation product the amount of 2 was reduced to 1 equivalent. Para- benzoquinone (p-BQ) was included in the reaction mixture as a palladium reoxidant since the computational results indicated that the formation of 5a results in the liberation of a palladium(0) complex. Different solvent sys- tems, reaction temperature, time, and ligands were screened (Table 2). The solvent system proved to be the most important factor for the synthesis of 5a. The best results with regard to selectivity as determined by crude 1H NMR were obtained using DMF: H2O (90:10) as the solvent. Using the car- boxylic acid in excess did not improve the reaction outcome but performing the reaction at 140 °C resulted in higher isolated yield.

Table 2. Selected examples from the optimisation of reaction conditions for the palladium(II)-catalysed decarboxylative synthesis of 5a.

Entry Ligand Solvent Time Temp. 5a:4a ratioa Yieldb 1 8m Dioxane 2 h 120 °C - - 2 8m DMSO 2 h 120 °C 3:1 - ° 3 8m DMSO: H2O (50:50) 2 h 120 C 13:1 - ° 4 8m [BMIM]BF4 2 h 120 C 10:1 - ° 5 8m H2O 2 h 120 C - - ° 6 8m DMF: H2O (75:25) 2 h 120 C 20:1 55% 7 8f DMF 2 h 120 °C 2.5:1 - ° 8 8f DMF: H2O (90:10) 2 h 120 C 28:1 63% ° a 9 8m DMF: H2O (90:10) 2 h 120 C 50:1 63% ° 10 8m DMF: H2O (90:10) 3 h 120 C 40:1 73% ° b 11 8m DMF: H2O (90:10) 3 h 120 C 50:1 69% ° 12 8m DMF: H2O (90:10) 2 h 140 C 28:1 74% ° 13 8m DMF: H2O (90:10) 3 h 140 C 28:1 70% a4 equivalents of the carboxylic acid was used. bNo sodium trifluoro- acetate was included in the reaction mixture.

48 With the optimised reaction conditions the internal arylation product 5a could be isolated in 74% yield. The reaction also showed impressive selec- tivity over the terminal arylation product (4a) with a 5a/4a ratio of 28:1 as determined by 1H NMR (entry 12). Other arylcarboxylic acids were investi- gated but only 2,4,6-trimethoxybenzoic acid 1f and 2,4-dimethoxybenzoic acid 1k gave the corresponding diarylated products 5f and 5k in low isolated yields (37% and 29%, respectively). These arylcarboxylic acids showed poor selectivity and favoured the terminal arylation product with a 5/4 ratio of 1:1.5 and 1:2.5, respectively. The reaction outcomes for the synthesis of 5 are shown in Scheme 21.

Scheme 21. Palladium(II)-catalysed decarboxylative synthesis of 5 under the opti- mised reaction conditions.

Though the decarboxylative synthesis of styrenes using vinyl acetate as the vinylating agent constitutes a desirable strategy, the palladium-catalysed protocol reported herein is limited to 2,6-alkoxy substituted arylcarboxylic acids. Diarylation can be effected under adjusted conditions, but this reaction also has a similar limited scope. The proposed reaction mechanisms for the decarboxylative Heck reaction with vinyl acetate 2 and subsequent internal arylation of the in situ generated styrene 3 are shown in Scheme 22.

49

Scheme 22. Proposed reaction mechanisms for the oxidative Heck reaction with vinyl acetate 2 and internal arylation of in situ generated styrene 3.

Palladium(II)-Catalysed Desulfitative Addition of Sodium Arylsulfinates to Nitriles (Paper II- III)

Background Though palladium-catalysed decarboxylative coupling reactions are a very desirable strategy for carbon-carbon bond formation, the current limitation to electron-rich ortho-substituted benzoic acids generally prevents this ap- proach from being truly synthetically viable. Decarboxylation of unsubstitut- ed benzoic acids have a high activation barrier and previous DFT studies on palladium-mediated decarboxylation reactions have suggested that the car- boxyl group is orthogonal to the aromatic ring in the decarboxylative transi- tion state and the conjugation between the carboxyl group and the aromatic ring must be broken in order to reach the transition state.159 Consequently, the energy required to reach the transition state for abstraction of an acid functionality should directly correlate with the rotational barrier of the acid moiety. Based on this assumption alternative aryl acids with lower rotational barriers may serve as potentially superior arylating agents as they are pre- dicted to undergo metalation more readily. Once the key aryl-palladium complex is formed, a large number of divergent transformations can be achieved. For instance, imines could be formed by the insertion of the aryl group into the polar triple bond of nitriles and subsequently hydrolysed to the corresponding ketone.127–130 Due to the versatility of the ketone function- ality many different methods for preparing ketones and aryl ketones have been developed. However, the synthesis of aryl ketones from carboxylic acids or acid derivatives using classic approaches generally requires harsh reaction conditions.278

Initial Investigation

Based on the hypothesis outlined above, phenyl sulfonic acid, phenyl phos- phonic acid and the sodium salt of phenyl sulfinic acid were employed in the protocol reported by Lindh for decarboxylative addition to nitriles.279 How-

51 ever, only sodium benzenesulfinate furnished the formation of the corre- sponding aryl ketone, albeit in trace amounts (see Scheme 23).

Scheme 23. Scouting reactions using benzenesulfonic acid, benzenephosphonic acid and benzenesulfinic acid as aryl-palladium precursors in the palladium-catalysed addition to acetonitrile under conditions from Lindh.279

The rotational barriers for the acid functionalities of the aryl acids of interest were calculated by performing dihedral drives for 2,6-dimethoxy-benzoic acid, benzoic acid and benzenesulfinic acid (see Figure 14). The energy re- quired to break the conjugation of the aromatic ring was lower for benzene- sulfinic acid compared to the carboxylic acid, hence the former should pos- sess a lower free energy requirement to reach the transition state for desul- fination than for the corresponding decarboxylation of arylcarboxylic acids. This was also supported by the preparative results above, as sodium 4- methyl-benzenesulfinate, bearing no ortho substituents, produced the desired product.

52 30 Benzenesulfinic acid

Benzoic acid

2,6-Dimethoxybenzoic 25 acid

20

) 15 -1

E (kJ mol 10 

5

0 0 30 60 90 120 150 180 Dihedral angle 1-2-3-4 (degrees) Figure 14. Dihedral drive energies for benzenesulfinic acid, benzoic acid and 2,6- dimethoxybenzoic acid.

A preparative investigation was initiated and the initial findings concerning the use of sodium arylsulfinates and nitriles for the Pd(II)-catalysed synthesis of aryl ketones were communicated93 simultaneously as the reports by Wang91 and Deng92. Subsequently, further efforts to investigate and optimise the microwave-assisted palladium(II)-catalysed addition of sodium aryl- sulfinates to organic nitriles and successive hydrolysis have been made and these results are discussed herein.

Optimisation of Reaction Conditions for the Synthesis of Aryl Ketones Encouraged by the results above, an optimisation of the reaction conditions for the palladium(II)-catalysed desulfitative addition of sodium 4- methylbenzene sulfinate 9a to acetonitrile 10a under microwave heating was pursued (see Table 3). Initially, the catalyst loading and the water component of the solvent system were increased. The presence of a strong acid to pro- mote hydrolysis was crucial for the reaction outcome. Inclusion of trifluoro- acetic acid (TFA) in the reaction mixture greatly improved the yield the de-

53 sired aryl ketone 11aa. This observation is attributed to the dependence of catalytic turnover on the hydrolysis of the intermediate ketimine. The solvent survey showed that dioxane worked as a substitute for THF in the solvent system with an equal proportion of water (entry 9), but using solvents with one equivalent of water did not provide satisfactory results (entries 6, 7 and 10-14). Also, neat water was not an efficient medium for the reaction.

Table 3. Optimisation of reaction conditions for palladium(II)-catalysed desulfita- tive addition of sodium 4-methylbenzene sulfinate to acetonitrile.

Entry Solvent Catalyst Acid Ligand Yielda

1 Water/THF (1:1) Pd(O2CF3)2 - 8f Traces 2 Water/THF (1:1) Pd(O2CF3)2 1 eq. TFA 8f 19% 3 Water/THF (1:1) Pd(O2CF3)2 5 eq. TFA 8f 72% 4 Water/THF (1:1) Pd(O2CF3)2 10 eq. TFA 8f 87% 5 Water/THF (1:1) Pd(O2CF3)2 20 eq. TFA 8f 83% 6 THF Pd(O2CF3)2 10 eq. TFA 8f 39% 7 Dioxane Pd(O2CF3)2 10 eq. TFA 8f 43% 8 Water Pd(O2CF3)2 10 eq. TFA 8f 28% 9 Water/dioxane (1:1) Pd(O2CF3)2 10 eq. TFA 8f 77% 10 Isobutanol Pd(O2CF3)2 10 eq. TFA 8f 64% 11 DMF Pd(O2CF3)2 10 eq. TFA 8f 54% 12 DME Pd(O2CF3)2 10 eq. TFA 8f 28% 13 NMP Pd(O2CF3)2 10 eq. TFA 8f 17% 14 DMSO Pd(O2CF3)2 10 eq. TFA 8f 5% 15 Water/THF (1:1) Pd(OAc)2 10 eq. TFA 8f 47% 16 Water/THF (1:1) Pd(OH)2 10 eq. TFA 8f 47% 17 Water/THF (1:1) Pd(dba)2 10 eq. TFA 8f 32% 18 Water/THF (1:1) PdCl2 10 eq. TFA 8f 5% 19 Water/THF (1:1) Pd(O2CF3)2 10 eq. TFA 8a 39% 20 Water/THF (1:1) Pd(O2CF3)2 10 eq. TFA 8h 8% 21 Water/THF (1:1) Pd(O2CF3)2 10 eq. TFA 8i 31% 22 Water/THF (1:1) Pd(O2CF3)2 10 eq. TFA 8m 8% 23 Water/THF (1:1) Pd(O2CF3)2 10 eq. TFA 8s - 24 Water/THF (1:1) Pd(O2CF3)2 10 eq. TFA 8r - 25 Water/THF (1:1) Pd(O2CF3)2 10 eq. TFA 8t, PPh3 15% 26 Water/THF (1:1) Pd(O2CF3)2 10 eq. TFA 8f, DMSO 10% 27 Water/THF (1:1) Pd(O2CF3)2 10 eq. TFA - Traces aIsolated yield.

54 The more reactive Pd(O2CCF3)2 proved to be more efficient than other palla- dium sources and the ligand (6-methyl-2,2’-bipyridyl, 8f) was crucial for catalytic activity. The structure of the ligands are shown in Figure 9. The use of similar nitrogen based bidentate ligands resulted in the formation of the desired product but in significantly lower yields. Under the optimised reac- tion conditions, aryl ketone 11aa was obtained in 87% isolated yield.

Investigation of the Reaction Mechanism using Electrospray Ionisation Mass Spectrometry In order to gain experimental insight into the reaction mechanism, stoichio- metric reactions with varying ligands and sodium sulfinates were set up and heated at 100 °C for 10 minutes, diluted 10 times with 10a and then directly analysed using electrospray ionisation mass spectrometry. The characteristic isotopic distribution of palladium was exploited to identify palladium com- plexes and the composition of these complexes was further verified via MS- MS and MS-MS-MS analyses. Concomitantly, neutral loss experiments monitoring the loss of acetonitrile and SO2 (41 and 64 Da, respectively) were also performed to verify the identity of some of the proposed palladium complexes. Only a small number of signals for non-palladium ions were observed. ESI-MS(+) spectra and the identified complexes are depicted in Figure 15. The major signal in all the ESI-MS(+) spectra corresponded to the aryl- palladium complex coordinated only by the ligand with a free palladium coordination site. This observation is likely due to the formation of the com- plex in the ion source of the mass spectrometer through loss of weakly coor- dinated neutral SO2 and nitrile ligands.

55 300 400 500m/z 600 700 800 Sulfinate: Sodium benzenesulfinate Sulfinate: Sodium 4-methylbenzenesulfinate Ligand: Phenanthroline Figure 15. ESI-MS spectra and identified cationic palladium complexes with nu- mering according to their role in the catalytic cycle.

It is important to note that the intensity of the ESI-MS signals only correlate to the stability and formation of the ions in the mass spectrometer and not to their abundance or role in the reaction mixture. Although the majority of the complexes observed (shown in Figure 13) supported the previously suggest- ed mechanism, further mechanistic investigation by an orthogonal method was warranted.

Investigation of the Reaction Mechanism using Density Functional Theory To further elucidate the mechanism of the desulfitative reaction developed herein, a computational study was performed. The reaction mechanism was investigated by DFT calculations and geometries were optimised at the B3LYP-LACVP* level of theory and single point energies were calculated at the B3LYP-LACVP**+ level of theory. The decarboxylative and desul- fitative pathways were compared using water as the solvation model because the calculations did not allow solvent mixtures. The free energy profiles for desulfination and decarboxylation are shown in Figure 16 and the free ener- gy profiles for 1,2-carbopalladation of 10a (acetonitrile) are shown in Figure 17. The starting point and reference energy level for the investigation was Pd(O2CCF3)2 (complex I) as it is the common lowest energy complex for all

56 three reactions. Stepwise dissociation of the trifluoroacetate ligands from the catalyst followed by association of a single aryl acid led to complex VI which has the lowest energy prior to the transition state for decarboxylation for 2,6-dimethoxybenzoate 1a. In contrast, the lowest energy complex for benzoate 1k is the neutral complex V bearing two aryl acid ligands. The lowest energy prior to desulfination of benzenesulfinic acid 9e was complex IVa. The binding mode of the aryl acid in IV is changed to allow an interaction between the aryl group and the palladium centre in complex VI which pre- cedes the desulfination/decarboxylation transition state (TS-I). The calculation of the free energy requirement for decarboxylation of benzoic acid involves comparison of neutral and cationic complexes, which is associated with a relatively large degree of uncertainty. However, the free energy requirement for decarboxylation is significantly higher (2.5 fold) than for the desulfitative process (117.9 kJ mol-1, Vb to TS-Ib, compared to 46.5 kJ mol-1, IVa to TS-Ia). The difference between the free energy require- ments is larger than the expected error in the calculations. This is in agree- ment with previous observations regarding the relative free energy require- ments in copper-mediated decarboxylation and desulfination reactions.204

Figure 16. Free energy profile for desulfination of benzenesulfinic acid and decarboxylation of benzoic acid and 2,6-dimethoxybenzoic acid.

57

Figure 17. Free energy profile for 1,2-carbopalladation of acetonitrile 10a.

The free energy requirement for the carbopalladation of acetonitrile with phenyl was calculated to 99.1 kJ mol-1 (VIIIa-b to TS-IIa-b). The free ener- gy requirement for the more electron-rich 2,6-dimethoxy-phenyl was consid- erably lower (81.6 kJ mol-1, VIIIc to TS-IIc), which is in agreement with previous studies on the carbopalladation of cyanamides.136 It should be noted that the free energy requirement backward from com- plex VIIIa to complex IVa (67.3 kJ mol-1) is lower compared to the car- bopalladation (Figure 18). Thus, the backward reaction should occur more easily than the forward reaction. However, further transformation of com- plex IXa to give the product 11aa and regeneration of the low energy com- plex Va is associated with a free energy gain of 17.0 kJ mol-1 which provides a thermodynamic driving force for the reaction.

58

Figure 18. The free energy requirement for the backward reaction from VIIIa to IVa is and for carbopalladation in the desulfitative reaction.

In the computational model, the imine and ammonia were released as neutral species, although the reaction is performed in the presence of 10 equivalents of TFA. The protonation states and subsequent hydrolysis was not computa- tionally investigated. Assuming that these processes occurred without signif- icant free energy requirements, the insertion step was rate determining for benzene sulfinic acid 1e and 2,6-dimethoxybenzoic acid 1a. Conversely, the decarboxylation step was rate determining for benzoic acid. The results from the computational and ESI-MS studies are in agreement with the previously proposed mechanism but provide greater insight and validity to the mechanism. The proposed reaction mechanism is shown in Scheme 24.

59

Scheme 24. Proposed reaction mechanism for the palladium(II)-catalysed desulfita- tive addition of sodium arylsulfinates to nitriles.

Investigation of the Scope of the Microwave-Assisted Organic Synthesis of Aryl Ketones The scope of the protocol developed herein was first investigated by employ- ing a range of sodium arylsulfinates, purchased commercially or synthesised in-house, with different electronic properties using acetonitrile 10a as the reaction partner. As seen in Table 4, the desulfitative reaction tolerated a much wider scope of aryl acids compared to the corresponding decarboxyla- tive reaction. Electron-rich and moderately electron-deficient substrates per- formed well and ortho substitution was not a requirement for productive reactions. In fact, sterically congested sodium 2,4,6-trimethylbenzene sul- finate 9k resulted in very low conversion. The corresponding carboxylic acid provided acceptable yield in the decarboxylative reaction, which illustrated the orthogonality of these protocols. The amide functionality of sodium 4-acetamidobenzene sulfinate was hy- drolysed under the optimised reaction conditions, but the desired aryl ketone product could be obtained in satisfying 57% yield by performing the reaction with 10% water in 10a as the solvent.

60 For selected entries, the reaction was performed under both conventional and microwave heating. In all these examples microwave heating proved more efficient with shorter reaction times and higher isolated yields. The reactions were performed in 2-5 ml microwave transparent borosilicate glass vials and were either heated in a dedicated batch microwave instrument or using a metal heating block with the temperature maintained at 100 °C. Un- fortunately, highly electron-deficient substrates resulted in homocoupling, probably due to sluggish carbopalladation, and poor yield of the desired aryl ketones were obtained regardless of the heating method used.

Table 4. Preparation of aryl ketones by palladium(II)-catalysed desulfitative addi- tion of various sodium arylsulfinates to 10a.

Isolated yields. aPerformed under microwave heating. bPerformed under conventional heating. cPerformed using 10% water in 10a as the solvent.

Next, the scope of nitriles was investigated. A wide range of aliphatic, ben- zylic and aromatic nitriles was employed and the reactions proceeded smoothly to produce the corresponding aryl ketones in moderate to excellent yields (see Table 5). Both electron-rich and electron-poor nitriles performed well as substrates, as illustrated by the yields obtained for the reactions em-

61 ploying 4-methoxy- and 4-acetyl-benzonitrile (68% and 51%, respectively). 4-Formyl substituted benzonitrile also yielded the desired aryl ketone, albeit in slightly lower yield. Reactions using different bromobenzonitriles did not show any bromine activation by the palladium catalyst which illustrates the orthogonality of the palladium(II)-catalysed reaction developed herein to conventional palladium(0) processes. In addition, ethyl 2-cyano-2- phenylacetate could be employed without hydrolysis of the ester functionali- ty, highlighting the functional group tolerability of the reaction. Selected entries were also performed under both conventional and microwave heating and the results were consistent with the previous observations.

Table 5. Preparation of aryl ketones by palladium(II)-catalysed desulfitative addi- tion of 9a to various nitriles.

Isolated yields. aPerformed under microwave heating. bPerformed under convention- al heating.

62 Next, heterocycles containing the nitrile functionality were investigated to access an array of heterocyclic ketones. Interestingly, the palladium(II)- catalysed desulfitative reaction of (2-hydroxyphenyl)acetonitrile led to the formation of 2-arylbenzofuran in 49% isolated yield through aryl ketone formation and subsequent intramolecular condensation (Scheme 25). Related desulfitative coupling with 2-(gemdibromovinyl)phenols280 and addition of arylboronic acids to nitriles133 followed by cyclisation to the corresponding benzofurans have previously been reported.

Scheme 25. Aryl ketone formation and subsequent intramolecular condensation.

The scope of the reaction was further investigated by combining an array of sodium arylsulfinates and nitriles to give 22 additional examples of diverse aryl ketones (see Table 6). All these reactions proceeded smoothly, in which high yields were obtained for the reaction between electron-rich to moder- ately electron-poor sodium arylsulfinates and aliphatic, benzylic and aro- matic nitriles. These results highlighted the consistency and versatility of the method developed herein.

63 Table 6. Preparation of aryl ketones by palladium(II)-catalysed desulfitative addi- tion of sodium arylsulfinates to nitriles.

Isolated yields. aPerformed under microwave heating. bPerformed under convention- al heating.

64 Continuous Flow Microwave-Assisted Organic Synthesis of Aryl Ketones The application of the palladium(II)-catalysed desulfitative addition of sodi- um arylsulfinates to nitriles was further extended to continuous flow under microwave heating. As described in the introduction to this thesis, the rapid heating associated with microwave technology makes it useful for flow ap- plications. The reaction mixture is directly heated when microwave transpar- ent reactors such as borosilicate glass reactors are used. Borosilicate glass has previously been used as a reactor material in microwave-assisted metal- catalysed flow reactions.281,282 However, microwave induced superheating due to metal precipitation leading to potential reactor failure constitutes a major problem. Due to their higher durability, silicon carbide reactors have been sug- gested as a safe alternative for microwave heated metal-catalysed flow reactions. These reactors have the ability to withstand the stress caused by microwave induced hot-spots at higher temperature that may form if metal is deposited on the internal surface of the reactor.289 However, un- like borosilicate glass, silicon carbide absorbs microwaves and the reaction mixture is heated in a more conventional manner using these reactors.283 Aluminium oxide is thermostable with a maximum operating temperature of 1950 °C and is microwave transparent. Thus, this material should serve as a safe alternative that has microwave properties more similar to borosilicate glass. Aluminium oxide is a common material used in standard flow chemistry applications but has not been exploited as a reactor material for microwave-assisted continuous flow processes. The continuous flow setup is shown in Figure 19 and Figure 20.

65 Figure 19. Schematic representation of the continuous flow setup.

Figure 20. Photograph of the continuous flow setup. Setup for CF-MAOS showing pump, generator with MW applicator on top and fraction collector (left), and close- up of microwave applicator with the front piece unmounted and a borosilicate glass tube reactor inside the helical antenna (right).

The temperature profiles for heating benzene in the continuous flow system (Figure 21) illustrate the properties of these materials under microwave heat- ing. Benzene absorbs very little of the microwave radiation which resulted in a slow increase in temperature for the microwave transparent materials. Sili- con carbide absorbs microwaves and the temperature increases more quick- ly. For microwave-absorbing solvents or reaction mixtures, the temperature profiles are similar. It must be noted that different sensors were used to measure the temperature for the different materials. A study correlating the measurements to a fiber-optic sensor has previously demonstrated that an

66 Optris CSmicro 3M IR sensor provides an accurate estimation of the temper- ature of the solution in borosilicate glass reactors.284 Hence, this sensor was used for temperature measurements of reactions performed in borosilicate glass reactors. For the aluminium oxide and silicon carbide reactors, an Optris CT IR sensor with a LT22 sensing head was used which measures the outside temperature of the reactor.

100 100

80 80

60 60 Borosilicate glass 40 40 Aluminium oxide Temperature (C) Temperature (C) Silicon carbide 20 20

0 0 0 60 120 180 0 60 120 180 Time (s) Time (s) Figure 21. Heating profiles with the set temperature 100 °C for different reactor materials pumping benzene (left) and THF:H2O (1:1) (right) at flow rate 1 ml/min.

In comparison to the batch protocol, the stock reaction mixture was diluted 10-fold to avoid excessive metal precipitation in the reactor. Reactions em- ploying hydrophobic nitriles resulted in biphasic reaction mixtures and were particularly challenging to perform in a continuous flow setup. However, upon vigorous stirring, the reaction mixture formed an emulsion that could be pumped into the reactor in a controlled fashion using an HPLC pump (see Figure 22).

Figure 22. Reaction mixture prepared for the synthesis of aryl ketone 11ad without (left) and with stirring (right).

67 The reaction conditions were optimised for the palladium(II)-catalysed addi- tion of sodium 4-methylbenzene sulfinate 9a to benzyl cyanide 10d (see Table 7). The temperature and flow rate were varied, and the highest isolated yield of 79% was obtained at 120 °C and a flow rate of 1 mL/min which corresponds to 50 s residence time in the microwave heated zone that covers 120 mm of the 200 mm tube reactor.

Table 7. Optimisation of process parameters for microwave-heated continuous flow palladium(II)-catalysed desulfitative synthesis of aryl ketone 11ad.

Entry Temperature Flow rate Residence time Isolated yield

1 80 °C 1 mL/min 50 s Trace 2 100 °C 1 mL/min 50 s 78% 3 120 °C 0.5 mL/min 100 s 66% 4 120 °C 1 mL/min 50 s 79% 5 120 °C 2 mL/min 25 s 74% 6 140 °C 1 mL/min 50 s 53% The temperature was measured using an Optris CSmicro 3M IR sen- sor.

Next, three different reactor materials (see Figure 23) were compared for the continuous flow microwave-assisted synthesis of aryl ketone 11ad. The re- action was performed at three different temperatures for each reactor materi- al (100 °C, 120 °C and 140 °C). The productivity of these reactions were determined by GC-MS using 2-methylnapthalene as an internal standard, and the product from the reaction with the highest product to internal stand- ard ratio was isolated.

68

Figure 23. Flow reactors made from aluminium oxide (top), borosilicate glass (mid- dle) and silicon carbide (bottom).

The borosilicate glass and aluminium oxide reactors provided equal yields of the aryl ketone product at 120 °C. The best result was achieved at 100 °C for the silicon carbide reactor and the isolated yield was comparable to the other reactors (see Table 8). It should be noted that the heated zone corresponds to the length of the microwave heated zone (120 mm) for the borosilicate glass and aluminium oxide reactors, but to the full length of the reactor (200 mm) for the silicon carbide reactor. For the aluminium oxide and silicon carbide reactors the temperature was measured using an Optris CT IR sensor with a LT22 sensing head which measures the outside temperature of the reactor. By using the microwave-absorbing silicon carbide reactor, conventional heating can be mimicked and be used to investigate a potential microwave activated liquid interface. However, the use of this reactor material gave comparable results and no non-thermal beneficial microwave effect was observed. The influence of a phase transfer catalyst (tetrabutylammonium hydro- gensulphate) was also investigated, but no improvement of the reaction out- come was observed.

Table 8. Comparison of flow reactors for the continuous flow microwave-assisted synthesis of aryl ketone 11ad.

Entry Reactor material Temperature Sensor Yielda 1 Borosilicate glass 120 °C Optris CSmicro 3M 79% 2 Silicon carbide 120 °C Optris CT IR sensor, LT22 sensing head 78% 3 Aluminium oxide 100 °C Optris CT IR sensor, LT22 sensing head 74% aIsolated yields.

69 Encouraged by the positive results obtained above, the scope of the palladi- um(II)-catalysed desulfitative ketone synthesis under continuous flow condi- tions further was explored using both the borosilicate glass and aluminium oxide reactors (see Table 9). The palladium(II)-catalysed addition of arylsulfinates 9a, 9b and 9e to 10a under continuous flow conditions afforded the corresponding aryl ke- tones in 55%/50%, 68%/62% and 61%/58% yields, respectively. The reac- tion employing sodium 4-acetamidobenzenesulfinate 9d showed no trace of the corresponding hydrolysed product and 11da was obtained in 60%/56% yield. However, the moderately electron-poor sodium 4- bromobenzenesulfinate 9r only afforded trace amounts of 11ra, likely due to the short reaction time of the CF-MAOS protocol. Reactions employing 10c and 10g as the nitrile furnished the corresponding aryl ketone in 71%/54% and 53%/50% yield, respectively. 3-thiophenecarbonitrile could be used to introduce a heterocyclic functionality and gave a satisfying 66%/58% yield of 11eq. No traces of side products from palladium(0)-catalysed activation of the phenyl bromide functionality was observed when 4-bromobenzonitrile 10h was employed as the nitrile. However, the CF-MAOS protocol only furnished 47%/56% isolated yield of the desired product for this substrate.

Table 9. Preparation of aryl ketones by microwave-heated continuous flow palladi- um(II)-catalysed desulfitative addition of sodium arylsulfinates to nitriles.

Pd(O2CCF3)2 O 8f O S + RCN CF-MAOS Ar ONa TFA, THF/H2O Ar R 1 mL/min, 120 °C 9 10 11 75 mM 5 eq. øi 3mm O O O O O Me Me Me Me N MeO H 11aa 11ba 11da 11ea 55%a/50%b 68%a/62%b 60%a/56%b 61%a/58%b O O O Me

Br 11ra 11bc 11bd Trace/Trace 71%a/54%b 79%a/63%b O O O

S Br 11bg 11bh 11bq 53%a/50%b 47%a/56%b 66%a/58%b Isolated yields. aPerformed in a borosilicate glass reactor and the temperature was measured using an Optris CSmicro 3M IR sensor. bPerformed in an aluminium oxide reactor and the temperature was measured using an Optris CT IR sensor with a LT22 sensing head.

70 Lower yield than for the batch protocol were obtained for all entries ex- cept for the reaction between 9a and 10d. However, when palladium(II)- catalysed addition reactions of 1b with 2a and 2e were performed in batch at the same diluted concentrations that were used for flow applications, identi- cal yields were obtained. The yields for reactions performed in the alumini- um oxide reactor were generally slightly lower compared to the results ob- tained using the borosilicate glass reactor. No problems with reactor failure due to metal precipitation and microwave induced superheating was ob- served for either reactor material.

71 Synthesis of Sodium Arylsulfinates Using a Solid Source of Sulphur Dioxide (Paper IV)

Background As outlined in the previous chapters, sodium arylsulfinates are an increasing- ly important class of building blocks, especially as aryl-palladium precursors in palladium(II)-catalysis. Potential applications include the preparation of pharmaceuticals, functional materials and bioactive compounds. The com- mercial availability of sulfinic acids and their corresponding salts is limited. However, they can be prepared by the methods described in the introduction, such as the reaction of organometallic reagents with sulphur dioxide. However, the use of toxic sulphur dioxide gas and the requirement for spe- cialised equipment is considered inconvenient. A solid source of sulphur dioxide would allow for safer and more convenient synthesis of these com- pounds. The complex of sulphur dioxide with DABCO (depicted in Scheme 26) was chosen as the gas surrogate, which has previously been exploited for in situ generation of sulfinates for further reactions.285–291

Scheme 26. Formation of the sulphur dioxide complex DABSO.

The aim of this project was to develop a convenient method for the synthesis and isolation of sodium arylsulfinates.

72 Development of the Protocol An initial reaction between commercially available phenylmagnesium chlo- ride and DABSO in THF was conducted. Gratifyingly, this resulted in al- most quantitative conversion to the magnesium chloride sulfinate 9b, how- ever purification of the crude sulfinate proved to be difficult due to rapid decomposition. Liquid-liquid extraction using 2 M HCl, attempted purifica- tion of the sulfinic acid by silica gel, aluminum oxide, or reverse phase chromatography resulted in almost quantitative conversion to the corre- sponding sulphonic acid. Due to the stability issues above and the promotion of disproportionation of sulfinic acids by HCl,292 an alternative purification strategy was devised in which the sulfinate salt was isolated by treatment with aqueous Na2CO3 and purification by liquid–liquid and solid–liquid extraction. Importantly, sulfu- ric acid was used in the liquid–liquid extraction and it does not promote dis- propotionation of sulfinic acids.170 This purification strategy proved success- ful as purification of crude 9b afforded the desired sodium arylsulfinate in 90% yield, without any traces of the corresponding sulphonic acid. Next, different approaches to the formation of the were investigated since the classical generation of the Grignard reagent proved sluggish. The halogen-metal exchange reagent iPrMgCl*LiCl has previously been used with great success for this process but often requires long reaction times for neutral or electron-rich aryl bromides.293 A previously developed microwave heated protocol proved to be a rapid and convenient strategy for the generation of the Grignard reagents.294

Investigation of the Scope of the Synthesis of Sodium Arylsulfinates To explore the scope the two-step synthesis of sodium sulfinates developed herein, a series of aryl halides were processed in the microwave protocol for the generation of the Grignard reagents and subsequently allowed to react with DABSO at 0 °C. As seen in Table 10, electron-poor substrates were more productive than electron-rich substrates which is illustrated by compar- ing the outcomes of the reactions with 4-methoxy- and 3- methoxybromobenzene (48% and 85%, respectively). Electron-poor 4-CF3- bromobenzene gave an excellent yield of the desired product 9h but intro- ducing a steric factor by moving the substituent to the ortho position resulted in a lower yield (91% and 75%, respectively). However, treatment of the sterically congested 2,4,6-trimethyl-substituted aryl bromide 12k was well- tolerated and the corresponding sodium arylsulfinate was obtained in 69% yield.

73 Table 10. Preparation of sodium arylsulfinates by the reaction of Grignard reagents with DABSO.

Isolated yields. aIodobenzene. bCommercial phenylmagnesium chloride. cBromobenzene. dIodobenzene.

Sodium 4-biphenyl sulfinate is inaccessible through the common protocol employing sulfonyl chlorides in an aqueous solution due to poor solubility of the starting material in water. However, the use of 4-bromobiphenyl in the developed protocol returned the desired sodium sulfinate 9q in 87% yield. Next, the corresponding reactions via lithium halogen exchange and sub- sequent reaction with DABSO were investigated (see Table 11). In general, the yields obtained were comparable to those using the Grignard reagents previously investigated. Electron-poor substrates were also favoured in the synthesis via lithiation, e.g. 3-methoxy substituted aryl halide 12r afforded higher yield than the 4-methoxy substituted regioisomer 12b (91% compared to 59%). However, contrary to the trend seen for the corresponding Grignard reagents, the use of 4-CF3-bromobenzene resulted in significantly lower yield than using 2-CF3-bromobenzene (48% compared to 99%). Notably, sodium sulfinate 9e could be obtained in better yield than via the corre- sponding protocol employing the generated Grignard reagent (85% com- pared to 72%). Sodium 4-biphenyl sulfinate 9q was obtained in low yield due to the poor solubility of the generated aryl lithium reagent and problem- atic addition to the reaction mixture.

74 Table 11. Preparation of sodium arylsulfinates by the reaction of lithium reagents with DABSO.

Isolated yields.

Treatment of 2-bromo-thiophene with n-butyl lithium at -78 °C resulted in the formation of butyl-thiophene sulfinate by direct substitution and subse- quent reaction with DABSO (see Scheme 27). This sodium sulfinate could be isolated in 68% yield.

Scheme 27. Formation of undesired product using n-butyl lithium. Isolated yield.

The prospect of direct deprotonation of heteroaryls and reaction with DAB- SO to generate sulfinates was also investigated.By using the more basic and less nucleophilic t-BuLi the desired thiophene sulfinate could be obtained. However, the isolated yield was poor due to the high polarity of the product resulting in problematic purification by liquid-liquid extraction. The reaction with furan furnished the corresponding sodium sulfinate in similar disap- pointing yield, but when the more hydrophobic benzofuran was employed the yield of the desired product was improved to 50% (see Table 12).

75 Table 12. Preparation of sodium arylsulfinates by the reaction of deprotonated het- eroaryl compounds with DABSO.

Isolated yields.

The protocols developed herein constitute convenient and safe routes to a wide range of sodium arylsulfinates, and is especially efficient for electron- poor aryl bromides which are often difficult to prepare using existing meth- ods. Both electron-poor and electron-rich aryl bromides can successfully be employed in the reaction between aryl Grignard or aryl lithium reagents and DABSO, without the need for handling gaseous SO2. Thus, a wide range of sodium arylsulfinates can be produced, which can then be deployed as sub- strates in reactions such as the palladium(II)-catalysed transformations de- scribed in Papers II-III.

76 Conclusions

Palladium(II) catalysis has developed over the last few decades and has be- come a useful synthetic tool that often complements other palladium(0)- catalysed reactions. In the first part of this thesis, the development of a pal- ladium(II)-catalysed Heck protocol for decarboxylative synthesis of styrenes is described. The reaction mechanism was investigated by means of DFT calculations and the theoretical investigation, which was in agreement with the experimental results, indicated that both the formation of the correspond- ing styrene and 1,1-diarylethene are possible outcomes of the reaction. Based on the computational study a one-pot synthetic protocol could be de- veloped to produce sterically congested di-ortho alkoxy-substituted 1,1- diarylethene. However, the transformation had a very limited substrate scope and high selectivity for the 1,1-diarylated product was only observed for 2,6- dimethoxybenzoic acid. Sodium sulfinates were identified as a suitable alternative to carboxylic acids as arylating agent. A palladium(II)-catalysed protocol for the synthesis of a wide range of aryl ketones by desulfitative addition to nitriles was de- veloped. Importantly, these substrates do not require 2,6-dialkoxy substitu- tion to be productive in the reaction. The desulfitative and decarboxylative processes of benzenesulfinic acid, benzoic acid and 2,6-methoxy-benzoic acid were investigated using DFT calculations. The results gave insights into the mechanisms of these processes and support that palladium(II)-catalysed desulfination of arylsulfinic acids is a viable route to aryl-palladium inter- mediates. Additional experimental support for the proposed reaction mecha- nism was provided by direct ESI-MS studies. The wide scope of the desul- fitative reaction was demonstrated by combining 15 different sodium aryl- sulfinates and 21 nitriles to give 56 examples of aryl ketones. An adapted protocol for continuous flow synthesis of aryl ketones was also developed. Though certain reaction mixtures were biphasic, acceptable to good yields of the desired products were obtained with a production capacity of up to 4.3 mmol/h. Microwave transparent borosilicate glass reactors performed well and aluminium oxide reactors were identified as safe alternatives. A gas-free synthesis of sodium arylsulfinates from Grignard and organo- lithium regents using the SO2 surrogate DABSO was developed. The devel- oped protocol is a convenient and safe route to a wide range of sodium aryl- sulfinates and especially the use of electron-poor substrates resulted in high yields.

77 Decarboxylative processes have great potential and are environmentally benign but current monometallic systems suffer from serious limitations. Sodium arylsulfinates have emerged as alternative and more versatile ary- lating agents in palladium(II) catalysis and are becoming increasingly im- portant. This thesis illustrates the interplay between experimental chemistry and mechanistic insight from theoretical investigations in the development of new useful synthetic tools.

78 Acknowledgements

This thesis is the result of several people’s hard work and would not have been possible without their contributions. Everyone at the department of organic pharmaceutical chemistry has contributed to a fantastic workplace and many have contributed directly or indirectly to the material presented herein. I would like to thank the following people in particular:

Professor Mats Larhed for taking me on as a PhD student, allowing me to work independently but not hesitating to answer research questions in the middle of the night while dining at the castle.

Professor Anders Karlén for supervising me during the medicinal chemistry projects that did not make it into this thesis.

Dr. Luke Odell for providing invaluable practical and theoretical guidance during the different projects and for setting an example for all organic chem- ists.

Dr. Christian Sköld for introducing me to the world of computational chem- istry.

Dr. Jonas Sävmarker for always being a helpful mentor regarding palladium and flow chemistry.

Dr. Ulrika Rosenström for support during my struggles in research, but mainly for all the support and confidence placed in me during teaching.

Jonas Rydfjord for remaining my faithful sidekick since high school, and making contributions to the majority of the projects described in this thesis.

Dr. Shiao Chow for entertaining discussions regarding all fields of science, help with graphical design and particularly all the help regarding this thesis.

Dr. Fredrik Svensson for discussions regarding reaction mechanisms, math- ematics, science and totally unrelated topics.

79 Dr. Ashkan Fardost for all the speculative discussions and having the cour- age to think big.

Dr. Sanjay Borhade who helped me perfect the art of column chromatog- raphy.

Dr. Anneli Nordqvist for taking the time to introduce me to the chemistry and teaching at the department, even during the preparations for her own thesis defence.

Dr. Rajiv Sawant for setting an example for how to work in the lab and for providing encyclopedic knowledge regarding organic chemistry.

Mohammed Al-Tikriti and Mostafa Haitham for their valuable contributions to research projects that were not included in this thesis.

Gunilla Eriksson all the help with small and big things. You always went the extra mile and made the department an amazing workplace.

Sorin Srbu for the swift assistance with IT stuff that went haywire.

Dr. Per Sjöberg for introducing me to ESI-MS and help with the elucidation of mountains of mass spectra.

Dr. Peter Brandt for critical feedback and support with calculations.

Dr. Johan Gising for support during the medicinal chemistry projects and Dr. Charles Hedgecock for always providing useful feedback and the invaluable perspective from the pharmaceutical industry.

Dr. Anja Sandström and Dr. Charlotta Wallinder for help with and discus- sions regarding teaching.

Dr. Marc Stevens for challenging me to keep up to date with the literature and collecting obscure knowledge.

Linda Åkerbladh for struggling even more than I with the synthesis of an- titubercular compounds, but always retaining a positive outlook.

Linda Wikberg for invaluable moral support and for providing a perspective from a non-chemist point of view.

80 Apotekarsocieteten, Apotekare C. D. Carlssons stiftelse and IFs stiftelse for giving me the opportunity to participate at international research confer- ences.

81 References

(1) Wöhler, F. Ueber Künstliche Bildung Des Harnstoffs. Ann. der Phys. und Chemie 1828, 88 (2), 253–256. (2) Ramberg, P. J. The Death of Vitalism and the Birth of Organic Chemistry: Wöhler’s Urea Synthesis and the Disciplinary Identity of Organic Chemistry. Ambix 2000, 47 (3), 170–195. (3) Nicolaou, K. C.; Bulger, P. G.; Sarlah, D. Palladium-Catalyzed Cross-Coupling Reactions in Total Synthesis. Angew. Chemie Int. Ed. 2005, 44 (29), 4442– 4489. (4) Glaser, C. Beiträge Zur Kenntnifs Des Acetenylbenzols. Berichte der Dtsch. Chem. Gesellschaft 1869, 2, 422 – 424. (5) Ullmann, F.; Bielecki, J. Ueber Synthesen in Der Biphenylreihe. Berichte der Dtsch. Chem. Gesellschaft 1901, 34 (2), 2174–2185. (6) Wurtz, A. Sur Une Nouvelle Classe de Radicaux Organiques. Ann. Chim. Phys. 1855, 44 (1), 275 – 312. (7) Wurtz, A. Ueber Eine Neue Klasse Organischer Radicale. Ann. der Chemie und Pharm. 1855, 96 (3), 364–375. (8) Fittig, R. Ueber Einige Derivate Des Phenyls. Ann. der Chemie und Pharm. 1862, 124 (2), 275–289. (9) Grignard, V. Sur Quelques Nouvelles Combinaisons Organométalliques Du Magnèsium et Leur Application À Des Synthèses D’alcools et D’hydrocarbures. Comptes rendus Hebd. des séances l’Académie 1900, 130 (2), 1322–1324. (10) Johansson Seechurn, C. C. C.; Kitching, M. O.; Colacot, T. J.; Snieckus, V. Palladium-Catalyzed Cross-Coupling: A Historical Contextual Perspective to the 2010 Nobel Prize. Angew. Chemie Int. Ed. 2012, 51 (21), 5062–5085. (11) Berzelius, J. Årsberättelsen Om Framsteg I Fysik Och Kemi; Norstedt, 1835. (12) Lindström, B.; Pettersson, L. J. A Brief History of Catalysis. Cattech 2003, 7 (4), 130–138. (13) Taylor, C. C. Autmomobile Catalytic Converters; Springer Berlin Heidelberg, 1984. (14) Haber, F. Thermodynamik Technischer Gasreaktionen; Verlag von R.Oldenbourg, 1905. (15) Smil, V. Detonator of the Population Explosion. Nature 1999, 400, 1999. (16) Erisman, J. W.; Sutton, M. a.; Galloway, J.; Klimont, Z.; Winiwarter, W. How a Century of Ammonia Synthesis Changed the World. Nat. Geosci. 2008, 1 (October 2008), 636–639. (17) Howarth, R. W. Coastal Nitrogen Pollution: A Review of Sources and Trends Globally and Regionally. Harmful Algae 2008, 8 (1), 14–20.

82 (18) Hegedus, L. S. Transition Metals in the Synthesis of Complex Organic Molecules; University Science Books, 1999. (19) Miyaura, N. Cross-Coupling Reactions; Miyaura, N., Ed.; Topics in Current Chemistry; Springer Berlin Heidelberg: Berlin, Heidelberg, 2002; Vol. 219. (20) Negishi, E. Handbook of Organopalladium Chemistry for Organic Synthesis, 2nd ed.; Negishi, E., Ed.; John Wiley & Sons, Inc., 2003. (21) de Meijere, A.; Diederich, F. Metal-Catalyzed Cross-Coupling Reactions, Second Completely Revised and Enlarged Edition; Volume 2; de Meijere, A., Diederich, F., Eds.; Wiley-VCH Verlag GmbH: Weinheim, Germany, 2004. (22) Phan, N. T. S.; Van Der Sluys, M.; Jones, C. W. On the Nature of the Active Species in Palladium Catalyzed Mizoroki-Heck and Suzuki-Miyaura Couplings - Homogeneous or Heterogeneous Catalysis, a Critical Review. Adv. Synth. Catal. 2006, 348 (6), 609–679. (23) Torborg, C.; Beller, M. Recent Applications of Palladium-Catalyzed Coupling Reactions in the Pharmaceutical, Agrochemical, and Fine Chemical Industries. Adv. Synth. Catal. 2009, 351 (18), 3027–3043. (24) Wollaston, W. H. On a New Metal, Found in Crude Platina. Philos. Trans. R. Soc. London 1804, 94, 419–430. (25) Smidt, J.; Hafner, W.; Jira, R. Katalytische Umsetzungen von Olefinen an Platinmetall Verbindungen Das Consortium-Verfahren Zur Herstellung von Acetaldehyd. Angew. Chemie 1959, 71 (5), 176–182. (26) Negishi, E. Fundamental Properties of Palladium and Patterns of the Reactions of Palladium and Its Complexes. In Handbook of Organopalladium Chemistry for Organic Synthesis; John Wiley & Sons, Inc.: New York, USA, 2002; pp 17–35. (27) Kim, J. S.; Lee, H. S.; Jeon, P. J.; Lee, Y. T.; Yoon, W.; Ju, S. Y.; Im, S. Multifunctional Schottky-Diode Circuit Comprising Palladium/molybdenum Disulfide Nanosheet. Small 2014, 10 (23), 4845–4850. (28) Perumal, R.; Cui, Z.; Gille, P.; Harmand, J.-C.; Yoh, K. Palladium Assisted Hetroepitaxial Growth of an InAs Nanowire by Molecular Beam Epitaxy. Semicond. Sci. Technol. 2014, 29 (11), 115005. (29) Arzubiaga, L.; Golmar, F.; Llopis, R.; Casanova, F.; Hueso, L. E. In Situ Electrical Characterization of Palladium-Based Single Electron Transistors Made by Electromigration Technique. AIP Adv. 2014, 4 (117126), 1–7. (30) Carrera-Cerritos, R.; Fuentes-Ramírez, R.; Cuevas-Muñiz, F. M.; Ledesma- García, J.; Arriaga, L. G. Performance and Stability of Pd Nanostructures in an Alkaline Direct Ethanol Fuel Cell. J. Power Sources 2014, 269, 370–378. (31) Hsu, C. H.; Chang, E. Y.; Chang, H. J.; Yu, H. W.; Nguyen, H. Q.; Chung, C. C.; Maa, J. S.; Pande, K. Gold-Free Fully Cu-Metallized InGaP/InGaAs/Ge Triple-Junction Solar Cells. IEEE Electron Device Lett. 2014, 35 (12), 1275– 1277. (32) Pranczk, J.; Jacewicz, D.; Wyrzykowski, D.; Chmurzy, L. Platinum ( II ) and Palladium ( II ) Complex Compounds as Anti-Cancer Drugs . Methods of Cytotoxicity Determination. Curr. Pharm. Anal. 2014, 10, 2–9. (33) Arentz, D.; Herbst, B. Platinum and Palladium Printing, 2nd ed.; Elsevier Focal Press: Amsterdam;Boston, 2005.

83 (34) Zapf, A.; Beller, M. Fine Chemical Synthesis with Homogeneous Palladium Catalysts: Examples, Status and Trends. Top. Catal. 2002, 19 (1), 101–109. (35) Garrett, C. E.; Prasad, K. The Art of Meeting Palladium Specifications in Active Pharmaceutical Ingredients Produced by Pd-Catalyzed Reactions. Adv. Synth. Catal. 2004, 346 (8), 889–900. (36) Watson, D. A. Legacy of Richard Heck. Organometallics 2016, 35 (9), 1177– 1178. (37) Heck, R. F. Allylation of Aromatic Compounds with Organopalladium Salts. J. Am. Chem. Soc. 1968, 90 (20), 5531–5534. (38) Heck, R. Acylation, Methylation, and Carboxyalkylation of Olefins by Group VIII Metal Derivatives. J. Am. Chem. Soc. 1968, 299 (3), 5518–5526. (39) Heck, R. F. The Palladium-Catalyzed Arylation of Enol Esters, Ethers, and Halides. A New Synthesis of 2-Aryl Aldehydes and Ketones. J. Am. Chem. Soc. 1968, 90 (20), 5535–5538. (40) Heck, R. F. A Synthesis of Diaryl Ketones from Arylmercuric Salts. J. Am. Chem. Soc. 1968, 90 (20), 5546–5548. (41) Heck, R. F. The Arylation of Allylic Alcohols with Organopalladium Compounds. A New Synthesis of 3-Aryl Aldehydes and Ketones. J. Am. Chem. Soc. 1968, 90 (20), 5526–5531. (42) Heck, R. F. The Addition of Alkyl- and Arylpalladium Chlorides to Conjugated . J. Am. Chem. Soc. 1968, 90 (20), 5542–5546. (43) Heck, R. F. Aromatic Haloethylation with Palladium and Copper Halides. J. Am. Chem. Soc. 1968, 90 (20), 5538–5542. (44) Fujiwara, Y.; Moritani, I.; Matsuda, M.; Teranishi, S. Aromatic Substitution of Olefin. IV Reaction with Palladium Metal and Silver Acetate. Tetrahedron Lett. 1968, 9 (35), 3863–3865. (45) Mizoroki, T.; Mori, K.; Ozaki, A. Arylation of Olefin with Aryl Iodide Catalyzed by Palladium. Bull. Chem. Soc. Jpn. 1971, 44 (2), 581–581. (46) Heck, R. F.; Nolley, J. P. Palladium-Catalyzed Vinylic Hydrogen Substitution Reactions with Aryl, Benzyl, and Styryl Halides. J. Org. Chem. 1972, 37 (14), 2320–2322. (47) Tamao, K.; Sumitani, K.; Kumada, M. Selective Carbon-Carbon Bond Formation by Cross-Coupling of Grignard Reagents with Organic Halides. Catalysis by Nickel-Phosphine Complexes. J. Am. Chem. Soc. 1972, 94 (12), 4374–4376. (48) Corriu, J. P.; Masse, J. P. Grignard Reagents by Transition-Metal Complexes. A New and Simple Synthesis of Trans-Stilbenes and Polyphenyls. Chem. Commun. 1972, 144. (49) Yamamura, M.; Moritani, I.; Murahashi, S. I. The Reaction of σ- Vinylpalladium Complexes with Alkyllithiums. Stereospecific Syntheses of Olefins from Vinyl Halides and Alkyllithiums. J. Organomet. Chem. 1975, 91 (2), 3–6. (50) Murahashi, S.; Yamamura, M.; Yanagisawa, K.; Mita, N.; Kondo, K. Stereoselective Synthesis of Alkenes and Alkenyl Sulfides from Alkenyl Halides Using Palladium and Ruthenium Catalysts. J. Org. Chem. 1979, 44 (14), 2408–2417.

84 (51) King, A.; Okukado, N.; Negishi, E.-I. Selective Carbon-Carbon Bond Formation via Transition Metal Catalysis. 3.’ A Highly Selective Synthesis of Unsymmetrical Biaryls and Diarylmethanes by the Nickel-O R Palladium- Catalyzed Reaction of Aryl-a N D Benzylzinc Derivatives with Aryl Halides. J. Org. Chem 1977, 42 (10), 1821–1823. (52) Sonogashira, K.; Tohda, Y.; Hagihara, N. A Convenient Synthesis of Acetylenes: Catalytic Substitutions of Acetylenic Hydrogen with Bromoalkenes, Iodoarenes and Bromopyridines. Tetrahedron Lett. 1975, 16 (50), 4467–4470. (53) Kosugi, M.; Shimizu, Y.; Migita, T. Alkylation, Arylation, and Vinylation of Acyl Chlorides by Means of Organotin Compounds in the Presence of Catalytic Amounts of tetrakis(triphenylphosphine)palladium(0). Chem. Lett. 1977, 6 (12), 1423–1424. (54) Milstein, D.; Stille, J. K. A General, Selective, and Facile Method for Ketone Synthesis from Acid Chlorides and Organotin Compounds Catalyzed by Palladium. J. Am. Chem. Soc. 1978, 100 (11), 3636–3638. (55) Miyaura, N.; Yamada, K. A New Stereospecific Cross-Coupling by the Palladium-Catalyzed Reaction of 1-Alkenylboranes with 1-Alkenyl or 1- Alkynyl Halides. Tetrahedron Lett. 1979, 20 (36), 3437–3440. (56) Miyaura, N.; Suzuki, A. Stereoselective Synthesis of Arylated (E)-Alkenes by the Reaction of Alk-1-Enylboranes with Aryl Halides in the Presence of Palladium Catalyst. J. Chem. Soc. Chem. Commun. 1979, No. 19, 866–867. (57) Hatanaka, Y.; Hiyama, T. Cross-Coupling of Organosilanes with Organic Halides Mediated by a Palladium Catalyst and Tris (Diethylamino) Sulfonium Difluorotrimethylsilicate. J. Org. Chem. 1988, 53 (4), 918–920. (58) Brown, D. G.; Boström, J. Analysis of Past and Present Synthetic Methodologies on Medicinal Chemistry: Where Have All the New Reactions Gone? J. Med. Chem. 2016, 59 (10), 4443–4458. (59) Amatore, C.; Pfluger, F. Mechanism of Oxidative Addition of palladium(0) with Aromatic Iodides in Toluene, Monitored at Ultramicroelectrodes. Organometallics 1990, 9 (8), 2276–2282. (60) Fleckenstein, C. a; Plenio, H. Sterically Demanding Trialkylphosphines for Palladium-Catalyzed Cross Coupling Reactions-Alternatives to PtBu3. Chem. Soc. Rev. 2010, 39 (2), 694–711. (61) Cárdenas, D. J. Towards Efficient and Wide-Scope Metal-Catalyzed Alkyl- Alkyl Cross- Coupling Reactions. Angew. Chemie - Int. Ed. 1999, 38 (20), 3018–3020. (62) Firmansjah, L.; Fu, G. C. Intramolecular Heck Reactions of Unactivated Alkyl Halides. J. Am. Chem. Soc. 2007, 129 (37), 11340–11341. (63) Saito, B.; Fu, G. C. Alkyl-Alkyl Suzuki Cross-Couplings of Unactivated Secondary Alkyl Halides at Room Temperature. J. Am. Chem. Soc. 2007, 129 (31), 9602–9603. (64) Saito, B.; Fu, G. C. Enantioselective Alkyl-Alkyl Suzuki Cross-Couplings of Unactivated Homobenzylic Halides. J. Am. Chem. Soc. 2008, 130 (21), 6694– 6695.

85 (65) Achonduh, G. T.; Hadei, N.; Valente, C.; Avola, S.; O’Brien, C. J.; Organ, M. G. On the Role of Additives in Alkyl-Alkyl Negishi Cross-Couplings. Chem. Commun. 2010, 46 (23), 4109–4111. (66) Casado, A. L.; Espinet, P. Mechanism of the . 1. The Transmetalation Step. Coupling of R1I and R2SnBu3 Catalyzed by Trans- [PdR1IL2] (R1 = C6Cl2F3; R2 = Vinyl, 4-Methoxyphenyl; L = AsPh3). J. Am. Chem. Soc. 1998, 120 (35), 8978–8985. (67) Espinet, P.; Echavarren, A. M. The Mechanisms of the Stille Reaction. Angew. Chemie - Int. Ed. 2004, 43 (36), 4704–4734. (68) Miyaura, N.; Suzuki, A. Palladium-Catalyzed Cross-Coupling Reactions. Chem. Rev. 1995, 95 (1), 2457–2483. (69) Shi, W.; Liu, C.; Lei, A. Transition-Metal Catalyzed Oxidative Cross-Coupling Reactions to Form C-C Bonds Involving Organometallic Reagents as Nucleophiles. Chem. Soc. Rev. 2011, 40 (5), 2761–2776. (70) Beletskaya, I. P.; Cheprakov, A. V. Heck Reaction as a Sharpening Stone of Palladium Catalysis. Chem. Rev. 2000, 100 (8), 3009–3066. (71) Whitcombe, N. J.; Kuok, K.; Hii, M.; Gibson, S. E. Advances in the Heck Chemistry of Aryl Bromides and Chlorides. Tetrahedron 2001, 57 (582), 7449–7476. (72) Knowles, J. P.; Whiting, A. The Heck-Mizoroki Cross-Coupling Reaction: A Mechanistic Perspective. Org. Biomol. Chem. 2007, 5 (August 2006), 31–44. (73) Heck, R. F. Electronic and Steric Effects in the Olefin Arylation and Carboalkoxylation Reactions with Organopalladium Compounds. J. Am. Chem. Soc. 1971, 93 (25), 6896–6901. (74) Daves, D.; Hallberg, A. 1,2-Additions to Heteroatom-Substituted Olefins by Organopalladium Reagents. Chem. Rev. 1989, 89 (1), 1433–1445. (75) Ruan, J.; Xiao, J. From α-Arylation of Olefins to Acylation with Aldehydes: A Journey in Regiocontrol of the Heck Reaction. Acc. Chem. Res. 2011, 44 (8), 614–626. (76) Cabri, W.; Candiani, I.; Bedeschi, A.; Santi, R. Palladium-Catalyzed Arylation of Unsymmetrical Olefins. Bidentate Phosphine Ligand Controlled Regioselectivity. J. Org. Chem. 1992, 57 (13), 3558–3563. (77) Cabri, W.; Candiani, I.; Bedeschi, A.; Penco, S.; Santi, R. A-Regioselectivity in Palladium-Catalyzed Arylation. J. Org. Chem. 1992, 57 (5), 1481–1486. (78) Nilsson, K.; Hallberg, A. Regioselective Palladium-Catalyzed Tandem. J. Org. Chem. 1990, 55 (8), 2464–2470. (79) Andersson, C. M.; Hallberg, A.; Daves, G. D. Regiochemistry of Palladium- Catalyzed Arylation Reactions of Enol Ethers. Electronic Control of Selection for .alpha.- or .beta.-Arylation. J. Org. Chem. 1987, 52 (16), 3529–3536. (80) Cabri, W.; Candiani, I. Recent Developments and New Perspectives in the Heck Reaction. Acc. Chem. Res. 1995, 28 (1), 2–7. (81) Nakamura, S.; Yasui, T. Formation of Palladous Acetate and Stability of Catalyst in Palladium-Metal-Catalyzed Synthesis of Vinyl Acetate from Ethylene. J. Catal. 1971, 23 (3), 315–320. (82) Smidt, J.; Haftner, W.; Jira, R.; Sieber, R.; Sedlmeier, J.; Sabel, A. The Oxidation of Olefins with Palladium Chloride Catalysts. Angew. Chemie Int. Ed. 1962, 1 (2), 80–88.

86 (83) Tsuji, J. Palladium Reagents and Catalysts; John Wiley & Sons, Ltd: Chichester, UK, 2004. (84) I. Moritani and Y. Fujiwara. Aromatic Substitution of Styrene-Palladium Chloride Complex. Tetrahedron Lett. 1967, No. 12, 1119–1122. (85) Fujiwara, Y.; Moritani, I.; Matsuda, M.; Teranishi, S. Aromatic Substitution of Styrene-Palladium Chloride Complex. II Effect of Metal Acetate. Tetrahedron Lett. 1968, 9 (5), 633–636. (86) Fujiwara, Y.; Noritani, I.; Danno, S.; Asano, R.; Teranishi, S. Aromatic Substitution of Olefins. VI. Arylation of Olefins with palladium(II) Acetate. J. Am. Chem. Soc. 1969, 91 (25), 7166–7169. (87) Chen, X.; Engle, K. M.; Wang, D. H.; Jin-Quan, Y. Palladium(II)-cataIyzed C- H aetivation/C-C Cross-Coupling Reactions: Versatility and Practicality. Angew. Chemie - Int. Ed. 2009, 48 (28), 5094–5115. (88) Kuhl, N.; Hopkinson, M. N.; Wencel-Delord, J.; Glorius, F. Beyond Directing Groups: Transition-Metal-Catalyzed C-H Activation of Simple Arenes. Angew. Chemie - Int. Ed. 2012, 51 (41), 10236–10254. (89) Myers, A. G.; Tanaka, D.; Mannion, M. R. Development of a Decarboxylative Palladation Reaction and Its Use in a Heck-Type Olefination of Arene Carboxylates. J. Am. Chem. Soc. 2002, 124 (38), 11250–11251. (90) Tanaka, D.; Myers, A. G. Heck-Type Arylation of 2-Cycloalken-1-Ones with Arylpalladium Intermediates Formed by Decarboxylative Palladation and by Aryl Iodide Insertion. Org. Lett. 2004, 6 (3), 433–436. (91) Miao, T.; Wang, G.-W. Synthesis of Ketones by Palladium-Catalysed Desulfitative Reaction of Arylsulfinic Acids with Nitriles. Chem. Commun. 2011, 47 (33), 9501–9503. (92) Liu, J.; Zhou, X.; Rao, H.; Xiao, F.; Li, C.-J.; Deng, G.-J. Direct Synthesis of Aryl Ketones by Palladium-Catalyzed Desulfinative Addition of Sodium Sulfinates to Nitriles. Chem. - A Eur. J. 2011, 17 (29), 7996–7999. (93) Behrends, M.; Sävmarker, J.; Sjöberg, P. J. R.; Larhed, M. Microwave-Assisted Palladium(II)-Catalyzed Synthesis of Aryl Ketones from Aryl Sulfinates and Direct ESI-MS Studies Thereof. ACS Catal. 2011, 1 (11), 1455–1459. (94) Dieck, H. A.; Heck, R. F. Palladium-Catalyzed Conjugated Synthesis from Vinylic Halides and Olefinic Compounds. J. Org. Chem. 1975, 40 (8), 1083–1090. (95) Oda, H.; Morishita, M.; Fugami, K.; Sano, H.; Kosugi, M. A Novel Diarylation Reaction of Alkynes by Using Aryltributylstannane in the Presence of Palladium Catalyst. Chemistry Letters. 1996, pp 811–812. (96) Hirabayashi, K.; Nishihara, Y.; Mori, A.; Hiyama, T. A Novel C-C Bond Forming Reaction of Aryl- and Alkenylsilanols. A Halogen-Free Mizoroki- Heck Type Reaction. Tetrahedron Lett. 1998, 39 (43), 7893–7896. (97) Matoba, K.; Motofusa, S.; Sik Cho, C.; Ohe, K.; Uemura, S. Palladium(II)- Catalyzed Phenylation of Unsaturated Compounds Using Phenylantimony Chlorides under Air. J. Organomet. Chem. 1999, 574 (1), 3–10. (98) Inoue, A.; Shinokubo, H.; Oshima, K. Oxidative Heck-Type Reaction Involving Cleavage of a Carbon-Phosphorus Bond of Arylphosphonic Acids. J. Am. Chem. Soc. 2003, 125 (6), 1484–1485.

87 (99) Jung, Y. C.; Mishra, R. K.; Yoon, C. H.; Jung, K. W. Oxygen-Promoted Pd(II) Catalysis for the Coupling of Organoboron Compounds and Olefins. Org. Lett. 2003, 5 (13), 2231–2234. (100) Andappan, M. M. S.; Nilsson, P.; Larhed, M. The First Ligand-Modulated Oxidative Heck Vinylation. Efficient Catalysis with Molecular Oxygen as Palladium (0) Oxidant. Chem. Commun. 2004, 2, 218–219. (101) Piera, J.; Bäckvall, J.-E. Catalytic Oxidation of Organic Substrates by Molecular Oxygen and Hydrogen Peroxide by Multistep Electron Transfer - A Biomimetic Approach. Angew. Chem. Int. Ed. Engl. 2008, 47 (19), 3506–3523. (102) Bäckvall, J.-E.; Gogoll, A. Palladium-Hydroquinone Catalysed Electrochemical 1,4-Oxidation of Conjugated Dienes. J. Chem. Soc. Chem. Commun. 1987, No. 16, 1236–1238. (103) Bäckvall, J. E.; Gogoll, A. Evidence for (Pi-allyl)palladium(II)(quinone) Complexes in the Palladium-Catalyzed 1,4-Diacetoxylation of Conjugated Dienes. Tetrahedron Lett. 1988, 29 (18), 2243–2246. (104) Amatore, C.; Cammoun, C.; Jutand, A. Electrochemical Recycling of Benzoquinone in the Pd/benzoqui-None-Catalyzed Heck-Type Reactions from Arenes. Adv. Synth. Catal. 2007, 349 (3), 292–296. (105) Antonsson, T.; Moberg, C.; Tottie, L.; Heumann, A. Palladium-Catalyzed Oxidative Cyclization of 1,5-Dienes - Influence of Different Substitution Patterns on the Regiochemistry and Stereochemistry of the Reaction. J. Org. Chem. 1989, 54 (20), 4914–4929. (106) Lei, A.; Zhang, X. Palladium-Catalyzed Homocoupling Reactions between Two Csp(3)- Csp(3) Centers. Org. Lett. 2002, 4 (14), 2285–2288. (107) Zhu, G.; Lu, X. Reactivity and Stereochemistry of Beta-Heteroatom Elimination. A Detailed Study through a Palladium-Catalyzed Cyclization Reaction Model. Organometallics 1995, 14 (10), 4899–4904. (108) Ramnauth, J.; Poulin, O.; Rakhit, S.; Maddaford, S. P. Palladium(II) Acetate Catalyzed Stereoselective C-Glycosidation of Peracetylated Glycals with Arylboronic Acids. Org. Lett. 2001, 3 (13), 2013–2014. (109) Zhang, Z.; Lu, X.; Xu, Z.; Zhang, Q.; Han, X. Role of Halide Ions in Divalent Palladium-Mediated Reactions: Competition between Beta-Heteroatom Elimination and Beta-Hydride Elimination of a Carbon-Palladium Bond. Organometallics 2001, 20 (17), 3724–3728. (110) Boele, M. D. K.; Van Strijdonck, G. P. F.; De Vries, A. H. M.; Kamer, P. C. J.; De Vries, J. G.; Van Leeuwen, P. W. N. M. Selective Pd-Catalyzed Oxidative Coupling of Anilides with Olefins through C-H Bond Activation at Room Temperature. J. Am. Chem. Soc. 2002, 124 (8), 1586–1587. (111) Stahl, S. S. Palladium Oxidase Catalysis: Selective Oxidation of Organic Chemicals by Direct Dioxygen-Coupled Turnover. Angew. Chemie - Int. Ed. 2004, 43 (26), 3400–3420. (112) Enquist, P. A.; Lindh, J.; Nilsson, P.; Larhed, M. Open-Air Oxidative Heck Reactions at Room Temperature. Green Chem. 2006, 8 (4), 338–343. (113) Gligorich, K. M.; Sigman, M. S. Recent Advancements and Challenges of palladium(II)-Catalyzed Oxidation Reactions with Molecular Oxygen as the Sole Oxidant. Chem. Commun. (Camb). 2009, 3854–3867.

88 (114) Molander, G.; Ellis, N. Organotrifluoroborates: Protected Boronic Acids That Expand the Versatility of the Suzuki Coupling Reaction. Acc. Chem. Res. 2007, 40 (4), 275–286. (115) Cho, C. S.; Uemura, S. Palladium-Catalyzed Cross-Coupling of Aryl and Alkenyl Boronic Acids with Alkenes via Oxidative Addition of a Carbonboron Bond to palladium(0). J. Organomet. Chem. 1994, 465 (1–2), 85–92. (116) Du, X.; Suguro, M.; Hirabayashi, K.; Mori, A.; Nishikata, T.; Hagiwara, N.; Kawata, K.; Okeda, T.; Wang, H. F.; Fugami, K.; Kosugi, M. Mizoroki-Heck Type Reaction of Organoboron Reagents with Alkenes and Alkynes. A Pd(II)- Catalyzed Pathway with Cu(OAc)2 as an Oxidant. Org. Lett. 2001, 3 (21), 3313–3316. (117) Parrish, J. P.; Jung, Y. C.; Shin, S. Il; Jung, K. W. Mild and Efficient Aryl- Alkenyl Coupling via Pd(II) Catalysis in the Presence of Oxygen or Cu(II) Oxidants. J. Org. Chem. 2002, 67 (20), 7127–7130. (118) Andappan, M. M. S.; Nilsson, P.; Von Schenck, H.; Larhed, M. Dioxygen- Promoted Regioselective Oxidative Heck Arylations of Electron-Rich Olefins with Arylboronic Acids. J. Org. Chem. 2004, 69 (16), 5212–5218. (119) Yamamoto, T.; Ohta, T.; Ito, Y. Palladium-Catalyzed Addition of Arylboronic Acids to Aldehydes. Org. Lett. 2005, 7 (19), 4153–4155. (120) Cho, C. S.; Motofusa, S.; Ohe, K.; Uemura, S.; Shim, S. C. A New Catalytic Activity of Antimony(III) Chloride in Palladium(0)-Catalyzed Conjugate Addition of Aromatics to .alpha.,.beta.-Unsaturated Ketones and Aldehydes with Sodium Tetraphenylborate and Arylboronic Acids. J. Org. Chem. 1995, 60 (0), 883–888. (121) Gooßen, L. J.; Ghosh, K. Palladium-Catalyzed Synthesis of Aryl Ketones from Boronic Acids and Carboxylic Acids or Anhydrides. Angew. Chemie Int. Ed. 2001, 40 (18), 3458–3460. (122) Gooßen, L. J.; Ghosh, K. Palladium-Catalyzed Synthesis of Aryl Ketones from Boronic Acids and Carboxylic Acids Activated in Situ by Pivalic Anhydride. Eur. J. Org. Chem 2002, No. 19, 3254–3267. (123) Goossen, L. J.; Koley, D.; Hermann, H. L.; Thiel, W. The Palladium-Catalyzed Cross-Coupling Reaction of Carboxylic Anhydrides with Arylboronic Acids: A DFT Study. J. Am. Chem. Soc. 2005, 127 (31), 11102–11114. (124) Goossen, L. J.; Koley, D.; Hermann, H. L.; Thiel, W. Palladium Monophosphine Intermediates in Catalytic Cross-Coupling Reactions: A DFT Study. Organometallics 2006, 25 (1), 54–67. (125) Tsukamoto, H.; Kondo, Y. Palladium(II)-Catalyzed Annulation of Alkynes with Ortho-Ester-Containing Phenylboronic Acids. Org. Lett. 2007, 9 (21), 4227–4230. (126) Zhao, L.; Lu, X. PdII-Catalyzed Cyclization of Alkynes Containing Aldehyde, Ketone, or Nitrile Groups Initiated by the Acetoxypalladation of Alkynes. Angew. Chemie 2002, 114 (22), 4519–4521. (127) Garves, K. Coupling, Carbonylation, and Vinylation Reactions of Aromatic Sulfinic Acids via Organopalladium Intermediates. J. Org. Chem. 1970, 35 (10), 3273–3275. (128) Larock, R. C.; Tian, Q.; Pletnev, A. A. Carbocycle Synthesis via Carbopalladation of Nitriles. J. Am. Chem. Soc. 1999, 121 (13), 3238–3239.

89 (129) Tian, Q.; Pletnev, A. A.; Larock, R. C. Carbopalladation of Nitriles: Synthesis of 3,4-Disubstituted 2-Aminonaphthalenes and 1,3-Benzoxazine Derivatives by the Palladium-Catalyzed Annulation of Alkynes by (2-Iodophenyl)acetonitrile. J. Org. Chem. 2003, 68 (2), 339–347. (130) Pletnev, A. A.; Tian, Q.; Larock, R. C. Carbopalladation of Nitriles: Synthesis of 2,3-Diarylindenones and Polycyclic Aromatic Ketones by the Pd-Catalyzed Annulation of Alkynes and Bicyclic Alkenes by 2-Iodoarenenitriles. J. Org. Chem. 2002, 67 (26), 9276–9287. (131) Zhao, B.; Lu, X. Palladium(II)-Catalyzed Addition of Arylboronic Acid to Nitriles. Tetrahedron Lett. 2006, 47 (38), 6765–6768. (132) Zhou, C.; Larock, R. C. Synthesis of Aryl Ketones by the Pd-Catalyzed C-H Activation of Arenes and Intermolecular Carbopalladation of Nitriles. J. Am. Chem. Soc. 2004, 126 (8), 2302–2303. (133) Zhao, B.; Lu, X. Cationic palladium(II)-Catalyzed Addition of Arylboronic Acids to Nitriles. One-Step Synthesis of Benzofurans from Phenoxyacetonitriles. Org. Lett. 2006, 8 (26), 5987–5990. (134) Ceder, R. M.; Muller, G.; Ordinas, M.; Ordinas, J. I. The Insertion Reaction of Acetonitrile on Aryl Nickel Complexes Stabilized by Bidentate N,N’-chelating Ligands. Dalt. Trans. 2007, No. 1, 83–90. (135) Zhou, C.; Larock, R. C. Synthesis of Aryl Ketones or Ketimines by Palladium- Catalyzed Arene C-H Addition to Nitriles. J. Org. Chem. 2006, 71 (9), 3551– 3558. (136) Rydfjord, J.; Svensson, F.; Trejos, A.; Sjöberg, P. J. R.; Sköld, C.; Sävmarker, J.; Odell, L. R.; Larhed, M. Decarboxylative palladium(II)-Catalyzed Synthesis of Aryl Amidines from Aryl Carboxylic Acids: Development and Mechanistic Investigation. Chem. A Eur. J. 2013, 19 (41), 13803–13810. (137) Straathof, A. J. J. Transformation of Biomass into Commodity Chemicals Using Enzymes or Cells. Chem. Rev. 2014, 114 (3), 1871–1908. (138) Pesci, L. Consutizione Dei Composti Organo-Mercurici Del’acido Benzoico. Atti della R. Accad. dei Lincei 1901, 1, 362–363. (139) Shepard, A. F.; Winslow, N. R.; Johnson, J. R. The Simple Halogen Derivatives of Furan. J. Am. Chem. Soc. 1930, 52 (5), 2083–2090. (140) Hunsdiecker, H.; Hunsdiecker, C.; Vogt, E. Halogen-containing Organic Compounds. US2176181, 1939. (141) Johnson, R. G.; Ingham, R. K. The Degradation Of Carboxylic Acid Salts By Means Of Halogen - The Hunsdiecker Reaction. Chem. Rev. 1956, 56 (2), 219– 269. (142) Nilsson, M.; Kulonen, E.; Sunner, S.; Frank, V.; Brunvoll, J.; Bunnenberg, E.; Djerassi, C.; Records, R. A New Biaryl Synthesis Illustrating a Connection between the Ullmann Biaryl Synthesis and Copper-Catalysed Decarboxylation. Acta Chem. Scand. 1966, 20 (2), 423–426. (143) Gooßen, L. J.; Deng, G.; Levy, L. M. Synthesis of Biaryls via Catalytic Decarboxylative Coupling. Science (80-. ). 2006, 313, 662–664. (144) Goosen, L.; Deng, G.-J. Method For Decarboxylating C-C Cross-Linking Of Carboxylic Acids With Carbon Electrophiles. DE102005022362, 2006.

90 (145) Hu, P.; Kan, J.; Su, W.; Hong, M. Pd(O2CCF3)2/benzoquinone: A Versatile Catalyst System for the Decarboxylative Olefination of Arene Carboxylic Acids. Org. Lett. 2009, 11 (11), 2341–2344. (146) Gooßen, L. J.; Rodríguez, N.; Gooßen, K. Carboxylic Acids as Substrates in Homogeneous Catalysis. Angew. Chemie - Int. Ed. 2008, 47 (17), 3100–3120. (147) Rodríguez, N.; Goossen, L. J. Decarboxylative Coupling Reactions: A Modern Strategy for C-C-Bond Formation. Chem. Soc. Rev. 2011, 40 (10), 5030–5048. (148) Voutchkova, A.; Coplin, A.; Leadbeater, N. E.; Crabtree, R. H. Palladium- Catalyzed Decarboxylative Coupling of Aromatic Acids with Aryl Halides or Unactivated Arenes Using Microwave Heating. Chem. Commun. 2008, No. 47, 6312. (149) Shen, Z.; Ni, Z.; Mo, S.; Wang, J.; Zhu, Y. Palladium-Catalyzed Intramolecular Decarboxylative Coupling of Arene Carboxylic Acids/esters with Aryl Bromides. Chem. - A Eur. J. 2012, 18 (16), 4859–4865. (150) Fu, Z.; Huang, S.; Su, W.; Hong, M. Pd-Catalyzed Decarboxylative Heck Coupling with Dioxygen as the Terminal Oxidant. Org. Lett. 2010, 12 (21), 4992–4995. (151) Zhao, Y.; Zhang, Y.; Wang, J.; Li, H.; Wu, L.; Liu, Z. Synthesis of Aryl- Substituted 1,4-Benzoquinone via palladium(II)-Catalyzed Decarboxylative Coupling of Arene Carboxylate with 1,4-Benzoquinone. Synlett 2010, 2010 (15), 2352–2356. (152) Gooßen, L. J.; Zimmermann, B.; Knauber, T. Pd-Catalyzed Decarboxylative Heck Vinylation of 2-Nitrobenzoates in the Presence of CuF2. Beilstein J. Org. Chem. 2010, 6 (43), 1–9. (153) Xiang, S.; Cai, S.; Zeng, J.; Liu, X. W. Regio- and Stereoselective Synthesis of 2-Deoxy-C-Aryl Glycosides via Palladium Catalyzed Decarboxylative Reactions. Org. Lett. 2011, 13 (17), 4608–4611. (154) Dickstein, J. S.; Mulrooney, C. A.; O’Brien, E. M.; Morgan, B. J.; Kozlowski, M. C. Development of a Catalytic Aromatic Decarboxylation Reaction. Org. Lett. 2007, 9 (13), 2441–2444. (155) Xue, L.; Su, W.; Lin, Z. A DFT Study on the Pd-Mediated Decarboxylation Process of Aryl Carboxylic Acids. Dalt. Trans. 2010, 39 (41), 9815–9822. (156) Svensson, F.; Mane, R. S.; Sävmarker, J.; Larhed, M.; Sköld, C. Theoretical and Experimental Investigation of palladium(II)-Catalyzed Decarboxylative Addition of Arenecarboxylic Acid to Nitrile. Organometallics 2013, 32 (2), 490–497. (157) Becht, J. M.; Le Drian, C. Biaryl Synthesis via Decarboxylative Pd-Catalyzed Reactions of Arenecarboxylic Acids and Diaryliodonium Triflates. Org. Lett. 2008, 10 (14), 3161–3164. (158) Goossen, L. J.; Rodríguez, N.; Linder, C.; Lange, P. P.; Fromm, A. Comparative Study of Copper- and Silver-Catalyzed Protodecarboxylations of Carboxylic Acids. ChemCatChem 2010, 2 (4), 430–442. (159) Goossen, L. J.; Rodríguez, N.; Melzer, B.; Linder, C.; Deng, G.; Levy, L. M. Biaryl Synthesis via Pd-Catalyzed Decarboxylative Coupling of Aromatic Carboxylates with Aryl Halides. J. Am. Chem. Soc. 2007, 129 (15), 4824–4833.

91 (160) Gooßen, L. J.; Zimmermann, B.; Knauber, T. Palladium/copper-Catalyzed Decarboxylative Cross-Coupling of Aryl Chlorides with Potassium Carboxylates. Angew. Chemie - Int. Ed. 2008, 47 (37), 7103–7106. (161) Gooßen, L. J.; Rodríguez, N.; Lange, P. P.; Linder, C. Decarboxylative Cross- Coupling of Aryl Tosylates with Aromatic Carboxylate Salts. Angew. Chemie - Int. Ed. 2010, 49 (6), 1111–1114. (162) Smiles, S.; Rossignol, R. Le. The Sulfination of Phenolic Ethers and the Influence of Substituents. J. Chem. Soc. 1908, 93, 745–762. (163) Krishna, S.; Singh, H. Estimation of—SOOH (Sulfinic) Group and Fe+++. J. Am. Chem. Soc. 1928, 50, 792–798. (164) Grothaus, C.; Dains, F. On the Reactions of Certain Methylene Hydrogen Derivatives Containing Cyanide, Thiocyanate or Sulfinate Radicals. J. Am. Chem. Soc. 1936, 959 (2), 1334–1336. (165) Jacob, C.; Holme, A. L.; Fry, F. H. The Sulfinic Acid Switch in Proteins. Org. Biomol. Chem. 2004, 2 (14), 1953–1956. (166) Kice, J. L.; Bowers, K. W. The Mechanism of the Disproportionation of Sulfinic Acids. J. Am. Chem. Soc. 1962, 84, 605–610. (167) Liu, L.; Chi, Y.; Jen, K. Copper-Catalyzed Additions of Sulfonyl Iodides to Simple and Cyclic Alkenes. J. Org. Chem. 1980, 45 (3), 406–410. (168) Fujiwara, Y.; Dixon, J. a; O’Hara, F.; Funder, E. D.; Dixon, D. D.; Rodriguez, R. a; Baxter, R. D.; Herlé, B.; Sach, N.; Collins, M. R.; Ishihara, Y.; Baran, P. S. Practical and Innate Carbon-Hydrogen Functionalization of Heterocycles. Nature 2012, 492 (7427), 95–99. (169) O’Hara, F.; Baxter, R. D.; O’Brien, A. G.; Collins, M. R.; Dixon, J. a; Fujiwara, Y.; Ishihara, Y.; Baran, P. S. Preparation and Purification of Zinc Sulfinate Reagents for Drug Discovery. Nat. Protoc. 2013, 8 (6), 1042–1047. (170) Pelzer, G.; Keim, W. Palladium-Catalyzed Synthesis of Sulfinic Acids from Aryldiazonium Tetrafluoroborates, Sulfur Dioxide and Hydrogen. J. Mol. Catal. A Chem. 1999, 139, 235–238. (171) Fu, H.; Liu, D.; Meng, L.; Luo, T.; Wei, F.; Wu, Y. Aromatic Sulfinic Acid Compound Preparation Method. CN102731349, 2012. (172) Allen, P. Aliphatic Sulfinic Acids. I. Analysis and Identification. J. Org. Chem. 1942, 7, 23–30. (173) Allen, P.; Rehl, R.; Fuchs, P. Sulfone Formation during Sulfination of the Alkyl Grignard Reagent. J. Org. Chem. 1955, 20, 1237–1239. (174) Umierski, N.; Manolikakes, G. Arylation of Lithium Sulfinates with Diaryliodonium Salts: A Direct and Versatile Access to Arylsulfones. Org. Lett. 2013, 15 (19), 4972–4975. (175) Aziz, J.; Messaoudi, S.; Alami, M.; Hamze, A. Sulfinate Derivatives: Dual and Versatile Partners in Organic Synthesis. Org. Biomol. Chem. 2014, 12 (48), 9743–9759. (176) Oae, S.; Fukushima, D.; Kim, Y. Novel Method of Activating Thiols by Their Conversion into Thionitrites with Dinitrogen Tetroxide. Chem. Commun. 1977, 12, 407–408. (177) Cox, J.; Ghosh, R. A Simple Synthesis of Sulphonyl Cyanides. Tetrahedron Lett. 1969, 39, 3351–3352.

92 (178) Oae, S.; Shinhama, K.; Kim, Y. Oxidation of Sulfinic Acids with Dinitrogen Tetraoxide: Isolation of Sulfonyl Nitrites. Tetrahedron Lett. 1979, 35, 3307– 3308. (179) Oae, S.; Togo, H.; Numata, T.; Fujimori, K. Facile Reduction of Sulfinic Acid to Disulfide with Thiol and Chlorotrimethylsilane. Chem. Lett. 1980, 9, 1193– 1196. (180) Firouzabadi, H.; Karimi, B. Efficient Deoxygenation of Sulfoxides to Thioethers and Reductive Coupling of Sulfonyl Chlorides to Disulfides with Tungsten Hexachloride. Synthesis (Stuttg). 1999, 3, 500–502. (181) Liu, C.-R.; Ding, L.-H. Byproduct Promoted Regioselective Sulfenylation of Indoles with Sulfinic Acids. Org. Biomol. Chem. 2015, 13 (8), 2251–2254. (182) Grayson, E. J.; Ward, S. J.; Hall, A. L.; Rendle, P. M.; Gamblin, D. P.; Batsanov, A. S.; Davis, B. G. Glycosyl Disulfides : Novel Glycosylating Reagents with Flexible Aglycon Alteration. J. Org. Chem. 2005, 70 (24), 9740–9754. (183) Yamamoto, K.; Miyatake, K.; Nishimura, Y.; Tsuchida, E. One-Pot Synthesis of Aryl Sulfoxides and Sulfonium Salts from Sulfinic Acid as a Novel Sulfurizing Agent. Chem. Commun. 1996, 17, 2099–2100. (184) Zhao, Y.; Huang, W.; Zhu, L.; Hu, J. Difluoromethyl 2-Pyridyl Sulfone: A New Gem-Difluoroolefination Reagent for Aldehydes and Ketones. Org. Lett. 2010, 12 (7), 1444–1447. (185) Zhao, J.; Xu, J.; Chen, J.; Wang, X.; He, M. Metal-Free Oxidative Coupling of Amines with Sodium Sulfinates: A Mild Access to Sulfonamides. RSC Adv. 2014, 4 (110), 64698–64701. (186) Nair, V.; Augustine, A.; George, T.; Nair, L. An Efficient One-Pot Synthesis of Vinyl Sulphones via CAN Mediated Reaction of Aryl Sulphinates and Alkenes. Tetrahedron Lett. 2001, 42, 6763–6765. (187) Truce, W. E.; Wolf, G. C. Adducts of Sulfonyl Iodides with Acetylenes. J. Org. Chem. 1971, 36 (13), 1727–1732. (188) Xu, Y.; Tang, X.; Hu, W.; Wu, W.; Jiang, H. Transition-Metal-Free Synthesis of Vinyl Sulfones via Tandem Cross-Decarboxylative/coupling Reactions of Sodium Sulfinates and Cinnamic Acids. Green Chem. 2014, 16 (8), 3720– 3723. (189) Wei, W.; Liu, X.; Yang, D.; Dong, R.; Cui, Y.; Yuan, F.; Wang, H. Direct Difunctionalization of Alkenes with Sulfinic Acids and NBS Leading to β- Bromo Sulfones. Tetrahedron Lett. 2015, 56 (14), 1808–1811. (190) Xu, Y.; Zhao, J.; Tang, X.; Wu, W.; Jiang, H. Chemoselective Synthesis of Unsymmetrical Internal Alkynes or Vinyl Sulfones via Palladium-Catalyzed Cross-Coupling Reaction of Sodium Sulfinates with Alkynes. Adv. Synth. Catal. 2014, 356 (9), 2029–2039. (191) Wang, G.-W.; Miao, T. Palladium-Catalyzed Desulfitative Heck-Type Reaction of Aryl Sulfinic Acids with Alkenes. Chem. - A Eur. J. 2011, 17 (21), 5787–5790. (192) Zhou, X.; Luo, J.; Liu, J.; Peng, S.; Deng, G.-J. Pd-Catalyzed Desulfitative Heck Coupling with Dioxygen as the Terminal Oxidant. Org. Lett. 2011, 13 (6), 1432–1435.

93 (193) Hu, S.; Xia, P.; Cheng, K.; Qi, C. Pd-Catalyzed Ligand-Free Desulfitative Heck Reaction with Arenesulfinic Acid Salts under Air. Appl. Organomet. Chem. 2013, 27 (3), 188–190. (194) Bal Raju, K.; Mari, V.; Nagaiah, K. Regioselective Palladium(II)-Catalyzed Desulfitative Heck-Type Reaction: Access to α-Benzyl-β-Keto Esters from Baylis-Hillman Adducts and Sodium Sulfinates. Synthesis (Stuttg). 2013, 45 (20), 2867–2874. (195) Rao, B.; Zhang, W.; Hu, L.; Luo, M. Catalytic Desulfitative Homocoupling of Sodium Arylsulfinates in Water Using PdCl2 as the Recyclable Catalyst and O2 as the Terminal Oxidant. Green Chem. 2012, 14 (12), 3436–3440. (196) Li, J.; Bi, X.; Wang, H.; Xiao, J. Palladium-Catalyzed Desulfitative C–P Coupling of Arylsulfinate Metal Salts and H-Phosphonates. RSC Adv. 2014, 4 (37), 19214–19217. (197) Wang, H.; Li, Y.; Zhang, R.; Jin, K.; Zhao, D.; Duan, C. Palladium-Catalyzed Desulfitative Conjugate Addition of Aryl Sulfinic Acids and Direct ESI-MS for Mechanistic Studies. J. Org. Chem. 2012, 77 (10), 4849–4853. (198) Chen, W.; Zhou, X.; Xiao, F.; Luo, J.; Deng, G.-J. Palladium-Catalyzed Desulfitative Addition of Sodium Sulfinates with Α,β-Unsaturated Carbonyl Compounds. Tetrahedron Lett. 2012, 53 (33), 4347–4350. (199) Chen, R.; Liu, S.; Liu, X.; Yang, L.; Deng, G.-J. Palladium-Catalyzed Desulfitative C-H Arylation of Azoles with Sodium Sulfinates. Org. Biomol. Chem. 2011, 9 (22), 7675–7679. (200) Wang, M.; Li, D.; Zhou, W.; Wang, L. A Highly Efficient Palladium-Catalyzed Desulfitative Arylation of Azoles with Sodium Arylsulfinates. Tetrahedron 2012, 68 (7), 1926–1930. (201) Wu, M.; Luo, J.; Xiao, F.; Zhang, S.; Deng, G.-J.; Luo, H.-A. Palladium- Catalyzed Direct and Site-Selective Desulfitative Arylation of Indoles with Sodium Sulfinates. Adv. Synth. Catal. 2012, 354 (2–3), 335–340. (202) Miao, T.; Li, P.; Wang, G.-W.; Wang, L. Microwave-Accelerated Pd- Catalyzed Desulfitative Direct C2-Arylation of Free (NH)-Indoles with Arylsulfinic Acids. Chem. - An Asian J. 2013, 8 (12), 3185–3190. (203) Miao, T.; Wang, L. Palladium-Catalyzed Desulfitative Cross-Coupling Reaction of Sodium Arylsulfinates with H-Phosphonate Diesters. Adv. Synth. Catal. 2014, 356 (5), 967–971. (204) O’Connor Sraj, L.; Khairallah, G. N.; da Silva, G.; O’Hair, R. A. J. Who Wins: Pesci, Peters, or Deacon? Intrinsic Reactivity Orders for Organocuprate Formation via Ligand Decomposition. Organometallics 2012, 31 (5), 1801– 1807. (205) Li, H.; Miao, T.; Wang, M.; Li, P.; Wang, L. Recent Advances in Exploring Diverse Decarbonylation, Decarboxylation and Desulfitation Coupling Reactions for Organic Transformations. Synlett 2016, 27 (11), 1635–1648. (206) Ortgies, D. H.; Hassanpour, A.; Chen, F.; Woo, S.; Forgione, P. Desulfination as an Emerging Strategy in Palladium-Catalyzed C-C Coupling Reactions. European J. Org. Chem. 2016, 2016 (3), 408–425. (207) Cramer, C. C. Essentials of Computational Chemistry: Theories and Models; Wiley-VCH Verlag GmbH: Weinheim, Germany, 2004.

94 (208) Schrödinger, E. An Undulatory Theory of the Mechanics of Atoms and Molecules. Phys. Rev. 1926, 28 (6), 1049–1070. (209) Becke, A. D. A New Mixing of Hartree–Fock and Local Density-Functional Theories. J. Chem. Phys. 1993, 98 (2), 1372–1377. (210) Becke, A. D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98 (7), 5648–5652. (211) Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98 (45), 11623–11627. (212) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132 (15), 154104. (213) Lockermann, G. The Centenary of the Bunsen Burner. J. Chem. Educ. 1956, 33 (1), 20. (214) Gedye, R.; Smith, F.; Westaway, K.; Ali, H.; Baldisera, L.; Laberge, L.; Rousell, J. The Use of Microwave Ovens for Rapid Organic Synthesis. Tetrahedron Lett. 1986, 27 (3), 279–282. (215) Giguere, R. J.; Bray, T. L.; Duncan, S. M.; Majetich, G. Application of Commercial Microwave Ovens to Organic Synthesis. Tetrahedron Lett. 1986, 27 (41), 4945–4948. (216) Larhed, M.; Hallberg, A. Microwave-Assisted High-Speed Chemistry: A New Technique in Drug Discovery. Drug Discov. Today 2001, 6 (8), 406–416. (217) Oliver Kappe, C. Microwave Dielectric Heating in Synthetic Organic Chemistry. Chem. Soc. Rev. 2008, 37 (6), 1127–1139. (218) Caddick, S.; Fitzmaurice, R. Microwave Enhanced Synthesis. Tetrahedron 2009, 65 (17), 3325–3355. (219) Gising, J.; Odell, L. R.; Larhed, M. Microwave-Assisted Synthesis of Small Molecules Targeting the Infectious Diseases Tuberculosis, HIV/AIDS, Malaria and Hepatitis C. Org. Biomol. Chem. 2012, 10 (14), 2713. (220) Sadler, S.; Moeller, A. R.; Jones, G. B. Microwave and Continuous Flow Technologies in Drug Discovery. Expert Opin. Drug Discov. 2012, 441 (October), 1–22. (221) Wathey, B.; Tierney, J.; Lidström, P.; Westman, J. The Impact of Microwave- Assisted Organic Chemistry on Drug Discovery. Drug Discov. Today 2002, 7 (6), 373–380. (222) Schanche, J.-S. Microwave Synthesis Solutions from Personal Chemistry. Mol. Divers. 2003, 7 (2–4), 291–298. (223) Kappe, C. O. Controlled Microwave Heating in Modern Organic. Angew. Chemie - Int. Ed. 2004, 43, 6250–6284. (224) Arvela, R. K.; Leadbeater, N. E.; Collins, M. J. Automated Batch Scale-up of Microwave-Promoted Suzuki and Heck Coupling Reactions in Water Using Ultra-Low Metal Catalyst Concentrations. Tetrahedron 2005, 61 (39), 9349– 9355. (225) Lidström, P.; Tierney, J.; Wathey, B.; Westman, J. Microwave Assisted Organic Synthesis—a Review. Tetrahedron 2001, 57 (45), 9225–9283.

95 (226) Mingos, D. M. P.; Baghurst, D. R. Applications of Microwave Dielectric Heating Effects to Synthetic Problems in Chemistry. Chem. Soc. Rev. 1991, 20, 1–47. (227) Gabriel, C.; Gabriel, S.; H. Grant, E.; H. Grant, E.; S. J. Halstead, B.; Michael P. Mingos, D. Dielectric Parameters Relevant to Microwave Dielectric Heating. Chem. Soc. Rev. 1998, 27 (3), 213–224. (228) Anderson, N. G. Practical Use of Continuous Processing in Developing and Scaling up Laboratory Processes. Org. Process Res. Dev. 2001, 5 (6), 613–621. (229) Watts, P.; Haswell, S. J. Continuous Flow Reactors for Drug Discovery. Drug Discov. Today 2003, 8 (13), 586–593. (230) Dye, J. L.; Lok, M. T.; Tehan, F. J.; Ceraso, J. M.; Voorhees, K. J. Flow Synthesis. A Substitute for the High-Dilution Steps in Cryptate Synthesis. J. Org. Chem. 1973, 38 (9), 1773–1775. (231) Atherton, E.; Brown, E.; Sheppard, R. C.; Rosevear, A. A Physically Supported Gel Polymer for Low Pressure, Continuous Flow Solid Phase Reactions. Application to Solid Phase Peptide Synthesis. J. Chem. Soc. Chem. Commun. 1981, No. 21, 1151. (232) Odell, L. R.; Lindh, J.; Gustafsson, T.; Larhed, M. Continuous Flow palladium(II)-Catalyzed Oxidative Heck Reactions with Arylboronic Acids. European J. Org. Chem. 2010, No. 12, 2270–2274. (233) Baumann, M.; Baxendale, I. R.; Ley, S. V; Nikbin, N.; Smith, C. D.; Tierney, J. P. A Modular Flow Reactor for Performing Curtius Rearrangements as a Continuous Flow Process. Org. Biomol. Chem. 2008, 6 (9), 1577. (234) Damm, M.; Glasnov, T. N.; Kappe, C. O. Translating High-Temperature Microwave Chemistry to Scalable Continuous Flow Processes. Org. Process Res. Dev. 2010, 14 (1), 215–224. (235) Ceylan, S.; Friese, C.; Lammel, C.; Mazac, K.; Kirschning, A. Inductive Heating for Organic Synthesis by Using Functionalized Magnetic Nanoparticles inside Microreactors. Angew. Chemie - Int. Ed. 2008, 47 (46), 8950–8953. (236) Ceylan, S.; Coutable, L.; Wegner, J.; Kirschning, A. Inductive Heating with Magnetic Materials inside Flow Reactors. Chem. - A Eur. J. 2011, 17 (6), 1884–1893. (237) Chen, S.-T.; Chiou, S.-H.; Wang, K.-T. Preparative Scale Organic Synthesis Using a Kitchen Microwave Oven. J. Chem. Soc. Chem. Commun. 1990, No. 11, 807–809. (238) He, P.; Haswell, S. J.; Fletcher, P. D. I. Microwave-Assisted Suzuki Reactions in a Continuous Flow Capillary Reactor. Appl. Catal. A Gen. 2004, 274 (1–2), 111–114. (239) Wilson, N. S.; Sarko, C. R.; Roth, G. P. Development and Applications of a Practical Continuous Flow Microwave Cell. Org. Process Res. Dev. 2004, 8 (3), 535–538. (240) Comer, E.; Organ, M. G. A Microreactor for Microwave-Assisted Capillary (Continuous Flow) Organic Synthesis. J. Am. Chem. Soc. 2005, 127 (22), 8160–8167.

96 (241) Bagley, M. C.; Jenkins, R. L.; Lubinu, M. C.; Mason, C.; Wood, R. A Simple Continuous Flow Microwave Reactor. J. Org. Chem. 2005, 70 (17), 7003– 7006. (242) Smith, C. J.; Iglesias-Sigüenza, F. J.; Baxendale, I. R.; Ley, S. V. Flow and Batch Mode Focused Microwave Synthesis of 5-Amino-4-Cyanopyrazoles and Their Further Conversion to 4-Aminopyrazolopyrimidines. Org. Biomol. Chem. 2007, 5 (17), 2758–2761. (243) Barnard, T. M.; Leadbeater, N. E.; Boucher, M. B.; Stencel, L. M.; Wilhite, B. A. Continuous-Flow Preparation of Biodiesel Using Microwave Heating. Energy and Fuels 2007, 21 (3), 1777–1781. (244) R. Baxendale, I.; J. Hayward, J.; V. Ley, S. Microwave Reactions Under Continuous Flow Conditions. Comb. Chem. High Throughput Screen. 2007, 10 (10), 802–836. (245) Cablewski, T.; Faux, F.; Strauss, C. R. Development and Application of a Continuous Microwave Reactor for Organic-Synthesis. J. Org. Chem. 1994, 59 (12), 3408–3412. (246) Wiles, C.; Watts, P. Recent Advances in Synthetic Micro Reaction Technology. Chem. Comm. 2011, 47, 6512–6535. (247) Fenn, J. B.; Mann, M.; Meng, C. K. A. I.; Wong, S. F.; Whitehouse, C. M. Electrospray Ionization for Mass Spectrometry of Large Biomolecules. Science (80-. ). 1989, 246, 64–71. (248) Laughlin, S.; David Wilson, W. May the Best Molecule Win: Competition ESI Mass Spectrometry. Int. J. Mol. Sci. 2015, 16 (10), 24506–24531. (249) Vikse, K. L.; Ahmadi, Z.; Scott McIndoe, J. The Application of Electrospray Ionization Mass Spectrometry to Homogeneous Catalysis. Coord. Chem. Rev. 2014, No. http://dx.doi.org/10.1016/j.ccr.2014.06.012. (250) Enquist, P.-A.; Nilsson, P.; Sjöberg, P.; Larhed, M. ESI-MS Detection of Proposed Reaction Intermediates in the Air-Promoted and Ligand-Modulated Oxidative Heck Reaction. J. Org. Chem. 2006, 71 (23), 8779–8786. (251) Svennebring, A.; Sjöberg, P. J. R.; Larhed, M.; Nilsson, P. A Mechanistic Study on Modern Palladium Catalyst Precursors as New Gateways to Pd(0) in Cationic Heck Reactions. Tetrahedron 2008, 64 (8), 1808–1812. (252) Brown, J. M.; Hii, K. K. Characterization of Reactive Intermediates in Palladium-Catalyzed Arylation of Methyl Acrylate (Heck Reaction). Angew. Chemie Int. Ed. 1996, 35 (6), 657–659. (253) Ripa, L.; Hallberg, A. Controlled Double-Bond Migration in Palladium- Catalyzed Intramolecular Arylation of Enamidines. J. Org. Chem. 1996, 61 (20), 7147–7155. (254) Sabino, A. A.; Machado, A. H. L.; Correia, C. R. D.; Eberlin, M. N. Probing the Mechanism of the Heck Reaction with Arene Diazonium Salts by Electrospray Mass and Tandem Mass Spectrometry. Angew. Chemie - Int. Ed. 2004, 43 (19), 2514–2518. (255) Aliprantis, A. O.; Canary, J. W. Observation of Catalytic Intermediates in the by Electrospray Mass Spectrometry. J. Am. Chem. Soc. 1994, 116 (15), 6985–6986.

97 (256) Oliveira, F. F. D.; Dos Santos, M. R.; Lalli, P. M.; Schmidt, E. M.; Bakuzis, P.; Lapis, A. A. M.; Monteiro, A. L.; Eberlin, M. N.; Neto, B. A. D. Charge- Tagged Acetate Ligands as Mass Spectrometry Probes for Metal Complexes Investigations: Applications in Suzuki and Heck Phosphine-Free Reactions. J. Org. Chem. 2011, 76 (24), 10140–10147. (257) Zhang, S.-L.; Fu, Y.; Shang, R.; Guo, Q.-X.; Liu, L. Theoretical Analysis of Factors Controlling Pd-Catalyzed Decarboxylative Coupling of Carboxylic Acids with Olefins. J. Am. Chem. Soc. 2010, 132 (2), 638–646. (258) Meima, G. R.; Menon, P. G. Catalyst Deactivation Phenomena in Styrene Production. Appl. Catal. A Gen. 2001, 212 (1–2), 239–245. (259) Wagner, T.-J. Thermische Und Photochemische Additionen von Dienophilen an Arene Sowie Deren Vinyloge Und Hetero-Analoge; II. Synthesis (Stuttg). 1980, 10, 769–798. (260) Alajlouni, A. M.; Espenson, J. H. Epoxidation of Styrenes By Hydrogen- Peroxide As Catalyzed By Methylrhenium Trioxide. J. Am. Chem. Soc. 1995, 117 (36), 9243–9250. (261) Chatterjee, A. K.; Choi, T. L.; Sanders, D. P.; Grubbs, R. H. A General Model for Selectivity in Olefin Cross Metathesis. J. Am. Chem. Soc. 2003, 125 (37), 11360–11370. (262) Chatterjee, T.; Dey, R.; Ranu, B. C. An Easy Access to Styrenes: Trans Aryl 1,3-, 1,4- and 1,5-Dienes, and 1,3,5-Trienes by Hiyama Cross-Coupling Catalyzed by Palladium Nanoparticles. New J. Chem. 2011, 35 (5), 1103. (263) Jin, M. J.; Lee, D. H. A Practical Heterogeneous Catalyst for the Suzuki, Sonogashira, and Stille Coupling Reactions of Unreactive Aryl Chlorides. Angew. Chemie Int. Ed. 2010, 49 (6), 1119–1122. (264) Lee, D.-H.; Taher, A.; Ahn, W.-S.; Jin, M.-J. Room Temperature Stille Cross- Coupling Reaction of Unreactive Aryl Chlorides and Heteroaryl Chlorides. Chem. Commun. 2010, 46 (3), 478–480. (265) Lu, G.; Cai, C.; Lipshutz, B. H. Stille Couplings in Water at Room Temperature. Green Chem. 2013, 15, 105–109. (266) Arai, I.; Daves, G. D. 5-Vinylpyrimidines. Synthesis via Organopalladium Intermediates. J. Heterocycl. Chem. 1978, 15 (2), 351–352. (267) Arai, I.; Daves, G. D. Palladium-Catalyzed Phenylation of Enol Ethers and Acetates. J. Org. Chem. 1979, 44 (1), 21–23. (268) Amatore, M.; Gosmini, C.; Périchon, J. Cobalt-Catalyzed Vinylation of Functionalized Aryl Halides with Vinyl Acetates. European J. Org. Chem. 2005, 2005 (6), 989–992. (269) Kormos, C. M.; Leadbeater, N. E. Preparation of Nonsymmetrically Substituted Stilbenes in a One-Pot Two-Step Heck Strategy Using Ethene as a Reagent. J. Org. Chem. 2008, 73 (10), 3854–3858. (270) Zheng, C.; Stahl, S. Regioselective Aerobic Oxidative Heck Reactions with Electronically Unbiased Alkenes: Efficient Access to Alpha-Alkyl Vinylarenes. Chem. Commun. 2015, 51, 12771–12774. (271) Lando, V. R.; Monteiro, A. L. Simple and Efficient Protocol for the Synthesis of Functionalized Styrenes from 1,2-Dibromoethane and Arylboronic Acids. Org. Lett. 2003, 5 (16), 2891–2894.

98 (272) Gøgsig, T. M.; Søbjerg, L. S.; Lindhardt, A. T.; Jensen, K. L.; Skrydstrup, T. Direct Vinylation and Difluorovinylation of Arylboronic Acids Using Vinyl- and 2 , 2-Difluorovinyl Tosylates via the Suzuki - Miyaura Cross Coupling. J. Org. Chem. 2008, 73 (9), 3404–3410. (273) Lindh, J.; Sävmarker, J.; Nilsson, P.; Sjöberg, P. J. R.; Larhed, M. Synthesis of Styrenes by Palladium(II)-Catalyzed Vinylation of Arylboronic Acids and Aryltrifluoroborates by Using Vinyl Acetate. Chem. A Eur. J. 2009, 15 (18), 4630–4636. (274) Fardost, A.; Lindh, J.; Sjöberg, P. J. R.; Larhed, M. Palladium(II)-Catalyzed Decarboxylative Heck Arylations of Acyclic Electron-Rich Olefins with Internal Selectivity. Adv. Synth. Catal. 2014, 356 (4), 870–878. (275) Bäcktorp, C.; Norrby, P.-O. A DFT Comparison of the Neutral and Cationic Heck Pathways. Dalt. Trans. 2011, 40 (42), 11308. (276) Fristrup, P.; Quement, S. L.; Tanner, D.; Norrby, P. O. Reactivity and Regioselectivity in the Heck Reaction: Hammett Study of 4-Substituted Styrenes. Organometallics 2004, 23 (26), 6160–6165. (277) Sköld, C.; Kleimark, J.; Trejos, A.; Odell, L. R.; Nilsson Lill, S. O.; Norrby, P. O.; Larhed, M. Transmetallation versus β-Hydride Elimination: The Role of 1,4-Benzoquinone in Chelation-Controlled Arylation Reactions with Arylboronic Acids. Chem. A Eur. J. 2012, 18 (15), 4714–4722. (278) Bradsher, C. K.; Webster, S. T. A New Base-Catalyzed Cyclization Reaction. J. Am. Chem. Soc. 1957, 3342 (1949), 6–8. (279) Lindh, J.; Sjöberg, P. J. R.; Larhed, M. Synthesis of Aryl Ketones by palladium(II)-Catalyzed Decarboxylative Addition of Benzoic Acids to Nitriles. Angew. Chemie 2010, 49 (42), 7733–7737. (280) Chen, W.; Li, P.; Miao, T.; Meng, L.-G.; Wang, L. An Efficient Tandem Elimination-Cyclization-Desulfitative Arylation of 2-(Gem- Dibromovinyl)phenols(thiophenols) with Sodium Arylsulfinates. Org. Biomol. Chem. 2013, 11 (3), 420–424. (281) Öhrngren, P.; Fardost, A.; Russo, F.; Schanche, J. S.; Fagrell, M.; Larhed, M. Evaluation of a Nonresonant Microwave Applicator for Continuous-Flow Chemistry Applications. Org. Process Res. Dev. 2012, 16 (5), 1053–1063. (282) Fardost, A.; Russo, F.; Larhed, M. A Non-Resonant Microwave Applicator Fully Dedicated to Continuous Flow Chemistry. Chim. Oggi 2012, 30, 14–16. (283) Obermayer, D.; Gutmann, B.; Oliver Kappe, C. Microwave Chemistry in Silicon Carbide Reaction Vials: Separating Thermal from Nonthermal Effects. Angew. Chemie - Int. Ed. 2009, 48 (44), 8321–8324. (284) Rydfjord, J.; Svensson, F.; Fagrell, M.; Sävmarker, J.; Thulin, M.; Larhed, M. Temperature Measurements with Two Different IR Sensors in a Continuous- Flow Microwave Heated System. Beilstein J. Org. Chem. 2013, 9, 2079–2087. (285) Woolven, H.; González-Rodríguez, C.; Marco, I.; Thompson, A. L.; Willis, M. C. DABCO-Bis (Sulfur Dioxide), DABSO, as a Convenient Source of Sulfur Dioxide for Organic Synthesis: Utility in Sulfonamide and Sulfamide Preparation. Org. Lett. 2011, 13 (7), 4876–4878. (286) Deeming, A. S.; Russell, C. J.; Hennessy, A. J.; Willis, M. C. DABSO-Based, Three-Component, One-Pot Sulfone Synthesis. Org. Lett. 2013, 16 (1), 150– 153.

99 (287) Deeming, A. S.; Russell, C. J.; Willis, M. C. Combining Organometallic Reagents, the Sulfur Dioxide Surrogate DABSO, and Amines: A One-Pot Preparation of Sulfonamides, Amenable to Array Synthesis. Angew. Chemie Int. Ed. 2015, 54 (4), 1168–1171. (288) Rocke, B. N.; Bahnck, K. B.; Herr, M.; Lavergne, S.; Mascitti, V.; Perreault, C.; Polivkova, J.; Shavnya, A. Synthesis of Sulfones from Organozinc Reagents, DABSO, and Alkyl Halides. Org. Lett. 2013, 16 (1), 154–157. (289) Waldmann, C.; Schober, O.; Haufe, G.; Kopka, K. A Closer Look at the Bromine-Lithium Exchange with Tert-Butyllithium in an Aryl Sulfonamide Synthesis. Org. Lett. 2013, 15 (12), 2954–2957. (290) Richards-Taylor, C. S.; Blakemore, D. C.; Willis, M. C. One-Pot Three- Component Sulfone Synthesis Exploiting Palladium-Catalysed Aryl Halide Aminosulfonylation. Chem. Sci. 2014, 5 (1), 222–228. (291) Chen, C. C.; Waser, J. One-Pot, Three-Component Arylalkynyl Sulfone Synthesis. Org. Lett. 2015, 17 (3), 736–739. (292) Kice, J. L.; Guaraldi, G.; Venier, C. G. The Mechanism of the Disproportionation of Sulfinic Acids. Rate and Equilibrium Constants for the Sulfinic Acid-Sulfinyl Sulfone (Sulfinic Anhydride) Equilibrium. J. Org. Chem. 1966, 31 (11), 3561–3567. (293) Krasovskiy, A.; Knochel, P. A LiCl-Mediated Br/Mg Exchange Reaction for the Preparation of Functionalized Aryl- and Heteroarylmagnesium Compounds from Organic Bromides. Angew. Chemie Int. Ed. 2004, 43 (25), 3333–3336. (294) Gold, H.; Larhed, M.; Nilsson, P. Microwave Irradiation as a High-Speed Tool for Activation of Sluggish Aryl Chlorides in Grignard Reactions. Synlett 2005, No. 10, 1596–1600.

100

Acta Universitatis Upsaliensis Digital Comprehensive Summaries of Uppsala Dissertations from the Faculty of Pharmacy 221 Editor: The Dean of the Faculty of Pharmacy

A doctoral dissertation from the Faculty of Pharmacy, Uppsala University, is usually a summary of a number of papers. A few copies of the complete dissertation are kept at major Swedish research libraries, while the summary alone is distributed internationally through the series Digital Comprehensive Summaries of Uppsala Dissertations from the Faculty of Pharmacy. (Prior to January, 2005, the series was published under the title “Comprehensive Summaries of Uppsala Dissertations from the Faculty of Pharmacy”.)

ACTA UNIVERSITATIS UPSALIENSIS Distribution: publications.uu.se UPPSALA urn:nbn:se:uu:diva-304746 2016