Skip to Main Content
Skip Nav Destination

Palladium-catalyzed cross-coupling reactions have emerged as an exceptionally powerful class of reactions for the creation of carbon–carbon and carbon–heteroatom bonds in both the academic and industrial sectors of research. This chapter provides a brief history, relevant background information and a preface to the important topics discussed in the remainder of the book.

Transition metal-catalyzed reactions play a vital role in the production of many industrially important chemicals, where homogeneous catalysis (reactions that take place in the same phase as the catalyst) is rapidly growing, as evidenced by the awarding of three distinct Nobel Prizes in Chemistry during the last decade – chiral catalysis (2001; Noyori, Sharpless and Knowles),1  olefin metathesis (2005; Grubbs, Chauvin and Schrock)2  and cross-coupling [2010; Heck (Figure 1.1), Suzuki (Figure 1.2) and Negishi (Figure 1.3)].3  The field of cross-coupling, well dominated by homogeneous catalysis, has undoubtedly turned into an area appreciated by all synthetic chemists, irrespective of their prominence in academia or industry.

Figure 1.1

Prof. Heck giving a lecture at Queens University, Canada, in 2006 using a transparency projector. Courtesy of Prof. Snieckus.

Figure 1.1

Prof. Heck giving a lecture at Queens University, Canada, in 2006 using a transparency projector. Courtesy of Prof. Snieckus.

Close modal
Figure 1.2

Prof. Suzuki (right) with the Editor, Dr Colacot (left), and Prof. Fu (middle). Courtesy of Prof. Dixneuf.

Figure 1.2

Prof. Suzuki (right) with the Editor, Dr Colacot (left), and Prof. Fu (middle). Courtesy of Prof. Dixneuf.

Close modal
Figure 1.3

Prof. Negishi at Stockholm during the 2010 Nobel Prize ceremony. Courtesy of Prof. Negishi.

Figure 1.3

Prof. Negishi at Stockholm during the 2010 Nobel Prize ceremony. Courtesy of Prof. Negishi.

Close modal

Recently, heterogeneous catalysis (reactions that take place in a different phase than the catalyst) has also been used for simple cross-coupling reactions, relying on metal leaching to mediate the desired reaction. However, the leached metal must subsequently be readsorbed in order not to contaminate the final product (“release-and-catch” strategy).4  This is not always ideal, depending on the target use of the product and from a reproducibility point of view. In addition, with reactions catalyzed heterogeneously, it is difficult to carry out reactions with high selectivity, in terms of stereo-, regio- or, in some cases, chemoselectivity. Pd-catalyzed cross-coupling has enriched the area of homogeneous catalysis, where rapid growth has been taking place in the past several years, as evidenced by the growing total number of publications/patents1 in the area (Table 1.1). Thus, Pd-catalyzed cross-coupling reactions comprise one of the most important classes of synthetic transformations in modern organic chemistry, providing chemists with an exceptionally powerful tool for the construction of carbon–carbon (C–C) and carbon–heteroatom (C–X) bonds. These and many related transformations have become ubiquitous in both industry and academia. Indeed, as mentioned above, the 2010 Nobel Prize in Chemistry was a monumental accomplishment for the assiduous contributions of Professors Richard F. Heck (University of Delaware), Akira Suzuki (University of Hokkaidō) and Ei-ichi Negishi (Purdue University) for their achievements within the area of Pd-catalyzed C–C bond-forming reactions.5  According to Negishi, the roots of cross-coupling can be traced all the way back to Victor Grignard (Figure 1.4). These early studies laid the foundation of what would become one of the most important and most studied classes of catalytic reactions. Intense research efforts would soon spawn several new C–C coupling reactions in addition to C–X coupling reactions as the chemistry evolved into what it has become today.

Table 1.1

Growth in the total number of publications and patents on cross-coupling reactions through April 2014.

 Pre-1990Through 2000Through 2010Through April 2014
Suzuki–Miyaura 20 824 10175 15883 
Heck 56 768 4029 5816 
Sonogashira 156 3623 5689 
Stille 470 2380 3537 
Negishi 13 429 737 
Buchwald–Hartwig 253 498 
Kumada–Corriu 136 298 
Hiyama 91 172 
Carbonyl α-arylation 13 33 113 193 
 Pre-1990Through 2000Through 2010Through April 2014
Suzuki–Miyaura 20 824 10175 15883 
Heck 56 768 4029 5816 
Sonogashira 156 3623 5689 
Stille 470 2380 3537 
Negishi 13 429 737 
Buchwald–Hartwig 253 498 
Kumada–Corriu 136 298 
Hiyama 91 172 
Carbonyl α-arylation 13 33 113 193 
Figure 1.4

‘It all started with Grignard’ – Negishi. Photograph courtesy of the Nobel Foundation.

Figure 1.4

‘It all started with Grignard’ – Negishi. Photograph courtesy of the Nobel Foundation.

Close modal

The aim of this book is to serve both academia and industry. In the following sections, some key parameters and basic concepts are introduced.

The turnover number is defined as the absolute number of passes through the catalytic cycle before the catalyst becomes deactivated. In general, industrial chemists are interested in both TON and turnover frequency (TOF) (see the next section). A large TON (e.g., 106–1010) indicates a stable, very long-lived catalyst. The TON can be calculated by dividing the amount of reactant (moles) by the amount of catalyst (moles):

graphic
This assumes a yield of the product of 100%, which is most often not the case. To calculate the true number of turnovers, the yield obtained needs to be taken into account. For example, if 10 mol of reactant and 2.5 mol of catalyst are used, then the TON becomes
graphic
If the yield of the product is 94%, then the actual number of turnovers is

Authors often report mole % of catalyst used. This refers to the fraction of catalyst used relative to the amount of limiting reactant present.

Turnover frequency is defined as the number of passes through the catalytic cycle per unit time (typically seconds, minutes or hours). This number is usually determined by dividing the TON by the time required to produce the given amount of product.

However, as with the TON, the actual yield of the product also needs to be taken into account. Continuing the example above, if the reaction in question was run for 7 h to obtain the 94% yield, the TOF is

graphic

The generally accepted simplified catalytic cycle for cross-coupling reactions is shown in Scheme 1.1, where LnPd(0), the active catalytic species, acts as a “matchmaker”. In Japanese language, “catalyst” is pronounced shoku bai and in Chinese it is chu mei (the same character as for matchmaker).6 

Scheme 1.1

General mechanism of cross-coupling reactions and the Heck–Mizoroki reaction.

Scheme 1.1

General mechanism of cross-coupling reactions and the Heck–Mizoroki reaction.

Close modal

In C–C bond-forming cross-coupling, there are two coupling partners: an aryl/vinyl halide or pseudohalide and an organometallic reagent such as a Grignard reagent. There are three basic steps in the catalytic cycle: oxidative addition, transmetallation and reductive elimination.

Here is an analogy: one of the partners with a family member or friend (R–X) establishes a connection with the matchmaker [LnPd(0)] with the profile of “R”. This is called the oxidative addition of an organic halide/pseudohalide, R–X, to LnPd(0) to generate an R–Pd(Ar)(X)(II) intermediate. In the second step, the other partner (R1) in the form of R1–M also forms a connection with the matchmaker so that R and R1 can communicate with each other through the Pd. This is the second step, called transmetallation, where M (a friend or family member of R1) forms a “bond” with X. In the third step, R and R1 are united and detach from the matchmaker (Pd catalyst) in an event called reductive elimination. The success of a matchmaker depends on how many challenging coupling partners are successfully coupled (get married) without any deleterious incidents, within a short time frame. This is related to the TON and TOF of the catalyst. Although Heck shared the 2010 Nobel Prize for Pd-catalyzed cross-coupling reactions with Suzuki and Negishi, some argue that the Heck–Mizoroki reaction (often shortened to the Heck reaction) is not a true cross-coupling reaction as it does not involve a transmetallation step. In the Heck reaction, the Pd(II)–R species undergoes a migratory insertion with the alkene substrate, followed by a syn-periplanar β-hydride elimination event to give the product. This step was well established by the work of Fu and Hartwig.7  Base is necessary to turn over palladium catalyst by inducing the reductive elimination of HX in the last step. Depending on the nature of substituents on the olefin, linear or branched coupled products are obtained, as these olefin substituents can influence the regioselectivity of the product. The general rule of thumb is that electron-withdrawing groups on the olefin favor linear products with neutral Pd complexes. Bidentate ligands such as dppf [1,1′-bis(diphenylphosphino)ferrocene] under cationic conditions8  and dnpf [1,1′-bis(dinaphthylphosphino)ferrocene] in presence of a polar solvent and TBAC (tetrabutylammonium chloride) additive9  produce branched products for electron-rich and electron-neutral olefins.

Since the original discoveries of cross-coupling reactions, there has been a great deal of effort in this area to better understand the reaction mechanism, where the role of the ligand is important. The electronic and steric nature of the ligand (L) and the coordination number of Pd can significantly influence two important steps of the cycle; oxidative addition and reductive elimination (Figure 1.5).10  The role of ligands in the transmetallation step is not as well understood; however, the groups of Hartwig, Amatore and Lloyd-Jones have carried out some impressive work in the area of Suzuki–Miyaura coupling.11  The groups of Beletskaya12  and Buchwald13  have shown that more electron-deficient ligands can increase the rate of C–N cross-coupling reactions involving ureas and amides, respectively, likely reflecting an increased rate of “transmetallation” (amide binding).13  Oxidative addition was considered to be the rate-limiting step, where the choice of the ligand is important. For example, it is proposed that electron-rich ligands make the Pd basic enough to do the oxidative addition of challenging aryl chlorides, while with aryl iodides and bromides oxidative addition is relatively facile, even with less electron-rich ligands such as Ph3P. Figure 1.5 shows the valence bond (VB) representations for the two components L2Pd and Ar–X and for a concerted, three-centered transition state of the oxidative addition process.14  The energy (ΔG) required to excite one electron into the antibonding (σ*) orbital of the Ar–X bond decreases in the series Ar–Cl>Ar–Br>Ar–I.

Figure 1.5

Valence bond (VB) representations for the two components L2Pd and Ar–X and for a concerted transition state of the oxidative addition process. Reproduced from Ref. 14.

Figure 1.5

Valence bond (VB) representations for the two components L2Pd and Ar–X and for a concerted transition state of the oxidative addition process. Reproduced from Ref. 14.

Close modal

The low reactivity of more challenging substrates such as unactivated aryl chlorides was often attributed to the large bond dissociation energy of the C–Cl bond (95 kcal mol−1) in comparison with Ar–Br (79 kcal mol−1) or Ar–I (64 kcal mol−1),15  which highlights the difficulty for an aryl chloride to add oxidatively to a less electron-rich LnPd(0) species.16 

Interestingly, in the transmetallation step, recent evidence suggests that the trend is the opposite: chloride complexes are transmetallated faster than those of bromides and iodides.13,17  The size of the ligand also plays an important role in the reductive elimination,10  in addition to stabilizing the coordinatively unsaturated LnPd(0).

The intent of this chapter is not to provide an exhaustive review of the history of cross-coupling reactions,18  but to identify the most notable milestones (Figure 1.6) and the genesis of some of the topics of the chapters presented here.

Figure 1.6

Simplified time-line of events in the early development of cross-coupling reactions.

Figure 1.6

Simplified time-line of events in the early development of cross-coupling reactions.

Close modal

Some argue that the history of the use of metals as catalysts to accomplish organic transformations was initiated by Fittig, who recorded sodium-mediated alkylations of halogenated arenes in 1862.19  In the early 1900s, Ullmann and Goldberg carried out extensive studies on copper-catalyzed C–C, C–O and C–N bond-forming reactions.20  Noteworthy is that the first person to combine successfully organometallic reagents with catalysis, in this case NiCl2, was the French chemist André Job.21  He reported that PhMgBr, in the presence of NiCl2, was able to absorb CO, NO, C2H4, C2H2 and H2. Since Job's underappreciated revolutionary discovery, nickel has been overshadowed by palladium in similar transformations. Since this early discovery, carbonylation has become an industrially important process and its modern version, carbonylative cross-coupling, is reviewed in detail by Xiao-Feng Wu and Christopher Barnard in Chapter 10.

Following Job's discoveries, the next notable milestone would be the reports by Kharasch on the metal-catalyzed homo-couplings of organomagnesium reagents.22  More specifically, he employed catalytic amounts of CoCl2, MnCl2, FeCl3 or NiCl2 in the presence of Grignard reagents and organic halides to affect this homo-coupling reaction (Scheme 1.2, top).

Scheme 1.2

Cobalt-catalyzed homo- and cross-coupling of Grignard reagents.

Scheme 1.2

Cobalt-catalyzed homo- and cross-coupling of Grignard reagents.

Close modal

The use of vinyl bromide in place of bromobenzene, under the same conditions, resulted not in the expected homo-coupling of the Grignard reagent, but in the first-ever reported catalytic cross-coupling reaction (Scheme 1.2, bottom).23  These findings, to some extent, make Kharasch (Figure 1.7) the “grandfather” of cross-coupling reactions.

Figure 1.7

Morris Kharasch. “The grandfather of metal-catalyzed cross-coupling reactions.” Photograph courtesy of http://www.ashoftruth.org/history.

Figure 1.7

Morris Kharasch. “The grandfather of metal-catalyzed cross-coupling reactions.” Photograph courtesy of http://www.ashoftruth.org/history.

Close modal

More than 20 years later came the next breakthrough in the Pd-catalyzed cross-coupling area. Heck reported in 1968 that arylations of alkenes could be achieved by using an organomercury arylating reagent and a palladium catalyst (Scheme 1.3).24 

Scheme 1.3

Palladium-catalyzed arylation of alkenes using an organomercury reagent.

Scheme 1.3

Palladium-catalyzed arylation of alkenes using an organomercury reagent.

Close modal

A modification of this Pd-catalyzed reaction was subsequently published by Moritani and Fujiwara. They disclosed the direct coupling between arenes and alkenes, first in the presence of stoichiometric amounts of palladium compounds25  and later using catalytic amounts (Scheme 1.4).26  This finding can be classified as one of the first direct C–C bond formations via C–H activation chemistry.

Scheme 1.4

Palladium-catalyzed direct coupling of arenes with alkenes.

Scheme 1.4

Palladium-catalyzed direct coupling of arenes with alkenes.

Close modal

Most of the early developments involved the use of prefunctionalized coupling partners in terms of organometallic reagents as nucleophiles and aryl halides as electrophiles. An alternative attractive approach would be (as Fujiwara and Moritani showed) the direct functionalization of arene C–H bonds, without the need for prefunctionalization. In addition to Fujiwara and Moritani's disclosure, a few examples of C–H activation were reported in the 1980s by Ames27  (intramolecular) and Ohta28  (intermolecular). During the past two decades, the development of palladium-catalyzed direct arylations has progressed enormously and these advances are discussed by Upendra Sharma, Atanu Modak, Soham Maity, Arun Maji and Debabrata Maiti in Chapter 12.

Building on Kharasch's cobalt-catalyzed cross-coupling reaction, Kochi accomplished an iron-catalyzed reaction between C(sp2)–Br electrophiles and Grignard reagents (Scheme 1.5).29 

Scheme 1.5

Cross-coupling of Grignard reagents with C(sp2)–bromides.

Scheme 1.5

Cross-coupling of Grignard reagents with C(sp2)–bromides.

Close modal

In the same year, Mizoroki and co-workers presented a related reaction to the one reported by Heck in 1968 that importantly did not require the use of toxic arylmercury, -tin or -lead reagents. The C–C bond formation between ethylene or monosubstituted alkenes and iodobenzene could be achieved using catalytic amounts of PdCl2 or heterogeneous Pd black (Scheme 1.6).30  Concurrently, Heck demonstrated independently the Pd-catalyzed reaction of aryl halides with alkenes in the presence of a hindered amine base.31  Heck's work on aryl and vinyl halide substrates led to the second most practiced reaction in cross-coupling, namely the Mizoroki–Heck reaction.32  Irina Beletskaya and Andrei Cheprakov in Chapter 9 discuss the role of modern Heck reactions in organic synthesis.

Scheme 1.6

Pd-catalyzed coupling of alkenes with aryl iodides.

Scheme 1.6

Pd-catalyzed coupling of alkenes with aryl iodides.

Close modal

So far, only simple metal salts had been employed as catalysts. Corriu and Masse33  and Tamao and Kumada34  independently described the nickel-catalyzed coupling reaction of Grignard reagents with aryl halides. Tamao and Kumada (Figure 1.8) thereby pioneered the area of cross-coupling by showing the effects of adding phosphine ligands to the catalysts.33–35 

Figure 1.8

Prof. Makoto Kumada. Photograph courtesy of Kyoto University.

Figure 1.8

Prof. Makoto Kumada. Photograph courtesy of Kyoto University.

Close modal

The benefit of using phosphine ligands was particularly striking in reactions with less reactive aryl chlorides. Chapter 2, authored by Andrew DeAngelis and Thomas Colacot, covers the emergence of the development and use of ligands in Pd-catalyzed cross-coupling reactions in detail, with some theoretical background in choosing the right ligands for specific reaction types. Only during the past 10–15 years has the importance of the steric and electronic properties of the ligands used been fully recognized and evaluated.

A few years after Tamao and Kumada's disclosure of the importance of added phosphine ligands, four groups made independent reports on the cross-coupling reactions between alkynes and C(sp2)–halide reagents.36  Cassar made use of Pd(PPh3)4 as catalyst, whereas Heck and Sonogashira (Figure 1.9) used Pd(PPh3)2(OAc)2 and Pd(PPh3)2Cl2, respectively. In addition, Sonogashira's studies showed that the presence of catalytic amounts of CuI enabled the reaction to proceed efficiently under significantly milder reaction conditions (Scheme 1.7). Recent work from a few laboratories, including ours, demonstrated that Cu is not required; in fact, it is detrimental for alkynylation reactions of aryl chlorides and unactivated aryl bromides where electron-rich Pd catalysts are required.37  An insight into this phenomenon has been proposed.37a 

Figure 1.9

Prof. Sonogashira (Pd/Cu-catalyzed sp–sp2 coupling).

Figure 1.9

Prof. Sonogashira (Pd/Cu-catalyzed sp–sp2 coupling).

Close modal
Scheme 1.7

Cu-mediated Pd-catalyzed coupling of alkenes with aryl iodides.

Scheme 1.7

Cu-mediated Pd-catalyzed coupling of alkenes with aryl iodides.

Close modal

Until 1976, the development of cross-coupling reactions had focused on the use of magnesium-based Grignard reagents as nucleophiles, with the exceptions of the discoveries of Mizoroki, Heck and Sonogashira. However, in 1976, Negishi showed that other organometallic reagents could be used as coupling partners.38  First, Ni- or Pd-catalyzed cross-coupling reactions with organoaluminium reagents as nucleophiles were disclosed,39  followed by the use of arylzinc reagents (Negishi reaction) (Scheme 1.8).40 

Scheme 1.8

Pd-catalyzed coupling of organozinc reagents with aryl iodides.

Scheme 1.8

Pd-catalyzed coupling of organozinc reagents with aryl iodides.

Close modal

Continuing the investigations of the use of alternative nucleophiles, Kosugi and Migita,41  followed by Stille,42  reported cross-coupling reactions involving organotin reagents (Scheme 1.9).

Scheme 1.9

Pd-catalyzed coupling of organotin reagents with aroyl chlorides.

Scheme 1.9

Pd-catalyzed coupling of organotin reagents with aroyl chlorides.

Close modal

Extending the range of possible organometallic coupling partners, following the report by Murahashi's on a transmetallation strategy for using trialkylboranes,43  Suzuki and Miyaura disclosed their findings on the beneficial effect of added bases in cross-coupling reactions of alkenylboranes with aryl halides catalyzed by palladium (Suzuki–Miyaura reaction) (Scheme 1.10).44 

Scheme 1.10

Pd-catalyzed coupling of organoboron reagents with aryl bromides.

Scheme 1.10

Pd-catalyzed coupling of organoboron reagents with aryl bromides.

Close modal

Since Suzuki and Miyaura's seminal report, this cross-coupling reaction has been extensively developed and its mechanism studied in great detail. Chapter 8, authored by Alastair Lennox and Guy Lloyd-Jones, describes the discovery, development and deployment of boron reagents in cross-coupling reactions, including mechanistic analysis. In Chapter 11, the topic of stereospecific and stereoselective Suzuki–Miyaura coupling reactions is discussed by Ben Glasspoole, Eric Keske and Cathleen Crudden.

Subsequent to the introduction of organoboron reagents as cross-coupling partners, Hiyama reported efficient cross-coupling reactions of arylsilanes by the use of fluoride additives (Scheme 1.11).45 

Scheme 1.11

Pd-catalyzed coupling of organosilicon reagents with aryl iodides.

Scheme 1.11

Pd-catalyzed coupling of organosilicon reagents with aryl iodides.

Close modal

By 1989, it was already possible to achieve C–C bond formation through cross-coupling reactions using a wide variety of organometallic reagents. The next breakthrough came with the independent discoveries by Buchwald and Hartwig that C–N bond formation could also be efficiently accomplished through palladium-catalyzed cross-coupling reactions of free amines with aryl halides (Buchwald–Hartwig amination) (Scheme 1.12).46  This methodology was also extended to cover C–O and C–S bond formation. Chapter 6, authored by James Stambuli, provides a specific discussion of the development of C–O and C–S bond-forming reactions using cross-coupling methodologies.

Scheme 1.12

Pd-catalyzed direct coupling of amines with aryl bromides.

Scheme 1.12

Pd-catalyzed direct coupling of amines with aryl bromides.

Close modal

At this stage, it may seem like most of the monumental discoveries had already been made; however, there is by no means an end to the advancement of cross-coupling reactions. Reports continue to appear in the literature regarding novel reactivity, enhanced substrate compatibility, new trends and solutions to previously problematic transformations.

Following the seminal discoveries made before 2000, the first decade of the new millennium would see the expansion of cross-coupling reactions to include the use of the previously very challenging aryl chlorides as coupling partners. This was achieved by the groups of Buchwald,47  Fu48  and Koie49  by means of using bulky, electron-rich monophosphines as ligands. In addition, enormous progress has been made to be able to employ C(sp3)-based coupling partners, such as alkyl–BR2 or alkyl halide reagents.50  The intent of the chapters in this book is to highlight and discuss the most recent advances and discoveries within the field of cross-coupling.

Chapter 3 by Carin Johansson Seechurn, Hongbo Li and Thomas Colacot discusses the advantages of the use of well-defined preformed Pd catalysts in cross-coupling reactions. This is an area that has only been recognized for its importance in the past 10 years.51 

Although Chapter 2 briefly touches the subject of ligands other than phosphine, Chapter 4, by Anthony Chartoire and Steven Nolan, details the advances in C–X coupling that has been made using NHC (N-heterocyclic carbene) ligands in conjunction with palladium.

Specific examples of the importance of the structural design of the ligands and its impact on the outcome of cross-coupling reactions are demonstrated in Chapter 5 by Mark Stradiotto. Here, selective monoarylation reactions of amines and carbonyl compounds are discussed.

In addition to more thorough mechanistic investigations and realizations and the continuous expansion of substrate scope for the named cross-coupling reactions, there are also examples of novel reactions appearing in the literature. An example is highlighted and discussed in Chapter 7 by David Petrone, Christine Le, Stephen Newman and Mark Lautens, where the recently disclosed Pd(0)-catalyzed carboiodination reaction, a reaction fundamentally rooted in the Heck reaction, is also discussed.

The importance of cross-coupling reactions is reflected in their frequent use within the chemical industry. To meet the criteria of the economic demands for cheaper and more efficient processes, the challenges within industry are somewhat different from those met within the academic community. In recent years, continuous flow technology has provided an attractive alternative to the more traditional batch preparations usually carried out in process chemistry. This technology has also reached the field of cross-coupling, an area which is discussed by Timothy Noël and Volker Hessel in Chapter 13. The chemical industry is constantly faced with stricter regulations regarding the environmental impact of the processes being carried out. Kevin Shaughnessy reviews the area of green chemistry via cross-coupling reactions in Chapter 14. Javier Magano and Joshua Dunetz demonstrate in Chapter 15 some recent applications of cross-coupling reactions in the large-scale synthesis of pharmaceuticals and showcase the problems encountered in a process chemistry environment. Chapter 16, authored by Kazunori Koide, highlights the problem with residual Pd in final APIs (active pharmaceutical ingredients) and discusses detection methods.

In addition to all the improvements developed to expand the field of cross-coupling involving the introduction of novel ligands, precatalysts, chemical engineering designs and ever-more in-depth understanding of reaction mechanisms, the area continues to grow as challenges in terms of incorporating problematic nucleophilic and electrophilic reaction partners are met with innovative solutions. As an example, shown in Scheme 1.13, Feringa and co-workers recently demonstrated the previously difficult52  direct coupling reaction of organolithium reagents with aryl halides (Murahashi coupling),53  demonstrating how important problems can be solved through the utilization of the ever-growing technology made available by the numerous research groups across the globe who are active within the field of cross-coupling.

Scheme 1.13

Direct coupling of organolithium reagents with aryl bromides.

Scheme 1.13

Direct coupling of organolithium reagents with aryl bromides.

Close modal

Substantial discoveries within the field of cross-coupling have not reached an end; new reactions are waiting to be discovered, as chemists will never run out of problems to solve. Even fundamental questions such as why Pd is significantly better than Ni or Pt in most cases are still not fully understood.

1

Based on SciFinder searches, May 2014.

Close Modal

or Create an Account

Close Modal
Close Modal