Cationic Pd(II)-catalyzed C–H activation/cross-coupling reactions at room temperature: synthetic and mechanistic studies

  1. ,
  2. ,
  3. and
Department of Chemistry & Biochemistry, University of California, Santa Barbara, CA 93106, USA
  1. Corresponding author email
Guest Editor: R. Sarpong
Beilstein J. Org. Chem. 2016, 12, 1040–1064. https://doi.org/10.3762/bjoc.12.99
Received 18 Jan 2016, Accepted 28 Apr 2016, Published 20 May 2016
Full Research Paper
cc by logo

Abstract

Cationic palladium(II) complexes have been found to be highly reactive towards aromatic C–H activation of arylureas at room temperature. A commercially available catalyst [Pd(MeCN)4](BF4)2 or a nitrile-free cationic palladium(II) complex generated in situ from the reaction of Pd(OAc)2 and HBF4, effectively catalyzes C–H activation/cross-coupling reactions between aryl iodides, arylboronic acids and acrylates under milder conditions than those previously reported. The nature of the directing group was found to be critical for achieving room temperature conditions, with the urea moiety the most effective in promoting facile coupling reactions at an ortho C–H position. This methodology has been utilized in a streamlined and efficient synthesis of boscalid, an agent produced on the kiloton scale annually and used to control a range of plant pathogens in broadacre and horticultural crops. Mechanistic investigations led to a proposed catalytic cycle involving three steps: (1) C–H activation to generate a cationic palladacycle; (2) reaction of the cationic palladacycle with an aryl iodide, arylboronic acid or acrylate, and (3) regeneration of the active cationic palladium catalyst. The reaction between a cationic palladium(II) complex and arylurea allowed the formation and isolation of the corresponding palladacycle intermediate, characterized by X-ray analysis. Roles of various additives in the stepwise process have also been studied.

Introduction

Transition metal-catalyzed, direct functionalization of aryl C–H bonds has made enormous progress over the past decade, and continues to attract a great deal of attention due to the highly efficient routes now available for elaborating aromatic rings. While reactions of this type have been known for decades [1-71], serious challenges remain in achieving reactivity and selectivity due to the inertness and ubiquity of C–H bonds. High temperatures are frequently required to realize aromatic C–H functionalization (>120 °C), increasing the potential for side reactions and functional group compatibility issues. Indeed, C–H activation transformations until recently have rarely proceeded at ambient temperature due to the typically low reactivity of these positions [72-79]. In the case of palladium-catalyzed C–H activation, the crucial, namesake “C–H activation” step typically involves a C–H to C–Pd refunctionalization, generating a reactive aryl-palladium species that is poised for further transformations.

Three approaches (Figure 1) have generally been employed to enhance the reactivity and promote the key metalation/C–H bond cleavage step: (1) tuning of the reaction conditions through inclusion of various additives such as metal salts [1-22], or strong acids such as TFA or HOAc, in addition to the application of heat, although it is not always clear which steps within the overall mechanism are most directly effected under these conditions; (2) in a major subset of C–H activation chemistry, internally chelating ortho-directing groups [71-85] have been found to effectively promote selective C–H activation, typically by aiding in the formation of a palladacycle intermediate. Careful tuning of the structure of the directing group, with functionalities including a variety of nitrogen-containing moieties, such as amides [86,87], N-heterocycles [88,89], imines [90,91], pyridine N-oxide [92], amines [93,94], as well as a variety of others [1-71], has been found to profoundly impact reactivity; (3) tuning of ligands around the transition metal catalyst center has emerged as an especially powerful means of enhancing and controlling reactivity in these processes [95-107].

[1860-5397-12-99-1]

Figure 1: Road map to enhanced C–H activation reactivity.

A fourth approach with considerable potential, and which appears to have received considerably less attention, involves tuning the cationicity of the transition metal catalyst [72-79]. Literature studies have suggested that certain anionic ligands on palladium, such as acetate or carbonate, may assist C–H bond cleavage by acting as internal bases as part of a concerted metalation–deprotonation (CMD) pathway, particularly in the case of less electron-rich arenes (Scheme 1, top) [34,108-119]. In other arrays, particularly those with more electron-rich substituents, evidence suggests an electrophilic aromatic substitution mechanism may be operative. In these instances, electron-poor catalysts, such as those generated from the reaction of Pd(II) and TFA, have in some cases been shown to be especially effective. We reasoned that substitution with a more distant coordinating anion would result in a highly Lewis acidic, dicationic palladium species that might be still more reactive in the electrophilic palladation step, potentially gaining entry to C–H activation under even milder reaction conditions for selected couplings than have previously been observed (Scheme 1, bottom) [120-122].

[1860-5397-12-99-i1]

Scheme 1: Concerted metalation–deprotonation and elelectrophilic palladation pathways for C–H activation.

Metal cations, in general, are well known to increase the reactivity of C–C and C–N double bonds due to their Lewis acidity. Cationic palladium complexes [123], in particular, possess a wide breadth of reactivity, having been used to catalyze Diels–Alder [124,125], aldol and Mannich reactions [126-128], Wacker oxidations [129], polymerizations of alkenes [130,131], and asymmetric 1,4-additions with arylboronic [132-134], arylbismuth [135], and arylsilicon [136] reagents. Although carbocations react with arenes through electrophilic aromatic hydrogen substitution in a Friedel–Crafts reaction, the potential for metal cations to participate in similar chemistry has been far less widely examined. A cationic palladium-catalyzed electrophilic aromatic C–H substitution without basic anions [137-142] would hold considerable promise as an alternative and potentially milder approach to achieving valued C–H activation/coupling reactions.

While several cationic palladium complexes are commercially available, they may also be generated in situ via a variety of routes (Scheme 2), including: (a) reaction of a palladium complex with a non-coordinating anion source, usually an acid or metal salt; (b) reaction of Pd(II) halide complexes and silver salts [143-145]; (c) electronic oxidation of Pd(0) [146]; and (d) chemical oxidation of Pd(0) with HBF4, Cu(BF4)2 or AgBF4 [136,147,148].

[1860-5397-12-99-i2]

Scheme 2: Routes for generation of cationic palladium(II) species.

We have previously reported that cationic palladium-catalyzed C–H arylations of arylureas [121,122] and Fujiwara–Moritani reactions of anilide derivatives [148] can be effected at room temperature. In this account we disclose, in addition to full details associated with this C–H activation chemistry, additional applications of room temperature Fujiwara–Moritani reactions including a synthesis of the herbicide boscalid, as well as spectroscopic and mechanistic studies.

Results and Discussion

C–H arylations of arylureas with aryl iodides and arylboronic acids

Among the most conceptually attractive approaches to aromatic C–H activation is the efficient synthesis of biaryls through direct arylation reactions. The widespread availability of aryl iodides and arylboronic acids make them appealing arylating agents [149-171]. Previously reported C–H activation reaction conditions employing these reagents, however, have typically required high temperatures to obtain the desired coupling products in good yields. In order to successfully carry out these reactions at ambient temperature, several considerations must be addressed, as illustrated in Figure 1. Optimization studies initially focused on the choice of an ortho-directing group together with a cationic palladium(II) catalyst. Although the combination of acetanilide together with a palladium(II) catalyst lead to the corresponding palladacycle, as reported by Tremont [172], in the presence of 2a, Pd(OAc)2, HBF4 and AgOAc at room temperature, acetyl or isopropyl anilides afford essentially no product. Only after heating to 50 °C did substrate 1a react with iodide 2a. The corresponding pivaloylanilide is also known to serve as an effective directing group at 130 °C, but at room temperature a poor yield was obtained. Only the dimethylurea analog gave satisfactory conversion to the desired biaryl, thus arriving at optimized conditions, as shown in Scheme 3 (left). By contrast, the Daugulis group and others [173,174] have demonstrated Pd-catalyzed ortho-arylations of anilides at temperatures typically greater than 100 °C, and the Sanford group has also studied similar transformations involving diaryliodonium salts [175]. Arylureas have recently been noticed to be more active coupling partners for C–H functionalizations as opposed to anilides, especially at lower temperatures [176]. A number of strong acids have been previously utilized in C–H activation reactions [1-22], e.g., HBF4 was found to be critical for generation of biaryl 3h in good yield. The structure of the ortho-arylated product was confirmed by X-ray analysis. While similar reactions, in addition to requiring high temperatures, have typically employed strong acids such as TFA as the organic solvent, here 2 wt % solutions of selected surfactants in water were found to be excellent reaction media, providing an additional environmentally appealing feature to this protocol. While good yields could be obtained using the first generation surfactant PTS (polyoxyethanyl α-tocopheryl sebacate) [121,177,178], several other amphiphiles that are both less costly and are items of commerce gave comparable results. Using commercialy available Brij 35 (2 wt %) in water [179-183] afforded the best levels of conversion and thus, overall yields, while in its absence (i.e.,”on water”), there was a noticeable drop in the extent of conversion. In addition, lower loadings of HBF4, silver salt, or the palladium catalyst also gave inferior results.

[1860-5397-12-99-i3]

Scheme 3: Optimized conditions for C–H arylations at room temperature.

In the case of C–H activation/Suzuki–Miyaura coupling reactions, the commercially available, pre-formed cationic Pd(II) catalyst [Pd(MeCN)4](BF4)2 [184], was found to efficiently catalyze the reaction between arylureas and arylboronic acids. On the other hand, C–H arylations with aryl iodides catalyzed by [Pd(MeCN)4](BF4)2 did not give any of the desired products (see mechanistic discussion; vide infra). Various neutral palladium catalysts were examined, such as Pd(OAc)2, PdCl2Ln, Pd2(dba)3, in the absence of added acid, but none led to cross-coupling at room temperature. 1,4-Benzoquinone (BQ) was found to be an effective additive in promoting the reaction, while addition of stoichiometric metal salts (e.g., silver or copper salts) was unnecessary. Moreover, in this case organic solvents were far more effective as the reaction medium than was water, possibly due to BQ solubility issues. EtOAc, rather than EtOH and THF was the most effective (Scheme 3, right), while other organic solvents (e.g., DMF) gave low-to-moderate yields of product 3a. Although reduced amounts of both phenylboronic acid (2b) and BQ still gave excellent yields, lower catalyst loadings caused slower reactions. A neutral palladium(II) complex, Pd(OAc)2, showed no catalytic acitivity, whereas catalytic Pd(OAc)2 in the presence of stoichiometric HBF4 reacted with an arylurea and arylboronic acid to afford the biaryl in high yield (Scheme 4).

[1860-5397-12-99-i4]

Scheme 4: Biaryl formation catalyzed by Pd(OAc)2.

Representative results for the reactions between aryl iodides and arylboronic acids are summarized in Figure 2. These arylations tolerate various combinations of substrates and reagents having electron-withdrawing or electron-donating groups, as well as sterically hindered aromatic rings, all taking place at room temperature. Suzuki–Miyaura-type C–H coupling reactions are typically more tolerant of electron-withdrawing groups (3d, 3f, 3k) and ortho-substitution (3g) on the aryl ring. On the other hand, the reaction with 4-methoxycarbonylphenyl iodide, for example, gave a low yield of product 3k. Arylureas having only an electron-withdrawing group showed no reactivity towards coupling under either set of conditions (3x), consistent with an electrophilic aromatic substitution pathway in the initial C–H activation step by a cationic palladium(II) species (vide infra). Arylureas with various alkyl substituents in the ortho-position, including cyclic arrays, gave good isolated yields (3t, 3v, 3w), whereas a 2-phenyl substituted arylurea did not participate in the C–H activation/coupling reaction under these conditions (3y). Overall, C–H Suzuki–Miyaura coupling reactions were applicable to a broader substrate scope than the corresponding reaction with aryl iodides, although the latter protocol remains appealing for a variety of cross-coupling combinations due to both the convenience of aryl iodides as substrates, and the use of water as the gross reaction medium.

[1860-5397-12-99-2]

Figure 2: C–H arylation results. Conditions A: Conducted at rt for 20 h in 2 wt % Brij 35/water (1 mL) with 10 mol % Pd(OAc)2, AgOAc (2 equiv), HBF4 (5 equiv), arylurea (1, 0.25 mmol), and ArI (2.0 equiv). Conditions B: Conducted at rt for 20 h in EtOAc (1 mL) with 10 mol % [Pd(MeCN)4](BF4)2, BQ (2 or 5 equiv), arylurea (1, 0.25 mmol), and ArB(OH)2 (1.5 or 3 equiv). aRun for 48 h. b2 equiv of BQ. cRun for 96 h. dRun for 72 h.

Especially noteworthy are the numerous examples of aniline derivatives lacking ortho- or meta-substitution, which have previously been shown to be prone to double arylation (Figure 3). Since literature conditions generally employ elevated temperatures, directed C–H arylations have often suffered from uncontrollable double arylation in symmetrical or unsubstituted substrates [1-22,173,174]. At ambient temperatures, however, coupling reactions on these more challenging substrate types underwent selective mono-arylations in water (Figure 3). In fact, doubly arylated products were rather difficult to generate under these room temperature conditions, not unexpected given the previously described low reactivity of ureas already possessing an ortho-aryl substituent [121,122].

[1860-5397-12-99-3]

Figure 3: Monoarylations in water at rt. Conditions A: Conducted at rt for 20 h in 2 wt % Brij 35/water with 10 mol % Pd(OAc)2, AgOAc (2 equiv), HBF4 (5 equiv), arylurea (1, 0.25 mmol), and ArI (2, 1.5 equiv). Conditions B: Conducted at rt for 20 h in EtOAc with 10 mol % [Pd(MeCN)4](BF4)2, BQ (5 equiv), arylurea (1, 0.25 mmol), and ArB(OH)2 (2, 1.5 equiv). The ratios of single:double arylation determined by 1H NMR are shown in the parentheses. aRun for 48 h. b1.2 equiv of ArB(OH)2. c1.2 equiv of ArI. dRun for 72 h.

A 1-naphthylurea also gave excellent selectivity at room temperature (Scheme 5). When this substrate was subjected to optimized conditions for the boronic acid C–H coupling, the corresponding singly ortho-arylated product was obtained in 97% yield solely as the 2-aryl isomer, as confirmed by 1H NMR and X-ray crystallography. Generally, it is difficult to efficiently control the selectivity between the C2 and C8 positions in naphthalene rings towards a single isomeric C–H activation product [163,174,185-189].

[1860-5397-12-99-i5]

Scheme 5: Selective arylation of a 1-naphthylurea derivative.

Fujiwara–Moritani reactions

Following these results on biaryl constructions via C–H activation at room temperature, we next sought to apply our cationic palladium(II) conditions to the venerable Fujiwara–Moritani reaction. As reported back in 1967, this direct aryl olefination reaction is among the first palladium-catalyzed C–H activation reactions to be described [190-192]. Subsequent studies have generally resorted to elevated temperatures (80–160 °C) and anhydrous conditions, and in many cases high pressures of CO or O2 are also required in order to carry out these Heck-like coupling reactions [1-22,193-203]. Additional progress of note includes coupling reactions with arenes containing an ortho-directing group [1-22,45-51,193-203], as well as a meta-selective Fujiwara–Moritani reaction [204,205]. A recent report employing arylureas as the C–H coupling partner achieved limited coupling at ambient temperature, with most examples requiring heating to 60 °C [74,206].

We have previously reported a methodology enabling Fujiwara–Moritani reactions to be run in water at room temperature using the cationic palladium catalyst [Pd(MeCN)4](BF4)2 (Figure 4, 5ac, conditions A). While this reaction proceeded with a number of alkyl anilide derivatives, as well as ureas as directing groups (5c), the substrate scope was otherwise somewhat limited; only anilides possessing a strongly donating alkoxy group meta to the directing group (para to the position where the C–H activation would occur) were reactive. However, we have since found that use of acetonitrile-free, in situ generated cationic palladium with arylureas as the directing group expanded the substrate scope to include reactions with 3-alkyl-substituted ureas, as well as a wider variety of acrylates and even some acrylamides (Figure 4, 5dh, conditions B). Many combinations of acrylates and more challenging arylureas, however, did not produce the desired product in satisfactory yields, and the reaction still required the use of stoichiometric silver salts in addition to benzoquinone.

[1860-5397-12-99-4]

Figure 4: Fujiwara–Moritani coupling rreactions in water. Conditions A: 10 mol % [Pd(MeCN)4](BF4)2, 1 equiv BQ, 2 equiv AgNO3, 2 wt % PTS/water, rt, 20 h. Conditions B: 10 mol % Pd(OAc)2, 1 equiv BQ, 2 equiv AgOAc, 5 equiv HBF4 aq, 2 wt % Brij 35/water, rt, 20 h.

To overcome these limitations, further optimization of the catalyst system was conducted (Figure 5). A combination of AgNO3 or AgOAc and BQ was critical to obtain good yields of the same products formed earlier in water (vide supra), but as seen previously in the corresponding Suzuki–Miyaura reactions, a switch to EtOAc obviated the need for a silver salt. In the presence of BQ and HBF4, the reaction of 1g and acrylate 4 was efficiently catalyzed by Pd(OAc)2 (Figure 5, runs 3 and 4). Lower loadings of BQ and HBF4 also gave good results (Figure 5, runs 5 and 7). Much lower loading of HBF4, however, afforded a low yield of product 5 (Figure 5, run 8). In the absence of acid or BQ, the product was not obtained (Figure 5, runs 6 and 9). The pre-formed cationic palladium(II) complex, [Pd(MeCN)4](BF4)2, was also found to effectively catalyze the reaction between 1g and ester 4 at room temperature, without additional acid, although somewhat longer reaction times were necessary (Figure 5, run 10). Despite the presence of two potentially reactive ortho-aromatic C–H bonds in 1g, the mono-acrylated product was obtained exclusively.

[1860-5397-12-99-5]

Figure 5: Optimization. Conducted at rt for 8 h or as otherwise noted in EtOAc with 10 mol % Pd catalyst, AgOAc, HBF4, arylurea (1g, 0.25 mmol), BQ and acrylate (2c, 2.0 equiv). a2 wt % Brij 35 in water instead of EtOAc.

Under optimized conditions, various acrylates and amides can be synthesized via C–H activation reactions (Figure 6). Methyl acrylate, which did not show good general reactivity with arylureas under previous conditions, could be coupled in excellent yields (5i, 5j, 5k). As previously mentioned, a drawback characteristic of several ortho-directed C–H activation cross-coupling approaches has been the undesired coupling at both sites ortho- to the directing group. These new conditions completely inhibited second-stage alkenylation, thereby generating singly derivatized arylureas in good yields (5l, 5m, 5n). Arylureas containing halogens, which are slightly electron-deficient but provide useful synthetic handles for subsequent functionalization, reacted cleanly to form the desired products (5p, 5q). Arylureas bearing ortho-alkyl substituents also gave excellent yields (5o, 5r), while acrylamides having simple amine or amino acid moieties also participated in cross-coupling reactions with the arylurea to produce the corresponding amide derivatives in moderate to good yields (5s, 5t).

[1860-5397-12-99-6]

Figure 6: Representative results in EtOAc. Conducted at rt in EtOAc with 10 mol % Pd(OAc)2, HBF4 (1 equiv), arylurea (1, 0.25 mmol), BQ (3 equiv) and acrylate (2, 2.0 equiv).

Total synthesis of boscalid® via C–H activation

The rationale behind the attention recently accorded C–H activation chemistry has been based, in part, on its potential to streamline routes towards valuable synthetic targets. As a demonstration of the utility of our C–H activation approach, we chose to synthesize boscalid®, a pesticide currently prepared on a yearly kiloton scale by means of a traditional Suzuki–Miyaura coupling. It is used to control a range of plant pathogens in broadacre and horticultural crops (Scheme 6) [207]. Felpin and co-workers have reported its synthesis starting from aryldiazonium salts [208], while the Heinrich group has employed a free-radical biaryl cross-coupling of diaminobenzene promoted by TiCl3 [209]. A number of additional syntheses can also be found in the literature [86,210]. Moreover, the BASF has patented routes using a traditional Suzuki–Miyaura cross-coupling in the presence of 0.5 mol % Pd catalyst to reach the same nitro-intermediate found in the Felpin route [211-213].

[1860-5397-12-99-i6]

Scheme 6: Previous syntheses of boscalid®.

Many of these syntheses require large excesses of iron or other stoichiometric metals to obtain high yields (Scheme 6). As shown in Scheme 7, a synthesis that proceeds via a C–H activation strategy, however, might provide a highly efficient, alternative route originating from just aniline. The corresponding phenylurea can be prepared in high yield (96%), which is then subjected to C–H Suzuki–Miyaura coupling at room temperature (91%). Sequential deprotection and acylation with 2-chloronicotinoyl chloride result in boscalid in four steps in an overall yield of 86%, which compares favorably with all known routes to this pesticide shown in Scheme 6 [86,208-213].

[1860-5397-12-99-i7]

Scheme 7: Synthesis of boscalid®. aConducted at rt for 20 h in EtOAc with 10 mol % [Pd(MeCN)4](BF4)2, BQ (5 equiv), arylurea (1, 0.25 mmol), and ArB(OH)2 (2, 1.5 equiv). bNaOH, dioxane/water, reflux, then 2-chloronicotinoyl chloride, Et3N, THF, rt.

Mechanistic insight

Although there have been a number of mechanistic studies on C–H activation reactions involving neutral palladium species [34,108-119], those catalyzed by cationic palladium have been much less thoroughly examined. We hypothesized that our catalytic cycles for the Fujiwara–Moritani, arylboronic acid, and aryl iodide coupling reactions catalyzed by cationic palladium complexes are composed of three key steps; (1) aromatic C–H activation by cationic palladium; (2) reaction of the resulting intermediate (a cationic palladacycle) with a corresponding reagent; and (3) re-generation of the active catalyst (Scheme 8). In order to test this hypothesis we explored the viability of each of these individual steps.

[1860-5397-12-99-i8]

Scheme 8: Hypothetical reaction sequence for cationic Pd(II)-catalyzed aromatic C–H activation reactions.

The C–H activation step

Although aromatic C–H bond activation through palladacycle [214] generation is a critical step in the ortho-directed, activation/cross-coupling sequence, many of its specific mechanistic features are still controversial. Previous studies with arylureas [73,206,215] have formulated a palladacycle as the likely initial intermediate associated with palladation and subsequent C–H bond cleavage.

In order to confirm palladacycle formation in our reactions with arylureas, the dicationic palladium complex [Pd(MeCN)4](BF4)2 was exposed to one equivalent of 3-methoxyphenylurea 1f at room temperature for 20 minutes (Scheme 9). This stoichiometric reaction led to the corresponding palladacycle 6 in 95% yield, without the aid of additives (i.e., no Ag salt or protic acid). In harmony, in situ-generated cationic palladium from the reaction of Pd(OAc)2 and HBF4 gave the same palladacycle upon addition of acetonitrile, as confirmed by NMR. The facile formation of this species supports the intermediacy of a palladacycle in the catalytic cycle. The structure of the isolated palladacycle was confirmed by X-ray analysis [216].

[1860-5397-12-99-i9]

Scheme 9: Palladacycle formation.

An ORTEP plot for palladacycle 6 is shown in Figure 7. The molecular structure consists of a Pd atom with an arylurea and two molecules of acetonitrile assembled in a square-planar geometry around the metal. The sum of the angles around Pd is 360.01°. The C(5)–Pd–O(2) angle (91.98°) is slightly larger than that of N(3)–Pd–N(4) (87.81°), but it is similar to the angles of neutral PdCl2(Ph2PCH2CH2CH2PPh2)(dppp)) (angle of P–Pd–P: 90.58°) having a six-membered ring conformation [217], and palladacycles reported previously [73,206,215]. The length of the Pd–N4 bond, (2.126 Å), is slightly longer than those of Pd–N(3), Pd–C(5), Pd–O(2) bonds, likely due to a trans effect of the strong σ-donor aryl group as has been observed in a related urea palladacycle [73,215]. The bond length of Pd–N is 1.96 Å in [Pd(MeCN)4](BF4)2.

[1860-5397-12-99-7]

Figure 7: X-ray structure of palladacycle 6 with thermal ellipsoids at the 50% probability level. BF4 and hydrogen atoms were omitted for clarity. Selected bond length (Å): Pd–C(5) = 1.980, Pd–N(3) = 1.995, Pd–N(4) = 2.126, Pd–O(2) = 1.988. Selected angles (°): C(5)–Pd–N(3) = 94.15, N(3)–Pd–N(4) = 87.81, N(4)–Pd–O(2) = 86.07, O(2)–Pd–C(5) = 91.98.

NMR spectroscopic studies on the reaction between a cationic Pd(II) complex and an arylurea to generate a palladacycle are illustrated in Figure 8. The pure palladacycle from pre-formed cationic palladium [Pd(MeCN)4](BF4)2 is shown as spectrum in Figure 8A. Generally, monocationic arylpalladium(II) complexes without strongly coordinating ligands are unstable even at low temperatures [218-220]; nonetheless, this cationic palladacycle, aided by the presence of strongly coordinating MeCN, was found to be quite stable at room temperature. While the in situ generated cationic palladium species from the reaction of Pd(OAc)2 and HBF4 gave the same palladacycle upon treatment with the arylurea (spectrum Figure 8B), the reaction in the absence of HBF4 did not result in palladacycle formation (spectrum Figure 8C). Here, essentially no conversion of the starting material was detected by 1H NMR in acetone-d6 (spectrum Figure 8D). Indeed, for reactions starting from Pd(OAc)2, no cross-coupling product was observed without adding a BF4 source for the Fujiwara–Moritani reaction (Figure 5, run 9), Suzuki–Miyaura coupling (Scheme 4), and arylation with aryl iodide [121,122,150]. HBF4 apparently acts as an acetate scavenger to generate the active cationic palladium(II) species (Scheme 10).

[1860-5397-12-99-8]

Figure 8: NMR studies. A: The reaction of [Pd(MeCN)4](BF4)2 and 3-MeOC6H4NHCONMe2 in acetone-d6. B: The reaction of Pd(OAc)2 and 3-MeOC6H4NHCONMe2 in the presence of HBF4 and MeCN in acetone-d6. C: The reaction of Pd(OAc)2 and 3-MeOC6H4NHCONMe2 in the presence of MeCN in acetone-d6. D: 3-MeOC6H4NHCONMe2 in acetone-d6.

[1860-5397-12-99-i10]

Scheme 10: The generation of cationic Pd(II) from Pd(OAc)2.

As discussed previously herein, there are several routes available for cyclopalladation and C–H bond cleavage, most notably the concerted metalation-deprotection (CMD) or electrophilic palladation pathways (Scheme 1) [221-224]. Although control experiments had previously indicated the importance of conditions involving cationic palladium for achieving overall reaction conversion, our studies of palladacycle formation suggest that a cationic palladium catalyst is specifically required for the initial C–H activation step itself. Since the crystal structure of 6 (Figure 7) is indicative of a monocationic palladacycle, the cationicity of the metal may still play a role as well in subsequent steps. However, the much higher reactivity of a cationic Pd species (even under acetate-free conditions), the lack of effectiveness of Pd(OAc)2 alone in palladacycle formation, and the observed reactivity trends that strongly favor more electron-rich arylureas, all appear to be most consistent with an electrophilic palladation pathway over a CMD mechanism (Scheme 11). However, it is appreciated that further study might provide additional insight on this point.

[1860-5397-12-99-i11]

Scheme 11: Electrophilic substitution of aromatic hydrogen by cationic palladium(II) species.

Reactions of palladacycle 6 with Ph–I, PhB(OH)2, and an acrylate

Having demonstrated the potential for facile palladacycle formation at room temperature, we next examined the reactivity of this intermediate with coupling partners for each of the three reaction types studied. Stoichiometric reactions between the isolated palladacycle 6 and an acrylate or arylboronic acid were first attempted at room temperature (Scheme 12). Initial experiments, however, resulted in no formation of the desired products. Although the palladacycles were subjected to various conditions in the presence of BQ and HBF4, the anticipated reaction did not proceed from isolated catalyst complexes containing the stabilizing ligand MeCN.

[1860-5397-12-99-i12]

Scheme 12: Attempted reactions of palladacycle 6.

Although the initial C–H activation step proceeded readily in the presence of MeCN in these stoichiometric reactions, subsequent reactions of the palladacycle with acrylates, arylboronic acids, and aryl iodides appeared to be significantly suppressed by the presence of stoichiometric MeCN. The inhibitory effect of this ligand had been previously observed in the coupling reactions of aryl iodides (in which even 40 mol % of MeCN was enough to almost completely shut down the reaction) [121]. In a cationic palladium(II) complex-catalyzed 1,4-addition of arylsilane, the nitrile-free cationic Pd(II) catalyst was much more effective than a PhCN-containing cationic palladium(II) complex towards transmetallations with arylsilicon-containing partners and insertion of mono-cationic arylpalladium(II) species into olefins [136]. The detrimental effect of MeCN under our C–H activation conditions was further established through a series of reactions as illustrated in Scheme 13. Under optimized conditions previously determined, where C–H functionalized products were obtained in good yields, in the presence of added MeCN (1 equiv relative to 1f) all three reactions were completely inhibited, in all likelihood due to its strong coordinating ability as a ligand on cationic palladium.

[1860-5397-12-99-i13]

Scheme 13: The impact of MeCN on C-H activation/coupling reactions.

On the other hand, when nitrile-free conditions were applied to urea 1f, with in situ-generated palladacycle (from Pd(OAc)2 and HBF4; Figure 8), followed by addition of the usual reagents, each reaction proceeded to give the anticipated acrylated/arylated product (Scheme 14; unoptimized yields). Notably, all three stoichiometric reactions now proceeded in the absence of other additives, such as BQ or Ag(I) salts, which are required for the catalytic versions to proceed efficiently.

[1860-5397-12-99-i14]

Scheme 14: Stoichiometric MeCN-free reactions. a2% Brij 35 was used instead of EtOAc.

Although Lloyd-Jones and Booker-Milburn also reported the reaction of a urea-derived palladacycle and arylboronic acid in the presence of base in THF under reflux conditions to produce the corresponding coupling product, our cationic palladacycle underwent coupling without added base (Scheme 15) [178]. In fact, it has been previously shown that cationic palladium species can undergo transmetalation with an arylboronic acid in the absence of base even at 0 °C [219,220]. Wu and co-workers have also reported the interesting reactivities of neutral palladacycles with arylboronic acids (Scheme 15). Under their conditions, BQ and high temperature were critical to obtain the product from their 5-membered isoxazoline-containing palladacycle [160,168,225]. Although BQ is sometimes used as a ligand for palladium to accelerate reductive elimination [103,226-230], its presence was not necessary in our stoichiometric reaction of a cationic 6-membered ring palladacycle.

[1860-5397-12-99-i15]

Scheme 15: The reactions of divalent palladacycles.

Regeneration of active catalyst; the roles of additives BQ, AgOAc, and HBF4

As shown in previous sections herein, both the formation of palladacycles from arylureas and their subsequent coupling reactions with acrylates, arylboronic acids, and aryl iodides proceed under stoichiometric palladium conditions in the absence of additives, such as BQ and AgOAc, which had been necessary in the corresponding optimized catalytic reactions. To establish the roles of these additives, the reaction of the palladacycle in the presence of excesses of both coupling partners was carried out (Scheme 16). In Fujiwara–Moritani and Suzuki–Miyaura coupling reactions, 41 and 39% of the products (isolated yields based on the palladacycle) were obtained, respectively, in the absence of BQ, indicating that no catalyst turnover was occurring without this additive. When 20 equivalents of BQ were added, along with the coupling partners, however, 847% (TON = 8.5) and 467% (TON = 4.7) yields of the products were obtained, respectively, supporting a key role for BQ in regeneration of the active dicationic species (PdL4(BF4)2).

[1860-5397-12-99-i16]

Scheme 16: Role of BQ in stoichiometric Fujiwara–Moritani and Suzuki–Miyaura coupling reactions. aYields based on Pd.

Benzoquinone (BQ) has been well studied as an oxidant for Pd(0) to Pd(II) processes, generating hydroquinone as a byproduct. For the Fujiwara–Moritani coupling, addition of the palladacycle 6 to an acrylate followed by β-hydride elimination and reductive elimination of HPd+BF4 would result in a Pd(0) species unable to participate in palladacycle formation until it is oxidized by BQ to Pd2+(BF4)2, whereupon it reacts with another equivalent of arylurea (Scheme 17) [231]. Similarly, BQ’s role in C–H coupling of boronic acids would likely be to oxidize Pd(0) to Pd(II) after the product forming step (Scheme 18). Transmetallation between the palladacycle and arylboronic acid followed by reductive elimination would give the expected product and Pd(0), where the metal can be subsequently oxidized with BQ.

[1860-5397-12-99-i17]

Scheme 17: Proposed role of BQ in Fujiwara–Moritani reactions.

[1860-5397-12-99-i18]

Scheme 18: Proposed role of BQ in Suzuki–Miyaura coupling reactions.

In the reaction with aryl iodides, when an excess of coupling partner was employed, a yield of ca. 90% (TON = 0.9) of the product was obtained with or without additional HBF4 (Scheme 19), in this case indicating that no catalyst turnover was occurring in the absence of a Ag(I) salt. Surprisingly, when 15 equivalents of AgOAc were added along with an excess of both coupling partners, but without the addition of HBF4 beyond the three equivalents required for initial palladacycle formation, no coupling product was observed. Under these conditions, a large excess of acetate anion relative to BF4 would exist, which may decrease the cationicity of the cationic palladium(II) species formed in the initial cyclopalladation (Scheme 20), or otherwise disrupt the reaction sequence subsequent to palladacycle formation. On the other hand, the reaction in the presence of both silver salt and excess of HBF4 gave the corresponding product in 342% yield relative to palladium (TON = 3.4) (Scheme 19).

[1860-5397-12-99-i19]

Scheme 19: Stoichiometric C–H arylation of iodobenzene. aYields based on Pd.

[1860-5397-12-99-i20]

Scheme 20: Impact of acetate on the cationicity of Pd.

Based on these results, the proposed roles of silver and acid are shown in Scheme 21. After the coupling of the palladacycle and aryl iodide, I–Pd+BF4 is generated, which is catalytically inactive. Then, I–Pd+BF4 reacts with AgOAc and HBF4 to regenerate active cationic Pd2+(BF4)2. Under this proposed sequence, AgOAc would primarily act, therefore, as an iodide scavenger.

[1860-5397-12-99-i21]

Scheme 21: Roles of additives in C–H arylation.

In our previous studies it was found that when AgBF4 was used in place of AgOAc under optimized conditions for C–H arylation with aryl iodides, HBF4 (or any other added acid) was unnecessary for the catalytic reaction to occur. Here, since AgBF4 apparently reacts with I–Pd+BF4 to produce catalytically active Pd2+(BF4)2 (Scheme 22), and there are no stoichiometric quantities of competing acetate anions, additional acid is not needed to produce and maintain active catalyst.

[1860-5397-12-99-i22]

Scheme 22: Cross-coupling in the presence of AgBF4.

Proposed mechanisms

Our results have demonstrated that a dicationic palladium complex effectively catalyzes C–H activation of arylureas at room temperature. Based on these studies of the major steps for each reaction, proposed catalytic cycles are illustrated below.

Fujiwara–Moritani reactions. At the first stage of the catalytic cycle, an active divalent cationic palladium species is generated from the reaction of a neutral complex, Pd(OAc)2, and HBF4 [232]. This reaction results in the formation of a monocationic arylpalladium(II) palladacycle, likely via a Wheland-like intermediate (Scheme 23) [233-237]. The product-forming portion of the cycle may proceed in a manner resembling a traditional Heck cross-coupling. In this case, addition of the palladacycle to an acrylate followed by β-hydride elimination yields the corresponding product 5. As previously demonstrated in palladium-catalyzed Mizoroki–Heck reactions [238-240], insertion of a alkenyl double bond into C–Pd+ present within the cationic palladacyle is facile, owing to the high Lewis acidity of the metal center. This is a noteworthy advantage associated with the use of cationic palladium(II) catalysts. Finally, BQ oxidizes the Pd(0) that is reductively eliminated from the HPd(II)+BF4 formed to regenerate the active cationic palladium species Pd2+(BF4)2.

[1860-5397-12-99-i23]

Scheme 23: A proposed catalytic cycle for Fujiwara–Moritani reactions.

C–H boronic acid coupling reactions. Unlike traditional Suzuki–Miyaura coupling reactions [241-243], C–H coupling reactions catalyzed by a cationic palladium(II) complex require an oxidant instead of a strong base, but otherwise likely share a number of features with this widely used C–C bond-forming process. The reaction also presumably starts from the generation of a cationic palladacycle, which may undergo a facile transmetalation with an arylboronic acid without prior activation by base (Scheme 24) [241-248]. This step is followed by reductive elimination of a diarylpalladium(II) species, affording the coupling product and Pd(0). The resulting Pd(0) is then oxidized with BQ to regenerate the dicationic palladium species, which can re-enter the catalytic cycle.

[1860-5397-12-99-i24]

Scheme 24: Proposed catalytic cycle of C–H activation/Suzuki–Miyaura coupling reactions.

Arylation with an aryl iodide. Coupling reactions of aryl iodides have fewer similarities in terms of traditional cross-coupling reactions compared with features characteristic of other room temperature C–H activations we have studied, and are more difficult to rationalize with a Pd(II)/Pd(0) catalytic cycle. Hence, a possible Pd(II)/Pd(IV) catalytic cycle, similar to that previously proposed by Daugulis [8], is proposed below (Scheme 25). A mono-cationic palladium intermediate reacts with the aryl iodide, albeit in a poorly understood step of the sequence. Stoichiometric studies reveal that this step occurs in the absence of AgOAc, and in fact its presence in excess relative to HBF4 inhibits the reaction (Scheme 19). In one possible pathway, an oxidative addition to the aryl iodide would provide a Pd(IV) intermediate, which could then rapidly reductively eliminate in the C–C bond forming step. The resulting I–Pd(II)–X species could then be converted back to the active cationic palladium species through reactions with the silver salt and HBF4 (or in situ generated AgBF4).

[1860-5397-12-99-i25]

Scheme 25: A proposed catalytic cycle for C–H arylation involving a Pd(IV) intermediate.

The specific nature of the reaction between the palladacycle and aryl iodide and resulting intermediate is lacking in details. It is known that divalent palladacycles react with alkyl iodides or diaryliodonium salts, and this process likely involves a Pd(IV) intermediate. Tremont and co-workers previously proposed a Pd(IV) intermediate in C–H alkylation of acetoanilides and alkyl iodides [172]. In this case the reaction of a divalent palladacycle and MeI readily occurred at room temperature and was shown not to proceed through a radical pathway. Notably, in 2011 Vicente and co-workers obtained the crystal structure of a Pd(IV) complex obtained by the room temperature oxidative addition of an internally-chelated Pd(II) species into an aryl iodide, and demonstrated this species’ competence as a pre-catalyst for C–H olefinations. Liu [73] and Sanford reported the stoichiometric reaction of palladacycles and aryliodonium salts at high temperature to give the corresponding products through Pd(II)/Pd(IV) or Pd(III)/Pd(III) dimeric species bridged by acetates (Scheme 26) [68,175,249,250]. Sequences involving Pd(III) intermediates have been suggested as alternatives to Pd(II)/Pd(IV) cycles in some cases, but most well-studied examples involve Pd(III)/Pd(III) dimers formed with the aid of bridging anionic ligands such as acetate or nitrate [251,252] that do not match as well with the cationic palladium conditions employed here. Silver-mediated one electron oxidations to form monomeric Pd(III) complexes have also been studied [251], but the successful implementation of silver-free conditions with stoichiometric palladium herein would appear to eliminate this as a key step. To the best of our knowledge, the existence of Pd(IV) complexes has yet to be conclusively demonstrated from the oxidative addition of anilide-derived, divalent palladacycles into aryl halides, although formation of octahedral Pd(IV) complexes from N-substituted biphenyl palladacycles that possess similar highly planar structures as found in the urea-derived palladacycle (Figure 7) have been well studied [253-255].

[1860-5397-12-99-i26]

Scheme 26: Selected reactions of divalent palladacycles.

Conclusion

In conclusion, we have demonstrated that a dicationic palladium complex enables facile C–H activation reactions of arylureas with aryl iodides, arylboronic acids, and acrylates at room temperature. In many cases nanomicelles in water can be used in place of organic solvents, allowing for some of the mildest and environmentally responsible conditions yet achieved in C–H activation chemistry. The practical value of this approach has been further demonstrated with an efficient, streamlined application to the synthesis of the herbicide boscalid. Mechanistic investigations revealed that a dicationic Pd(II) complex reacts readily with an arylurea to rapidly produce a mono-cationic palladacycle at room temperature, and this likely, catalytically competent species has been characterized by X-ray crystallography. Experiments revealed that a highly cationic palladium complex is required for the formation of this palladacycle at room temperature. While some key steps, including the precise nature of the reaction between this cationic palladacycle and aryl iodides, require further clarification, studies using stoichiometric palladium have provided insight into the roles of additives HBF4, BQ, and AgOAc, as well as the crucial steps of each reaction’s catalytic cycle. Lastly, this study highlights the advantages of dicationic palladium complexes in synthesis, and their tolerance to aqueous media. Such cationic reagent/medium combinations, potentially applicable to other group 10 metals, and Ni in particular, may well offer related synthetic opportunities.

Supporting Information

Supporting Information File 1: Experimental procedures and characterization of all new compounds.
Format: PDF Size: 2.1 MB Download
Supporting Information File 2: Crystal structure of 6.
Format: CIF Size: 14.7 KB Download
Supporting Information File 3: Crystal structure of 6 No 2.
Format: PDF Size: 213.6 KB Download

Acknowledgements

We warmly thank the NIH for support of this work (GM 86485), and Johnson Matthey (Dr. Thomas Colacot) for supplying the Pd(OAc)2 used in this work.

References

  1. Godula, K.; Sames, D. Science 2006, 312, 67–72. doi:10.1126/science.1114731
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  2. Deprez, N. R.; Sanford, M. S. Inorg. Chem. 2007, 46, 1924–1935. doi:10.1021/ic0620337
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  3. Alberico, D.; Scott, M. E.; Lautens, M. Chem. Rev. 2007, 107, 174–238. doi:10.1021/cr0509760
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  4. Park, Y. J.; Park, J.-W.; Jun, C.-H. Acc. Chem. Res. 2008, 41, 222–234. doi:10.1021/ar700133y
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  5. Kakiuchi, F.; Kochi, T. Synthesis 2008, 3013–3039. doi:10.1055/s-2008-1067256
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  6. Thansandote, P.; Lautens, M. Chem. – Eur. J. 2009, 15, 5874–5883. doi:10.1002/chem.200900281
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  7. Chen, X.; Engle, K. M.; Wang, D.-H.; Yu, J.-Q. Angew. Chem., Int. Ed. 2009, 48, 5094–5115. doi:10.1002/anie.200806273
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  8. Daugulis, O.; Do, H.-Q.; Shabashov, D. Acc. Chem. Res. 2009, 42, 1074–1086. doi:10.1021/ar9000058
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7] [8]
  9. Giri, R.; Shi, B.-F.; Engle, K. M.; Maugel, N.; Yu, J.-Q. Chem. Soc. Rev. 2009, 38, 3242–3272. doi:10.1039/b816707a
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  10. Storr, T. E.; Baumann, C. G.; Thatcher, R. J.; De Ornellas, S.; Whitwood, A. C.; Fairlamb, I. J. S. J. Org. Chem. 2009, 74, 5810–5821. doi:10.1021/jo9012282
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  11. Zhang, M. Adv. Synth. Catal. 2009, 351, 2243–2270. doi:10.1002/adsc.200900426
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  12. Vedernikov, A. N. Chem. Commun. 2009, 4781–4790. doi:10.1039/b907036b
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  13. Muñiz, K. Angew. Chem., Int. Ed. 2009, 48, 9412–9423. doi:10.1002/anie.200903671
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  14. Colby, D. A.; Bergman, R. G.; Ellman, J. A. Chem. Rev. 2010, 110, 624–655. doi:10.1021/cr900005n
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  15. Bellina, F.; Rossi, R. Chem. Rev. 2010, 110, 1082–1146. doi:10.1021/cr9000836
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  16. Scheuermann, C. J. Chem. – Asian J. 2010, 5, 436–451. doi:10.1002/asia.200900487
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  17. Mkhalid, I. A. I.; Barnard, J. H.; Marder, T. B.; Murphy, J. M.; Hartwig, J. F. Chem. Rev. 2010, 110, 890–931. doi:10.1021/cr900206p
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  18. Lyons, T. W.; Sanford, M. S. Chem. Rev. 2010, 110, 1147–1169. doi:10.1021/cr900184e
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  19. Jazzar, R.; Hitce, J.; Renaudat, A.; Sofack-Kreutzer, J.; Baudoin, O. Chem. – Eur. J. 2010, 16, 2654–2672. doi:10.1002/chem.200902374
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  20. Mousseau, J. J.; Charette, A. B. Acc. Chem. Res. 2013, 46, 412–424. doi:10.1021/ar300185z
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  21. Neufeldt, S. R.; Sanford, M. S. Acc. Chem. Res. 2012, 45, 936–946. doi:10.1021/ar300014f
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  22. Guo, X.-X.; Gu, D.-W.; Wu, Z.; Zhang, W. Chem. Rev. 2015, 115, 1622–1651. doi:10.1021/cr500410y
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  23. Aihara, Y.; Chatani, N. J. Am. Chem. Soc. 2014, 136, 898–901. doi:10.1021/ja411715v
    Return to citation in text: [1] [2]
  24. Rouquet, G.; Chatani, N. Angew. Chem., Int. Ed. 2013, 52, 11726–11743. doi:10.1002/anie.201301451
    Return to citation in text: [1] [2]
  25. Chan, K. S. L.; Wasa, M.; Chu, L.; Laforteza, B. N.; Miura, M.; Yu, J.-Q. Nat. Chem. 2014, 6, 146–150. doi:10.1038/nchem.1836
    Return to citation in text: [1] [2]
  26. Thuy-Boun, P. S.; Villa, G.; Dang, D.; Richardson, P.; Su, S.; Yu, J.-Q. J. Am. Chem. Soc. 2013, 135, 17508–17513. doi:10.1021/ja409014v
    Return to citation in text: [1] [2]
  27. Lewis, J. C.; Bergman, R. G.; Ellman, J. A. J. Am. Chem. Soc. 2007, 129, 5332–5333. doi:10.1021/ja070388z
    Return to citation in text: [1] [2]
  28. Tan, K. L.; Bergman, R. G.; Ellman, J. A. J. Am. Chem. Soc. 2002, 124, 3202–3203. doi:10.1021/ja017351d
    Return to citation in text: [1] [2]
  29. Tan, K. L.; Bergman, R. G.; Ellman, J. A. J. Am. Chem. Soc. 2002, 124, 13964–13965. doi:10.1021/ja0281129
    Return to citation in text: [1] [2]
  30. Kakiuchi, F.; Sonoda, M.; Tsujimoto, T.; Chatani, N.; Murai, S. Chem. Lett. 1999, 28, 1083–1088. doi:10.1246/cl.1999.1083
    Return to citation in text: [1] [2]
  31. Cook, A. K.; Emmert, M. H.; Sanford, M. S. Org. Lett. 2013, 15, 5428–5431. doi:10.1021/ol4024248
    Return to citation in text: [1] [2]
  32. Cheng, X.-F.; Li, Y.; Su, Y.-M.; Yin, F.; Wang, J.-Y.; Sheng, J.; Vora, H. U.; Wang, X.-S.; Yu, J.-Q. J. Am. Chem. Soc. 2013, 135, 1236–1239. doi:10.1021/ja311259x
    Return to citation in text: [1] [2]
  33. Desai, L. V.; Malik, H. A.; Sanford, M. S. Org. Lett. 2006, 8, 1141–1144. doi:10.1021/ol0530272
    Return to citation in text: [1] [2]
  34. Desai, L. V.; Stowers, K. J.; Sanford, M. S. J. Am. Chem. Soc. 2008, 130, 13285–13293. doi:10.1021/ja8045519
    Return to citation in text: [1] [2] [3] [4]
  35. Kim, H.; Shin, K.; Chang, S. J. Am. Chem. Soc. 2014, 136, 5904–5907. doi:10.1021/ja502270y
    Return to citation in text: [1] [2]
  36. Foo, K.; Sella, E.; Thomé, I.; Eastgate, M. D.; Baran, P. S. J. Am. Chem. Soc. 2014, 136, 5279–5282. doi:10.1021/ja501879c
    Return to citation in text: [1] [2]
  37. Matsubara, T.; Asako, S.; Ilies, L.; Nakamura, E. J. Am. Chem. Soc. 2014, 136, 646–649. doi:10.1021/ja412521k
    Return to citation in text: [1] [2]
  38. Xiao, B.; Gong, T.-J.; Xu, J.; Liu, Z.-J.; Liu, L. J. Am. Chem. Soc. 2011, 133, 1466–1474. doi:10.1021/ja108450m
    Return to citation in text: [1] [2]
  39. Sun, K.; Li, Y.; Xiong, T.; Zhang, J.; Zhang, Q. J. Am. Chem. Soc. 2011, 133, 1694–1697. doi:10.1021/ja1101695
    Return to citation in text: [1] [2]
  40. Kawano, T.; Hirano, K.; Satoh, T.; Miura, M. J. Am. Chem. Soc. 2010, 132, 6900–6901. doi:10.1021/ja101939r
    Return to citation in text: [1] [2]
  41. Houlden, C. E.; Bailey, C. D.; Ford, J. G.; Gagné, M. R.; Lloyd-Jones, G. C.; Booker-Milburn, K. I. J. Am. Chem. Soc. 2008, 130, 10066–10067. doi:10.1021/ja803397y
    Return to citation in text: [1] [2]
  42. Dick, A. R.; Remy, M. S.; Kampf, J. W.; Sanford, M. S. Organometallics 2007, 26, 1365–1370. doi:10.1021/om061052l
    Return to citation in text: [1] [2]
  43. He, J.; Wasa, M.; Chan, K. S. L.; Yu, J.-Q. J. Am. Chem. Soc. 2013, 135, 3387–3390. doi:10.1021/ja400648w
    Return to citation in text: [1] [2]
  44. Tobisu, M.; Ano, Y.; Chatani, N. Org. Lett. 2009, 11, 3250–3252. doi:10.1021/ol901049r
    Return to citation in text: [1] [2]
  45. Vora, H. U.; Silvestri, A. P.; Engelin, C. J.; Yu, J.-Q. Angew. Chem., Int. Ed. 2014, 53, 2683–2686. doi:10.1002/anie.201310539
    Return to citation in text: [1] [2] [3]
  46. García-Rubia, A.; Urones, B.; Arrayás, R. G.; Carretero, J. C. Angew. Chem., Int. Ed. 2011, 50, 10927–10931. doi:10.1002/anie.201105611
    Return to citation in text: [1] [2] [3]
  47. Nakao, Y.; Yada, A.; Ebata, S.; Hiyama, T. J. Am. Chem. Soc. 2007, 129, 2428–2429. doi:10.1021/ja067364x
    Return to citation in text: [1] [2] [3]
  48. Satoh, T.; Nishinaka, Y.; Miura, M.; Nomura, M. Chem. Lett. 1999, 28, 615–618. doi:10.1246/cl.1999.615
    Return to citation in text: [1] [2] [3]
  49. Ueura, K.; Satoh, T.; Miura, M. Org. Lett. 2007, 9, 1407–1409. doi:10.1021/ol070406h
    Return to citation in text: [1] [2] [3]
  50. Lee, G. T.; Jiang, X.; Prasad, K.; Repič, O.; Blacklock, T. J. Adv. Synth. Catal. 2005, 347, 1921–1924. doi:10.1002/adsc.200505202
    Return to citation in text: [1] [2] [3]
  51. Lim, S.-G.; Lee, J. H.; Moon, C. W.; Hong, J.-B.; Jun, C.-H. Org. Lett. 2003, 5, 2759–2761. doi:10.1021/ol035083d
    Return to citation in text: [1] [2] [3]
  52. Furukawa, T.; Tobisu, M.; Chatani, N. J. Am. Chem. Soc. 2015, 137, 12211–12214. doi:10.1021/jacs.5b07677
    Return to citation in text: [1] [2]
  53. Cho, S. H.; Hartwig, J. F. J. Am. Chem. Soc. 2013, 135, 8157–8160. doi:10.1021/ja403462b
    Return to citation in text: [1] [2]
  54. Kawamorita, S.; Murakami, R.; Iwai, T.; Sawamura, M. J. Am. Chem. Soc. 2013, 135, 2947–2950. doi:10.1021/ja3126239
    Return to citation in text: [1] [2]
  55. Kawamorita, S.; Ohmiya, H.; Hara, K.; Fukuoka, A.; Sawamura, M. J. Am. Chem. Soc. 2009, 131, 5058–5059. doi:10.1021/ja9008419
    Return to citation in text: [1] [2]
  56. Boebel, T. A.; Hartwig, J. F. J. Am. Chem. Soc. 2008, 130, 7534–7535. doi:10.1021/ja8015878
    Return to citation in text: [1] [2]
  57. Boller, T. M.; Murphy, J. M.; Hapke, M.; Ishiyama, T.; Miyaura, N.; Hartwig, J. F. J. Am. Chem. Soc. 2005, 127, 14263–14278. doi:10.1021/ja053433g
    Return to citation in text: [1] [2]
  58. Ishiyama, T.; Takagi, J.; Ishida, K.; Miyaura, N.; Anastasi, N. R.; Hartwig, J. F. J. Am. Chem. Soc. 2002, 124, 390–391. doi:10.1021/ja0173019
    Return to citation in text: [1] [2]
  59. Chen, H.; Schlecht, S.; Semple, T. C.; Hartwig, J. F. Science 2000, 287, 1995. doi:10.1126/science.287.5460.1995
    Return to citation in text: [1] [2]
  60. Cho, J.-Y.; Iverson, C. N.; Smith, M. R., III. J. Am. Chem. Soc. 2000, 122, 12868–12869. doi:10.1021/ja0013069
    Return to citation in text: [1] [2]
  61. Tse, M. K.; Cho, J.-Y.; Smith, M. R., III. Org. Lett. 2001, 3, 2831–2833. doi:10.1021/ol0162668
    Return to citation in text: [1] [2]
  62. Zhao, X.; Dimitrijević, E.; Dong, V. M. J. Am. Chem. Soc. 2009, 131, 3466–3467. doi:10.1021/ja900200g
    Return to citation in text: [1] [2]
  63. Yu, W.-Y.; Sit, W. N.; Lai, K.-M.; Zhou, Z.; Chan, A. S. C. J. Am. Chem. Soc. 2008, 130, 3304–3306. doi:10.1021/ja710555g
    Return to citation in text: [1] [2]
  64. Chu, L.; Wang, X.-C.; Moore, C. E.; Rheingold, A. L.; Yu, J.-Q. J. Am. Chem. Soc. 2013, 135, 16344–16347. doi:10.1021/ja408864c
    Return to citation in text: [1] [2]
  65. Wang, X.; Mei, T.-S.; Yu, J.-Q. J. Am. Chem. Soc. 2009, 131, 7520–7521. doi:10.1021/ja901352k
    Return to citation in text: [1] [2]
  66. Li, J.-J.; Mei, T.-S.; Yu, J.-Q. Angew. Chem., Int. Ed. 2008, 47, 6452–6455. doi:10.1002/anie.200802187
    Return to citation in text: [1] [2]
  67. Whitfield, S. R.; Sanford, M. S. J. Am. Chem. Soc. 2007, 129, 15142–15143. doi:10.1021/ja077866q
    Return to citation in text: [1] [2]
  68. Hull, K. L.; Anani, W. Q.; Sanford, M. S. J. Am. Chem. Soc. 2006, 128, 7134–7135. doi:10.1021/ja061943k
    Return to citation in text: [1] [2] [3]
  69. Powers, D. C.; Ritter, T. Nat. Chem. 2009, 1, 302–309. doi:10.1038/nchem.246
    Return to citation in text: [1] [2]
  70. Jia, X.; Zhang, S.; Wang, W.; Luo, F.; Cheng, J. Org. Lett. 2009, 11, 3120–3123. doi:10.1021/ol900934g
    Return to citation in text: [1] [2]
  71. Ma, S.; Villa, G.; Thuy-Boun, P. S.; Homs, A.; Yu, J.-Q. Angew. Chem., Int. Ed. 2014, 53, 734–737. doi:10.1002/anie.201305388
    Return to citation in text: [1] [2] [3]
  72. Hesp, K. D.; Bergman, R. G.; Ellman, J. A. J. Am. Chem. Soc. 2011, 133, 11430–11433. doi:10.1021/ja203495c
    Return to citation in text: [1] [2] [3]
  73. Xiao, B.; Fu, Y.; Xu, J.; Gong, T.-J.; Dai, J.-J.; Yi, J.; Liu, L. J. Am. Chem. Soc. 2010, 132, 468–469. doi:10.1021/ja909818n
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  74. Brand, J. P.; Charpentier, J.; Waser, J. Angew. Chem., Int. Ed. 2009, 48, 9346–9349. doi:10.1002/anie.200905419
    Return to citation in text: [1] [2] [3] [4]
  75. 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. J. Am. Chem. Soc. 2002, 124, 1586–1587. doi:10.1021/ja0176907
    Return to citation in text: [1] [2] [3]
  76. Ishiyama, T.; Takagi, J.; Hartwig, J. F.; Miyaura, N. Angew. Chem., Int. Ed. 2002, 41, 3056–3058. doi:10.1002/1521-3773(20020816)41:16<3056::AID-ANIE3056>3.0.CO;2-#
    Return to citation in text: [1] [2] [3]
  77. Campeau, L.-C.; Bertrand-Laperle, M.; Leclerc, J.-P.; Villemure, E.; Gorelsky, S.; Fagnou, K. J. Am. Chem. Soc. 2008, 130, 3276–3277. doi:10.1021/ja7107068
    Return to citation in text: [1] [2] [3]
  78. Chary, B. C.; Kim, S.; Park, Y.; Kim, J.; Lee, P. H. Org. Lett. 2013, 15, 2692–2695. doi:10.1021/ol4009987
    Return to citation in text: [1] [2] [3]
  79. Yang, F.; Song, F.; Li, W.; Lan, J.; You, J. RSC Adv. 2013, 3, 9649–9652. doi:10.1039/c3ra41981a
    Return to citation in text: [1] [2] [3]
  80. Kakiuchi, F.; Murai, S. Org. Synth. 2003, 80, 104–110. doi:10.15227/orgsyn.080.0104
    Return to citation in text: [1]
  81. Sonoda, M.; Kakiuchi, F.; Chatani, N.; Murai, S. Bull. Chem. Soc. Jpn. 1997, 70, 3117–3128. doi:10.1246/bcsj.70.3117
    Return to citation in text: [1]
  82. Kakiuchi, F.; Sekine, S.; Tanaka, Y.; Kamatani, A.; Sonoda, M.; Chatani, N.; Murai, S. Bull. Chem. Soc. Jpn. 1995, 68, 62–83. doi:10.1246/bcsj.68.62
    Return to citation in text: [1]
  83. Sonoda, M.; Kakiuchi, F.; Chatani, N.; Murai, S. J. Organomet. Chem. 1995, 504, 151–152. doi:10.1016/0022-328X(95)05607-Q
    Return to citation in text: [1]
  84. Murai, S.; Kakiuchi, F.; Sekine, S.; Tanaka, Y.; Kamatani, A.; Sonoda, M.; Chatani, N. Pure Appl. Chem. 1994, 66, 1527–1534. doi:10.1351/pac199466071527
    Return to citation in text: [1]
  85. Murai, S.; Kakiuchi, F.; Sekine, S.; Tanaka, Y.; Kamatani, A.; Sonoda, M.; Chatani, N. Nature 1993, 366, 529–531. doi:10.1038/366529a0
    Return to citation in text: [1]
  86. Yang, S.; Li, B.; Wan, X.; Shi, Z. J. Am. Chem. Soc. 2007, 129, 6066–6067. doi:10.1021/ja070767s
    Return to citation in text: [1] [2] [3]
  87. Li, B.-J.; Tian, S.-L.; Fang, Z.; Shi, Z.-J. Angew. Chem., Int. Ed. 2008, 47, 1115–1118. doi:10.1002/anie.200704092
    Return to citation in text: [1]
  88. Ackermann, L.; Vicente, R.; Althammer, A. Org. Lett. 2008, 10, 2299–2302. doi:10.1021/ol800773x
    Return to citation in text: [1]
  89. Kakiuchi, F.; Sato, T.; Yamauchi, M.; Chatani, N.; Murai, S. Chem. Lett. 1999, 28, 19–20. doi:10.1246/cl.1999.19
    Return to citation in text: [1]
  90. Desai, L. V.; Hull, K. L.; Sanford, M. S. J. Am. Chem. Soc. 2004, 126, 9542–9543. doi:10.1021/ja046831c
    Return to citation in text: [1]
  91. Thu, H.-Y.; Yu, W.-Y.; Che, C.-M. J. Am. Chem. Soc. 2006, 128, 9048–9049. doi:10.1021/ja062856v
    Return to citation in text: [1]
  92. Campeau, L.-C.; Schipper, D. J.; Fagnou, K. J. Am. Chem. Soc. 2008, 130, 3266–3267. doi:10.1021/ja710451s
    Return to citation in text: [1]
  93. Lazareva, A.; Daugulis, O. Org. Lett. 2006, 8, 5211–5213. doi:10.1021/ol061919b
    Return to citation in text: [1]
  94. Cai, G.; Fu, Y.; Li, Y.; Wan, X.; Shi, Z. J. Am. Chem. Soc. 2007, 129, 7666–7673. doi:10.1021/ja070588a
    Return to citation in text: [1]
  95. Kuninobu, Y.; Ida, H.; Nishi, M.; Kanai, M. Nat. Chem. 2015, 7, 712–717. doi:10.1038/nchem.2322
    Return to citation in text: [1]
  96. He, J.; Li, S.; Deng, Y.; Fu, H.; Laforteza, B. N.; Spangler, J. E.; Homs, A.; Yu, J.-Q. Science 2014, 343, 1216–1220. doi:10.1126/science.1249198
    Return to citation in text: [1]
  97. Iwai, T.; Harada, T.; Hara, K.; Sawamura, M. Angew. Chem., Int. Ed. 2013, 52, 12322–12326. doi:10.1002/anie.201306769
    Return to citation in text: [1]
  98. Ishiyama, T.; Isou, H.; Kikuchi, T.; Miyaura, N. Chem. Commun. 2010, 46, 159–161. doi:10.1039/B910298A
    Return to citation in text: [1]
  99. Wang, D.-H.; Engle, K. M.; Shi, B.-F.; Yu, J.-Q. Science 2010, 327, 315–319. doi:10.1126/science.1182512
    Return to citation in text: [1]
  100. Stang, E. M.; White, M. C. Nat. Chem. 2009, 1, 547–551. doi:10.1038/nchem.351
    Return to citation in text: [1]
  101. Lafrance, M.; Lopointe, D.; Fagnou, K. Tetrahedron 2008, 64, 6015–6020. doi:10.1016/j.tet.2008.01.057
    Return to citation in text: [1]
  102. Reed, S. A.; White, M. C. J. Am. Chem. Soc. 2008, 130, 3316–3318. doi:10.1021/ja710206u
    Return to citation in text: [1]
  103. Ackermann, L.; Althammer, A.; Born, R. Angew. Chem., Int. Ed. 2006, 45, 2619–2622. doi:10.1002/anie.200504450
    Return to citation in text: [1] [2]
  104. Saiki, T.; Nishio, Y.; Ishiyama, T.; Miyaura, N. Organometallics 2006, 25, 6068–6073. doi:10.1021/om050968+
    Return to citation in text: [1]
  105. Chen, M. S.; Prabagaran, N.; Labenz, N. A.; White, M. C. J. Am. Chem. Soc. 2005, 127, 6970–6971. doi:10.1021/ja0500198
    Return to citation in text: [1]
  106. Tsukada, N.; Mitsuboshi, T.; Setoguchi, H.; Inoue, Y. J. Am. Chem. Soc. 2003, 125, 12102–12103. doi:10.1021/ja0375075
    Return to citation in text: [1]
  107. Gray, J. B.; Cook, A. K.; Sanford, M. S. ACS Catal. 2013, 3, 700–703. doi:10.1021/cs300786j
    Return to citation in text: [1]
  108. Balcells, D.; Clot, E.; Eisenstein, O. Chem. Rev. 2010, 110, 749–823. doi:10.1021/cr900315k
    Return to citation in text: [1] [2]
  109. Powers, D. C.; Geibel, M. A. L.; Klein, J. E. M. N.; Ritter, T. J. Am. Chem. Soc. 2009, 131, 17050–17051. doi:10.1021/ja906935c
    Return to citation in text: [1] [2]
  110. Hull, K. L.; Sanford, M. S. J. Am. Chem. Soc. 2009, 131, 9651–9653. doi:10.1021/ja901952h
    Return to citation in text: [1] [2]
  111. García-Cuadrado, D.; de Mendoza, P.; Braga, A. A. C.; Maseras, F.; Echavarren, A. M. J. Am. Chem. Soc. 2007, 129, 6880–6886. doi:10.1021/ja071034a
    Return to citation in text: [1] [2]
  112. Lafrance, M.; Fagnou, K. J. Am. Chem. Soc. 2006, 128, 16496–16497. doi:10.1021/ja067144j
    Return to citation in text: [1] [2]
  113. Campeau, L.-C.; Fagnou, K. Chem. Commun. 2006, 1253–1264. doi:10.1039/b515481m
    Return to citation in text: [1] [2]
  114. García-Cuadrado, D.; Braga, A. A. C.; Maseras, F.; Echavarren, A. M. J. Am. Chem. Soc. 2006, 128, 1066–1067. doi:10.1021/ja056165v
    Return to citation in text: [1] [2]
  115. Davies, D. L.; Donald, S. M. A.; Macgregor, S. A. J. Am. Chem. Soc. 2005, 127, 13754–13755. doi:10.1021/ja052047w
    Return to citation in text: [1] [2]
  116. Maleckis, A.; Kampf, J. W.; Sanford, M. S. J. Am. Chem. Soc. 2013, 135, 6618–6625. doi:10.1021/ja401557m
    Return to citation in text: [1] [2]
  117. Dick, A. R.; Kampf, J. W.; Sanford, M. S. J. Am. Chem. Soc. 2005, 127, 12790–12791. doi:10.1021/ja0541940
    Return to citation in text: [1] [2]
  118. Biswas, B.; Sugimoto, M.; Sakaki, S. Organometallics 2000, 19, 3895–3908. doi:10.1021/om000002s
    Return to citation in text: [1] [2]
  119. Horino, H.; Inoue, N. J. Org. Chem. 1981, 46, 4416–4422. doi:10.1021/jo00335a019
    Return to citation in text: [1] [2]
  120. Zudin, V. N.; Chinakov, V. D.; Nekipelov, V. M.; Rogov, V. A.; Likholobov, V. A.; Yermakov, Yu. I. J. Mol. Catal. 1989, 52, 27–48. doi:10.1016/0304-5102(89)80080-X
    Return to citation in text: [1]
  121. Nishikata, T.; Abela, A. R.; Lipshutz, B. H. Angew. Chem., Int. Ed. 2010, 49, 781–784. doi:10.1002/anie.200905967
    Return to citation in text: [1] [2] [3] [4] [5] [6]
  122. Nishikata, T.; Abela, A. R.; Huang, S.; Lipshutz, B. H. J. Am. Chem. Soc. 2010, 132, 4978–4979. doi:10.1021/ja910973a
    Return to citation in text: [1] [2] [3] [4]
  123. Yamamoto, A. J. Organomet. Chem. 1995, 500, 337–348. doi:10.1016/0022-328X(95)00521-Q
    Return to citation in text: [1]
  124. Oi, S.; Kashiwagi, K.; Inoue, Y. Tetrahedron Lett. 1998, 39, 6253–6256. doi:10.1016/S0040-4039(98)01288-X
    Return to citation in text: [1]
  125. Oi, S.; Kashiwagi, K.; Terada, E.; Ohuchi, K.; Inoue, Y. Tetrahedron Lett. 1996, 37, 6351–6354. doi:10.1016/0040-4039(96)01369-X
    Return to citation in text: [1]
  126. Sodeoka, M.; Hamashima, Y. Chem. Commun. 2009, 5787–5798. doi:10.1039/b911015a
    Return to citation in text: [1]
  127. Sodeoka, M.; Hamashima, Y. Bull. Chem. Soc. Jpn. 2005, 78, 941–956. doi:10.1246/bcsj.78.941
    Return to citation in text: [1]
  128. Fujii, A.; Hagiwara, E.; Sodeoka, M. J. Am. Chem. Soc. 1999, 121, 5450–5458. doi:10.1021/ja9902827
    Return to citation in text: [1]
  129. Gaunt, M. J.; Spencer, J. B. Org. Lett. 2001, 3, 25–28. doi:10.1021/ol0066882
    Return to citation in text: [1]
  130. Chen, C.; Luo, S.; Jordan, R. F. J. Am. Chem. Soc. 2010, 132, 5273–5284. doi:10.1021/ja100491y
    Return to citation in text: [1]
  131. Brookhart, M.; Wagner, M. I.; Balavoine, G. G. A.; Haddou, H. A. J. Am. Chem. Soc. 1994, 116, 3641–3642. doi:10.1021/ja00087a077
    Return to citation in text: [1]
  132. Yamamoto, Y.; Nishikata, T.; Miyaura, N. Pure Appl. Chem. 2008, 80, 807–817. doi:10.1351/pac200880050807
    Return to citation in text: [1]
  133. Miyaura, N. Synlett 2009, 2039–2042. doi:10.1055/s-0029-1217555
    Return to citation in text: [1]
  134. Nishikata, T.; Yamamoto, Y.; Miyaura, N. Angew. Chem., Int. Ed. 2003, 42, 2768–2770. doi:10.1002/anie.200350888
    Return to citation in text: [1]
  135. Nishikata, T.; Yamamoto, Y.; Miyaura, N. Chem. Commun. 2004, 1822–1823. doi:10.1039/b407272c
    Return to citation in text: [1]
  136. Nishikata, T.; Yamamoto, Y.; Miyaura, N. Chem. Lett. 2003, 32, 752–753. doi:10.1246/cl.2003.752
    Return to citation in text: [1] [2] [3]
  137. Lebrasseur, N.; Larrosa, I. J. Am. Chem. Soc. 2008, 130, 2926–2927. doi:10.1021/ja710731a
    Return to citation in text: [1]
  138. Stuart, D. R.; Fagnou, K. Science 2007, 316, 1172–1175. doi:10.1126/science.1141956
    Return to citation in text: [1]
  139. Deprez, N. R.; Kalyani, D.; Krause, A.; Sanford, M. S. J. Am. Chem. Soc. 2006, 128, 4972–4973. doi:10.1021/ja060809x
    Return to citation in text: [1]
  140. Hughes, C. C.; Trauner, D. Angew. Chem., Int. Ed. 2002, 41, 1569–1572. doi:10.1002/1521-3773(20020503)41:9<1569::AID-ANIE1569>3.0.CO;2-8
    Return to citation in text: [1]
  141. Jia, C.; Piao, D.; Oyamada, J.; Lu, W.; Kitamura, T.; Fujiwara, Y. Science 2000, 287, 1992–1995. doi:10.1126/science.287.5460.1992
    Return to citation in text: [1]
  142. Pivsa-Art, S.; Satoh, T.; Kawamura, Y.; Miura, M.; Nomura, M. Bull. Chem. Soc. Jpn. 1998, 71, 467–473. doi:10.1246/bcsj.71.467
    Return to citation in text: [1]
  143. Albano, V. G.; Di Serio, M.; Monari, M.; Orabona, I.; Panunzi, A.; Ruffo, F. Inorg. Chem. 2002, 41, 2672–2677. doi:10.1021/ic011059p
    Return to citation in text: [1]
  144. Liston, D. J.; Lee, Y. J.; Scheidt, W. R.; Reed, C. A. J. Am. Chem. Soc. 1989, 111, 6643–6648. doi:10.1021/ja00199a025
    Return to citation in text: [1]
  145. Davis, J. A.; Hartley, E. R.; Muray, S. G. J. Chem. Soc., Dalton Trans. 1980, 2246–2249. doi:10.1039/dt9800002246
    Return to citation in text: [1]
  146. Amatore, C.; Jutand, A.; Mederious, M. J.; Mottier, L. J. Electroanal. Chem. 1997, 422, 125–132. doi:10.1016/S0022-0728(96)04901-7
    Return to citation in text: [1]
  147. Grushin, V. V. Chem. Rev. 1996, 96, 2011–2034. doi:10.1021/cr950272y
    Return to citation in text: [1]
  148. Nishikata, T.; Lipshutz, B. H. Org. Lett. 2010, 12, 1972–1975. doi:10.1021/ol100331h
    Return to citation in text: [1] [2]
  149. Ricci, P.; Krämer, K.; Cambeiro, X. C.; Larrosa, I. J. Am. Chem. Soc. 2013, 135, 13258–13261. doi:10.1021/ja405936s
    Return to citation in text: [1]
  150. Join, B.; Yamamoto, T.; Itami, K. Angew. Chem., Int. Ed. 2009, 48, 3644–3647. doi:10.1002/anie.200806358
    Return to citation in text: [1] [2]
  151. Campeau, L.-C.; Stuart, D. R.; Leclerc, J.-P.; Bertrand-Laperle, M.; Villemure, E.; Sun, H.-Y.; Lasserre, S.; Guimond, N.; Lecavallier, M.; Fagnou, K. J. Am. Chem. Soc. 2009, 131, 3291–3306. doi:10.1021/ja808332k
    Return to citation in text: [1]
  152. Yang, F.; Wu, Y.; Li, Y.; Wang, B.; Zhang, J. Tetrahedron 2009, 65, 914–919. doi:10.1016/j.tet.2008.11.001
    Return to citation in text: [1]
  153. Scarborough, C. C.; McDonald, R. I.; Hartmann, C.; Sazama, G. T.; Bergant, A.; Stahl, S. S. J. Org. Chem. 2009, 74, 2613–2615. doi:10.1021/jo802632v
    Return to citation in text: [1]
  154. Kim, J.; Jo, M.; So, W.; No, Z. Tetrahedron Lett. 2009, 50, 1229–1362. doi:10.1016/j.tetlet.2009.01.010
    Return to citation in text: [1]
  155. Gorelsky, S. I.; Lapointe, D.; Fagnou, K. J. Am. Chem. Soc. 2008, 130, 10848–10849. doi:10.1021/ja802533u
    Return to citation in text: [1]
  156. Li, D.-D.; Yuan, T.-T.; Wang, G.-W. J. Org. Chem. 2012, 77, 3341–3347. doi:10.1021/jo300126n
    Return to citation in text: [1]
  157. Chen, F.; Min, Q.-Q.; Zhang, X. J. Org. Chem. 2012, 77, 2992–2998. doi:10.1021/jo300036d
    Return to citation in text: [1]
  158. Liang, Z.; Feng, R.; Yin, H.; Zhang, Y. Org. Lett. 2013, 15, 4544–4547. doi:10.1021/ol402207g
    Return to citation in text: [1]
  159. Shi, B.-S.; Zhang, Y.-H.; Lam, J. K.; Wang, D.-H.; Yu, J.-Q. J. Am. Chem. Soc. 2010, 132, 460–461. doi:10.1021/ja909571z
    Return to citation in text: [1]
  160. Chu, J.-H.; Tsai, S.-L.; Wu, M.-J. Synthesis 2009, 3757–3764. doi:10.1055/s-0029-1217014
    Return to citation in text: [1] [2]
  161. Shi, B.-F.; Maugel, N.; Zhang, Y.-H.; Yu, J.-Q. Angew. Chem., Int. Ed. 2008, 47, 4882–4886. doi:10.1002/anie.200801030
    Return to citation in text: [1]
  162. Wang, D.-H.; Wasa, M.; Giri, R.; Yu, J.-Q. J. Am. Chem. Soc. 2008, 130, 7190–7191. doi:10.1021/ja801355s
    Return to citation in text: [1]
  163. Wang, D.-H.; Mei, T.-S.; Yu, J.-Q. J. Am. Chem. Soc. 2008, 130, 17676–17677. doi:10.1021/ja806681z
    Return to citation in text: [1] [2]
  164. Kirchberg, S.; Vogler, T.; Studer, A. Synlett 2008, 2841–2845. doi:10.1055/s-0028-1083546
    Return to citation in text: [1]
  165. Vogler, T.; Studer, A. Org. Lett. 2008, 10, 129–131. doi:10.1021/ol702659a
    Return to citation in text: [1]
  166. Giri, R.; Maugel, N.; Li, J.-J.; Wang, D.-H.; Breazzano, S. P.; Saunders, L. B.; Yu, J.-Q. J. Am. Chem. Soc. 2007, 129, 3510–3511. doi:10.1021/ja0701614
    Return to citation in text: [1]
  167. Chen, X.; Goodhue, C. E.; Yu, J.-Q. J. Am. Chem. Soc. 2006, 128, 12634–12635. doi:10.1021/ja0646747
    Return to citation in text: [1]
  168. Kakiuchi, F.; Kan, S.; Igi, K.; Chatani, N.; Murai, S. J. Am. Chem. Soc. 2003, 125, 1698–1699. doi:10.1021/ja029273f
    Return to citation in text: [1] [2]
  169. Liang, Z.; Yao, J.; Wang, K.; Li, H.; Zhang, Y. Chem. – Eur. J. 2013, 19, 16825–16831. doi:10.1002/chem.201301229
    Return to citation in text: [1]
  170. Meng, X.; Kim, S. J. Org. Chem. 2013, 78, 11247–11251. doi:10.1021/jo401716p
    Return to citation in text: [1]
  171. Yamaguchi, K.; Kondo, H.; Yamaguchi, J.; Itami, K. Chem. Sci. 2013, 4, 3753–3757. doi:10.1039/c3sc51206a
    Return to citation in text: [1]
  172. Tremont, S. J.; Ur Rahman, H. J. Am. Chem. Soc. 1984, 106, 5759–5760. doi:10.1021/ja00331a073
    Return to citation in text: [1] [2]
  173. Shabashov, D.; Daugulis, O. J. Org. Chem. 2007, 72, 7720–7725. doi:10.1021/jo701387m
    Return to citation in text: [1] [2]
  174. Daugulis, O.; Zaitsev, V. G. Angew. Chem., Int. Ed. 2005, 44, 4046–4048. doi:10.1002/anie.200500589
    Return to citation in text: [1] [2] [3]
  175. Kalyani, D.; Deprez, N. R.; Desai, L. V.; Sanford, M. S. J. Am. Chem. Soc. 2005, 127, 7330–7331. doi:10.1021/ja051402f
    Return to citation in text: [1] [2]
  176. Houlden, C. E.; Hutchby, M.; Bailey, C. D.; Ford, J. G.; Tyler, S. N. G.; Gagné, M. R.; Lloyd-Jones, G. C.; Booker-Milburn, K. I. Angew. Chem., Int. Ed. 2009, 48, 1830–1833. doi:10.1002/anie.200805842
    Return to citation in text: [1]
  177. Nishikata, T.; Lipshutz, B. H. J. Am. Chem. Soc. 2009, 131, 12103–12105. doi:10.1021/ja905082c
    Return to citation in text: [1]
  178. Nishikata, T.; Lipshutz, B. H. Org. Lett. 2009, 11, 2377–2379. doi:10.1021/ol900235s
    Return to citation in text: [1] [2]
  179. Lipshutz, B. H.; Ghorai, S. Aldrichimica Acta 2008, 41, 58–72.
    Return to citation in text: [1]
  180. Ohnmacht, S. A.; Mamone, P.; Culshaw, A. J.; Greaney, M. F. Chem. Commun. 2008, 1241–1243. doi:10.1039/b719466h
    Return to citation in text: [1]
  181. Flegeau, E. F.; Popkin, M. E.; Greaney, M. F. Org. Lett. 2008, 10, 2717–2720. doi:10.1021/ol800869g
    Return to citation in text: [1]
  182. Turner, G. L.; Morris, J. A.; Greaney, M. F. Angew. Chem., Int. Ed. 2007, 46, 7996–8000. doi:10.1002/anie.200702141
    Return to citation in text: [1]
  183. Herrerías, C. I.; Yao, X.; Li, Z.; Li, C.-J. Chem. Rev. 2007, 107, 2546–2562. doi:10.1021/cr050980b
    Return to citation in text: [1]
  184. Mitsudo, K.; Kaide, T.; Nakamoto, E.; Yoshida, K.; Tanaka, H. J. Am. Chem. Soc. 2007, 129, 2246–2247. doi:10.1021/ja069043r
    Return to citation in text: [1]
  185. Oi, S.; Aizawa, E.; Ogino, Y.; Inoue, Y. J. Org. Chem. 2005, 70, 3113–3119. doi:10.1021/jo050031i
    Return to citation in text: [1]
  186. Bedford, R. B.; Limmert, M. E. J. Org. Chem. 2003, 68, 8669–8682. doi:10.1021/jo030157k
    Return to citation in text: [1]
  187. Oi, S.; Watanabe, S.-i.; Fukita, S.; Inoue, Y. Tetrahedron Lett. 2003, 44, 8665–8668. doi:10.1016/j.tetlet.2003.09.151
    Return to citation in text: [1]
  188. Oi, S.; Fukita, S.; Hirata, N.; Watanuki, N.; Miyano, S.; Inoue, Y. Org. Lett. 2001, 3, 2579–2581. doi:10.1021/ol016257z
    Return to citation in text: [1]
  189. Oi, S.; Fukita, S.; Inoue, Y. Chem. Commun. 1998, 2439–2440. doi:10.1039/a806790b
    Return to citation in text: [1]
  190. Fujiwara, Y.; Moritani, I.; Danno, S.; Asano, R.; Teranishi, S. J. Am. Chem. Soc. 1969, 91, 7166–7169. doi:10.1021/ja01053a047
    Return to citation in text: [1]
  191. Fujiwara, Y.; Moritani, I.; Matsuda, M.; Teranishi, S. Tetrahedron Lett. 1968, 9, 633–636. doi:10.1016/S0040-4039(01)98820-3
    Return to citation in text: [1]
  192. Moritani, S.; Fujiwara, Y. Tetrahedron Lett. 1967, 8, 1119–1122. doi:10.1016/S0040-4039(00)90648-8
    Return to citation in text: [1]
  193. Jiao, L.-Y.; Oestreich, M. Org. Lett. 2013, 15, 5374–5377. doi:10.1021/ol402687t
    Return to citation in text: [1] [2]
  194. Wang, L.; Liu, S.; Li, Z.; Yu, Y. Org. Lett. 2011, 13, 6137–6139. doi:10.1021/ol202738j
    Return to citation in text: [1] [2]
  195. Rauf, W.; Thompson, A. L.; Brown, J. M. Dalton Trans. 2010, 39, 10414–10421. doi:10.1039/c0dt00378f
    Return to citation in text: [1] [2]
  196. Wu, J.; Cui, X.; Chen, L.; Jiang, G.; Wu, Y. J. Am. Chem. Soc. 2009, 131, 13888–13889. doi:10.1021/ja902762a
    Return to citation in text: [1] [2]
  197. Aouf, C.; Thiery, E.; Le Bras, J.; Muzart, J. Org. Lett. 2009, 11, 4096–4099. doi:10.1021/ol901570p
    Return to citation in text: [1] [2]
  198. Cheng, D.; Gallagher, T. Org. Lett. 2009, 11, 2639–2641. doi:10.1021/ol900627q
    Return to citation in text: [1] [2]
  199. Cho, S. H.; Hwang, S. J.; Chang, S. J. Am. Chem. Soc. 2008, 130, 9254–9256. doi:10.1021/ja8026295
    Return to citation in text: [1] [2]
  200. Dams, M.; De Vos, D. E.; Celen, S.; Jacobs, P. A. Angew. Chem., Int. Ed. 2003, 42, 3512–3515. doi:10.1002/anie.200351524
    Return to citation in text: [1] [2]
  201. Yokota, T.; Tani, M.; Sakaguchi, S.; Ishii, Y. J. Am. Chem. Soc. 2003, 125, 1476–1477. doi:10.1021/ja028903a
    Return to citation in text: [1] [2]
  202. Weissman, H.; Song, X.; Milstein, D. J. Am. Chem. Soc. 2001, 123, 337–338. doi:10.1021/ja003361n
    Return to citation in text: [1] [2]
  203. Meng, X.; Kim, S. Org. Lett. 2013, 15, 1910–1913. doi:10.1021/ol400565r
    Return to citation in text: [1] [2]
  204. Dai, H.-X.; Li, G.; Zhang, X.-G.; Stephan, A. F.; Yu, J.-Q. J. Am. Chem. Soc. 2013, 135, 7567–7571. doi:10.1021/ja400659s
    Return to citation in text: [1]
  205. Zhang, Y.-H.; Shi, B.-F.; Yu, J.-Q. J. Am. Chem. Soc. 2009, 131, 5072–5074. doi:10.1021/ja900327e
    Return to citation in text: [1]
  206. Rauf, W.; Thompson, A. L.; Brown, J. M. Chem. Commun. 2009, 3874–3876. doi:10.1039/b905717j
    Return to citation in text: [1] [2] [3]
  207. Rouhi, A. M. Chem. Eng. News 2004, 82, 49–58.
    Return to citation in text: [1]
  208. Felpin, F.-X.; Fouquet, E.; Zakri, C. Adv. Synth. Catal. 2009, 351, 649–650. doi:10.1002/adsc.200800783
    Return to citation in text: [1] [2]
  209. Wetzel, A.; Ehrhardt, V.; Heinrich, M. R. Angew. Chem., Int. Ed. 2008, 47, 9130–9133. doi:10.1002/anie.200803785
    Return to citation in text: [1] [2]
  210. Spivey, A. C.; Tseng, C.-C.; Hannah, J. P.; Gripton, C. J. P.; de Fraine, P.; Parr, N. J.; Scicinski, J. J. Chem. Commun. 2007, 2926–2928. doi:10.1039/B707517K
    Return to citation in text: [1] [2]
  211. Ehrenfreund, J.; Lamberth, C.; Tobler, H.; Walter, H. Biphenyl derivatives and their use as fungicides. WO Patent WO 2004058723, July 15, 2004.
    Return to citation in text: [1] [2]
  212. Eicken, K.; Rack, M.; Wetterich, F.; Ammermann, E.; Lorenz, G.; Strathmann, S. Biphenylamide. German Patent DE 19735224, Feb 18, 1999.
    Return to citation in text: [1] [2]
  213. Eicken, K.; Rang, H.; Harreus, A.; Goetz, N.; Ammermann, E.; Lorenz, G.; Strathmann, S. Bisphenylamide. German Patent DE19531813, March 6, 1997.
    Return to citation in text: [1] [2]
  214. Urriolabeitia, E. P. Oxidative Addition and Transmetallation. In Palladacycles: Synthesis, Characterization and Applications; Dupont, J.; Pfeffer, M., Eds.; Wiley-VCH: Weinheim, Germany, 2008.
    Return to citation in text: [1]
  215. Roiban, G.-D.; Serrano, E.; Soler, T.; Contel, M.; Grosu, I.; Cativiela, C.; Urriolabeitia, E. P. Organometallics 2010, 29, 1428–1435. doi:10.1021/om901068f
    Return to citation in text: [1] [2] [3]
  216. The palladacycle is very stable and did not release the corresponding arylurea after 2–3 weeks in wet solvent.
    Return to citation in text: [1]
  217. Steffen, W. L.; Palenik, G. J. Inorg. Chem. 1976, 15, 2432–2439. doi:10.1021/ic50164a025
    Return to citation in text: [1]
  218. Ludwig, M.; Strömberg, S.; Svensson, M.; Åkermark, B. Organometallics 1999, 18, 970–975. doi:10.1021/om9803265
    Return to citation in text: [1]
  219. Nishikata, T.; Yamamoto, Y.; Gridnev, I. D.; Miyaura, N. Organometallics 2005, 24, 5025–5032. doi:10.1021/om050678t
    Return to citation in text: [1] [2]
  220. Nishikata, T.; Yamamoto, Y.; Miyaura, N. Organometallics 2004, 23, 4317–4324. doi:10.1021/om0498044
    Return to citation in text: [1] [2]
  221. Albrecht, M. Chem. Rev. 2010, 110, 576–623. doi:10.1021/cr900279a
    Return to citation in text: [1]
  222. Chatani, N., Ed. Directed Metallation; Topics in Organometallic Chemistry, Vol. 24; Springer: Berlin, Germany, 2007.
    Return to citation in text: [1]
  223. Yu, J.-Q.; Giri, R.; Chen, X. Org. Biomol. Chem. 2006, 4, 4041–4047. doi:10.1039/B611094K
    Return to citation in text: [1]
  224. Dupont, J.; Consorti, C. S.; Spencer, J. Chem. Rev. 2005, 105, 2527–2572. doi:10.1021/cr030681r
    Return to citation in text: [1]
  225. Chu, J.-H.; Chen, C.-C.; Wu, M.-J. Organometallics 2008, 27, 5173–5176. doi:10.1021/om800606c
    Return to citation in text: [1]
  226. Pérez-Rodríguez, M.; Braga, A. A. C.; Garcia-Melchor, M.; Pérez-Temprano, M. H.; Casares, J. A.; Ujaque, G.; de Lera, A. R.; Alvarez, R.; Maseras, F.; Espinet, P. J. Am. Chem. Soc. 2009, 131, 3650–3657. doi:10.1021/ja808036j
    Return to citation in text: [1]
  227. Albéniz, A. C.; Espinet, P.; Martín-Ruiz, B. Chem. – Eur. J. 2001, 7, 2481–2485. doi:10.1002/1521-3765(20010601)7:11<2481::AID-CHEM24810>3.0.CO;2-2
    Return to citation in text: [1]
  228. Szabó, K. J. Organometallics 1998, 17, 1677–1686. doi:10.1021/om9800344
    Return to citation in text: [1]
  229. Bäckvall, J. E.; Byström, S. E.; Nordberg, R. E. J. Org. Chem. 1984, 49, 4619–4631. doi:10.1021/jo00198a010
    Return to citation in text: [1]
  230. Temple, J. S.; Riediker, M.; Schwartz, J. J. Am. Chem. Soc. 1982, 104, 1310–1315. doi:10.1021/ja00369a028
    Return to citation in text: [1]
  231. 1,4-Dihydroxybenzene was confirmed by 1H, and 13C NMR and GC–MS analysis.
    Return to citation in text: [1]
  232. Some previous C–H activation reactions catalyzed by Pd(OAc)2 require high temperature to generate a cationic species. TFA or p-TsOH led to room temperature C–H activation only when catalyzed by Pd(OAc)2 probably due to generation of a cationic Pd(II) species.
    Return to citation in text: [1]
  233. Crociani, B.; Di Bianca, F.; Uguagliati, P.; Canovese, L.; Berton, A. J. Chem. Soc., Dalton Trans. 1991, 71–79. doi:10.1039/DT9910000071
    Return to citation in text: [1]
  234. Eaborn, C.; Odell, K. J.; Pidcock, A. J. Chem. Soc., Dalton Trans. 1978, 357–368. doi:10.1039/DT9780000357
    Return to citation in text: [1]
  235. Foresee, L. N.; Tunge, J. A. Organometallics 2005, 24, 6440–6444. doi:10.1021/om0507225
    Return to citation in text: [1]
  236. Jia, C.; Lu, W.; Oyamada, J.; Kitamura, T.; Matsuda, K.; Irie, M.; Fujiwara, Y. J. Am. Chem. Soc. 2000, 122, 7252–7263. doi:10.1021/ja0005845
    Return to citation in text: [1]
  237. We have not yet obtained strong proof of this species based on either a kinetic or an isotope effect study.
    Return to citation in text: [1]
  238. Kawataka, F.; Shimizu, I.; Yamamoto, A. Bull. Chem. Soc. Jpn. 1995, 68, 654–660. doi:10.1246/bcsj.68.654
    Return to citation in text: [1]
  239. Karabelas, K.; Hallberg, A. J. Org. Chem. 1986, 51, 5286–5290. doi:10.1021/jo00376a044
    Return to citation in text: [1]
  240. Karabelas, K.; Westerlund, C.; Hallberg, A. J. Org. Chem. 1985, 50, 3896–3900. doi:10.1021/jo00220a042
    Return to citation in text: [1]
  241. Molander, G. A.; Canturk, B. Angew. Chem., Int. Ed. 2009, 48, 9240–9261. doi:10.1002/anie.200904306
    Return to citation in text: [1] [2]
  242. Martin, R.; Buchwald, S. L. Acc. Chem. Res. 2008, 41, 1461. doi:10.1021/ar800036s
    Return to citation in text: [1] [2]
  243. Miyaura, N.; Suzuki, A. Chem. Rev. 1995, 95, 2457–2483. doi:10.1021/cr00039a007
    Return to citation in text: [1] [2]
  244. Sicre, C.; Braga, A. A. C.; Maseras, F.; Magdalena Cid, M. Tetrahedron 2008, 64, 7437–7650. doi:10.1016/j.tet.2008.05.018
    Return to citation in text: [1]
  245. Braga, A. A. C.; Morgon, N. H.; Ujaque, G.; Maseras, F. J. Am. Chem. Soc. 2005, 127, 9298–9307. doi:10.1021/ja050583i
    Return to citation in text: [1]
  246. Matos, K.; Soderquist, J. A. J. Org. Chem. 1998, 63, 461–470. doi:10.1021/jo971681s
    Return to citation in text: [1]
  247. Pantcheva, I.; Nishihara, Y.; Osakada, K. Organometallics 2005, 24, 3815–3817. doi:10.1021/om049050t
    Return to citation in text: [1]
  248. Zhao, P.; Incarvito, C. D.; Hartwig, J. F. J. Am. Chem. Soc. 2007, 129, 1876–1877. doi:10.1021/ja068587q
    Return to citation in text: [1]
  249. Deprez, N. R.; Sanford, M. S. J. Am. Chem. Soc. 2009, 131, 11234–11241. doi:10.1021/ja904116k
    Return to citation in text: [1]
  250. Cotton, F. A.; Koshevoy, I. O.; Lahuerta, P.; Murillo, C. A.; Sanaú, M.; Ubeda, M. A.; Zhao, Q. J. Am. Chem. Soc. 2006, 128, 13674–13675. doi:10.1021/ja0656595
    Return to citation in text: [1]
  251. Powers, D. C.; Ritter, T. Palladium(III) in Synthesis and Catalysis. In Higher Oxidation State Organopalladium and Platinum Chemistry; Canty, A. J., Ed.; Topics in Organometallic Chemistry, Vol. 35; Springer: Berlin, Germany, 2011; pp 129–156. doi:10.1007/978-3-642-17429-2_6
    Return to citation in text: [1] [2]
  252. Powers, D. C.; Xiao, D. Y.; Geibel, M. A. L.; Ritter, T. J. Am. Chem. Soc. 2010, 132, 14530–14536. doi:10.1021/ja1054274
    Return to citation in text: [1]
  253. Sehnal, P.; Taylor, R. J. K.; Fairlamb, I. J. S. Chem. Rev. 2010, 110, 824–889. doi:10.1021/cr9003242
    Return to citation in text: [1]
  254. Racowski, J. M.; Dick, A. R.; Sanford, M. S. J. Am. Chem. Soc. 2009, 131, 10974–10983. doi:10.1021/ja9014474
    Return to citation in text: [1]
  255. Catellani, M.; Motti, E.; Della Ca’, N. Acc. Chem. Res. 2008, 41, 1512–1522. doi:10.1021/ar800040u
    Return to citation in text: [1]
Other Beilstein-Institut Open Science Activities