Active Palladium Colloids via Palladacycle Degradation as Efficient

Jan 24, 2017 - (23) The Fourier transforms (FTs) of the EXAFS data for the Pd colloids are presented in Figure 7. In the R-space, two main peaks ... (...
0 downloads 3 Views 7MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article http://pubs.acs.org/journal/acsodf

Active Palladium Colloids via Palladacycle Degradation as Efficient Catalysts for Oxidative Homocoupling and Cross-Coupling of Aryl Boronic Acids Vaibhav Sable,† Karan Maindan,† Anant R. Kapdi,*,† Pushkar Sudhakar Shejwalkar,‡ and Kenji Hara‡ †

Department of Chemistry, Institute of Chemical Technology, Nathalal Parekh Road, Matunga, Mumbai 400019, India Department of Applied Chemistry, School of Engineering, Tokyo University of Technology, Hachioji, Tokyo 192-0982, Japan



S Supporting Information *

ABSTRACT: Active palladium colloids formed upon degradation of a palladacyclic complex (Herrmann−Beller 1) have been isolated for the first time and thoroughly characterized with techniques such as transmission electron microscopy (TEM), high-resolution TEM, X-ray photoelectron spectroscopy (XPS), and extended X-ray absorption fine structure spectroscopy. The synthesized palladium colloids have been utilized as efficient catalysts for the oxidative homocoupling of aryl boronic acids. Cross-coupling of two different aryl boronic acids has also been made possible using these active palladium colloids. This is the first report of this kind of coupling between aryl boronic acids.



Pd(0).9 However, one of the strongest arguments to date, based on experimental evidence, is about the presence of soluble palladium colloids/nanoparticles.10 Loss of activity for an immobilized palladacyclic catalyst by Nowotny,11a along with the observation of induction period in the catalytic reaction followed by sigmoidal kinetic profile by Beletskaya and co-workers,11b has offered support for such assumptions. Several others have also provided evidence to highlight nanoparticular/colloidal involvement, notably, Gladysz and co-workers,11c,d through their report on transmission electron microscopy (TEM) confirmation of Pd colloids. These initial reports on the presence of Pd nanoparticles/colloids brought about a flurry of activity, leading to the development of palladacyclic complexes, which could act as sources of highly active palladium colloids. Accordingly, a recent review by Alonso and Najera12 elegantly highlights the advantages and applicability of oxime-derived palladacycles as sources of active palladium colloids/nanoparticles. Besides these reports, pioneering work by Dupont on the kinetic and mechanistic aspects of the Heck reaction using the palladacycle {Pd[κ1-C, κ1-N-Cd(C6H5)C(Cl)CH2NMe2](μCl)}2 has also revealed the presence of highly active Pd(0) colloids.13 The presence of Pd colloids was confirmed by performing catalyst-poisoning studies, TEM analysis, and timeresolved ultraviolet−visible analysis. Although these studies have provided evidence based on TEM, mass spectroscopy (MS), and sometimes extended X-ray absorption fine structure (EXAFS) analysis of the catalytic

INTRODUCTION In the quest for developing highly efficient and robust Pd-based catalysts for C−C bond formation via cross-coupling reactions,1 palladacycles have played a significant role and, in recent years, found a wide variety of applications, ranging from catalysts2 for cross-coupling and related reactions to their more recent application as anticancer agents.3 Early examples for the use of palladacyclic complexes in catalysis employed azobenzene and hydrazobenzene as coordinating ligands, finding applications in processes such as selective reduction of alkenes, alkynes, or nitroalkanes.4 Over the years, with the increasing demand for more robust, air-stable, and catalytically active catalysts for addressing synthetically challenging processes, palladacyclic complexes have come forth as a possible solution. Since the reports by Herrmann and Beller (the Herrmann−Beller complex is henceforth referred to as palladacycle 1) of a highly active palladacyclic complex formed by the reaction of Pd(OAc)2 and P(o-Tol)3, an exponential growth in examples related to their use in cross-coupling reactions has been observed.5 Although the reactivity exhibited by the palladacycles has been exceptional, the current challenge is to apply them toward solving more important synthetic problems. Although the synthetic utility of palladacycles has been well explored, for many years, the unique reactivity profile of these catalytic systems was attributed to the presence of either the Pd(II)/Pd(IV)5a,6 or Pd(0)/Pd(II)7 catalytic cycle. However, the noteworthy work by Jutand and co-workers8 needs a special mention in this regard, providing useful insight into the possible mechanistic intermediates in these processes. Another proposal that has gained lot of interest in recent years is the possibility of palladacycles acting as reservoirs for © 2017 American Chemical Society

Received: October 19, 2016 Accepted: January 9, 2017 Published: January 24, 2017 204

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217

ACS Omega

Article

Scheme 1. Screening Study for Diarylmethanol Synthesis

solutions,14 no data to date has been presented on the isolation of colloids as well as their applications. To address this gap in the literature, we present herein the isolation of highly active palladium colloids/nanoparticles stabilized on carbon, formed through the degradation of palladacycle 1. Support for the formation of colloids has been provided using techniques such as TEM, high-resolution TEM (HRTEM), X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), and EXAFS. The isolated palladium colloids have demonstrated excellent activity in the oxidative homocoupling and cross-coupling of aryl boronic acids.

Although, the best results were obtained on changing the solvent from PhMe to CH3CN/H2O (for a complete screening study, see the Supporting Information). In all cases, a catalytic amount of CHCl3 was added, in line with recent reports on the beneficial effects of its addition through activation of the Pdcatalyst toward such types of reactions.17 Chloroform when used as the solvent also gave comparable yields (86%) to those of CH3CN/H2O, and it could also be used as the reaction solvent. To assess the applicability of the developed protocol, a scope study was undertaken using palladacycle 1 in the CH3CN/H2O solvent system at 80 °C (Scheme 2). First, we investigated the effect of electronics on the aryl aldehyde on the catalytic reaction. Electron-accepting substituents favored arylation of the corresponding aryl aldehyde, smoothly furnishing the desired products in good yields (4a−d; Scheme 2). Interestingly, electron-releasing substituents led to complete reduction of product formation, possibly due to the poor reactivity of the respective aldehydes. In contrast, employing a variety of differently substituted aryl boronic acids proved to be beneficial in obtaining good yields of diarylmethanols (4e−j; Scheme 2). Colloidal/Nanoparticular Pathway Investigation. While following the catalytic reactions, it was observed that



RESULTS AND DISCUSSION Our interest in exploring the possibility of conversion of palladacycles to active colloids came about as a part of our studies on the application of Herrmann−Beller palladacycle 1 for diarylmethanol synthesis. Most of these processes are catalyzed by transition metals, namely, Fe, Cu, Ti, Rh, and Pd, with the latter showing enhanced reactivity.15,16 Our study on diarylmethanol synthesis from palladacycle 1 was initiated by performing a screening study for the reaction of an aryl aldehyde with aryl boronic acid using different catalyst systems in PhMe (Scheme 1). 205

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217

ACS Omega

Article

Scheme 2. Substrate Scope for Palladium-Catalyzed Diarylmethanol Synthesisa,b

a

Aldehyde (1.0 mmol), aryl boronic acid (2.0 mmol), palladacycle 1 (1.0 mol %), CH3CN/H2O (1:1 v/v, 2.0 mL), K2CO3 (2.0 equiv.), CHCl3 (0.1 mL) at 80 °C for 24 h. bIsolated yields after column chromatography.

24 h. On completion of the reaction, analysis of the catalytic products showed slight inhibition (39% product obtained), suggesting the possible presence of homogeneous Pd species or soluble Pd colloids (Scheme 3). All of the above results strongly suggest the presence of active palladium nanoparticles or colloids. However, the presence of homotopic Pd catalysts also cannot be excluded, and further analysis is needed to ascertain the exact nature of the catalytic species. TEM Studies of Catalytic Solution. To ascertain the involvement of palladium colloids/nanoparticles, a detailed TEM study was undertaken. This was done by directly subjecting the general catalytic solution to TEM analysis after a reaction time of 24 h, represented in Scheme 2, involving the reaction between 4-chlorobenzaldehyde and phenylboronic acid, by simple deposition of the solution on carbon-coated copper grids. Palladium colloids approximately 2−4 nm in size were observed to be present in the catalytic solution (Figure 1a). Further confirmation for the involvement of palladium colloids was obtained by trapping these colloids in the catalytic reaction with PEG21 (added at the beginning of the reaction). Because of the stabilized nature of the colloids, they appear to be well-dispersed compared with the nonstabilized colloidal particles. The size of the colloidal particles was found to be in the approximate range of 3−5 nm (Figure 1b). Similar observations were also made during the HRTEM analysis of mercury-engulfed Pd colloids obtained after the mercury drop test. Energy-dispersive X-ray analysis clearly indicated the presence of mercury, with the size of the colloids increasing up to ∼4.0 nm (Figure 1c). Isolation and TEM Characterization of Palladium Colloids. Next, we turned our attention toward the possible

in almost all cases black particulate matter was formed, pointing toward the possible formation of palladium colloids/nanoparticles. To investigate this further, we performed a series of catalyst-poisoning experiments. The first of this series was an experiment that is popular and was commonly employed by Whitesides in the early 80s. It encompasses poisoning the catalytic solutions with Hg, which results in amalgamation of the colloidal palladium species.18 To our excitement, complete inhibition of the catalytic reaction was observed. This alone cannot be taken as absolute evidence for a nanoparticular pathway, and further analysis was undertaken to ascertain this assumption. Carbon disulfide (addition of excess of CS2) has also been employed as catalyst poison,18b,c resulting in total inhibition of the catalytic reaction. The addition of tetrabutylammonium bromide18c as a nanoparticular stabilizer gave identical yields of the coupling product. The Collman test19 has been the method of choice for detecting the nanoparticular involvement in palladacyclic systems for the past several years. However, the nonavailability of polyethyleneglycol-bound benzaldehydes or aryl boronic acids prompted us to run the Crabtree test instead.20 This test involves the specific use of a homogeneous species inhibitor, dibenzo[a,e]cyclooctatetraene (DCT), as the catalyst poison. The employment of such a sterically demanding poison would enable selective inhibition of any homogeneous species present in the catalytic solution via formation of a strong Pd−DCT complex, whereas in heterogeneous systems, coordination would be unfavorable. The Crabtree test was carried out by addition of DCT (2.0 equiv. with respect to the Pd concentration) to the catalytic reaction 2 h prior to the addition of all substrates. The substrates were then added, and the reaction was continued for 206

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217

ACS Omega

Article

Scheme 3. (a) Poison Tests for Confirming the Nanoparticular or Colloidal Catalyst and (b) Crabtree Test

Confirmation of the presence of Pd colloids was further obtained by EDS. The spectra thus recorded allowed the identification of characteristic Pd lines at Lα1 = 2.838 keV and Lβ1 = 3.03 keV, which are commonly assigned to palladium metal, confirming the successful formation of the palladium colloids. A closer look at the EDS spectra also reveals the presence of elemental carbon (Kα = 0.277), which could possibly be obtained through the degradation pathway of palladacycle 1 or the aldehyde. Copper lines are observed as a part of the grid used for the analysis. To further strengthen our claim for the presence of Pd colloids in the catalytic solution, powder XRD analysis was carried out on the isolated black powder. XRD analysis (Figure 3) revealed the palladium colloids to be present in a single phase (111 of Bragg reflection), obtained at a 2θ value of 40. A slight hump in the region of 20−30 (2θ) also points toward the presence of carbon, although its exact nature cannot be ascertained at this moment and the employment of high-end characterization techniques, such as XPS or EXAFS, could help shed light on this aspect. XPS and EXAFS Characterization of Pd Colloids. To verify the observation of carbon, to understand its origin, as well as to accurately determine the composition of the isolated

isolation of active palladium colloids from the catalytic solution. For this to be achieved, we first tried to find the substrate combination that led to the formation of active palladium colloids. Experiments were performed with palladacycle 1, used in stoichiometric amounts under different conditions, which have been summarized in Table 1. Only the combination of palladacycle 1 with 4-chlorobenzaldehyde in CHCl3 and added K2CO3 base at 80 °C resulted in the formation of a black powder. The solution was then centrifuged to obtain the black powder, which was washed several times with deionized water and after suspension in dry ethanol was collected and dried in vacuo (see isolation reaction in Table 1). The powder thus obtained was characterized thoroughly by TEM, HRTEM, and energy-dispersive X-ray spectroscopy (EDS) (Figure 2). TEM analysis of the powder (A) suggested the average size of colloids to be in the range of 2−4 nm, which is in good accordance with the values obtained from TEM analysis of the catalytic solution. To obtain a more precise value for the size of the colloidal particles, HRTEM analysis was performed on the powder sample, with image (B) of the palladium colloids indicating the formation of uniformly distributed spherical structures of ∼2 nm diameter. 207

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217

ACS Omega

Article

Pd(0) (22%), and PdX2 (halides possibly coming from the aldehyde). Next, X-ray absorption near-edge structure analysis and comparison of Pd colloids with standards, using the Athena22 program, suggest that the colloids are mostly composed of Pd(0) species along with some Pd(II) (Figures 5 and 6). The comparison of Pd(OAc)2 and Pd2(DBA)3 with Pd foil and palladacycle 1 suggest that E0 for the colloids (24341 eV) lies within the region of Pd(0) (24350 eV) and Pd(II) (24353.3 eV). Thus, the sample of colloids can be considered as mixture of Pd (0) and Pd(II). The EXAFS region of Pd colloids in Rspace is similar to that of Pd foil, except that the amplitude was reduced. This can be attributed to the termination effect.23 The Fourier transforms (FTs) of the EXAFS data for the Pd colloids are presented in Figure 7. In the R-space, two main peaks were observed: one in the range of 1.2−2.0 Å and the other, more prominent one, in the range of 2.3−2.7 Å. The latter can be attributed to the Pd−Pd shell. For the former, it is difficult to distinguish between the Pd−O, Pd−C, and Pd−P shells in the region of 1.2−2.0 Å. Our mechanistic investigations from other methods prompted us to use the oxygen shell for fitting purposes. FT spectra and best fits are represented in Figure 7. The fitting of Pd colloids using Pd foil and Pd−O shell of Pd(OAc)2 as theory yielded a good fit, with a residual factor (Rvalue) of 0.006. During the fitting, the data and theory were kweighted and Fourier-transformed using a Hanning window function in the k-range from 3 to 13 Å−1. On the basis of the fitting values, it is fair to assume that the prior peak around 2.0 Å arises from the Pd−O shell. However, the possibility of a carbon atom cannot be ruled out at this step. From the fitting, it was possible to calculate the coordination numbers with respect to the central Pd atom. It was found that there were roughly six Pd atoms present around the central Pd at the distance of about 2.78 ± 0.005 Å. About 1.4 ± 0.5 atoms of oxygen (or carbon) were also present in the lattice at a distance of 1.97 ± 0.013 Å. FTs of Pd colloids show characteristic signatures of the Pd foil in the FCC structure, which was used as a reference and theory for calculation purposes. The percent composition of Pd(0)/Pd(II) can be obtained by the calibration method using EXAFS analysis. However, discussion on such experiments is beyond the scope of this article. We are currently carrying out detailed studies on the subjected and will be reporting the results of the same shortly. Besides the two main peaks mentioned above, an additional peak below 1.2 Å was observed, which mostly comes from background. It was not possible to completely subtract this

Figure 1. (a) TEM image of the catalytic solution, (b) TEM image of the poly(ethylene) glycol (PEG)-stabilized colloids. (c) HRTEM image of mercury-engulfed Pd colloids.

Pd colloids, XPS was utilized. The XPS spectral analysis revealed only palladium, carbon, and oxygen, whereas no phosphorus species were observed, suggesting the absence of phosphine on the surface of Pd colloids (Figure 4). The overall composition of the nanoparticular material was revealed to be 71.5% C, 22.4% O, and 3.4% Pd. On closer inspection of the palladium peaks (deconvolution of the peaks), it was observed that Pd is present as a mixture of PdO (66%),

Table 1. Experiments for Ascertaining Substrate Combination for Active Palladium Colloid Formation

1

ald

boronic acid

CHCl3

K2CO3

CH3CN/H2O

observation

√ √ √ √

√ √ × ×

× × √ ×

√ × √ √

√ √ √ √

× √ × ×

black particles no change no change no change

208

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217

ACS Omega

Article

Figure 2. (A) TEM image of Pd colloids, (B) HRTEM image of Pd colloids (∼2 nm diameter), and (C) EDS analysis of Pd colloids.

Figure 3. Powder XRD analysis of Pd colloids.

With the well-characterized active palladium colloids in hand, we performed a catalytic reaction to obtain diarylmethanol, to prove that the same species is responsible for the catalytic activity (Scheme 4). To our excitement, the catalytic reaction with palladium colloids proceeded smoothly and furnished the coupled product in comparable yields (4b) in some cases, whereas the catalyst performed better than palladacycle 1 in certain other cases (4c and 4g). On the basis of the composition determined by XPS analysis of the carbon-

peak using AUTOBKG without affecting the amplitude of the peak for the Pd−Pd shell; hence, it was left untouched. All of the above characterization details have enabled us to predict with greater certainty the formation of highly active Pd colloids through the possible degradation/conversion of palladacycle 1. Although this is the first time that the isolation and complete characterization of colloids has been undertaken, it is still not clear how or by which pathway the palladacycle degrades/converts to active colloids. 209

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217

ACS Omega

Article

Figure 4. X-ray photoelectron spectral analysis of the synthesized nanoparticles.

Figure 5. Comparison of Pd compounds and Pd nanoparticles plotted in R-space.

transformation under base- and ligand-free conditions.28 However, most of these reports have been found to employ homogeneous catalysts that are difficult to recover, and the catalyst loadings have also been high. In this regard, nanoparticular catalysis29 has provided researchers with a sustainable alternative in terms of catalyst concentration and recoverability. However, only a few reports have emerged in the literature on their application in oxidative homocoupling of aryl boronic acids.30 Among other catalysts, Au nanosized catalysts have also been employed in recent years. However, the relatively expensive nature of Au+ and Au3+ salts31 make the protocol synthetically less attractive.32 On closer inspection of the these reports, it is also evident that the protocols have been restricted to aryl boronic acids, whereas heteroaryl boronic acids provided poor yields. Similarly, electron-withdrawing substituents on the aryl boronic acids

stabilized Pd colloids, 10 mg (0.32 mol% of Pd) of the catalyst was able to efficiently catalyze the transformation at a relatively lower Pd concentration as compared to palladacycle 1. Oxidative Homocoupling of Aryl Boronic Acids. With the success obtained with the isolated Pd colloids in catalyzing the synthesis of diarylmethanol, we envisaged their use for tackling a more challenging synthetic process, oxidative homocoupling of aryl boronic acids. Homocoupling of aryl boronic acids has proven to be a very useful synthetic methodology for obtaining symmetrical biaryls, with several reports published over the last decade detailing the homocoupling of aryl boronic acids as the most convenient pathway.24,25 Although, this transformation has been carried out using different metals, palladium-catalyzed oxidative homocoupling is at the forefront.26,27 Recent advancements have also highlighted the possibility of performing the catalytic 210

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217

ACS Omega

Article

Figure 6. Comparison of Pd compounds and Pd nanoparticles plotted in energy.

Figure 7. FT spectra for palladium colloids.

Scheme 4. Diarylmethanol Synthesis Using Pd Colloidsa

a

Yields in brackets correspond to those obtained using 1.0 mol% palladacycle 1 under identical conditions

have proven to be sluggish substrates for homocoupling in general. With this background, we envisaged the possibility of

employing isolated palladium colloids as catalysts to address these issues with the homocoupling process. 211

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217

ACS Omega

Article

Scheme 5. Screening Studies for Oxidative Homocoupling of Aryl Boronic Acids

As a starting point of the study, 3-nitrophenyl boronic acid was chosen as the substrate and an extensive screening study was performed (Scheme 5). Initially, the same conditions as that used for diarylmethanol synthesis were employed (Scheme 4), without the addition of aldehyde as the coupling partner (40%). Encouraging results obtained under the conditions prompted us to explore the reaction further. Eventually, optimum product formation was observed on employment of air as the oxidant in the CH3CN/H2O solvent system at 80 °C. It is to be noted that the removal of CHCl3 (that was added in catalytic amounts for diarylmethanol synthesis) promoted the homocoupling reaction to proceed efficiently. However, at this moment, the exact reason for such an observation is not known. With these results in hand, palladium-catalyzed oxidative homocoupling of a wide variety of aryl and heteroaryl boronic acids was undertaken. From the table given below (Scheme 6), it can be observed that the palladium colloids are able to catalyze oxidative homocoupling of aryl boronic acids in very good yields. Electronic effects on the aryl boronic acids were found to have no effect on the catalytic activity of the colloids, with electron-withdrawing as well as electron-donating substituted boronic acids exhibiting comparable results. Heteroaryl boronic acids when employed as substrates also furnished good yields of the homocoupled products. Similar observations were made on using challenging, bulkier aryl boronic acids, with good-to-excellent yields of the homocoupled products obtained under the catalytic conditions. Interestingly, the more reactive alkyl boronic acid, cyclohexyl boronic acid, which is prone to protodeborylation, also underwent the homocoupling reaction, furnishing dicylohexane in decent yields. Interestingly, 1,4-phenylenediboronic acid when subjected to homocoupling conditions furnished a polymeric material, which is insoluble in most organic solvents (characterization was not possible). Such a species has been described in the literature to be formed as a part of the

homocoupling process, and it has been assumed that the material in hand is that reported.33 Another interesting result was obtained during the oxidative homocoupling of 2-bromo phenylboronic acid under oxidative coupling conditions (Scheme 7). The homocoupling reaction when monitored by gas chromatography−mass spectrometry suggested the formation of three different products, with complete comparison of the starting boronic acid. Instead of exclusively furnishing the desired 2,2′-dibromo biphenyl, the coupling process provided two more products, namely, 2bromobiphenyl and 1,1′ and 2′,1″-terphenyl in a ratio of 10:30:60 (the possible pathway for 6c has been represented in Scheme 7). Although, this result reiterates the active nature of the palladium colloids, its efficiency in catalyzing a subsequent cross-coupling reaction via C−Br activation could have further synthetic relevance. The possibility of formation of products 6a and 6c could be understood from the proposed mechanism that has been put forth on the basis of some literature evidence for protodeboronation34 and protodehalogenation35 pathways that are known to take place during the catalytic processes. Cross-Coupling of Aryl Boronic Acids. The success obtained with oxidative homocoupling of aryl boronic acids (including electron-poor substituents) prompted us to further investigate the applicability of the isolated palladium colloids. Crossed homocoupling of substrates also provides researchers with an attractive alternative to the general cross-coupling processes. Success in such processes could be achieved by the judicious choice of electronically different substrates. However, to date crossed homocoupling has seldom been reported, given the difficulty in preferentially obtaining the cross-coupled product over the commonly observed homocoupling product. An example for oxidative cross-coupling of two different Grignard reagents was recently reported by Cahiez and coworkers under manganese-catalyzed conditions.36 Similarly, gold(I)-catalyzed cross-coupling of two different alkynes has also been reported by Shi and co-workers for the synthesis of 212

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217

ACS Omega

Article

Scheme 6. Palladium Colloids for Oxidative Homocoupling of Aryl Boronic Acids

unsymmetrical 1,3-diynes.37 Herein, we present the first report on cross-coupling of aryl boronic acids using synthesized palladium colloids under aerobic conditions. At the outset of our studies, we decided to employ one of the coupling partners as an electron-poor boronic acid (3-nitro phenylboronic acid) that could serve as the limiting reagent. A more reactive electron-rich boronic acid (4-methoxyphenyl boronic acid) was then added in excess (1:1.3) under aerobic conditions using oxygen as the oxidant. To our surprise, the cross-coupled product (4-methoxy-3-nitrobiphenyl) was obtained in good yields (68%), providing the possibility of obtaining differently substituted biaryls via palladium colloidcatalyzed oxidative coupling of air-stable and cheaply available aryl boronic acids (Scheme 8). Therefore, with the proper selection of coupling partners it could be possible to achieve good selectivity toward promoting the cross-coupling process over the commonly observed homocoupling of aryl boronic acids.

This is an attractive synthetic alternative for obtaining unsymmetrical biaryls compared with using hazardous Grignard reagents. Although at this moment only representative examples have been highlighted, the full potential of the synthetic protocol along with mechanistic studies will be reported elsewhere.



CONCLUSIONS In conclusion, herein, we report evidence for the degradation/ conversion of palladacycle 1 to active palladium colloids, which were isolated and thoroughly characterized by techniques such as TEM, HRTEM, XPS, and EXAFS. The distribution of these palladium colloids, with an average size of ∼2 nm, on carbon was established by the above-mentioned techniques. The isolated palladium colloids were shown to be active catalysts in a variety of coupling processes, including the synthesis of diarylmethanol and oxidative homocoupling of aryl boronic acids. One of the first reports on the cross-coupling of aryl 213

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217

ACS Omega

Article

Scheme 7. Oxidative Homocoupling of 2-Bromophenyl Boronic Acid

Scheme 8. Cross-Coupling of Different Aryl Boronic Acids for Accessing Unsymmetrical Biaryls

boronic acids was also made possible by the unique reactivity exhibited by the palladium colloids.



Notes

The authors declare no competing financial interest.



ASSOCIATED CONTENT

ACKNOWLEDGMENTS A.R.K. would like to thank Department of Science and Technology for the DST Inspire Faculty Award (IFA12-CH22) and DST Fast Track Fellowship (SR/FT/CS-58/2010) as well as DST FIST for support to the Department of Chemistry (ICT, Mumbai). The Alexander von Humboldt Foundation is also thanked for the Equipment grant. The authors would also like to acknowledge the SPring-8 synchrotron facility at Japan for the collection of EXAFS data.

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.6b00326. Synthetic procedures, catalytic process details, and complete characterization data, including 1H and 13C NMR data along with TEM, HRTEM, and XRD data and other relevant details (PDF)





AUTHOR INFORMATION

ABBREVIATIONS TEM, transmission electron microscopy; HRTEM, highresolution transmission electron microscopy; XRD, X-ray diffraction; EXAFS, extended X-ray absorption fine structure; XPS, X-ray photoelectron spectroscopy

Corresponding Author

*E-mail: [email protected], [email protected]. ORCID



Anant R. Kapdi: 0000-0001-9710-8146 Pushkar Sudhakar Shejwalkar: 0000-0002-6342-9299 Author Contributions

REFERENCES

(1) (a) Anastasia, L.; Negishi, E. Palladium-Catalyzed Aryl−Aryl Coupling. In Handbook of Organopalladium Chemistry for Organic Synthesis, Negishi, E., de Meijere, A., Eds.; Wiley: New York, 2003; pp 311−334. (b) Miyaura, N. Metal-Catalyzed Cross-Coupling Reactions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. 214

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217

ACS Omega

Article

of Organoboron Compounds with Organic Halides. In MetalCatalyzed Cross-Coupling Reactions, 2nd ed.; de Meijere, A.; Diederich, F., Eds.; Wiley-VCH: Weinheim, 2004; pp 41−123. (2) (a) Dehand, J.; Pfeffer, M. Cyclometallated compounds. Coord. Chem. Rev. 1976, 18, 327−352. (b) Dupont, J.; Consorti, C. S.; Spencer, J. The Potential of palladacycles: More than just precatalysts. Chem. Rev. 2005, 105, 2527−2572. (c) Parshall, G. W. Intramolecular aromatic substitution in transition metal complexes. Acc. Chem. Res. 1970, 3, 139−144. (d) Bruce, M. I. Cyclometallation reactions. Angew. Chem., Int. Ed. Engl. 1977, 16, 73−86. (e) Omae, I. Organometallic intramolecular-coordination compounds. Recent aspects in the study of sulfur donor ligands. Coord. Chem. Rev. 1979, 28, 97−115. (f) Omae, I. Organometallic intramolecular-coordination compounds containing a nitrogen donor ligand. Chem. Rev. 1979, 79, 287−321. (g) Omae, I. Organometallic intramolecular-coordination compounds containing a phosphorus donor ligand. Coord. Chem. Rev. 1980, 32, 235−271. (h) Newkome, G. R.; Puckett, W. E.; Gupta, V. K.; Kiefer, G. E. Cyclometalation of the platinum metals with nitrogen and alkyl, alkenyl, and benzyl carbon donors. Chem. Rev. 1986, 86, 451−489. (i) Pfeffer, M. Selected applications to organic synthesis of intramolecular C−H activation reactions by transition metals. Pure Appl. Chem. 1992, 64, 335−342. (j) Herrmann, W. A.; Bohm, V. W. P.; Reisinger, C. P. Application of palladacycles in Heck type reactions. J. Organomet. Chem. 1999, 576, 23−41. (k) Albrecht, M.; van Koten, G. Platinum Group Organometallics Based on “Pincer” Complexes: Sensors, Switches, and Catalysts. Angew. Chem., Int. Ed. 2001, 40, 3750−3781. (l) Beletskaya, I. P.; Cheprakov, A. V. Palladacycles in catalysis − a critical survey. J. Organomet. Chem. 2004, 689, 4055− 4082. (3) For a comprehensive review on the application of palladacycles as possible anticancer agents, see: Kapdi, A. R.; Fairlamb, I. J. S. Anticancer palladium complexes: a focus on PdX2L2, palladacycles and related complexes. Chem. Soc. Rev. 2014, 43, 4751−4777. (4) (a) Bose, A.; Saha, C. R. Orthometalated palladium(II) complexcatalyzed reduction of nitroalkanes and nitriles. J. Mol. Catal. 1989, 49, 271−283. (b) Santra, P. K.; Saha, C. R. Dihydrogen reduction of nitroaromatics, alkenes, alkynes and aromatic carbonyls by orthometalated Pd(II) complexes in homogeneous phase. J. Mol. Catal. 1987, 39, 279−292. (5) (a) Herrmann, W. A.; Brossmer, C.; Ofele, K.; Reisinger, C. P.; Priermeier, T.; Beller, M.; Fischer, H. Palladacycles as Structurally Defined Catalysts for the Heck Olefination of Chloro- and Bromoarenes. Angew. Chem., Int. Ed. Engl. 1995, 34, 1844−1848. (b) Herrmann, W. A.; Brossmer, C.; Reisinger, C. P.; Riermeier, T. H.; Ofele, K.; Beller, M. Palladacycles: Efficient New Catalysts for the Heck Vinylation of Aryl Halides. Chem. Eur. J. 1997, 3, 1357−1364. (c) Bohm, V. P. W.; Herrmann, W. A. Mechanism of the Heck Reaction Using a Phosphapalladacycle as the Catalyst: Classical versus Palladium(IV) Intermediates. Chem. Eur. J. 2001, 7, 4191−4197. (6) For early examples on Pd(II)/Pd(IV) mechanistic proposal for palladacycles, see: (a) Shaw, B. L. Speculations on new mechanisms for Heck reactions. New. J. Chem. 1998, 22, 77−79. (b) Shaw, B. L.; Parera, S. D.; Staley, E. A. Highly active, stable, catalysts for the Heck reaction; further suggestions on the mechanism. Chem. Commun. 1998, 1361−1362. (7) For early examples on Pd(0)/Pd(II) mechanistic proposal for palladacycles, see: (a) Louie, J.; Hartwig, J. F. A Route to Pdo from PdII Metallacycles in Animation and Cross-Coupling Chemistry. Angew. Chem., Int. Ed. Engl. 1996, 35, 2359−2361. (b) Beller, M.; Riermeier, T. H. Phosphapalladacycle-Catalyzed Heck Reactions for Efficient Synthesis of Trisubstituted Olefins: Evidence for Palladium(0) Intermediates. Eur. J. Inorg. Chem. 1998, 29−35. (8) (a) Amatore, C.; Jutand, A. Mechanistic and kinetic studies of palladium catalytic systems. J. Organomet. Chem. 1999, 576, 254−278. (b) Amatore, C.; Jutand, A. Anionic Pd(0) and Pd(II) Intermediates in Palladium-Catalyzed Heck and Cross-Coupling Reactions. Acc. Chem. Res. 2000, 33, 314−321. (c) Cruz, S. C.; Aarnoutse, P. J.; Rothenberg, G.; Westerhuis, J. A.; Smilde, A. K.; Bliek, A. Kinetic and mechanistic studies on the Heck reaction using real-time near infrared spectros-

copy. Phys. Chem. Chem. Phys. 2003, 5, 4455−4460. (d) Buback, M.; Perkovic, T.; Redlich, S.; de Meijere, A. The Effect of High Pressure on the Heck Reaction − A Contribution to a Deeper Understanding of the Mechanism. Eur. J. Org. Chem. 2003, 2375−2382. (e) d’Orlyé, F.; Jutand, A. In situ formation of palladium(0) from a P,C-palladacycle. Tetrahedron 2005, 61, 9670−9678. (9) For more details on the role of palladacycles as reservoirs of Pd(0), see: (a) de Vries, A. H. M.; Mulders, J. M. C. A.; Mommers, J. H. M.; Henderickx, H. J. W.; de Vries, J. G. Homeopathic Ligand-Free Palladium as a Catalyst in the Heck Reaction. A Comparison with a Palladacycle. Org. Lett. 2003, 5, 3285−3288. (b) Bedford, R. Palladacyclic Precatalysts for Suzuki Coupling, Buchwald−Hartwig Amination and Related Reaction. In Palladacycles: Synthesis, Characterization and Applications, Dupont, J., Pfeffer, M., Eds.; Wiley-VCH: Weinheim, 2008; pp 209−225. (10) For a review on the role of colloids or nanoparticles in palladium catalysis, see: de Vries, J. G. A unifying mechanism for all high-temperature Heck reactions. The role of palladium colloids and anionic species. Dalton Trans. 2006, 421−429. (11) (a) Nowotny, M.; Hanefeld, U.; van Koningsveld, H.; Maschmeyer, T. Cyclopalladated imine catalysts in Heck arylation: search for the catalytic species. Chem. Commun. 2000, 1877−1878. (b) Beletskaya, I. P.; Kashin, A. N.; Karlstedt, N. B.; Mitin, A. V.; Cheprakov, A. V.; Kazankov, G. M. NC-palladacycles as highly effective cheap precursors for the phosphine-free Heck reactions. J. Organomet. Chem. 2001, 622, 89−96. (c) Rocaboy, C.; Gladysz, J. A. Highly Active Thermomorphic Fluorous Palladacycle Catalyst Precursors for the Heck Reaction; Evidence for a Palladium Nanoparticle Pathway. Org. Lett. 2002, 4, 1993−1996. (d) Rocaboy, C.; Gladysz, J. A. Thermomorphic fluorous imine and thioether palladacycles as precursors for highly active Heck and Suzuki catalysts; evidence for palladium nanoparticle pathways. New. J. Chem. 2003, 27, 39−49. (12) (a) Alacid, E.; Alonso, D. A.; Botella, L.; Najera, C.; Pacheco, M. C. Oxime palladacycles revisited: stone-stable complexes nonetheless very active catalysts. Chem. Rec. 2006, 6, 117−132. (b) Alonso, D. A.; Najera, C. Oxime-derived palladacycles as source of palladium nanoparticles. Chem. Soc. Rev. 2010, 39, 2891−2902. (13) Consorti, C. S.; Flores, F. R.; Dupont, J. Kinetics and Mechanistic Aspects of the Heck Reaction Promoted by a CN− Palladacycle. J. Am. Chem. Soc. 2005, 127, 12054−12065. (14) For more examples for the formation of palladium nanoparticles or colloids from palladacycles, see: (a) Grigg, R.; Zhang, L.; Collard, S.; Ellis, P.; Keep, A. Facile generation and morphology of Pd nanoparticles from palladacycles and carbon monoxide. J. Organomet. Chem. 2004, 689, 170−173. (b) Astruc, D.; Lu, F.; Aranzaes, J. R. Nanoparticles as Recyclable Catalysts: The Frontier between Homogeneous and Heterogeneous Catalysis. Angew. Chem., Int. Ed. 2005, 44, 7852−7872. (c) Chen, C.-L.; Liu, Y.-H.; Peng, S.-M.; Liu, S.T. Air- and Moisture-Stable Cyclopalladated Complexes as Efficient Catalysts for Suzuki-Miyaura Coupling Reaction. Organometallics 2005, 24, 1075−1081. (d) Alacid, E.; Najera, C. Palladated Kaiser Oxime Resin as Precatalyst for the Heck Reaction in Organic and Aqueous Media. Synlett 2006, 2006, 2959−2964. (e) Alacid, E.; Najera, C. Acrolein Diethyl Acetal: A Three-Carbon Homologating Reagent for the Synthesis of β-Arylpropanoates and Cinnamaldehydes by Heck Reaction Catalyzed by a Kaiser Oxime Resin Derived Palladacycle. Eur. J. Org. Chem. 2008, 2008, 3102−3106. (f) Rao, G. K.; Kumar, A.; Ahmed, J.; Singh, A. K. Palladacycle containing nitrogen and selenium: highly active pre-catalyst for the Suzuki−Miyaura coupling reaction and unprecedented conversion into nano-sized Pd17Se15. Chem. Commun. 2010, 46, 5954−5956. (g) Rao, G. K.; Kumar, A.; Kumar, S.; Dupare, U. B.; Singh, A. K. Catalytic Activity of Nanoparticles. Organometallics 2013, 32, 2452−2458. (h) Karami, K.; Moghadam, Z. K.; Hosseini-Kharat, M. Polyethylene glycol-supported recyclable NC palladacycle catalyst for Heck cross-coupling reactions. Catal. Commun. 2014, 43, 25−28. (15) For reports on the metal-mediated diarylmethanol syntheses, see: (a) Sakai, M.; Ueda, M.; Miyaura, N. Rhodium-Catalyzed 215

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217

ACS Omega

Article

Addition of Organoboronic Acids to Aldehydes. Angew. Chem., Int. Ed. 1998, 37, 3279−3281. (b) Dahmen, S.; Lormann, M. Triarylborane Ammonia Complexes as Ideal Precursors for Arylzinc Reagents in Asymmetric Catalysis. Org. Lett. 2005, 7, 4597−4600. (c) Liao, Y.-X.; Xing, C.-H.; He, P.; Hu, Q.-S. Orthoplatinated Triarylphosphite as a Highly Efficient Catalyst for Addition Reactions of Arylboronic Acids with Aldehydes: Low Catalyst Loading Catalysis and a New Tandem Reaction Sequence. Org. Lett. 2008, 10, 2509−2512. (d) Trindade, A. F.; Gois, P. M. P.; Veiros, L. F.; Andre, V.; Duarte, M. T.; Afonso, C. A. M.; Caddick, S.; Cloke, F. G. N. Axial Coordination of NHC Ligands on Dirhodium(II) Complexes: Generation of a New Family of Catalysts. J. Org. Chem. 2008, 73, 4076−4086. (e) Kuriyama, M.; Shimazawa, R.; Shirai, R. Efficient 1,2-Addition of Aryl- and Alkenylboronic Acids to Aldehydes Catalyzed by the Palladium/ Thioether−Imidazolinium Chloride System. J. Org. Chem. 2008, 73, 1597−1600. (f) Zheng, H.; Zhang, Q.; Chen, J.; Liu, M.; Cheng, S.; Ding, J.; Wu, H.; Su, W. Copper(II) Acetate-Catalyzed Addition of Arylboronic Acids to Aromatic Aldehydes. J. Org. Chem. 2009, 74, 943−945. (g) Chen, W.; Baghbanzadeh, M.; Kappe, C. O. Microwaveassisted nickel(II) acetylacetonate-catalyzed arylation of aldehydes with arylboronic acids. Tetrahedron Lett. 2011, 52, 1677−1679. (h) Liao, Y. X.; Xing, C.-H.; Hu, Q.-S. Rhodium(I)/Diene-Catalyzed Addition Reactions of Arylborons with Ketones. Org. Lett. 2012, 14, 1544−1547. (16) For reports on the palladium-mediated diarylmethanol syntheses, see: (a) He, P.; Lu, Y.; Dong, C.-G.; Hu, Q.-S. Anionic Four-Electron Donor-Based Palladacycles as Catalysts for Addition Reactions of Arylboronic Acids with α,β-Unsaturated Ketones, Aldehydes, and α-Ketoesters. Org. Lett. 2007, 9, 343−346. (b) He, P.; Lu, Y.; Hu, Q.-S. Phosphinite- and phosphite-based type I palladacycles as highly active catalysts for addition reactions of arylboronic acids with aldehydes, α,β-unsaturated ketones, αketoesters, and aldimines. Tetrahedron Lett. 2007, 48, 5283−5288. (c) Qin, C.; Wu, H.; Cheng, J.; Chen, X.; Liu, M.; Zhang, W.; Su, W.; Ding, J. The Palladium-Catalyzed Addition of Aryl- and Heteroarylboronic Acids to Aldehydes. J. Org. Chem. 2007, 72, 4102−4107. (d) Lin, S.; Lu, X. Cationic Pd(II)/Bipyridine-Catalyzed Addition of Arylboronic Acids to Arylaldehydes. One-Pot Synthesis of Unsymmetrical Triarylmethanes. J. Org. Chem. 2007, 72, 9757−9760. (e) Liu, T.-P.; Liao, Y. X.; Xing, C.-H.; Hu, Q.-S. Fluorenone Synthesis by Palladacycle-Catalyzed Sequential Reactions of 2-Bromobenzaldehydes with Arylboronic Acids. Org. Lett. 2011, 13, 2452−2455. (17) Yamamoto, T.; Ohta, T.; Ito, Y. Palladium-Catalyzed Addition of Arylboronic Acids to Aldehydes. Org. Lett. 2005, 7, 4153−4155. (18) (a) Whitesides, G. M.; Hackett, M.; Brainard, R. L.; Lavalleye, J.P. P. M.; Sowinski, A. F.; Izumi, A. N.; Moore, S. S.; Brown, D. W.; Staudt, E. M. Suppression of unwanted heterogeneous platinum(0)catalyzed reactions by poisoning with mercury(0) in systems involving competing homogeneous reactions of soluble organoplatinum compounds: thermal decomposition of bis(triethylphosphine)3,3,4,4-tetramethylplatinacyclopentane. Organometallics 1985, 4, 1819−1830. (b) Crabtree, R. H. Resolving Heterogeneity Problems and Impurity Artifacts in Operationally Homogeneous Transition Metal Catalysts. Chem. Rev. 2012, 112, 1536−1554. (c) Khake, S. M.; Soni, V.; Gonnade, R. G.; Punji, B. Design and development of POCN-pincer palladium catalysts for C−H bond arylation of azoles with aryl iodides. Dalton Trans. 2014, 43, 16084−16096. (d) Prajapati, D.; Schulzke, C.; Kindermann, M. K.; Kapdi, A. R. Selective palladiumcatalysed arylation of 2,6-dibromopyridine using N-heterocyclic carbene ligands. RSC Adv. 2015, 5, 53073−53085. (19) (a) Collman, J. P.; Kosydar, K. M.; Bressan, M.; Lamanna, W.; Garrett, T. Polymer-bound substrates: a method to distinguish between homogeneous and heterogeneous catalysis. J. Am. Chem. Soc. 1984, 106, 2569−2579. (b) Davies, I. W.; Matty, L.; Hughes, D. L.; Reider, P. J. Are Heterogeneous Catalysts Precursors to Homogeneous Catalysts? J. Am. Chem. Soc. 2001, 123, 10139−10140. (20) Anton, D. R.; Crabtree, R. H. Dibenzo[a,e]cyclooctatetraene in a proposed test for heterogeneity in catalysts formed from soluble platinum-group metal complexes. Organometallics 1983, 2, 855−859.

(21) For reports on the use of poly(ethylene)glycol as a stabilizer and support for palladium nanoparticles or colloids, see: (a) Feng, B.; Hou, Z.; Yang, H.; Wang, X.; Hu, Y.; Li, H.; Qiao, Y.; Zhao, X.; Huang, Q. Functionalized Poly(ethylene glycol)-Stabilized Water-Soluble Palladium Nanoparticles: Property/Activity Relationship for the Aerobic Alcohol Oxidation in Water. Langmuir 2010, 26, 2505−2513. (b) Harraz, F. A.; El-Hout, S. E.; Killa, H. M.; Ibrahim, I. A. Palladium nanoparticles stabilized by polyethylene glycol: Efficient, recyclable catalyst for hydrogenation of styrene and nitrobenzene. J. Catal. 2012, 286, 184−192. (c) Borah, R. K.; Saikia, H. J.; Mahanta, A.; Das, V. K.; Bora, U.; Thakur, A. J. Biosynthesis of poly(ethylene glycol)-supported palladium nanoparticles using Colocasia esculenta leaf extract and their catalytic activity for Suzuki−Miyaura cross-coupling reactions. RSC Adv. 2015, 5, 72453−72457. (22) Ravel, B.; Newville, M. Athena, Artemis, Hephastus: data analysis for X-ray absorption spectroscopy using IFEFFIT. J. Synchrotron Radiat. 2005, 12, 537−541. (23) Calvin, S.; Miller, M. M.; Goswami, R.; Cheng, S. F.; Mulvaney, S. P.; Whitman, H. J.; Harris, V. G. Determination of crystallite size in a magnetic nanocomposite using extended x-ray absorption fine structure. J. Appl. Phys. 2003, 94, 778−783. (24) (a) Wang, L.; Zhang, W.; Su, D. S.; Meng, X.; Xiao, F.-S. Supported Au nanoparticles as efficient catalysts for aerobic homocoupling of phenylboronic acid. Chem. Commun. 2012, 48, 5476−5478. and references therein. (b) Cheng, G.; Luo, M. Homocoupling of Arylboronic Acids Catalyzed by CuCl in Air at Room Temperature. Eur. J. Org. Chem. 2011, 2011, 2519−2523 and references therein.. (25) (a) Kaboudin, B.; Abedi, Y.; Yokomatsu, T. CuII−β-Cyclodextrin Complex as a Nanocatalyst for the Homo- and Cross-Coupling of Arylboronic Acids under Ligand- and Base-Free Conditions in Air: Chemoselective Cross-Coupling of Arylboronic Acids in Water. Eur. J. Org. Chem. 2011, 2011, 6656−6662. (b) Xu, Z.; Mao, J.; Zhang, Y. Pd(OAc)2-catalyzed room temperature homocoupling reaction of arylboronic acids under air without ligand. Catal. Commun. 2008, 9, 97−100. (c) Parrish, J. P.; Jung, Y. C.; Floyd, R. J.; Jung, K. W. Oxidative dimerization: Pd(II) catalysis in the presence of oxygen using aqueous media. Tetrahedron Lett. 2002, 43, 7899−7902. (26) (a) Prastaro, A.; Ceci, P.; Chiancone, E.; Boffi, A.; Fabrizi, G.; Cacchi, S. Homocoupling of arylboronic acids and potassium aryltrifluoroborates catalyzed by protein-stabilized palladium nanoparticles under air in water. Tetrahedron Lett. 2010, 51, 2550−2552. (b) Chen, J.-S.; Krogh-Jespersen, K.; Khinast, J. G. Base- and ligandfree heterogeneously catalyzed homocoupling of arylboronic acids. J. Mol. Catal. A: Chem. 2008, 285, 14−19. (27) (a) Punna, S.; Diaz, D. D.; Finn, M. G. Palladium-Catalyzed Homocoupling of Arylboronic Acids and Esters Using Fluoride in Aqueous Solvents. Synlett 2004, 2351−2354. (b) Jin, Z.; Guo, S. X.; Gu, X. P.; Qiu, L. L.; Song, H. B.; Fang, J. X. Highly Active, WellDefined (Cyclopentadiene)(N-heterocyclic carbene)palladium Chloride Complexes for Room-Temperature Suzuki−Miyaura and Buchwald−Hartwig Cross-Coupling Reactions of Aryl Chlorides and Deboronation Homocoupling of Arylboronic Acids. Adv. Synth. Catal. 2009, 351, 1575−1585. (c) Yadav, J. S.; Gayathri, K. U.; Ather, H.; Rehman, H.; Prasad, A. R. The Pd(OAc)2-catalyzed homocoupling of arylboronic acids in water and ionic liquid. J. Mol. Catal. A: Chem. 2007, 271, 25−27. (d) Wong, M. S.; Zhang, X. L. Ligand promoted palladium-catalyzed homo-coupling of arylboronic acids. Tetrahedron Lett. 2001, 42, 4087−4089. (e) Cravotto, G.; Beggiato, M.; Penoni, A.; Palmisano, G.; Tollari, S.; Lévêque, J. M.; Bonrath, W. High-intensity ultrasound and microwave, alone or combined, promote Pd/C-catalyzed aryl−aryl couplings. Tetrahedron Lett. 2005, 46, 2267−2271. (f) Amatore, C.; Cammoun, C.; Jutand, A. Pd(OAc)2/p-Benzoquinone-Catalyzed Anaerobic Electrooxidative Homocoupling of Arylboronic Acids, Arylboronates and Aryltrifluoroborates in DMF and/or Water. Eur. J. Org. Chem. 2008, 2008, 4567− 4570. (g) Cheng, K.; Xin, B.; Zhang, Y. The Pd(OAc)2-catalyzed homocoupling of arylboronic acids in water and ionic liquid. J. Mol. Catal. A: Chem. 2007, 273, 240−243. (h) Mitsudo, K.; Shiraga, T.; 216

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217

ACS Omega

Article

Kagen, D.; Shi, D.; Becker, J. Y.; Tanaka, H. Pd/TEMPO-catalyzed electrooxidative synthesis of biaryls from arylboronic acids or arylboronic esters. Tetrahedron 2009, 65, 8384−8388. (28) (a) Yamamoto, Y.; Suzuki, R.; Hattori, K.; Nishiyama, H. Baseand Phosphine-Free Palladium-Catalyzed Homocoupling of Arylboronic Acid Derivatives under Air. Synlett 2006, 2006, 1027−1030. (b) Kapdi, A. R.; Dhangar, G.; Serrano, J. L.; De Haro, J. A.; Lozano, P.; Fairlamb, I. J. S. [Pd(Phbz)(X)(PPh3)] palladacycles promote the base-free homocoupling of arylboronic acids in air at room temperature. RSC Adv. 2014, 4, 55305−55312. (29) Xia, Y.; Yang, H.; Campbell, C. T. Nanoparticles for Catalysis. Acc. Chem. Res. 2013, 46, 1671−1672. (30) (a) Zheng, J.; Lin, S.; Zhu, X.; Jiang, B.; Yang, Z.; Pan, Z. Reductant-directed formation of PS−PAMAM-supported gold nanoparticles for use as highly active and recyclable catalysts for the aerobic oxidation of alcohols and the homocoupling of phenylboronic acids. Chem. Commun. 2012, 48, 6235−6237. (b) Kaboudin, B.; Mostafalu, R.; Yokomatsu, T. Fe3O4 nanoparticle-supported Cu(II)-β-cyclodextrin complex as a magnetically recoverable and reusable catalyst for the synthesis of symmetrical biaryls and 1,2,3-triazoles from aryl boronic acids. Green Chem. 2013, 15, 2266−2274. (c) Raul, P. K.; Mahanta, A.; Bora, U.; Thakur, A. J.; Veer, V. In water homocoupling of arylboronic acids using nano-rod shaped and reusable copper oxide(II) catalyst at room temperature. Tetrahedron Lett. 2015, 56, 7069−7073. (31) (a) Carrettin, S.; Guzman, J.; Corma, A. Supported Gold Catalyzes the Homocoupling of Phenylboronic Acid with High Conversion and Selectivity. Angew. Chem., Int. Ed. 2005, 44, 2242− 2245. (b) Gonzalez-Arellano, C.; Corma, A.; Iglesias, M.; Sanchez, F. Homogeneous and heterogenized Au(III) Schiff base-complexes as selective and general catalysts for self-coupling of aryl boronic acids. Chem. Commun. 2005, 1990−1992. (c) Carrettin, S.; Corma, A.; Iglesias, M.; Sanchez, F. Stabilization of Au(III) on heterogeneous catalysts and their catalytic similarities with homogeneous Au(III) metal organic complexes. Appl. Catal., A 2005, 291, 247−252. (d) González-Arellano, C.; Corma, A.; Iglesias, M.; Sanchez, F. Gold (I) and (III) catalyze Suzuki cross-coupling and homocoupling. J. Catal. 2006, 238, 497−501. (e) Cheng, G.; Luo, M. Homocoupling of Arylboronic Acids Catalyzed by CuCl in Air at Room Temperature. Eur. J. Org. Chem. 2011, 2011, 2519−2523. (32) (a) Kirai, N.; Yamamoto, Y. Homocoupling of Arylboronic Acids Catalyzed by 1,10-Phenanthroline-Ligated Copper Complexes in Air. Eur. J. Org. Chem. 2009, 2009, 1864−1867. (b) Yoshida, H.; Yamaryo, Y.; Ohshita, J.; Kunai, A. Base-free oxidative homocoupling of arylboronic esters. Tetrahedron Lett. 2003, 44, 1541−1544. (33) Xu, S.-Y.; Wang, H.-C.; Flower, S. E.; Fossey, J. S.; Jiang, Y.-B.; James, T. D. Suzuki homo-coupling reaction based fluorescent sensors for monosaccharides. RSC Adv. 2014, 4, 35238−35241. (34) For more information on the protodeboronation pathway in palladium-catalyzed coupling reactions, see: Kuivila, H. G.; Reuwer, J. F.; Mangravite, J. A. Electrophilic displacement reactions. XVI Metal ion catalysis in the protodeboronation of areneboronic acids. J. Am. Chem. Soc. 1964, 86, 2666−2670. (35) For more information on the protodeboronation pathway in palladium-catalyzed coupling reactions, see: Ahmadi, Z.; McIndoe, J. S. A mechanistic investigation of hydrodehalogenation using ESI-MS. Chem. Commun. 2013, 49, 11488−11490. (36) Cahiez, G.; Duplais, C.; Buendia, J. Manganese-Catalyzed Oxidative Cross-Coupling of Grignard Reagents with Oxygen as an Oxidant. Angew. Chem., Int. Ed. 2009, 48, 6731−6734. (37) Peng, H.; Xi, Y.; Ronaghi, N.; Dong, B.; Akhmedov, N. G.; Shi, X. Gold-Catalyzed Oxidative Cross-Coupling of Terminal Alkynes: Selective Synthesis of Unsymmetrical 1,3-Diynes. J. Am. Chem. Soc. 2014, 136, 13174−13177.

217

DOI: 10.1021/acsomega.6b00326 ACS Omega 2017, 2, 204−217