Nuances in Fundamental Suzuki–Miyaura Cross-Couplings Employing [Pd(PPh3)4]: Poor Reactivity of Aryl Iodides at Lower TemperaturesClick to copy article linkArticle link copied!
- Curtis C. HoCurtis C. HoSchool of Natural Sciences − Chemistry, University of Tasmania, Hobart, Tasmania 7001, AustraliaMore by Curtis C. Ho
- Angus OldingAngus OldingSchool of Natural Sciences − Chemistry, University of Tasmania, Hobart, Tasmania 7001, AustraliaMore by Angus Olding
- Jason A. SmithJason A. SmithSchool of Natural Sciences − Chemistry, University of Tasmania, Hobart, Tasmania 7001, AustraliaMore by Jason A. Smith
- Alex C. Bissember*Alex C. Bissember*E-mail for A.C.B.: [email protected]School of Natural Sciences − Chemistry, University of Tasmania, Hobart, Tasmania 7001, AustraliaMore by Alex C. Bissember
Abstract
We have explored fundamental Pd-catalyzed Csp2–Csp2 Suzuki–Miyaura cross-couplings of aryl iodides (Ar–I) employing “classical” Pd/PPh3 catalyst systems. Surprisingly, we observed particularly inefficient couplings of these ostensibly reactive electrophiles in a range of conventional solvent mixtures at lower temperatures (∼50 °C), which was in stark contrast to analogous reactions featuring the equivalent aryl bromides. This feature of well-established Pd/PPh3-mediated Suzuki–Miyaura reactions has received scant attention in the literature. Most significantly, our studies suggest that the inefficient coupling of aryl iodides at lower temperatures derives from the unexpectedly poor turnover of the key on-cycle intermediate trans-[Pd(PPh3)2(Ar)(I)] (or related PdII–I species) in the presence of PPh3.
Introduction
Scheme 1
Results and Discussion


Determined via gas chromatography (GC) with the aid of a calibrated internal standard (average of two experiments). Yields after 0.25 h are given in parentheses.
THF used instead of n-PrOH.
Acetone used instead of n-PrOH.
DMF used instead of n-PrOH.
K2CO3 used instead of Na2CO3.
NaOH used instead of Na2CO3.
Figure 1

yield (%)a | ||||
---|---|---|---|---|
entry | Ar′ | Ar″ | using Ar–I | using Ar–Br |
1 | o-tol | Ph | 8 (6) | 59 (40) |
2 | m-tol | Ph | 5 (3) | 52 (35) |
3 | p-OMeC6H4 | Ph | 10 (8) | 55 (38) |
4 | p-NO2C6H4 | Ph | 17 (11) | 74 (53) |
5 | Ph | p-tol | 12 (10) | 59 (49) |
6 | Ph | p-OMeC6H4 | 15 (14) | 52 (41) |
7b | Ph | p-NO2C6H4 | 16 (13) | 50 (39) |
Determined via GC with the aid of a calibrated internal standard (average of two experiments). Yields after 0.25 h are given in parentheses.
Reaction performed at 70 °C.

yield (%)a | ||||
---|---|---|---|---|
entry | nucleophile | temp (°C) | using 1 | using 2 |
1b | Ph-B(OH)3Na | 50 | 15 (13) | 57 (57) |
2 | Ph-Bpin | 100 | 89 (89) | 95 (81) |
3 | Ph-Bpin | 80 | 87 (62) | 85 (79) |
4 | Ph–Bpin | 50 | 7 (7) | 61 (43) |
5 | Ph-BF3K | 100 | 77 (72) | 67 (43) |
6 | Ph-BF3K | 80 | 47 (34) | 50 (25) |
7 | Ph-BF3K | 50 | <2 (<2) | 12 (2) |
Determined via GC with the aid of a calibrated internal standard (average of two experiments). Yields after 0.25 h are provided in parentheses.
Na2CO3 not added.

yield (%)a | |||||
---|---|---|---|---|---|
entry | ligande | θ (deg)b | ν (cm–1)c | using 1 | using 2 |
1 | PPh3 | 145 | 2068.9 | 11 (9) | 68 (57) |
2 | P(o-tol)3 | 194 | 2066.6 | 66 (50) | 61 (56) |
3 | TTMPP | 184 | 6 (5) | 47 (27) | |
4 | AsPh3 | 142 | 2067.9 | 57 (50) | 22 (18) |
5 | P(p-tol)3 | 145 | 2066.7 | 43 (35) | 51 (39) |
6 | P(2-furyl)3 | 133 | 2078.4 | 51 (44) | 9 (9) |
7 | P(OPh)3 | 130 | 2075.9 | 66 (53) | 54 (49) |
8 | PBn3 | 139.5 | 2066.1 | 39 (26) | 42 (34) |
9 | dppf | 42 (20) | 28 (9) | ||
10d | dppf | 47 (31) | 25 (9) | ||
11 | dppe | 3 (3) | <2 (<2) | ||
12d | dppe | <2 (<2) | <2 (<2) |
Determined via gas chromatography (GC) with the aid of a calibrated internal standard (average of two experiments). Yields after 0.25 h are given in parentheses.
Tolman cone angle. (9,20)
Tolman electronic parameter (IR frequency of Ni(CO)3L). (9,20a,21)
1% ligand used.
Abbreviations: TTMPP, tris(2,4,6-trimethoxyphenyl)phosphine; dppf, 1,1′-ferrocenediylbis(diphenylphosphine); dppe, bis(diphenylphosphino)ethane.


Figure 2
Figure 2. Two fundamental pathways leading to pretransmetalation intermediate C in Suzuki–Miyaura couplings.


Determined via GC with the aid of a calibrated internal standard (average of two experiments). Yields after 0.25 h are given in parentheses.
0.5% [Pd] employed.
Scheme 2
aDetermined via GC with the aid of a calibrated internal standard (average of two experiments). Yields after 0.25 h are given in parentheses.
Figure 3
Figure 3. Formation of biaryl 3 in reactions employing trans-[Pd(PPh3)2(p-tol)(I)] (4a) (solid red line), [Pd–I] 4a + 6 equiv of PPh3 (dashed red line), trans-[Pd(PPh3)2(p-tol)(Br)] (4b) (solid blue line), [Pd–Br] 4b + 6 equiv PPh3 (dashed blue line), [Pd–OH] 5 (solid black line), and [Pd–OH] 5 + 6 equiv PPh3 (dashed black line). A 0.5 equiv amount of [Pd(PPh3)(p-tol)(μ-OH)]2 was used in these experiments (and no Na2CO3 was used in these reactions). Yields were determined via GC with the aid of a calibrated internal standard (average of two experiments).

Figure 4
Figure 4. Formation of biaryl 3 in reactions employing: 6% PPh3 (red), 6% P(2-furyl)3 (blue), and 3% PPh3 and 3% P(2-furyl)3 (black). Yields were determined via GC with the aid of a calibrated internal standard (average of two experiments).
Scheme 3
aDetermined via GC with the aid of a calibrated internal standard (average of two experiments). Yields after 0.25 h are given in parentheses.
Conclusions
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.organomet.8b00189.
Experimental procedures and compound characterization data (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgments
The authors acknowledge the University of Tasmania School of Natural Sciences − Chemistry for financial support, the University of Tasmania Central Science Laboratory for access to NMR spectroscopy services, Mr. Brendon Schollum for assistance with gas chromatography, and Prof. Allan Canty and Prof. Justin Mohr for helpful discussions. A.O. thanks the University of Tasmania for a Dean’s Summer Research Scholarship.
References
This article references 34 other publications.
- 1
See, for example:
(a) Choi, J.; Fu, G. C. Transition metal-catalyzed alkyl-alkyl bond formation: Another dimension in cross-coupling chemistry. Science 2017, 356, eaaf7230, DOI: 10.1126/science.aaf7230Google ScholarThere is no corresponding record for this reference.(b) Metal-Catalyzed Cross-Coupling Reactions and More, 3rd ed.; de Meijere, A., Bras̈e, S., Oestreich, M., Eds.; Wiley-VCH: Weinheim, Germany, 2014.Google ScholarThere is no corresponding record for this reference.(c) Magano, J.; Dunetz, J. R. Large-Scale Applications of Transition Metal-Catalyzed Couplings for the Synthesis of Pharmaceuticals. Chem. Rev. 2011, 111, 2177– 2250, DOI: 10.1021/cr100346gGoogle Scholar1cLarge-Scale Applications of Transition Metal-Catalyzed Couplings for the Synthesis of PharmaceuticalsMagano, Javier; Dunetz, Joshua R.Chemical Reviews (Washington, DC, United States) (2011), 111 (3), 2177-2250CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review was given summarizing large-scale applications of transition metal-catalyzed coupling reactions for the manuf. of drug components in pharmaceutical industry. In addn. to showcasing bond-forming chemistries, focus was given on reaction work-ups and purifications that purge metal catalysts to provide material of sufficient purity. Most of the examples presented originated on process chem. groups in pharmaceutical companies and were selected based on 2 criteria: (a) the transformation was realized on a large scale (at least 100 mmol); (b) the article contains a detailed exptl. procedure. A list of publications reporting coupling reactions for the large-scale synthesis of pharmaceuticals, but do not provide exptl. procedures was also included.(d) Torborg, C.; Beller, M. Recent Applications of Palladium-Catalyzed Coupling Reactions in the Pharmaceutical, Agrochemical, and Fine Chemical Industries. Adv. Synth. Catal. 2009, 351, 3027– 3043, DOI: 10.1002/adsc.200900587Google Scholar1dRecent Applications of Palladium-Catalyzed Coupling Reactions in the Pharmaceutical, Agrochemical, and Fine Chemical IndustriesTorborg, Christian; Beller, MatthiasAdvanced Synthesis & Catalysis (2009), 351 (18), 3027-3043CODEN: ASCAF7; ISSN:1615-4150. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Palladium-catalyzed coupling reactions have become a central tool for the synthesis of biol. active compds. both in academia and industry. Most of these transformations make use of easily available substrates and allow for a shorter and more selective prepn. of substituted arenes and heteroarenes compared to non-catalytic pathways. Notably, mol.-defined palladium catalysts offer high chemoselectivity and broad functional group tolerance. Considering these advantages, it is not surprising that several palladium-catalyzed coupling reactions have been implemented in the last decade into the industrial manuf. of pharmaceuticals and fine chems. In this review different examples from 2001-2008 are highlighted, which have been performed at least on a kilogram scale in the chem. and pharmaceutical industries.(e) Corbet, J.-P.; Mignani, G. Selected Patented Cross-Coupling Reaction Technologies. Chem. Rev. 2006, 106, 2651– 2710, DOI: 10.1021/cr0505268Google Scholar1eSelected patented cross-coupling reaction technologiesCorbet, Jean-Pierre; Mignani, GerardChemical Reviews (Washington, DC, United States) (2006), 106 (7), 2651-2710CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review dealing with cross-coupling reactions : carbon-carbon coupling reactions (Miyaura-Suzuki reaction, Corriu-Kumada-Tamao reaction, Negishi reaction, etc.) and carbon-nitrogen coupling reactions. - 2
For selected reviews on the Suzuki–Miyaura cross-coupling reactions, see:
(a) Hussain, I.; Capricho, J.; Yawer, M. A. Synthesis of Biaryls via Ligand-Free Suzuki–Miyaura Cross-Coupling Reactions: A Review of Homogeneous and Heterogeneous Catalytic Developments. Adv. Synth. Catal. 2016, 358, 3320– 3349, DOI: 10.1002/adsc.201600354Google Scholar2aSynthesis of Biaryls via Ligand-Free Suzuki-Miyaura Cross-Coupling Reactions: A Review of Homogeneous and Heterogeneous Catalytic DevelopmentsHussain, Ibrar; Capricho, Jaworski; Yawer, Mirza A.Advanced Synthesis & Catalysis (2016), 358 (21), 3320-3349CODEN: ASCAF7; ISSN:1615-4150. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The recent developments in the field of Suzuki-Miyaura cross-coupling reaction, in particular with regard to ligand-free catalysis were described. The catalysts outlined here allowed convenient and green synthetic pathways specifically for the construction of C-C bonds. They enabled the synthesis of biaryls by the coupling arylboronic acids with aryl halides. The literature reporting ligand-free synthesis of biaryls from 2010 to May 2015 was reviewed.(b) Lennox, A. J. J.; Lloyd-Jones, G. C. Selection of boron reagents for Suzuki–Miyaura coupling. Chem. Soc. Rev. 2014, 43, 412– 443, DOI: 10.1039/C3CS60197HGoogle Scholar2bSelection of boron reagents for Suzuki-Miyaura couplingLennox, Alastair J. J.; Lloyd-Jones, Guy C.Chemical Society Reviews (2014), 43 (1), 412-443CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Suzuki-Miyaura (SM) cross-coupling is arguably the most widely-applied transition metal catalyzed carbon-carbon bond forming reaction to date. Its success originates from a combination of exceptionally mild and functional group tolerant reaction conditions, with a relatively stable, readily prepd. and generally environmentally benign organoboron reagent. A variety of such reagents were developed for the process, with properties that were tailored for application under specific SM coupling conditions. This review analyses the seven main classes of boron reagent that were developed. The general phys. and chem. properties of each class of reagent are evaluated with special emphasis on the currently understood mechanisms of transmetalation. The methods to prep. each reagent are outlined, followed by example applications in SM coupling.(c) Suzuki, A. Carbon–carbon bonding made easy. Chem. Commun. 2005, 4759– 4763, DOI: 10.1039/b507375hGoogle Scholar2cCarbon-carbon bonding made easySuzuki, AkiraChemical Communications (Cambridge, United Kingdom) (2005), (38), 4759-4763CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)A review. The palladium-catalyzed cross-coupling reaction between organoboranes and org. electrophiles in the presence of base was first developed 30 years ago. It offers a powerful and general methodol. for forming carbon-carbon bonds. The scope of the reaction has continued to evolve and broaden to meet modern synthetic requirements. The role of base additives and the reaction mechanism were discussed. Coupling reactions of arom. halides with boranes were mentioned. Problems involving sterically hindered substrates were addressed. Coupling reactions of alkyl compds. and alkynyl borane derivs. were also discussed. The applicability of Suzuki coupling reactions toward environmentally friendly processes was also pointed out. - 3Johansson Seechurn, C. C. C.; Kitching, M. O.; Colacot, T. J.; Snieckus, V. Palladium-catalyzed cross-coupling: a historical contextual perspective to the 2010 Nobel Prize. Angew. Chem., Int. Ed. 2012, 51, 5062– 5085, DOI: 10.1002/anie.201107017Google Scholar3Palladium-Catalyzed Cross-Coupling: A Historical Contextual Perspective to the 2010 Nobel PrizeJohansson Seechurn, Carin C. C.; Kitching, Matthew O.; Colacot, Thomas J.; Snieckus, VictorAngewandte Chemie, International Edition (2012), 51 (21), 5062-5085CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. In 2010, Richard Heck, Ei-ichi Negishi, and Akira Suzuki joined the prestigious circle of Nobel Laureate chemists for their roles in discovering and developing highly practical methodologies for C-C bond construction. From their original contributions in the early 1970s the landscape of the strategies and methods of org. synthesis irreversibly changed for the modern chemist, both in academia and in industry. In this review, we attempt to trace the historical origin of these powerful reactions, and outline the developments from the seminal discoveries leading to their eminent position as appreciated and applied today.
- 4
For selected reviews on the applications of alkylphosphane ligands in Pd-catalyzed C–C cross-coupling reactions, see:
(a) Rossi, R.; Bellina, F.; Lessi, M.; Manzini, C.; Marianetti, G.; Perego, L. A. Recent Applications of Phosphane-based Palladium Catalysts in Suzuki-Miyaura Reactions Involved in Total Syntheses of Natural Products. Curr. Org. Chem. 2015, 19, 1302– 1409, DOI: 10.2174/1385272819666150506230050Google Scholar4aRecent Applications of Phosphane-based Palladium Catalysts in Suzuki-Miyaura Reactions Involved in Total Syntheses of Natural ProductsRossi, Renzo; Bellina, Fabio; Lessi, Marco; Manzini, Chiara; Marianetti, Giulia; Perego, Luca A.Current Organic Chemistry (2015), 19 (14), 1302-1409CODEN: CORCFE; ISSN:1385-2728. (Bentham Science Publishers Ltd.)A review with refs. on This review with 800 refs. illustrates applications of Suzuki-Miyaura (S.-M.) reactions in the total syntheses of 147 natural products that were made in the period 2010-2013. The review has been organized on the basis of the seven classes of phosphane-based Pd catalysts that have been used in the reported total syntheses. Emphasis has been given to describe and discuss the exptl. conditions of the Pd-catalyzed (S.-M.) cross-coupling reactions also outlining the methods used to prep. the reactants. A focus has also been set on the biol. and pharmacol. properties of the reported natural products as well as on the most significant steps of the reported total syntheses.(b) Li, H.; Johansson Seechurn, C. C. C.; Colacot, T. J. Development of Preformed Pd Catalysts for Cross-Coupling Reactions, Beyond the 2010 Nobel Prize. ACS Catal. 2012, 2, 1147– 1164, DOI: 10.1021/cs300082fGoogle Scholar4bDevelopment of Preformed Pd Catalysts for Cross-Coupling Reactions, Beyond the 2010 Nobel PrizeLi, Hongbo; Johansson Seechurn, Carin C. C.; Colacot, Thomas J.ACS Catalysis (2012), 2 (6), 1147-1164CODEN: ACCACS; ISSN:2155-5435. (American Chemical Society)A review. Strategies for the development of Pd catalysts based on utilization of L2Pd and LPd (L = ligand) species, beyond the contributions of the 2010 Nobel Laureates Richard Heck, Ei-ichi Negishi, and Akira Suzuki, along with their contemporaries, are reviewed. These well-defined, preformed Pd catalysts improve the selectivity and activity of selected cross-coupling reactions by reducing the metal loading and the ligand-to-metal ratios. This review describes predominantly the development of Pd precatalysts over the last 10 years and highlights the benefits often obsd. when using well-defined preformed catalysts relative to those generated in situ.(c) Fu, G. C. The Development of Versatile Methods for Palladium-Catalyzed Coupling Reactions of Aryl Electrophiles through the Use of P(t-Bu)3 and PCy3 as Ligands. Acc. Chem. Res. 2008, 41, 1555– 1564, DOI: 10.1021/ar800148fGoogle Scholar4cThe Development of Versatile Methods for Palladium-Catalyzed Coupling Reactions of Aryl Electrophiles through the Use of P(t-Bu)3 and PCy3 as LigandsFu, Gregory C.Accounts of Chemical Research (2008), 41 (11), 1555-1564CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review. Metal-catalyzed coupling reactions of aryl electrophiles with organometallics and with olefins serve as unusually effective tools for forming new carbon-carbon bonds. By 1998, researchers had developed catalysts that achieved reactions of aryl iodides, bromides, and triflates. Nevertheless, many noteworthy challenges remained; among them were couplings of aryl iodides, bromides, and triflates under mild conditions (at room temp., for example), couplings of hindered reaction partners, and couplings of inexpensive aryl chlorides. This Account highlights some of the progress that has been made in our lab. over the past decade, largely through the appropriate choice of ligand, in achieving these synthetic objectives. In particular, we have established that palladium in combination with a bulky trialkylphosphine accomplishes a broad spectrum of coupling processes, including Suzuki, Stille, Negishi, and Heck reactions. These methods have been applied in a wide array of settings, such as natural-product synthesis, materials science, and bioorg. chem. - 5
For selected studies on oxidative addition to Pd(0) complexes featuring phosphine ligands, see:
(a) Goossen, L. J.; Koley, D.; Hermann, H. L.; Thiel, W. Mechanistic Pathways for Oxidative Addition of Aryl Halides to Palladium(0) Complexes: A DFT Study. Organometallics 2005, 24, 2398– 2410, DOI: 10.1021/om0500220Google ScholarThere is no corresponding record for this reference.(b) Senn, H. M.; Ziegler, T. Oxidative Addition of Aryl Halides to Palladium(0) Complexes: A Density-Functional Study Including Solvation. Organometallics 2004, 23, 2980– 2988, DOI: 10.1021/om049963nGoogle ScholarThere is no corresponding record for this reference.(c) 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, DOI: 10.1021/ar980063aGoogle Scholar5cAnionic Pd(0) and Pd(II) Intermediates in Palladium-Catalyzed Heck and Cross-Coupling ReactionsAmatore, Christian; Jutand, AnnyAccounts of Chemical Research (2000), 33 (5), 314-321CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review with 17 refs. The anions of PdCl2L2 and Pd(OAc)2, precursors of palladium(0) used in cross-coupling and Heck reactions, play a crucial role in these reactions. Tricoordinated anionic complexes Pd0L2Cl- and Pd0L2(OAc)- are the effective catalysts instead of the usually postulated Pd0L2 complex. The anion ligated to the palladium(0) affects the kinetics of the oxidative addn. to ArI as well as the structure and reactivity of the arylpalladium(II) complexes produced in this reaction. Thus, pentacoordinated anionic complexes are formed, ArPdI(Cl)L2- or ArPdI(OAc)L2-, the precursor of neutral trans-ArPd(OAc)L2, instead of the usually postulated trans-ArPdIL2 complex (L = PPh3).(d) Casado, A. L.; Espinet, P. On the Configuration Resulting from Oxidative Addition of RX to Pd(PPh3)4 and the Mechanism of the cis-to-trans Isomerization of [PdRX(PPh3)2] Complexes (R = Aryl, X = Halide). Organometallics 1998, 17, 954– 959, DOI: 10.1021/om9709502Google Scholar5dOn the Configuration Resulting from Oxidative Addition of RX to Pd(PPh3)4 and the Mechanism of the cis-to-trans Isomerization of [PdRX(PPh3)2] Complexes (R = Aryl, X = Halide)Casado, Arturo L.; Espinet, PabloOrganometallics (1998), 17 (5), 954-959CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)The oxidative addn. of RI to Pd(0) and further cis-to-trans isomerization, which are involved in the Stille reaction and other Pd-catalyzed syntheses, have been studied. C6Cl2F3I (C6Cl2F3 = 3,5-dichlorotrifluorophenyl) adds to Pd(PPh3)4 in THF at room temp. giving cis-[Pd(C6Cl2F3)I(PPh3)2] (2), which could be isolated before isomerization to the more stable trans-[Pd(C6Cl2F3)I(PPh3)2] (3). A 19F NMR kinetic study of the isomerization of 2 in THF at 322.6 K reveals a first-order law riso = kiso[2], with kiso = f + g[2]0 + (h + i[2]0)/([PPh3] + j) (f = (1.66 ± 0.03) × 10-4 s-1, g = (2.5 ± 0.2) × 10-3 mol-1 L s-1, h = (1.3 ± 0.7) × 10-8 mol L-1 s-1, i = (4 ± 2) × 10-4 s-1, and j = (1.4 ± 0.7) × 10-5 mol L-1). A four-pathway mechanism accounts for these results: two are assigned to the associative replacements of PPh3 coordinated to 2 by an iodide ligand of I-[Pd] (I-[Pd] = 2 or 3), both THF-assisted (coeff. h) or direct (coeff. i), leading to a monoiodide-bridged intermediate cis-{Pd(C6Cl2F3)I(PPh3)(μ-I)-[Pd]}. The later rearranges via terminal-for-bridging iodide exchange to trans-{Pd(C6Cl2F3)I(PPh3)(μ-I)-[Pd]}, which is finally cleaved by PPh3 yielding complex 3. The other two concurrent pathways are assigned to the isomerization via two consecutive Berry pseudorotations in the pentacoordinated species derived from 2 by coordination of THF (coeff. f) or I-[Pd] (coeff. g). The apparent activation entropy assocd. with kiso is neg. (ΔS⧧ = -21 ± 3 J K-1 mol-1), in agreement with the proposed bimol. mechanisms.(e) Hartwig, J. F.; Paul, F. Oxidative Addition of Aryl Bromide after Dissociation of Phosphine from a Two-Coordinate Palladium(0) Complex, Bis(tri-o-tolylphosphine)Palladium(0). J. Am. Chem. Soc. 1995, 117, 5373– 5374, DOI: 10.1021/ja00124a026Google Scholar5eOxidative Addition of Aryl Bromide after Dissociation of Phosphine from a Two-Coordinate Palladium(0) Complex, Bis(tri-o-tolylphosphine)Palladium(0)Hartwig, John F.; Paul, FredericJournal of the American Chemical Society (1995), 117 (19), 5373-4CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The 14-electron Pd(0) complex [(o-tol)3P]2Pd (1) undergoes oxidative addn. reactions with aryl bromides that occur by a dissociative mechanism. The products of aryl bromide addn. are dimeric with only one phosphine bound to each palladium. In contrast to typical oxidative addns. to Pt(0), Pd(0), and Ni(0) phosphine compds. that occur to intermediates with two or three phosphine ligands, the addn. to 1 occurs after phosphine dissocn. and generation of a monophosphine intermediate. Clean first order kinetics were obsd. when monitoring the disappearance of 1. Inverse first order rate behavior in phosphine demonstrated that the reaction is dissociative. Reaction rates in benzene, toluene, and xylene were essentially identical, suggesting that solvent coordination does not stabilize the unsatd. intermediate. Rather, displacement of a phosphine by coordinated aryl bromide before irreversible oxidative addn., or coordination of a ligand C-H bond appears to stabilize the monophosphine complex. This mechanism is unusual, but falls in accord with those for the opposite reaction, reductive elimination from Pd(II). Many of these elimination reactions occur after phosphine dissocn. and, therefore, lead to initial products that contain only one phosphine ligand. This initial product from reductive elimination is analogous to the intermediate generated from 1 that undergoes oxidative addn.(f) Paul, F.; Patt, J.; Hartwig, J. F. Structural Characterization and Simple Synthesis of {Pd[P(o-Tol)3]2}. Spectroscopic Study and Structural Characterization of the Dimeric Palladium(II) Complexes Obtained by Oxidative Addition of Aryl Bromides and Their Reactivity with Amines. Organometallics 1995, 14, 3030– 3039, DOI: 10.1021/om00006a053Google Scholar5fStructural Characterization and Simple Synthesis of {Pd[P(o-Tol)3]2}. Spectroscopic Study and Structural Characterization of the Dimeric Palladium(II) Complexes Obtained by Oxidative Addition of Aryl Bromides and Their Reactivity with AminesPaul, Frederic; Patt, Joe; Hartwig, John F.Organometallics (1995), 14 (6), 3030-9CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)The Pd(0) complex {Pd[P(o-Tol)3]2} was prepd. by addn. of P(o-Tol)3 to crude [Pd(DBA)2], which is an approx. equimolar mixt. of Pd2(DBA)3 and Pd(DBA)3, followed by crystn. from the reaction medium by addn. of ether. The formation of {Pd[P(o-Tol)3]2} appeared to be driven by its insoly. in the benzene/ether solvent mixt. Benzene solns. of [Pd(DBA)2] and P(o-Tol)3 did not contain amts. of the L2Pd compd. that could be detected by 31P NMR spectroscopy. {Pd[P(o-Tol)3]2} was characterized crystallog. and showed an exactly linear geometry. Similar Pd(0) compds. {Pd[P(2,4-dimethylphenyl)3]2}, {Pd[P(2-methyl-4-fluorophenyl)3]2}, and the low-coordinate trialkylphosphine complex {Pd[P(t-Bu)3]2} were also prepd. by this method, but [Pd(PCy3)2(DBA)] was produced from reactions involving PCy3 and [Pd(DBA)2], and [Pd(TMPP)(DBA)2] was isolated after addn. of tris(1,3,5-trimethoxyphenyl)phosphine (TMPP) to [Pd(DBA)2]. The oxidative addn. of aryl halides to {Pd[P(o-Tol)3]2} at room temp. led to dimeric {Pd[P(o-Tol)3](Ar)(Br)}2. An example of these compds. was characterized crystallog. as well as by soln. mol. wt. anal. This aryl halide complex is dimeric in the solid state as well as in soln. The NMR spectra of the large triarylphosphine complexes showed temp. dependent behavior, presumably due to isomerizations and ligand rotations that occurred on the NMR time scale. The aryl halide compds. did not form four-coordinate monometallic species in the presence of excess P(o-Tol)3, but they did undergo cleavage to four-coordinate monometallic complexes upon addn. of primary and secondary amines.(g) Amatore, C.; Jutand, A.; Suarez, A. Intimate mechanism of oxidative addition to zerovalent palladium complexes in the presence of halide ions and its relevance to the mechanism of palladium-catalyzed nucleophilic substitutions. J. Am. Chem. Soc. 1993, 115, 9531– 9541, DOI: 10.1021/ja00074a018Google Scholar5gIntimate mechanism of oxidative addition to zerovalent palladium complexes in the presence of halide ions and its relevance to the mechanism of palladium-catalyzed nucleophilic substitutionsAmatore, Christian; Jutand, Anny; Suarez, AlejandraJournal of the American Chemical Society (1993), 115 (21), 9531-41CODEN: JACSAT; ISSN:0002-7863.The mechanism of oxidative addn. of aryl halides to low-ligated zerovalent palladium species obtained by redn. of their divalent precursor complexes, PdIICl2(PR3)2, was investigated on the basis of 31P NMR and electrochem. In strong contrast to usual expectations, the reaction proceeds via a complex sequence of steps: (i) fast addn. of the aryl halide to a halide ligated zerovalent palladium center, so as to afford a pentacoordinated anionic arylpalladium(II) center; (ii) a fast but reversible uphill elimination from the pentacoordinated anionic arylpalladium(II) center of a halide ion ligand, possibly through its substitution by a solvent ligand; (iii) rearrangement of this second short-lived intermediate into the thermodynamically stable trans-arylpalladium(II) product of the reaction. On the basis of this detailed mechanism, a new mechanism is proposed for the catalysis of nucleophilic substitutions by palladium complexes. In contrast to the catalytic cycles that are usually considered, this new catalytic cycle accounts for the well-used effects of halide ions as well as that of small metal cations and rationalizes their role in the overall efficiency of palladium-catalyzed nucleophilic substitutions.(h) Amatore, C.; Pflüger, F. Mechanism of oxidative addition of palladium(0) with aromatic iodides in toluene, monitored at ultramicroelectrodes. Organometallics 1990, 9, 2276– 2282, DOI: 10.1021/om00158a026Google Scholar5hMechanism of oxidative addition of palladium(0) with aromatic iodides in toluene, monitored at ultramicroelectrodesAmatore, Christian; Pfluger, FernandoOrganometallics (1990), 9 (8), 2276-82CODEN: ORGND7; ISSN:0276-7333.The mechanism and rates of oxidative addn. of substituted iodobenzenes to tetrakis(triphenylphosphine)palladium(0) is investigated in toluene by means of electrochem. at ultramicroelectrodes. The mechanism is identical with that previously reported in THF (F. F. Fauvarque et al., 1981) and to correspond to a similar slope for Hammett's correlation (ρ = +2 in THF vs +2.3 ± 0.2 in toluene). Moreover, the important change in polarity (εs ≈ 7.58 in THF and 2.38 in toluene) does not affect the enthalpy of activation and leads to similar activation entropies in both solvents. This suggests that the transition state of addn. of aryl halides to the coordinately unsatd. zerovalent palladium complex Pd0(PPh3)2 has no significant ionic character.(i) Fauvarque, J.-F.; Pflüger, F. Kinetics of oxidative addition of zerovalent palladium to aromatic iodides. J. Organomet. Chem. 1981, 208, 419– 427, DOI: 10.1016/S0022-328X(00)86726-1Google Scholar5iKinetics of oxidative addition of zerovalent palladium to aromatic iodidesFauvarque, Jean Francois; Pfluger, Fernando; Troupel, MichelJournal of Organometallic Chemistry (1981), 208 (3), 419-27CODEN: JORCAI; ISSN:0022-328X.The kinetics of oxidative addn. of Pd(0) to arom. iodides in THF contg. the ligand Ph3P were studied by amperometry on the Pd oxidn. wave. The reaction is 1st order each in arom. iodide and Pd(0), and the rate const. is inversely proportional to the free-ligand concn. The reactive Pd(0) intermediate is Pd(PPh3)2. With substituted arom. iodides, the rate consts. give a linear Hammett relationship with ρ = +2. The mechanism is related to an arom. nucleophilic substitution with some assistance from halogen-Pd interaction.(j) Stille, J. K.; Lau, K. S. Y. Acc. Chem. Res. 1977, 10, 434, DOI: 10.1021/ar50120a002Google ScholarThere is no corresponding record for this reference.(k) Fitton, P.; Rick, E. A. The addition of aryl halides to tetrakis(triphenylphosphine)palladium(0). J. Organomet. Chem. 1971, 28, 287– 291, DOI: 10.1016/S0022-328X(00)84578-7Google Scholar5kAddition of aryl halides to tetrakis(triphenylphosphine)palladium(0)Fitton, P.; Rick, Edward A.Journal of Organometallic Chemistry (1971), 28 (2), 287-91CODEN: JORCAI; ISSN:0022-328X.The order of reactivity of halobenzenes towards oxidative addn. to Pd(PPh3)4 was PhI > PhBr > PhCl. Iodobenzene will react with Pd(PPh3)4 in benzene at room temp., and bromobenzene will react at 80°. Chlorobenzene, however, is unreactive even at 135°. Aryl chlorides substituted with electron-donating groups similarly are unreactive, but aryl chlorides substituted with electron-withdrawing groups are reactive. This increased reactivity, and order of reactivity 4-O2NC6H4Cl > 4-CNC6H4Cl > 4-PhCOC6H4Cl > PhCl suggests that the mechanism of the oxidative addn. is similar to that of a bimol. nucleophilic aromatic displacement reaction in which breaking of the bond to the leaving group is involved in the rate detg. step. - 6Organotransition Metal Chemistry From Bonding to Catalysis; Hartwig, J. F., Ed.; University Science Books: Sausalito, CA, 2010; Chapter 19.Google ScholarThere is no corresponding record for this reference.
- 7Pullen, R.; Olding, A.; Smith, J. A.; Bissember, A. C. Manuscript in preparation.Google ScholarThere is no corresponding record for this reference.
- 8Wallow, T. I.; Novak, B. M. Highly Efficient and Accelerated Suzuki Aryl Couplings Mediated by Phosphine-Free Palladium Sources. J. Org. Chem. 1994, 59, 5034– 5037, DOI: 10.1021/jo00096a056Google Scholar8Highly efficient and accelerated Suzuki aryl couplings mediated by phosphine-free palladium sourcesWallow, Thomas I.; Novak, Bruce M.Journal of Organic Chemistry (1994), 59 (17), 5034-7CODEN: JOCEAH; ISSN:0022-3263.Suzuki aryl cross-couplings employing aryl bromides and aryl iodides proceed under mild conditions (65°) with high efficiency (substrate-to-catalyst ratios >500) in the presence of phosphine-free Pd catalysts derived from Pd acetate, Pd2(dba)3·C6H6 (dba = dibenzylideneacetone), and [(η3-C3H5)PdCl]2. Phosphine inhibition is shown to play a key role in limiting catalytic efficiency; qual. comparison studies show that the phosphine-free systems are 1-2 orders of magnitude more active than phosphine-supported catalytic systems. Pd[P(Ph)3]4 proved to be the least active of the catalytic species screened. The phosphine-free methodol. appears to be generally applicable; cross-couplings of aryl iodides yielding biaryls p-MeOC6H4C6H4CF3-p 6 and p-MeOC6H4C6H4Me-o 7 proceed without noticeable steric or electronic effects. Cross-couplings employing aryl bromides are insensitive to electronic effects in the synthesis of 6 but are slowed by steric hindrance in the synthesis of 7. Acceleration of cross-coupling is obsd. in the presence of polar cosolvents and at high pH.
- 9
The poor reactivity of aryl iodides with vinyltributyltin in Stille couplings employing Pd/PPh3 at ∼50 °C has been reported; see:
Farina, V.; Krishnan, B. Large rate accelerations in the stille reaction with tri-2-furylphosphine and triphenylarsine as palladium ligands: mechanistic and synthetic implications. J. Am. Chem. Soc. 1991, 113, 9585– 9595, DOI: 10.1021/ja00025a025Google Scholar9Large rate accelerations in the stille reaction with tri-2-furylphosphine and triphenylarsine as palladium ligands: mechanistic and synthetic implicationsFarina, Vittorio; Krishnan, BalaJournal of the American Chemical Society (1991), 113 (25), 9585-95CODEN: JACSAT; ISSN:0002-7863.The effect of changing the palladium ligands on the rates of typical Stille cross-coupling reactions was studied. Large rate enhancements (typically 102-103 over triphenylphosphine-based catalyst) were obsd. with tri-2-furylphosphine and triphenylarsine, which are recommended as the new ligands of choice in the palladium-catalyzed coupling between olefinic stannanes and electrophiles. On the basis of the evidence presented, the transmetalation, which is the rate-detg. step in the catalytic cycle, is postulated to involve a π-complex between the metal and the stannane double bond. In general, ligands that readily dissoc. from Pd(II) and allow ready formation of this π-complex are the ones that produce the fastest coupling rates. The utility of the new ligands is demonstrated with several synthetic examples. - 10Huff, B. E.; Koenig, T. M.; Mitchell, D.; Staszak, M. A. Synthesis of unsymmetrical biaryls using a modified Suzuki cross-coupling: 4-biphenylcarboxaldehyde. Org. Synth. 1998, 75, 53, DOI: 10.1002/0471264180.os075.07Google ScholarThere is no corresponding record for this reference.
- 11
In general, iodide consumption was comparable to product yield.
There is no corresponding record for this reference. - 12
Analogous results were obtained in experiments employing higher [Pd(PPh3)4] loadings (2.5% and 5%) (see the Supporting Information).
There is no corresponding record for this reference. - 13
We also investigated the effect of incrementally varying the Na2CO3 loading on the outcome of the reaction at 70 °C. Our results suggest that the Na2CO3 loading (1–5 equiv) has little effect on the efficiency of reactions employing p-iodotoluene (see the Supporting Information). Substituting either Cs2CO3 or NMe4OH for Na2CO3 also provided inefficient couplings of p-iodotoluene (see the Supporting Information).
There is no corresponding record for this reference. - 14
We also performed experiments employing higher PPh3 loadings (see the Supporting Information).
There is no corresponding record for this reference. - 15
It has been established that PPh3 can effect the reduction of Pd(OAc)2 to generate catalytically active Pd(0) species. See, for example: ref (5c).
(a) Carole, W. A.; Colacot, T. J. Understanding Palladium Acetate from a User Perspective. Chem. - Eur. J. 2016, 22, 7686– 7695, DOI: 10.1002/chem.201601450Google Scholar15aUnderstanding Palladium Acetate from a User PerspectiveCarole, William A.; Colacot, Thomas J.Chemistry - A European Journal (2016), 22 (23), 7686-7695CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The behavior of palladium acetate is reviewed with respect to its synthesis, characterization, structure (in both soln. and solid state), and activation pathways. In addn., comparisons of catalytic activities between pure palladium acetate and two common byproducts, Pd3(OAc)5(NO2) and polymeric [Pd(OAc)2]n, typically present in com. available material are reviewed. Hence, this minireview serves as a concise guide for the users of palladium acetate from both academia and industry.(b) Amatore, C.; Carré, E.; Jutand, A.; M’Barki, M. Rates and Mechanism of the Formation of Zerovalent Palladium Complexes from Mixtures of Pd(OAc)2 and Tertiary Phosphines and Their Reactivity in Oxidative Additions. Organometallics 1995, 14, 1818– 1826, DOI: 10.1021/om00004a039Google Scholar15bRates and Mechanism of the Formation of Zerovalent Palladium Complexes from Mixtures of Pd(OAc)2 and Tertiary Phosphines and Their Reactivity in Oxidative AdditionsAmatore, Christian; Carre, Emmanuelle; Jutand, Anny; M'Barki, Mohamed AmineOrganometallics (1995), 14 (4), 1818-26CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)Mixts. of Pd(OAc)2 and tertiary phosphines spontaneously afford Pd(0) complexes. Kinetic studies demonstrate that this reaction proceeds from the complex Pd(OAc)2(PR3)2 via an inner-sphere redn. which is the rate-detg. step of the overall reaction. The phosphine is thus oxidized to the corresponding phosphine oxide. The formation of the Pd(0) complex is sensitive to electronic and steric factors. The more the triarylphosphine is substituted by electron-withdrawing groups, the faster the reaction. The Pd(0) complex thus formed reacts with Ph iodide via an oxidn. addn., and this reaction is faster when the phosphine is more electron-rich.(c) Amatore, C.; Jutand, A.; M’Barki, M. Evidence of the formation of zerovalent palladium from Pd(OAc)2 and triphenylphosphine. Organometallics 1992, 11, 3009– 3013, DOI: 10.1021/om00045a012Google Scholar15cEvidence of the formation of zerovalent palladium from Pd(OAc)2 and triphenylphosphineAmatore, Christian; Jutand, Anny; M'Barki, Mohamed AmineOrganometallics (1992), 11 (9), 3009-13CODEN: ORGND7; ISSN:0276-7333.The mixt. Pd(OAc)2 + nPPh3 (n ≥2) and Pd(OAc)2(PPh3)2, commonly used as catalysts in reactions involving aryl and vinyl halides, aryl triflates or allylic acetates, spontaneously generate a zerovalent Pd complex that reacts with PhI. PPh3 reduces the PdII from Pd(OAc)2(PPh3)2 by an intramol. reaction and is oxidized to OPPh3. In the presence of an excess of PPh3, the zerovalent Pd complex generated in situ has the same 31P NMR and cyclic voltammetry properties as those of Pd0(PPh3)4. - 16
Pd(OAc)2 was used in preference to Pdx(dba)y, as challenges in establishing the exact speciation and purity of the latter have been identified. See, for example:
(a) Amatore, C.; Jutand, A. Role of dba in the reactivity of palladium(0) complexes generated in situ from mixtures of Pd(dba)2 and phosphines. Coord. Chem. Rev. 1998, 178–180, 511– 528, DOI: 10.1016/S0010-8545(98)00073-3Google Scholar16aRole of dba in the reactivity of palladium(0) complexes generated in situ from mixtures of Pd(dba)2 and phosphinesAmatore, Christian; Jutand, AnnyCoordination Chemistry Reviews (1998), 178-180 (Pt. 1), 511-528CODEN: CCHRAM; ISSN:0010-8545. (Elsevier Science S.A.)A review with 98 refs. The ligand dba plays a crucial role both on the structure and on the reactivity of Pd(0) complexes generated in situ in mixts. of Pd(dba)2 and phosphine ligands. Whatever the ligand, the major complex is always Pd(dba)L2 where L is a monodentate phosphine ligand or Pd(dba)(L-L) where L-L is a bidentate phosphine ligand. In all cases, the most reactive species in the oxidative addn. with Ph iodide is the lowest ligated complex PdL2 or Pd(L-L) in equil. with the major complex and dba. However, Pd(dba)(L-L) also reacts with Ph iodide. The presence of the major complexes ligated by dba diminishes the concn. of the more reactive species PdL2 or Pd(L-L) and consequently controls the rate of the overall reaction. The overall reactivity is governed both by the intrinsic reactivity of the reactive species and its concn., two factors which can be antagonistic and a nonlinear Hammett correlation of the oxidative addn. with the basicity of the phosphine is obsd. for monodentate ligands.(b) Zalesskiy, S. S.; Ananikov, V. P. Pd2(dba)3 as a Precursor of Soluble Metal Complexes and Nanoparticles: Determination of Palladium Active Species for Catalysis and Synthesis. Organometallics 2012, 31, 2302– 2309, DOI: 10.1021/om201217rGoogle Scholar16bPd2(dba)3 as a Precursor of Soluble Metal Complexes and Nanoparticles: Determination of Palladium Active Species for Catalysis and SynthesisZalesskiy, Sergey S.; Ananikov, Valentine P.Organometallics (2012), 31 (6), 2302-2309CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)Tris(dibenzylideneacetone)dipalladium (Pd2(dba)3) is ubiquitously used as a source of sol. Pd species for catalysis and as a precursor in the synthesis of more complex Pd structures. In spite of the massive usage of this convenient Pd complex, its nature in soln. was not revealed in detail and the applications rely on the assumed state and purity of the compd. A convenient NMR procedure is developed to reveal the nature of Pd2(dba)3 and to det. the purity of the complex. Surprisingly, com. available samples of Pd2(dba)3 may readily contain up to 40% of Pd nanoparticles in a wide range of sizes (10-200 nm). The routinely accepted practice of use of Pd2(dba)3 without anal. of the purity (both com. available and prepd. by common procedures) can introduce significant errors in the estn. of catalyst efficiency and lead to incorrect values of TON, TOF, and reported mol % values in the catalytic procedures. The presence of Pd nanoparticles in the catalyst precursor provides an opportunity for heterogeneous catalytic systems of different nature to be directly accessible from Pd2(dba)3. A modified procedure for the synthesis of Pd2(dba)3·CHCl3 with 99% purity is reported. - 17
See, for example:
Bumagin, N. A.; Bykov, V. V.; Beletskaya, I. P. Synthesis of diaryls from phenylboric acid and aryl iodides in an aqueous medium. Bull. Acad. Sci. USSR, Div. Chem. Sci. 1989, 38, 2206, DOI: 10.1007/BF00962156Google ScholarThere is no corresponding record for this reference.Bumagin, N. A.; Bykov, V. V.; Beletskaya, I. P. Bull. Acad. Sci. USSR, Div. Chem. Sci. (Engl. Transl.) 1989, 38, 2206, DOI: 10.1007/BF00962156Google ScholarThere is no corresponding record for this reference. - 18
We also performed experiments investigating the effect of halide additives (20% NH4I, NH4Br, NH4F, NMe4I, NMe4Br, NMe4Cl, LiBr, LiCl, KBr, KCl, CsBr, or CsCl) on Suzuki–Miyaura couplings (see the Supporting Information).
There is no corresponding record for this reference. - 19
It is generally acknowledged that defining key transmetalation intermediates is even more complicated with aryltrifluoroborate nucleophiles. See, for example:
Lennox, A. J. J.; Lloyd-Jones, G. C. The Slow-Release Strategy in Suzuki–Miyaura Coupling. Isr. J. Chem. 2010, 50, 664– 674, and references cited therein DOI: 10.1002/ijch.201000074Google Scholar19The slow-release strategy in Suzuki-Miyaura couplingLennox, Alastair J. J.; Lloyd-Jones, Guy C.Israel Journal of Chemistry (2010), 50 (5-6), 664-674CODEN: ISJCAT; ISSN:0021-2148. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Despite great advances in metal-catalyzed cross-coupling reactions, their efficacy is often compromised by side reactions, reducing the yield, or requiring a large excess of one component. Suzuki-Miyaura cross-coupling is no exception, as the boronic acid functionality can be susceptible to a range of undesired processes. A no. of methods were developed to mitigate these side reactions, and herein we focus on the slow-release strategy. These conditions involve deployment of a masking reagent that protects the vulnerable boronic acid functionality from degrdn., particularly protodeboronation, while simultaneously facilitating controlled liberation of the active reagent into the catalytic milieu. Under suitably tailored conditions, this dual-action approach ensures that the concn. of the free boronic acid is minimized, thus attenuating its degrdn. but still facilitating transmetallation of the organoboron species with the key organopalladium intermediate. - 20(a) Andersen, N. G.; Keay, B. A. 2-Furyl Phosphines as Ligands for Transition-Metal-Mediated Organic Synthesis. Chem. Rev. 2001, 101, 997– 1030, DOI: 10.1021/cr000024oGoogle Scholar20a2-Furyl Phosphines as Ligands for Transition-Metal-Mediated Organic SynthesisAndersen, Neil G.; Keay, Brian A.Chemical Reviews (Washington, D. C.) (2001), 101 (4), 997-1030CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review with 177 refs.(b) Niemeyer, Z. L.; Milo, A.; Hickey, D. P.; Sigman, M. S. Parameterization of phosphine ligands reveals mechanistic pathways and predicts reaction outcomes. Nat. Chem. 2016, 8, 610– 617, DOI: 10.1038/nchem.2501Google Scholar20bParameterization of phosphine ligands reveals mechanistic pathways and predicts reaction outcomesNiemeyer, Zachary L.; Milo, Anat; Hickey, David P.; Sigman, Matthew S.Nature Chemistry (2016), 8 (6), 610-617CODEN: NCAHBB; ISSN:1755-4330. (Nature Publishing Group)The mechanistic foundation behind the identity of a phosphine ligand that best promotes a desired reaction outcome is often nonintuitive, and thus was addressed in numerous exptl. and theor. studies. Multivariate correlations of reaction outcomes using 38 different phosphine ligands were combined with classic potentiometric analyses to study a Suzuki reaction, for which the site selectivity of oxidative addn. is highly dependent on the nature of the phosphine. These studies shed light on the generality of hypotheses regarding the structural influence of different classes of phosphine ligands on the reaction mechanism(s), and deliver a methodol. that should prove useful in future studies of phosphine ligands.
- 21(a) Tolman, C. A. Electron donor-acceptor properties of phosphorus ligands. Substituent additivity. J. Am. Chem. Soc. 1970, 92, 2953– 2956, DOI: 10.1021/ja00713a006Google Scholar21aElectron donor-acceptor properties of phosphorus ligands. Substituent additivityTolman, Chadwick A.Journal of the American Chemical Society (1970), 92 (10), 2953-6CODEN: JACSAT; ISSN:0002-7863.A rapid method is described for detg. electron donor-acceptor properties of triply connected P ligands based on the A1 carbonyl stretching frequency of Ni(CO)3L in CH2Cl2. Data are given for 70 ligands and a substituent additivity rule is proposed. Forty-seven substituent parameters χi are derived and found to correlate well with Kabachnik's σ parameters, based on ionization consts. of P acids.(b) Otto, S.; Roodt, A. Quantifying the electronic cis effect of phosphine, arsine and stibine ligands by use of rhodium(I) Vaska-type complexes. Inorg. Chim. Acta 2004, 357, 1– 10, DOI: 10.1016/S0020-1693(03)00436-5Google Scholar21bQuantifying the electronic cis effect of phosphine, arsine and stibine ligands by use of rhodium(I) Vaska-type complexesOtto, Stefanus; Roodt, AndreasInorganica Chimica Acta (2004), 357 (1), 1-10CODEN: ICHAA3; ISSN:0020-1693. (Elsevier Science B.V.)The cis effects of phosphine, arsine and stibine ligands were evaluated by measuring the IR stretching frequency in dichloromethane of the carbonyl ligand in Rh(I) Vaska-type complexes, trans-[RhCl(CO)(L)2]. These data were correlated with those obtained by Tolman for the electronic trans influences in [Ni(L)(CO)3]. The electronic contribution, χFc, of ferrocenyl was detd. as 0.8 from these plots by evaluating PPh2Fc as ligand. To accommodate arsine and stibine ligands an addnl. correction term, to compensate for differences in the donor atom, was added to Tolman's equation for calcn. of the Tolman electronic parameter of phosphine ligands. In the resulting equation: v(CONi) = 2056.1 + Σ3i=1 χi + CL values for CL of CP = 0, CAs = -1.5 and CSb = -3.1 are suggested for phosphine, arsine and stibine ligands, resp. The crystal and mol. structures of trans-[RhCl(CO)(PPh2Fc)2]·2C6H6, trans-[RhCl(CO){P(NMe2)3}2] and trans-[RhCl(CO)(AsPh3)2] are reported. The Tolman cone angles for PPh2Fc and P(NMe2)3 were detd. as 169° and 166°, while the effective cone angles for PPh2Fc, P(NMe2)3 and AsPh3 were detd. as 171°, 168° and 147°, resp.(c) Ackermann, M.; Pascariu, A.; Höcher, T.; Siehl, H.-U.; Berge, S. Electronic Properties of Furyl Substituents at Phosphorus and Their Influence on 31P NMR Chemical Shifts. J. Am. Chem. Soc. 2006, 128, 8434– 8440, DOI: 10.1021/ja057085uGoogle Scholar21cElectronic Properties of Furyl Substituents at Phosphorus and Their Influence on 31P NMR Chemical ShiftsAckermann, Marco; Pascariu, Aurelia; Hoecher, Thomas; Siehl, Hans-Ullrich; Berger, StefanJournal of the American Chemical Society (2006), 128 (26), 8434-8440CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The electronic properties of 2-furyl and 3-furyl substituents attached to phosphines and phosphonium salts were studied by IR spectroscopy and exptl. and computational 31P NMR spectroscopy. The heteroarom. systems proved to be electron withdrawing with respect to Ph substituents. However, phosphorus atoms with attached furyl substituents are strongly shielded in NMR. The reason for this phenomenon was studied by solid state 31P MAS NMR expts. The chem. shift tensor was extd., and the orientation within the mols. was detd. The tensor component σ33, which is effected the most by furyl systems, is oriented perpendicular to the P-C bonds of the substituents. P-furyl bonds are shorter than P-Ph bonds. We assume therefore a lower ground-state energy of the mols., because of the electron withdrawing properties of the 2-furyl systems. The σpara component of the 31P NMR magnetic shielding is therefore smaller, which results in an overall increase of the magnetic shielding.
- 22
The observation that oxidative addition is not the turnover-limiting step in this reaction is consistent with previous studies. See, for example:
Amatore, C.; Jutand, A.; Le Duc, G. Kinetic Data for the Transmetalation/Reductive Elimination in Palladium-Catalyzed Suzuki–Miyaura Reactions: Unexpected Triple Role of Hydroxide Ions Used as Base. Chem. - Eur. J. 2011, 17, 2492– 2503, and references cited therein DOI: 10.1002/chem.201001911Google Scholar22Kinetic Data for the Transmetalation/Reductive Elimination in Palladium-Catalyzed Suzuki-Miyaura Reactions: Unexpected Triple Role of Hydroxide Ions Used as BaseAmatore, Christian; Jutand, Anny; Le Duc, GaetanChemistry - A European Journal (2011), 17 (8), 2492-2503CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The mechanism of the reaction of trans-[ArPdX(PPh3)2] (Ar = p-Z-C6H4; Z = CN, F, H; X = I, Br, Cl) with Ar'B(OH)2 (Ar' = p-Z'-C6H4; Z' = CN, H, OMe) has been established in DMF in the presence of the base OH- in the context of real palladium-catalyzed Suzuki-Miyaura reactions. The formation of the cross-coupling product ArAr' and [Pd0(PPh3)3] has been followed through the application of electrochem. techniques. Kinetic data have been obtained for the first time, with detn. of the obsd. rate const., kobs, of the overall reaction. trans-[ArPdX(PPh3)2] is not reactive in the absence of the base. The base OH- plays three roles. It favors the reaction: (1) by formation of trans-[ArPd(OH)(PPh3)2], a key complex which, in contrast to trans-[ArPdX(PPh3)2], reacts with Ar'B(OH)2 (rate-detg. transmetalation), and (2) by unexpected promotion of the reductive elimination from the intermediate trans-[ArPdAr'(PPh3)2], which generates ArAr' and a Pd0 species. Conversely, the base OH- disfavors the reaction by formation of the unreactive anionic Ar'B(OH)3-. As a consequence of these antagonistic effects of OH-, the overall reactivity is controlled by the concn. of OH- and passes through a max. as the concn. of OH- is increased. Therefore, the base favors the rate-detg. transmetalation and unexpectedly also the reductive elimination. - 23
It was reported that trans-[Pd(PPh3)2I2] was formed in reactions with [Pd(PPh3)4] and (iodoethynyl)benzene (in addition to the expected oxidative addition product). See:
(a) Weigelt, M.; Becher, D.; Poetsch, E.; Bruhn, C.; Steinborn, D. Zur oxidativen Addition von 1-Halogenalk-1-inen – Synthese und Struktur von Phenylalkinylpalladium-Komplexen. Z. Anorg. Allg. Chem. 1999, 625, 1542– 1547, DOI: 10.1002/(SICI)1521-3749(199909)625:9<1542::AID-ZAAC1542>3.0.CO;2-9Google ScholarThere is no corresponding record for this reference.The formation of trans-[Pd(PPh3)2I2] from trans-[Pd(PPh3)2(R)(I)] has also been reported:
(b) Gulia, N.; Pigulski, B.; Szafert, S. Palladium End-Capped Polyynes via Oxidative Addition of 1-Haloalkynes to Pd(PPh3)4. Organometallics 2015, 34, 673– 682, DOI: 10.1021/om501105wGoogle Scholar23bPalladium End-Capped Polyynes via Oxidative Addition of 1-Haloalkynes to Pd(PPh3)4Gulia, Nurbey; Pigulski, Bartlomiej; Szafert, SlawomirOrganometallics (2015), 34 (4), 673-682CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)Reported here is the use of 1-haloacetylenes and 1-halopolyynes as synthons for the prepn. of new palladium(II) end-capped polyynes. The 1-haloalkynes were obtained in a series of transformations from para-substituted bromoarenes that included Sonogashira coupling followed by halogenation and chain elongation via Cadiot-Chodkiewicz protocol. The key step for the synthesis of metal complexes was oxidative addn. of 1-haloalkynes to Pd(PPh3)4, which allowed obtaining a series of metal compds. 1-5-CnX with carbon chains up to hexatriyne in 75-100% yield. All the compds. were characterized by NMR and HRMS or elemental anal. The 13C spectra of the 1-haloalkynes showed interesting, although expected, shifts of the carbon chain atoms close to the halogen termini. X-ray crystal structures were obtained for three polyynes-two butadiynes (2-C4[Pd]Br and 3-C4[Pd]Br) and one hexatriyne (1-C6[Pd]Br)-and the latter is the first reported X-ray crystal structure of palladium end-capped hexatriyne.trans-[Pd(PPh3)2I2] was also formed in reactions with [Pd(PPh3)4] and 9-iodo-m-carborane. See:
(c) Marshall, W. J.; Young, R. J., Jr.; Grushin, V. V. Mechanistic Features of Boron–Iodine Bond Activation of B–Iodocarboranes. Organometallics 2001, 20, 523– 533, DOI: 10.1021/om0008575Google ScholarThere is no corresponding record for this reference. - 24
We were also able to isolate and characterize trans-[Pd(PPh3)2I2]. The spectroscopic data obtained on this compound were consistent both with equivalent data reported in the literature and also with an authentic sample of trans-[Pd(PPh3)2I2] that we prepared via a literature procedure:
Hahn, F. E.; Lugger, T.; Beinhoff, M. Palladium(II) Complexes with Benzoxazol-2-ylidene Ligands: Crystal Structures of trans-Chloro(benzoxazol-2-ylidene)bis(triphenylphosphine)palladium(II) Chloride and cis-Diiodo(benzoxazol-2-ylidene)(triphenylphosphine)palladium(II). Z. Naturforsch. B Chem. Sci. 2004, 59, 196– 201Google Scholar24Palladium(II) complexes with benzoxazol-2-ylidene ligands: crystal structures of trans-chloro(benzoxazol-2-ylidene)bis(triphenylphosphine)palladium(II) chloride and cis-diiodo(benzoxazol-2-ylidene)(triphenylphosphine)palladium(II)Hahn, F. Ekkehardt; Luegger, Thomas; Beinhoff, MatthiasZeitschrift fuer Naturforschung, B: Chemical Sciences (2004), 59 (2), 196-201CODEN: ZNBSEN; ISSN:0932-0776. (Verlag der Zeitschrift fuer Naturforschung)The Pd(II) complexes trans-[PdCl(L)(PPh3)2]Cl, 5, and cis-[PdI2(L)PPh3], 7, (L = benzoxazol-2-ylidene) were synthesized by treatment of trans-[PdX2(PPh3)2] (4: X = Cl, 6: X = iodo) with 2-(trimethylsiloxy)phenyl isocyanide 1, and subsequent hydrolysis of the Si-O bond. The crystal structures of 5 and 7·CH2Cl2 were established by x-ray diffraction. NMR and IR studies indicate that the unexpected cis-configuration of 7 obtained from trans-[PdI2(PPh3)2] is not the result of a soln. equil. between the cis and the trans isomers. - 25
When the study shown in eq 2 was performed in DMF/H2O instead of n-PrOH/H2O, similar results were obtained. When the experiment shown in eq 2 was performed using p-bromotoluene instead of p-iodotoluene, 31P NMR spectroscopy indicated that trans-[Pd(PPh3)2Br2] was not present in the reaction mixture. trans-[Pd(PPh3)2Br2] was also not observed in equivalent experiments performed in DMF/H2O.
There is no corresponding record for this reference. - 26Carrow, B. P.; Hartwig, J. F. Distinguishing Between Pathways for Transmetalation in Suzuki–Miyaura Reactions. J. Am. Chem. Soc. 2011, 133, 2116– 2119, DOI: 10.1021/ja1108326Google Scholar26Distinguishing between pathways for transmetalation in Suzuki-Miyaura reactionsCarrow, Brad P.; Hartwig, John F.Journal of the American Chemical Society (2011), 133 (7), 2116-2119CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Transmetalation of arylpalladium complexes proceeds rather via palladium hydroxo complex and arylboronic acid pathway, than via arylpalladium halide and aryl(trihydroxo)borate. A systematic study of the stoichiometric reactions of isolated arylpalladium hydroxo and halide complexes with arylboronic acids and aryltrihydroxyborates is performed to evaluate the relative rates of the two reaction pathways commonly proposed to account for transmetalation in the Suzuki-Miyaura reaction. On the basis of the relative populations of the palladium and organoboron species generated under conditions common for the catalytic process and the obsd. rate consts. for the stoichiometric reactions between the two classes of reaction components, we conclude that the reaction of a palladium hydroxo complex with boronic acid, not the reaction of a palladium halide complex with trihydroxyborate, accounts for transmetalation in catalytic Suzuki-Miyaura reactions conducted with weak base and aq. solvent mixts.
- 27
For a recent review on transmetalation in the Suzuki–Miyaura cross-coupling, see:
Lennox, A. J. J.; Lloyd-Jones, G. C. Transmetalation in the Suzuki-Miyaura coupling: the fork in the trail. Angew. Chem., Int. Ed. 2013, 52, 7362– 7370, DOI: 10.1002/anie.201301737Google Scholar27Transmetalation in the Suzuki-Miyaura Coupling: The Fork in the TrailLennox, Alastair J. J.; Lloyd-Jones, Guy C.Angewandte Chemie, International Edition (2013), 52 (29), 7362-7370CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The Suzuki-Miyaura coupling is one of the few transition-metal-catalyzed C-C bond-forming reactions that were used in applications ranging from discovery chem. to manufg. processes. Although coupling proceeds through the generic three-stage oxidative addn., transmetalation, reductive elimination' sequence, there are a no. of features that differentiate the Suzuki-Miyaura process from other transition-metal-catalyzed cross-couplings. Most of these features are centered around, or are a consequence of, activation of the boron reagent for transmetalation through one or both of two distinct pathways. This review focuses on the evidence that was presented for this fork in the trail, and the potential to apply such mechanistic insight to the design of reaction conditions. - 28
For selected studies on transmetalation in the Suzuki–Miyaura cross-coupling, see: refs (22and26).
(a) Thomas, A. A.; Wang, H.; Zahrt, A. F.; Denmark, S. E. Structural, Kinetic, and Computational Characterization of the Elusive Arylpalladium(II)boronate Complexes in the Suzuki–Miyaura Reaction. J. Am. Chem. Soc. 2017, 139, 3805– 3821, DOI: 10.1021/jacs.6b13384Google Scholar28aStructural, Kinetic, and Computational Characterization of the Elusive Arylpalladium(II)boronate Complexes in the Suzuki-Miyaura ReactionThomas, Andy A.; Wang, Hao; Zahrt, Andrew F.; Denmark, Scott E.Journal of the American Chemical Society (2017), 139 (10), 3805-3821CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The existence of the oft-invoked intermediates contg. the crucial Pd-O-B subunit, the "missing link", has been established in the Suzuki-Miyaura cross-coupling reaction. The use of low-temp., rapid injection NMR spectroscopy (RI-NMR), kinetic studies, and computational anal. has enabled the generation, observation, and characterization of these highly elusive species. The ability to confirm the intermediacy of Pd-O-B-contg. species provided the opportunity to clarify mechanistic aspects of the transfer of the org. moiety from boron to palladium in the key transmetalation step. Specifically, these studies establish the identity of two different intermediates contg. Pd-O-B linkages, a tri-coordinate (6-B-3) boronic acid complex and a tetra-coordinate (8-B-4) boronate complex, both of which undergo transmetalation leading to the cross-coupling product. Two distinct mechanistic pathways have been elucidated for stoichiometric reactions of these complexes: (1) transmetalation via an unactivated 6-B-3 intermediate that dominates in the presence of an excess of ligand, and (2) transmetalation via an activated 8-B-4 intermediate that takes place with a deficiency of ligand.(b) Thomas, A. A.; Denmark, S. E. Pre-transmetalation intermediates in the Suzuki-Miyaura reaction revealed: The missing link. Science 2016, 352, 329– 332, DOI: 10.1126/science.aad6981Google Scholar28bPre-transmetalation intermediates in the Suzuki-Miyaura reaction revealed: The missing linkThomas, Andy A.; Denmark, Scott E.Science (Washington, DC, United States) (2016), 352 (6283), 329-332CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Despite the widespread application of Suzuki-Miyaura cross-coupling to forge carbon-carbon bonds, the structure of the reactive intermediates underlying the key transmetalation step from the boron reagent to the palladium catalyst remains uncertain. Here we report the use of low-temp. rapid injection NMR spectroscopy and kinetic studies to generate, observe, and characterize these previously elusive complexes. Specifically, this work establishes the identity of three different species contg. palladium-oxygen-boron linkages, a tricoordinate boronic acid complex, and two tetracoordinate boronate complexes with 2:1 and 1:1 stoichiometry with respect to palladium. All of these species transfer their boron-bearing aryl groups to a coordinatively unsatd. palladium center in the crit. transmetalation event.(c) Ortuño, M. A.; Lledós, A.; Maseras, F.; Ujaque, G. The Transmetalation Process in Suzuki–Miyaura Reactions: Calculations Indicate Lower Barrier via Boronate Intermediate. ChemCatChem 2014, 6, 3132– 3138, DOI: 10.1002/cctc.201402326Google Scholar28cThe transmetalation process in Suzuki-Miyaura reactions: Calculations indicate lower barrier via boronate IntermediateOrtuno, Manuel A.; Lledos, Agusti; Maseras, Feliu; Ujaque, GregoriChemCatChem (2014), 6 (11), 3132-3138CODEN: CHEMK3; ISSN:1867-3880. (Wiley-VCH Verlag GmbH & Co. KGaA)Recent exptl. reports have called into question the validity of the boronate mechanism (through a R-B(OH)3- intermediate) for the transmetalation step in the Suzuki-Miyaura cross-coupling, favoring instead the palladium hydroxo pathway (through an [LnPd(R')(OH)] intermediate). Herein we report DFT calcns. with the M06 functional performed on realistic model systems, including a combination of explicit solvent mols. along with a continuum method. These computational results support the boronate mechanism. The mechanistic proposal is shown to be compatible with the available exptl. evidence.(d) Lima, C. F. R. A. C.; Rodrigues, A. S. M. C.; Silva, V. L. M.; Silva, A. M. S.; Santos, L. M. N. B. F. Role of the Base and Control of Selectivity in the Suzuki–Miyaura Cross-Coupling Reaction. ChemCatChem 2014, 6, 1291– 1302, DOI: 10.1002/cctc.201301080Google Scholar28dRole of the Base and Control of Selectivity in the Suzuki-Miyaura Cross-Coupling ReactionLima, Carlos F. R. A. C.; Rodrigues, Ana S. M. C.; Silva, Vera L. M.; Silva, Artur M. S.; Santos, Luis M. N. B. F.ChemCatChem (2014), 6 (5), 1291-1302CODEN: CHEMK3; ISSN:1867-3880. (Wiley-VCH Verlag GmbH & Co. KGaA)The outcome of the Suzuki-Miyaura cross-coupling for the direct competition reaction between two boronic acids was evaluated under routine synthesis conditions. The reaction selectivity depends on the amt. of the base used, with fewer bases favoring the reactivity of the boronic acid with lower pKa (stronger acid). The dependence of the reaction selectivity on base stoichiometry increases with the increase in the difference in the pKa values of the competing boronic acids. These results confirm a relation between acid-base chem. and the Suzuki-Miyaura reaction catalytic cycle. Also, under these specific conditions, the most reactive organoboron species toward transmetalation is the borate anion R-B(OH)3- instead of the neutral boronic acid R-B(OH)2. Hence, the main role of the base in the reaction mechanism is to increase the reactivity of the boronic acid toward the Pd-halide complex by converting it into the resp. organoborate. Boric acid, an important reaction byproduct, affects the selectivity in the Suzuki reaction because its gradual formation in the reaction medium disturbs the acid-base equil.(e) Amatore, C.; Jutand, A.; Le Duc, G. Mechanistic Origin of Antagonist Effects of Usual Anionic Bases (OH–, CO32–) as Modulated by their Countercations (Na+, Cs+, K+) in Palladium-Catalyzed Suzuki–Miyaura Reactions. Chem. - Eur. J. 2012, 18, 6616– 6625, DOI: 10.1002/chem.201200516Google Scholar28eMechanistic Origin of Antagonist Effects of Usual Anionic Bases (OH-, CO32-) as Modulated by their Countercations (Na+, Cs+, K+) in Palladium-Catalyzed Suzuki-Miyaura ReactionsAmatore, Christian; Jutand, Anny; Le Duc, GaetanChemistry - A European Journal (2012), 18 (21), 6616-6625, S6616/1-S6616/8CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The mechanism of the reaction of trans-ArPdBrL2 (Ar = p-Z-C6H4, Z = CN, H; L = PPh3) with Ar'B(OH)2 (Ar' = p-Z'-C6H4, Z' = H, CN, MeO), which is a key step in the Suzuki-Miyaura process, was established in DMF with two bases, acetate (Bu4NOAc) or carbonate (Cs2CO3) and compared with that of hydroxide (Bu4NOH). As anionic bases are inevitably introduced with a countercation M+ (e.g., M+OH-), the role of cations in the transmetalation/reductive elimination was 1st studied. Cations M+ (Na+, Cs+, K+) are not innocent since they induce an unexpected decelerating effect in the transmetalation via their complexation to the OH ligand in the reactive ArPd(OH)L2, partly inhibiting its transmetalation with Ar'B(OH)2. A decreasing reactivity order is obsd. when M+ is assocd. with OH-: Bu4N+> K+> Cs+> Na+. Acetates give trans-ArPd(OAc)L2, which does not undergo transmetalation with Ar'B(OH)2. This explains why acetates are not used as bases in Suzuki-Miyaura reactions that involve Ar'B(OH)2. Carbonates (Cs2CO3) give rise to slower reactions than those performed from Bu4NOH at the same concn., even if the reactions are accelerated in the presence of H2O due to the generation of OH-. The mechanism of the reaction with carbonates is then similar to that established for Bu4NOH, involving ArPd(OH)L2 in the transmetalation with Ar'B(OH)2. Due to the low concn. of OH- generated from CO32- in H2O, both transmetalation and reductive elimination result slower than those performed from Bu4NOH at equal concns. as Cs2CO3. Therefore, the overall reactivity is finely tuned by the concn. of the common base OH- and the ratio [OH-]/[Ar'B(OH)2]. Hence, the anionic base (pure OH- or OH- generated from CO32-) assocd. with its countercation (Na+, Cs+, K+) plays four antagonist kinetic roles: acceleration of the transmetalation by formation of the reactive ArPd(OH)L2, acceleration of the reductive elimination, deceleration of the transmetalation by formation of unreactive Ar'B(OH)3- and by complexation of ArPd(OH)L2 by M+.(f) Schmidt, A. F.; Kurokhtina, A. A.; Larina, E. V. Role of a base in Suzuki-Miyaura reaction. Russ. J. Gen. Chem. 2011, 81, 1573– 1574, DOI: 10.1134/S1070363211070334Google Scholar28fRole of a base in Suzuki-Miyaura reactionSchmidt, A. F.; Kurokhtina, A. A.; Larina, E. V.Russian Journal of General Chemistry (2011), 81 (7), 1573-1574CODEN: RJGCEK; ISSN:1070-3632. (MAIK Nauka/Interperiodica)There is no expanded citation for this reference.(g) Butters, M.; Harvey, J.; Jover, J.; Lennox, A.; Lloyd-Jones, G.; Murray, P. Aryl Trifluoroborates in Suzuki–Miyaura Coupling: The Roles of Endogenous Aryl Boronic Acid and Fluoride. Angew. Chem., Int. Ed. 2010, 49, 5156– 5160, DOI: 10.1002/anie.201001522Google Scholar28gAryl Trifluoroborates in Suzuki-Miyaura Coupling: The Roles of Endogenous Aryl Boronic Acid and FluorideButters, Mike; Harvey, Jeremy N.; Jover, Jesus; Lennox, Alastair J. J.; Lloyd-Jones, Guy C.; Murray, Paul M.Angewandte Chemie, International Edition (2010), 49 (30), 5156-5160, S5156/1-S5156/68CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)The Suzuki-Miyaura coupling (SM) of 4-FC6H4BF3- K+ (I) with 3,5-(CF3)2C6H3Br was studied in a toluene/ water biphasic soln., and in a THF/ water soln.; both systems are commonly employed for the SM of trifluoroborates. The reactions in toluene/water, failed to go to completion: turnover ceased after 6 h, affording 55% of the base-catalyzed protodeboronation product PhF and ≤ 32% of coupling product 4-[3,5-(CF3)2C6H3]C6H4F (II). In aq. THF the reaction proceeded much more efficiently (5.5 h; > 95% yield of II), with few side products (< 0.1-2% ), even when the reaction was performed in air. But reaction of 4-C6H4B(OH)2 under identical conditions, gave II in variable yield, and afforded substantially more of side products 4-(4-FC6H4)C6H44-F and 4-FC6H4OH (2-40% ), compared to trifluoroborate substrate I. The reaction intermediates were established and a mechanism, supported by B3LYP calcns. was proposed for the SM.(h) Huang, Y.-L.; Weng, C.-M.; Hong, F.-E. Density Functional Studies on Palladium-Catalyzed Suzuki–Miyaura Cross-Coupling Reactions Assisted by N- or P-Chelating Ligands. Chem. - Eur. J. 2008, 14, 4426– 4434, DOI: 10.1002/chem.200800011Google Scholar28hDensity functional studies on palladium-catalyzed Suzuki-Miyaura cross-coupling reactions assisted by N- or P-chelating ligandsHuang, Yi-Luen; Weng, Chia-Ming; Hong, Fung-E.Chemistry - A European Journal (2008), 14 (14), 4426-4434CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)DFT studies with the B3LYP functional have been carried out on the Suzuki-Miyaura cross-coupling reactions of Ph chloride and phenylboronic acid catalyzed by palladium complexes with N- or P-chelating ligands. The full catalytic cycle, from the addn. of reactants to the catalyst to the release of the cross-coupled product from the complexed intermediate, has been examd. The stages within the cycle, such as oxidative addn., transmetalation, and reductive elimination, were validated by linking the mechanistically relevant intermediates and transition states. Various derivs. of diimine, diphosphine, and diamine were considered as potential model ligands. The catalytic reaction employing diimine as the chelating ligand has been verified as the one with the most energetically feasible route.(i) Braga, A. A. C.; Ujaque, G.; Maseras, F. A DFT Study of the Full Catalytic Cycle of the Suzuki–Miyaura Cross-Coupling on a Model System. Organometallics 2006, 25, 3647– 3658, DOI: 10.1021/om060380iGoogle Scholar28iA DFT Study of the Full Catalytic Cycle of the Suzuki-Miyaura Cross-Coupling on a Model SystemBraga, Ataualpa A. C.; Ujaque, Gregori; Maseras, FeliuOrganometallics (2006), 25 (15), 3647-3658CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)A computational study with the Becke3LYP DFT functional is carried out on the cross-coupling reaction of vinyl bromide, H2C:CHBr and vinylboronic acid, H2C:CHB(OH)2 catalyzed by Pd diphosphine [Pd(PH3)2] in the presence of an excess of OH- base. The full catalytic cycle is computed, starting from the sepd. reactants and the catalyst and finishing with the cross-coupled product and the regeneration of the catalyst. The different stages in the cycle (oxidative addn., isomerization, transmetalation, reductive elimination) were characterized through calcn. of the corresponding intermediates and transition states. Different alternative mechanisms are considered, depending on the no. of phosphine ligands at Pd, and on the cis or trans isomery around the metal center. The results indicate the existence of a no. of competitive pathways of reasonably low energy.(j) Braga, A. A. C.; Morgon, N. H.; Ujaque, G.; Lledós, A.; Maseras, F. Computational study of the transmetalation process in the Suzuki–Miyaura cross-coupling of aryls. J. Organomet. Chem. 2006, 691, 4459– 4466, DOI: 10.1016/j.jorganchem.2006.02.015Google Scholar28jComputational study of the transmetalation process in the Suzuki-Miyaura cross-coupling of arylsBraga, Ataualpa A. C.; Morgon, Nelson H.; Ujaque, Gregori; Lledos, Agusti; Maseras, FeliuJournal of Organometallic Chemistry (2006), 691 (21), 4459-4466CODEN: JORCAI; ISSN:0022-328X. (Elsevier Ltd.)The transmetalation step of the Suzuki-Miyaura cross-coupling between aryl groups is analyzed by DFT calcns. with the Becke3LYP functional. The halide considered is PhBr, and the organoboronic acid is PhB(OH)2. The model catalyst is Pd(PH3)2, and the base, OH-. The transmetalation is considered to start from the Pd(Ph)(PH3)2Br complex, the product of the oxidative addn. The results are compared with those of a previous study on the analogous reaction with vinyl groups, and it is shown that the reaction mechanism is very similar.(k) Braga, A. A. C.; Morgon, N. H.; Ujaque, G.; Maseras, F. Computational Characterization of the Role of the Base in the Suzuki–Miyaura Cross-Coupling Reaction. J. Am. Chem. Soc. 2005, 127, 9298– 9307, DOI: 10.1021/ja050583iGoogle Scholar28kComputational characterization of the role of the base in the Suzuki-Miyaura cross-coupling reactionBraga, Ataualpa A. C.; Morgon, Nelson H.; Ujaque, Gregori; Maseras, FeliuJournal of the American Chemical Society (2005), 127 (25), 9298-9307CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The role of the base in the transmetalation step of the Suzuki-Miyaura cross-coupling reaction is analyzed computationally by means of DFT calcns. with the Becke3LYP functional. The model system studied consists of Pd(CH:CH2)(PH3)2Br as the starting catalyst complex, CH2:CHB(OH)2 as the organoboronic acid, and OH- as the base. The two main mechanistic proposals, consisting of the base attacking first either the palladium complex or the organoboronic acid, are evaluated through geometry optimization of the corresponding intermediates and transition states. Supplementary calcns. are carried out on the uncatalyzed reaction and on a process where the starting complex is Pd(CH:CH2)(PH3)2(OH). These calcns., considered together with available exptl. data, strongly suggest that the main mechanism of transmetalation in the catalytic cycle starts with the reaction of the base and the organoboronic acid.(l) Miyaura, N. Cross-coupling reaction of organoboron compounds via base-assisted transmetalation to palladium(II) complexes. J. Organomet. Chem. 2002, 653, 54– 57, DOI: 10.1016/S0022-328X(02)01264-0Google Scholar28lCross-coupling reaction of organoboron compounds via base-assisted transmetalation to palladium(II) complexesMiyaura, NorioJournal of Organometallic Chemistry (2002), 653 (1-2), 54-57CODEN: JORCAI; ISSN:0022-328X. (Elsevier Science B.V.)A review. Until recently, organoboronic acids have been of limited use in org. synthesis due to their low reactivity for ionic reactions. During the past two decades, however, it has become increasingly clear that they are valuable reagents capable of undergoing many carbon-carbon bond formations in org. syntheses. In 1979, catalytic carbon-carbon bond formation via transmetalation between organoboron compds. and palladium(II) halides was found to proceed in the presence of a base.(m) Matos, K.; Soderquist, J. A. Alkylboranes in the Suzuki–Miyaura Coupling: Stereochemical and Mechanistic Studies. J. Org. Chem. 1998, 63, 461– 470, DOI: 10.1021/jo971681sGoogle ScholarThere is no corresponding record for this reference.(n) Aliprantis, A. O.; Canary, J. W. Observation of Catalytic Intermediates in the Suzuki Reaction by Electrospray Mass Spectrometry. J. Am. Chem. Soc. 1994, 116, 6985– 6986, DOI: 10.1021/ja00094a083Google Scholar28nObservation of Catalytic Intermediates in the Suzuki Reaction by Electrospray Mass SpectrometryAliprantis, Antonios O.; Canary, James W.Journal of the American Chemical Society (1994), 116 (15), 6985-6CODEN: JACSAT; ISSN:0002-7863.Pyridyl bromides, e.g., I (R = H, CHO, X = 2, 3-Br), couple with phenylboronic acids II (R1 = H, Me, R2 = H, Me) in the presence of Pd[PPh3]4 catalyst to give (2- and 3-aryl-substituted)pyridines III. Intermdiate formation of palladium complexes for this coupling reaction, e.g., [(pyrH)(R1R2C6H3)Pd(PPh3)2]+, are obsd. by electrospray Mass Spectrometry.(o) Smith, G. B.; Dezeny, G. C.; Hughes, D. L.; King, A. O.; Verhoeven, T. Mechanistic Studies of the Suzuki Cross-Coupling Reaction. J. Org. Chem. 1994, 59, 8151– 8156, DOI: 10.1021/jo00105a036Google Scholar28oMechanistic Studies of the Suzuki Cross-Coupling ReactionSmith, George B.; Dezeny, George C.; Hughes, David L.; King, Anthony O.; Verhoeven, Thomas R.Journal of Organic Chemistry (1994), 59 (26), 8151-6CODEN: JOCEAH; ISSN:0022-3263. (American Chemical Society)The key step in the synthesis of the drug losartan is a palladium-catalyzed cross-coupling reaction of an aryl bromide and a boronic acid. The reaction scheme was defined in kinetic studies using HPLC, and computer simulation served to depict the time dependence of the concns. of palladium species, which were not obsd. exptl. Two catalyst poisons were identified and characterized. One was an isomeric impurity of the aryl bromide; the other was formed in the reaction mixt. upon hydrolysis of the boronic acid and two of its impurities.(p) Thomas, A. A.; Zahrt, A. F.; Delaney, C. P.; Denmark, S. E. Elucidating the Role of the Boronic Esters in the Suzuki–Miyaura Reaction: Structural, Kinetic, and Computational Investigations. J. Am. Chem. Soc. 2018, 140, 4401– 4416, DOI: 10.1021/jacs.8b00400Google Scholar28pElucidating the Role of the Boronic Esters in the Suzuki-Miyaura Reaction: Structural, Kinetic, and Computational InvestigationsThomas, Andy A.; Zahrt, Andrew F.; Delaney, Connor P.; Denmark, Scott E.Journal of the American Chemical Society (2018), 140 (12), 4401-4416CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The Suzuki-Miyaura reaction is the most practiced palladium-catalyzed, cross-coupling reaction because of its broad applicability, low toxicity of the metal (B), and the wide variety of com. available boron substrates. A wide variety of boronic acids and esters, each with different properties, have been developed for this process. Despite the popularity of the Suzuki-Miyaura reaction, the precise manner in which the org. fragment is transferred from boron to palladium has remained elusive for these reagents. Herein, we report the observation and characterization of pretransmetalation intermediates generated from a variety of commonly employed boronic esters. The ability to confirm the intermediacy of pretransmetalation intermediates provided the opportunity to clarify mechanistic aspects of the transfer of the org. moiety from boron to palladium in the key transmetalation step. A series of structural, kinetic, and computational investigations revealed that boronic esters can transmetalate directly without prior hydrolysis. Furthermore, depending on the boronic ester employed, significant rate enhancements for the transfer of the B-aryl groups were obsd. Overall, two crit. features were identified that enable the transfer of the org. fragment from boron to palladium: (1) the ability to create an empty coordination site on the palladium atom and (2) the nucleophilic character of the ipso carbon bound to boron. Both of these features ultimately relate to the electron d. of the oxygen atoms in the boronic ester. - 29
Lloyd-Jones and Lennox note that “Elucidation of the dominant pathway to transmetalation is not at all straightforward: one must establish the kinetically active boron and palladium intermediates, and they may not necessarily be the most abundant species present in the medium.”. (27)
There is no corresponding record for this reference. - 30
Detailed studies exploring the transmetalation of trans-[Pd(PPh3)2(Ph)(I)] with p-tol-B(OH)3K (and 18-crown-6) and Pd(PPh3)(Ph)(μ-OH)]2 with p-tol-B(OH)2 under stoichiometric conditions were performed in THF/H2O solvent mixtures. (26) Extensive studies employing electrochemical methods to investigate the transmetalation of trans-[Pd(PPh3)2(Ar)(X)] with Ar-B(OH)2 (n-Bu4NOH as the base) under catalytic conditions were performed in a DMF/MeOH solvent mixture. (22)
There is no corresponding record for this reference. - 31
Germane observations regarding the adverse effect of PPh3 have also been reported in Pd-catalyzed Stille cross-couplings; see ref (9).
(a) Farina, V.; Baker, S. R.; Benigni, D. A.; Hauck, S. I.; Sapino, C., Jr Palladium catalysis in cephalosporin chemistry: general methodology for the synthesis of cephem side chains. J. Org. Chem. 1990, 55, 5833– 5847, DOI: 10.1021/jo00310a014Google Scholar31aPalladium catalysis in cephalosporin chemistry: general methodology for the synthesis of cephem side chainsFarina, Vittorio; Baker, Stephen R.; Benigni, Daniel A.; Hauck, Sheila I.; Sapino, Chester, Jr.Journal of Organic Chemistry (1990), 55 (23), 5833-47CODEN: JOCEAH; ISSN:0022-3263.The Pd-catalyzed coupling of 3-(triflyloxy)cephems with organotin compds.leads to 3-alkenyl-, 3-alkynyl-, and 3-arylcephems under exceptionally mild conditions. While this approach was not satisfactory for 3-allylcephems, the related coupling of easily available 3-(chloromethyl)cephems with stannanes provided a high-yielding route to such 3-allylcephems and also to 3-benzyl- and 3-homoallylcephems. The choice of the catalyst was crucial in both cases. PPh3-based catalysts were quite unsatisfactory. A much better ligand in this respect was tri(2-furyl)phosphine, which is introduced as a useful new ligand in organopalladium chem. The effect of this ligand on coupling rates is discussed. It is suggested that reduced electron d. at Pd enhances the rate of the transmetalation, considered to be the rate-detg. step in these coupling reactions. Cases of unexpected transfer order among unsym. substituted stannanes were obsd., as well as a complex coupling reaction between (chloromethyl)cephems and alkynylstannanes. These new facets of the Stille coupling, as well as the potential of this chem. for the development of new cephalosporin antibiotics, are discussed.(b) Scott, W. J.; Stille, J. K. Palladium-catalyzed coupling of vinyl triflates with organostannanes. Synthetic and mechanistic studies. J. Am. Chem. Soc. 1986, 108, 3033– 3040, DOI: 10.1021/ja00271a037Google Scholar31bPalladium-catalyzed coupling of vinyl triflates with organostannanes. Synthetic and mechanistic studiesScott, William J.; Stille, J. K.Journal of the American Chemical Society (1986), 108 (11), 3033-40CODEN: JACSAT; ISSN:0002-7863.The palladium-catalyzed coupling reaction of vinyl triflates, e.g., (I), with acetylenic, vinyl, allyl, and alkyl tin reagents in the presence of LiCl or another suitable salt takes place in high yields under mild reaction conditions; however, benzyl and Ph tin reagents give poor yield of coupled product. The utilization of a tin or silicon hydride reagent in place of the organotin partner yields the alkene by reductive cleavage of the triflate group. The palladium-catalyzed reaction of vinyl triflates with Me3SnSnMe3 gives vinyl stannanes in high yields. Regioselectively formed vinyl triflates can be used to produce 1,3-dienes as the regioisomeric coupled products. - 33
For a study focused on the reactivity of Pd(0) complexes derived from mixtures of Pd(dba)2 and PPh3 and P(2-furyl)3 ligands, see for example:
Amatore, C.; Jutand, A.; Meyer, G.; Atmani, H.; Khalil, F.; Chahdi, F. O. Comparative Reactivity of Palladium(0) Complexes Generated in Situ in Mixtures of Triphenylphosphine or Tri-2-furylphosphine and Pd(dba)2. Organometallics 1998, 17, 2958– 2964, DOI: 10.1021/om971064uGoogle Scholar33Comparative Reactivity of Palladium(0) Complexes Generated in Situ in Mixtures of Triphenylphosphine or Tri-2-furylphosphine and Pd(dba)2Amatore, Christian; Jutand, Anny; Meyer, Gilbert; Atmani, Hamid; Khalil, Fouad; Chahdi, Fouad OuazzaniOrganometallics (1998), 17 (14), 2958-2964CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)For PPh3, mixts. of Pd(dba)2 and nTFP (TFP = tri-2-furylphosphine, n ≥ 2) in DMF and THF (S) lead to the formation of Pd(dba)(TFP)2, SPd(TFP)3 in equil. with SPd(TFP)2. The substitution of dba by the phosphine in Pd(dba)L2 to form SPdL3 is easier for L = TPF than for L = PPh3. The less ligated complex SPd(TFP)2 is the reactive species in the oxidative addn. with Ph iodide. In THF, {Pd(dba)2 + nTFP}, a mixt. often used as a catalyst promoter in several synthetic org. reactions, is found to be less reactive than {Pd(dba)2 + nPPh3} for small values of n (n = 2 or 4) whereas it is more reactive for higher values of n (n > 6). Conversely, in DMF, {Pd(dba)2 + nTFP} is always found to be more reactive than {Pd(dba)2 + nPPh3} whatever n (n ≥ 2). - 34
Farina and Krishnan also state “When a large excess of P(2-furyl)3 was added to a solution of trans-[Pd(PPh3)2(Ph)(I)] and PPh3, no trace of trans-[Pd{P(2-furyl)3}2(Ph)(I)] was observed, the corresponding signal for P(2-furyl)3 being the only new peak in the spectrum.”. (9)
There is no corresponding record for this reference.
Cited By
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by ACS Publications if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
This article is cited by 24 publications.
- Angus Olding, Curtis C. Ho, Nigel T. Lucas, Brian F. Yates, Allan J. Canty, Alex C. Bissember. Suzuki–Miyaura Cross-Couplings of Alkylboranes: Transmetalation Studies and Synthesis of Model Organopalladium Pretransmetalation Species. ACS Catalysis 2024, 14
(21)
, 15946-15955. https://doi.org/10.1021/acscatal.4c05053
- Mirja Md Mahamudul Hassan, Biplab Mondal, Sukriti Singh, Chabush Haldar, Jagriti Chaturvedi, Ranjana Bisht, Raghavan B. Sunoj, Buddhadeb Chattopadhyay. Ir-Catalyzed Ligand-Free Directed C–H Borylation of Arenes and Pharmaceuticals: Detailed Mechanistic Understanding. The Journal of Organic Chemistry 2022, 87
(6)
, 4360-4375. https://doi.org/10.1021/acs.joc.2c00046
- Tao Wang, Imran Shah, Zengming Yang, Weidong Yin, Shaoxiong Zhang, Yuan Yang, Pei Yin, Hengchang Ma. Incorporating Thiourea into Fluorescent Probes: A Reliable Strategy for Mitochondrion-Targeted Imaging and Superoxide Anion Tracking in Living Cells. Analytical Chemistry 2020, 92
(3)
, 2824-2829. https://doi.org/10.1021/acs.analchem.9b05320
- José
J. Fuentes-Rivera, Mary E. Zick, M. Alexander Düfert, Phillip J. Milner. Overcoming Halide Inhibition of Suzuki–Miyaura Couplings with Biaryl Monophosphine-Based Catalysts. Organic Process Research & Development 2019, 23
(8)
, 1631-1637. https://doi.org/10.1021/acs.oprd.9b00255
- Curtis
C. Ho, Alireza Ariafard, Christopher J. T. Hyland, Alex C. Bissember. Phosphine-Scavenging Cationic Gold(I) Complexes: Alternative Applications of Gold Cocatalysis in Fundamental Palladium-Catalyzed Cross-Couplings. Organometallics 2019, 38
(13)
, 2683-2688. https://doi.org/10.1021/acs.organomet.9b00294
- Reyne Pullen, Angus Olding, Jason A. Smith, Alex C. Bissember. Capstone Laboratory Experiment Investigating Key Features of Palladium-Catalyzed Suzuki–Miyaura Cross-Coupling Reactions. Journal of Chemical Education 2018, 95
(11)
, 2081-2085. https://doi.org/10.1021/acs.jchemed.7b00964
- Jonas W. Meringdal, Dirk Menche. Suzuki–Miyaura (hetero-)aryl cross-coupling: recent findings and recommendations. Chemical Society Reviews 2025, 54
(12)
, 5746-5765. https://doi.org/10.1039/D4CS01108B
- Allan J. Canty, Alex C. Bissember. Suzuki‐Miyaura Cross‐Couplings: Juxtaposing the Transmetalation of Arylboronic Acids and Alkylboranes. Chemistry – A European Journal 2025, 31
(33)
https://doi.org/10.1002/chem.202501157
- Mario Pajić, Dajana Barišić, Darko Babić, Manda Ćurić, Marina Juribašić Kulcsár. From
In Situ
Monitoring to Better Understanding of the Suzuki‐Miyaura Cross Coupling in the Solid State. Chemistry–Methods 2025, 5
(6)
https://doi.org/10.1002/cmtd.202400025
- Neeraj Bhatt, Anupma Tyagi, Suraj Purohit, Arun Kumar. Click Chemistry for the Generation of Combination of Triazole Core and Thioether Donor Site in Organosulfur Ligands: Applications of Metal Complexes in Catalysis. Chemistry – An Asian Journal 2024, 13 https://doi.org/10.1002/asia.202400379
- Xu Jiang, Mehdi Kiasadegh. Preparation and characterization of Pd (II) based on 2,5-pyridinedicarboxylic acid as novel recyclable metal-organic frameworks for Suzuki–Miyaura reaction. Journal of Organometallic Chemistry 2023, 1001 , 122889. https://doi.org/10.1016/j.jorganchem.2023.122889
- Chandra Shekhar Tiwari, Praseetha Mathoor Illam, Arnab Rit. ‘Click’‐Derived 1,2,3‐Triazolium Salts Featuring a 4‐Halo‐phenyl Group: Precursors for Heterobimetallic Complexes. Chemistry – An Asian Journal 2023, 18
(15)
https://doi.org/10.1002/asia.202300484
- Carlos F. R. A. C. Lima, Marco A. L. Lima, J. Ricardo M. Pinto, M. Gabriela T. C. Ribeiro, Artur M. S. Silva, Luís M. N. B. F. Santos. Highlights on the General Preference for Multi-Over Mono-Coupling in the Suzuki–Miyaura Reaction. Catalysts 2023, 13
(6)
, 928. https://doi.org/10.3390/catal13060928
- Angus Olding, Curtis C. Ho, Debabrata Maiti, Alex C. Bissember. Structural authentication of intermediates of mechanistic significance in palladium- and nickel-catalysed cross-couplings: case studies. Chemical Communications 2023, 59
(35)
, 5144-5155. https://doi.org/10.1039/D3CC00882G
- Laila M. Nhari, Elham N. Bifari, Aisha R. Al-Marhabi, Fatimah A.M. Al-Zahrani, Huda A. Al-Ghamdi, Sameera N. Al-Ghamdi, Abdullah M. Asiri, Reda M. El-Shishtawy. Synthesis of novel phenothiazine, phenoxazine and carbazole derivatives via Suzuki-Miyaura reaction. Journal of Organometallic Chemistry 2023, 989 , 122648. https://doi.org/10.1016/j.jorganchem.2023.122648
- Gyandshwar Kumar Rao, Anirban Dutta, Nisha Yadav, Preeti Oswal, Arun Kumar, Ajai K. Singh. Palladium(
ii
)-complexes of bi- and tri-dentate phosphine ligands: precursor for palladium–phosphorous nanoparticles and activity towards Suzuki–Miyaura coupling. New Journal of Chemistry 2023, 47
(9)
, 4218-4227. https://doi.org/10.1039/D2NJ05357H
- Sheng-Bo Hong, Lan-Chang Liang. Aqueous Suzuki couplings mediated by a hydrophobic catalyst. RSC Advances 2022, 12
(44)
, 28862-28866. https://doi.org/10.1039/D2RA05230J
- Sepideh Samiee, Ahmadreza Shiralinia, Elham Hoveizi, Robert W. Gable. A new Pd(II) complex containing acetophenone oxime and 1,3-Bis(diphenylphosphino)propane ligands; crystal structure, catalytic activity, molecular docking and in vitro cytotoxic evaluation. Inorganica Chimica Acta 2022, 538 , 120964. https://doi.org/10.1016/j.ica.2022.120964
- Suvam Bhattacharjee, Sudip Laru, Alakananda Hajra. Remote difunctionalization of 2
H
-indazoles using Koser's reagents. Chemical Communications 2022, 58
(7)
, 981-984. https://doi.org/10.1039/D1CC06129A
- Kohtaro Osakada, Yasushi Nishihara. Transmetalation of boronic acids and their derivatives: mechanistic elucidation and relevance to catalysis. Dalton Transactions 2022, 51
(3)
, 777-796. https://doi.org/10.1039/D1DT02986J
- M. R. Crampton. Nucleophilic Aromatic Substitution. 2021, 167-189. https://doi.org/10.1002/9781119531975.ch5a
- Angus Olding, Curtis C. Ho, Allan J. Canty, Nigel T. Lucas, James Horne, Alex C. Bissember. Synthesis of Arylpalladium(II) Boronates: Confirming the Structure and Chemical Competence of Pre‐transmetalation Intermediates in the Suzuki–Miyaura Reaction. Angewandte Chemie 2021, 133
(27)
, 15023-15027. https://doi.org/10.1002/ange.202104802
- Angus Olding, Curtis C. Ho, Allan J. Canty, Nigel T. Lucas, James Horne, Alex C. Bissember. Synthesis of Arylpalladium(II) Boronates: Confirming the Structure and Chemical Competence of Pre‐transmetalation Intermediates in the Suzuki–Miyaura Reaction. Angewandte Chemie International Edition 2021, 60
(27)
, 14897-14901. https://doi.org/10.1002/anie.202104802
- Amine Bourouina, Valérie Meille, Claude de Bellefon. About Solid Phase vs. Liquid Phase in Suzuki-Miyaura Reaction. Catalysts 2019, 9
(1)
, 60. https://doi.org/10.3390/catal9010060
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
Abstract
Scheme 1
Scheme 1. Results of Pd-Catalyzed Suzuki–Miyaura Couplings Reported by Novak and Wallow in 1994 (8)Figure 1
Figure 2
Figure 2. Two fundamental pathways leading to pretransmetalation intermediate C in Suzuki–Miyaura couplings.
Scheme 2
Scheme 2. Investigating the Chemical Competence of trans-[Pd(PPh3)2(p-tol)(Br)] in Suzuki–Miyaura CouplingsaDetermined via GC with the aid of a calibrated internal standard (average of two experiments). Yields after 0.25 h are given in parentheses.
Figure 3
Figure 3. Formation of biaryl 3 in reactions employing trans-[Pd(PPh3)2(p-tol)(I)] (4a) (solid red line), [Pd–I] 4a + 6 equiv of PPh3 (dashed red line), trans-[Pd(PPh3)2(p-tol)(Br)] (4b) (solid blue line), [Pd–Br] 4b + 6 equiv PPh3 (dashed blue line), [Pd–OH] 5 (solid black line), and [Pd–OH] 5 + 6 equiv PPh3 (dashed black line). A 0.5 equiv amount of [Pd(PPh3)(p-tol)(μ-OH)]2 was used in these experiments (and no Na2CO3 was used in these reactions). Yields were determined via GC with the aid of a calibrated internal standard (average of two experiments).
Figure 4
Figure 4. Formation of biaryl 3 in reactions employing: 6% PPh3 (red), 6% P(2-furyl)3 (blue), and 3% PPh3 and 3% P(2-furyl)3 (black). Yields were determined via GC with the aid of a calibrated internal standard (average of two experiments).
Scheme 3
Scheme 3. Investigating the Chemical Competence of trans-[Pd{P(2-furyl)3}2(p-tol)(I)] in Suzuki–Miyaura CouplingsaDetermined via GC with the aid of a calibrated internal standard (average of two experiments). Yields after 0.25 h are given in parentheses.
References
This article references 34 other publications.
- 1
See, for example:
(a) Choi, J.; Fu, G. C. Transition metal-catalyzed alkyl-alkyl bond formation: Another dimension in cross-coupling chemistry. Science 2017, 356, eaaf7230, DOI: 10.1126/science.aaf7230There is no corresponding record for this reference.(b) Metal-Catalyzed Cross-Coupling Reactions and More, 3rd ed.; de Meijere, A., Bras̈e, S., Oestreich, M., Eds.; Wiley-VCH: Weinheim, Germany, 2014.There is no corresponding record for this reference.(c) Magano, J.; Dunetz, J. R. Large-Scale Applications of Transition Metal-Catalyzed Couplings for the Synthesis of Pharmaceuticals. Chem. Rev. 2011, 111, 2177– 2250, DOI: 10.1021/cr100346g1cLarge-Scale Applications of Transition Metal-Catalyzed Couplings for the Synthesis of PharmaceuticalsMagano, Javier; Dunetz, Joshua R.Chemical Reviews (Washington, DC, United States) (2011), 111 (3), 2177-2250CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review was given summarizing large-scale applications of transition metal-catalyzed coupling reactions for the manuf. of drug components in pharmaceutical industry. In addn. to showcasing bond-forming chemistries, focus was given on reaction work-ups and purifications that purge metal catalysts to provide material of sufficient purity. Most of the examples presented originated on process chem. groups in pharmaceutical companies and were selected based on 2 criteria: (a) the transformation was realized on a large scale (at least 100 mmol); (b) the article contains a detailed exptl. procedure. A list of publications reporting coupling reactions for the large-scale synthesis of pharmaceuticals, but do not provide exptl. procedures was also included.(d) Torborg, C.; Beller, M. Recent Applications of Palladium-Catalyzed Coupling Reactions in the Pharmaceutical, Agrochemical, and Fine Chemical Industries. Adv. Synth. Catal. 2009, 351, 3027– 3043, DOI: 10.1002/adsc.2009005871dRecent Applications of Palladium-Catalyzed Coupling Reactions in the Pharmaceutical, Agrochemical, and Fine Chemical IndustriesTorborg, Christian; Beller, MatthiasAdvanced Synthesis & Catalysis (2009), 351 (18), 3027-3043CODEN: ASCAF7; ISSN:1615-4150. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Palladium-catalyzed coupling reactions have become a central tool for the synthesis of biol. active compds. both in academia and industry. Most of these transformations make use of easily available substrates and allow for a shorter and more selective prepn. of substituted arenes and heteroarenes compared to non-catalytic pathways. Notably, mol.-defined palladium catalysts offer high chemoselectivity and broad functional group tolerance. Considering these advantages, it is not surprising that several palladium-catalyzed coupling reactions have been implemented in the last decade into the industrial manuf. of pharmaceuticals and fine chems. In this review different examples from 2001-2008 are highlighted, which have been performed at least on a kilogram scale in the chem. and pharmaceutical industries.(e) Corbet, J.-P.; Mignani, G. Selected Patented Cross-Coupling Reaction Technologies. Chem. Rev. 2006, 106, 2651– 2710, DOI: 10.1021/cr05052681eSelected patented cross-coupling reaction technologiesCorbet, Jean-Pierre; Mignani, GerardChemical Reviews (Washington, DC, United States) (2006), 106 (7), 2651-2710CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review dealing with cross-coupling reactions : carbon-carbon coupling reactions (Miyaura-Suzuki reaction, Corriu-Kumada-Tamao reaction, Negishi reaction, etc.) and carbon-nitrogen coupling reactions. - 2
For selected reviews on the Suzuki–Miyaura cross-coupling reactions, see:
(a) Hussain, I.; Capricho, J.; Yawer, M. A. Synthesis of Biaryls via Ligand-Free Suzuki–Miyaura Cross-Coupling Reactions: A Review of Homogeneous and Heterogeneous Catalytic Developments. Adv. Synth. Catal. 2016, 358, 3320– 3349, DOI: 10.1002/adsc.2016003542aSynthesis of Biaryls via Ligand-Free Suzuki-Miyaura Cross-Coupling Reactions: A Review of Homogeneous and Heterogeneous Catalytic DevelopmentsHussain, Ibrar; Capricho, Jaworski; Yawer, Mirza A.Advanced Synthesis & Catalysis (2016), 358 (21), 3320-3349CODEN: ASCAF7; ISSN:1615-4150. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The recent developments in the field of Suzuki-Miyaura cross-coupling reaction, in particular with regard to ligand-free catalysis were described. The catalysts outlined here allowed convenient and green synthetic pathways specifically for the construction of C-C bonds. They enabled the synthesis of biaryls by the coupling arylboronic acids with aryl halides. The literature reporting ligand-free synthesis of biaryls from 2010 to May 2015 was reviewed.(b) Lennox, A. J. J.; Lloyd-Jones, G. C. Selection of boron reagents for Suzuki–Miyaura coupling. Chem. Soc. Rev. 2014, 43, 412– 443, DOI: 10.1039/C3CS60197H2bSelection of boron reagents for Suzuki-Miyaura couplingLennox, Alastair J. J.; Lloyd-Jones, Guy C.Chemical Society Reviews (2014), 43 (1), 412-443CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Suzuki-Miyaura (SM) cross-coupling is arguably the most widely-applied transition metal catalyzed carbon-carbon bond forming reaction to date. Its success originates from a combination of exceptionally mild and functional group tolerant reaction conditions, with a relatively stable, readily prepd. and generally environmentally benign organoboron reagent. A variety of such reagents were developed for the process, with properties that were tailored for application under specific SM coupling conditions. This review analyses the seven main classes of boron reagent that were developed. The general phys. and chem. properties of each class of reagent are evaluated with special emphasis on the currently understood mechanisms of transmetalation. The methods to prep. each reagent are outlined, followed by example applications in SM coupling.(c) Suzuki, A. Carbon–carbon bonding made easy. Chem. Commun. 2005, 4759– 4763, DOI: 10.1039/b507375h2cCarbon-carbon bonding made easySuzuki, AkiraChemical Communications (Cambridge, United Kingdom) (2005), (38), 4759-4763CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)A review. The palladium-catalyzed cross-coupling reaction between organoboranes and org. electrophiles in the presence of base was first developed 30 years ago. It offers a powerful and general methodol. for forming carbon-carbon bonds. The scope of the reaction has continued to evolve and broaden to meet modern synthetic requirements. The role of base additives and the reaction mechanism were discussed. Coupling reactions of arom. halides with boranes were mentioned. Problems involving sterically hindered substrates were addressed. Coupling reactions of alkyl compds. and alkynyl borane derivs. were also discussed. The applicability of Suzuki coupling reactions toward environmentally friendly processes was also pointed out. - 3Johansson Seechurn, C. C. C.; Kitching, M. O.; Colacot, T. J.; Snieckus, V. Palladium-catalyzed cross-coupling: a historical contextual perspective to the 2010 Nobel Prize. Angew. Chem., Int. Ed. 2012, 51, 5062– 5085, DOI: 10.1002/anie.2011070173Palladium-Catalyzed Cross-Coupling: A Historical Contextual Perspective to the 2010 Nobel PrizeJohansson Seechurn, Carin C. C.; Kitching, Matthew O.; Colacot, Thomas J.; Snieckus, VictorAngewandte Chemie, International Edition (2012), 51 (21), 5062-5085CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. In 2010, Richard Heck, Ei-ichi Negishi, and Akira Suzuki joined the prestigious circle of Nobel Laureate chemists for their roles in discovering and developing highly practical methodologies for C-C bond construction. From their original contributions in the early 1970s the landscape of the strategies and methods of org. synthesis irreversibly changed for the modern chemist, both in academia and in industry. In this review, we attempt to trace the historical origin of these powerful reactions, and outline the developments from the seminal discoveries leading to their eminent position as appreciated and applied today.
- 4
For selected reviews on the applications of alkylphosphane ligands in Pd-catalyzed C–C cross-coupling reactions, see:
(a) Rossi, R.; Bellina, F.; Lessi, M.; Manzini, C.; Marianetti, G.; Perego, L. A. Recent Applications of Phosphane-based Palladium Catalysts in Suzuki-Miyaura Reactions Involved in Total Syntheses of Natural Products. Curr. Org. Chem. 2015, 19, 1302– 1409, DOI: 10.2174/13852728196661505062300504aRecent Applications of Phosphane-based Palladium Catalysts in Suzuki-Miyaura Reactions Involved in Total Syntheses of Natural ProductsRossi, Renzo; Bellina, Fabio; Lessi, Marco; Manzini, Chiara; Marianetti, Giulia; Perego, Luca A.Current Organic Chemistry (2015), 19 (14), 1302-1409CODEN: CORCFE; ISSN:1385-2728. (Bentham Science Publishers Ltd.)A review with refs. on This review with 800 refs. illustrates applications of Suzuki-Miyaura (S.-M.) reactions in the total syntheses of 147 natural products that were made in the period 2010-2013. The review has been organized on the basis of the seven classes of phosphane-based Pd catalysts that have been used in the reported total syntheses. Emphasis has been given to describe and discuss the exptl. conditions of the Pd-catalyzed (S.-M.) cross-coupling reactions also outlining the methods used to prep. the reactants. A focus has also been set on the biol. and pharmacol. properties of the reported natural products as well as on the most significant steps of the reported total syntheses.(b) Li, H.; Johansson Seechurn, C. C. C.; Colacot, T. J. Development of Preformed Pd Catalysts for Cross-Coupling Reactions, Beyond the 2010 Nobel Prize. ACS Catal. 2012, 2, 1147– 1164, DOI: 10.1021/cs300082f4bDevelopment of Preformed Pd Catalysts for Cross-Coupling Reactions, Beyond the 2010 Nobel PrizeLi, Hongbo; Johansson Seechurn, Carin C. C.; Colacot, Thomas J.ACS Catalysis (2012), 2 (6), 1147-1164CODEN: ACCACS; ISSN:2155-5435. (American Chemical Society)A review. Strategies for the development of Pd catalysts based on utilization of L2Pd and LPd (L = ligand) species, beyond the contributions of the 2010 Nobel Laureates Richard Heck, Ei-ichi Negishi, and Akira Suzuki, along with their contemporaries, are reviewed. These well-defined, preformed Pd catalysts improve the selectivity and activity of selected cross-coupling reactions by reducing the metal loading and the ligand-to-metal ratios. This review describes predominantly the development of Pd precatalysts over the last 10 years and highlights the benefits often obsd. when using well-defined preformed catalysts relative to those generated in situ.(c) Fu, G. C. The Development of Versatile Methods for Palladium-Catalyzed Coupling Reactions of Aryl Electrophiles through the Use of P(t-Bu)3 and PCy3 as Ligands. Acc. Chem. Res. 2008, 41, 1555– 1564, DOI: 10.1021/ar800148f4cThe Development of Versatile Methods for Palladium-Catalyzed Coupling Reactions of Aryl Electrophiles through the Use of P(t-Bu)3 and PCy3 as LigandsFu, Gregory C.Accounts of Chemical Research (2008), 41 (11), 1555-1564CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review. Metal-catalyzed coupling reactions of aryl electrophiles with organometallics and with olefins serve as unusually effective tools for forming new carbon-carbon bonds. By 1998, researchers had developed catalysts that achieved reactions of aryl iodides, bromides, and triflates. Nevertheless, many noteworthy challenges remained; among them were couplings of aryl iodides, bromides, and triflates under mild conditions (at room temp., for example), couplings of hindered reaction partners, and couplings of inexpensive aryl chlorides. This Account highlights some of the progress that has been made in our lab. over the past decade, largely through the appropriate choice of ligand, in achieving these synthetic objectives. In particular, we have established that palladium in combination with a bulky trialkylphosphine accomplishes a broad spectrum of coupling processes, including Suzuki, Stille, Negishi, and Heck reactions. These methods have been applied in a wide array of settings, such as natural-product synthesis, materials science, and bioorg. chem. - 5
For selected studies on oxidative addition to Pd(0) complexes featuring phosphine ligands, see:
(a) Goossen, L. J.; Koley, D.; Hermann, H. L.; Thiel, W. Mechanistic Pathways for Oxidative Addition of Aryl Halides to Palladium(0) Complexes: A DFT Study. Organometallics 2005, 24, 2398– 2410, DOI: 10.1021/om0500220There is no corresponding record for this reference.(b) Senn, H. M.; Ziegler, T. Oxidative Addition of Aryl Halides to Palladium(0) Complexes: A Density-Functional Study Including Solvation. Organometallics 2004, 23, 2980– 2988, DOI: 10.1021/om049963nThere is no corresponding record for this reference.(c) 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, DOI: 10.1021/ar980063a5cAnionic Pd(0) and Pd(II) Intermediates in Palladium-Catalyzed Heck and Cross-Coupling ReactionsAmatore, Christian; Jutand, AnnyAccounts of Chemical Research (2000), 33 (5), 314-321CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review with 17 refs. The anions of PdCl2L2 and Pd(OAc)2, precursors of palladium(0) used in cross-coupling and Heck reactions, play a crucial role in these reactions. Tricoordinated anionic complexes Pd0L2Cl- and Pd0L2(OAc)- are the effective catalysts instead of the usually postulated Pd0L2 complex. The anion ligated to the palladium(0) affects the kinetics of the oxidative addn. to ArI as well as the structure and reactivity of the arylpalladium(II) complexes produced in this reaction. Thus, pentacoordinated anionic complexes are formed, ArPdI(Cl)L2- or ArPdI(OAc)L2-, the precursor of neutral trans-ArPd(OAc)L2, instead of the usually postulated trans-ArPdIL2 complex (L = PPh3).(d) Casado, A. L.; Espinet, P. On the Configuration Resulting from Oxidative Addition of RX to Pd(PPh3)4 and the Mechanism of the cis-to-trans Isomerization of [PdRX(PPh3)2] Complexes (R = Aryl, X = Halide). Organometallics 1998, 17, 954– 959, DOI: 10.1021/om97095025dOn the Configuration Resulting from Oxidative Addition of RX to Pd(PPh3)4 and the Mechanism of the cis-to-trans Isomerization of [PdRX(PPh3)2] Complexes (R = Aryl, X = Halide)Casado, Arturo L.; Espinet, PabloOrganometallics (1998), 17 (5), 954-959CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)The oxidative addn. of RI to Pd(0) and further cis-to-trans isomerization, which are involved in the Stille reaction and other Pd-catalyzed syntheses, have been studied. C6Cl2F3I (C6Cl2F3 = 3,5-dichlorotrifluorophenyl) adds to Pd(PPh3)4 in THF at room temp. giving cis-[Pd(C6Cl2F3)I(PPh3)2] (2), which could be isolated before isomerization to the more stable trans-[Pd(C6Cl2F3)I(PPh3)2] (3). A 19F NMR kinetic study of the isomerization of 2 in THF at 322.6 K reveals a first-order law riso = kiso[2], with kiso = f + g[2]0 + (h + i[2]0)/([PPh3] + j) (f = (1.66 ± 0.03) × 10-4 s-1, g = (2.5 ± 0.2) × 10-3 mol-1 L s-1, h = (1.3 ± 0.7) × 10-8 mol L-1 s-1, i = (4 ± 2) × 10-4 s-1, and j = (1.4 ± 0.7) × 10-5 mol L-1). A four-pathway mechanism accounts for these results: two are assigned to the associative replacements of PPh3 coordinated to 2 by an iodide ligand of I-[Pd] (I-[Pd] = 2 or 3), both THF-assisted (coeff. h) or direct (coeff. i), leading to a monoiodide-bridged intermediate cis-{Pd(C6Cl2F3)I(PPh3)(μ-I)-[Pd]}. The later rearranges via terminal-for-bridging iodide exchange to trans-{Pd(C6Cl2F3)I(PPh3)(μ-I)-[Pd]}, which is finally cleaved by PPh3 yielding complex 3. The other two concurrent pathways are assigned to the isomerization via two consecutive Berry pseudorotations in the pentacoordinated species derived from 2 by coordination of THF (coeff. f) or I-[Pd] (coeff. g). The apparent activation entropy assocd. with kiso is neg. (ΔS⧧ = -21 ± 3 J K-1 mol-1), in agreement with the proposed bimol. mechanisms.(e) Hartwig, J. F.; Paul, F. Oxidative Addition of Aryl Bromide after Dissociation of Phosphine from a Two-Coordinate Palladium(0) Complex, Bis(tri-o-tolylphosphine)Palladium(0). J. Am. Chem. Soc. 1995, 117, 5373– 5374, DOI: 10.1021/ja00124a0265eOxidative Addition of Aryl Bromide after Dissociation of Phosphine from a Two-Coordinate Palladium(0) Complex, Bis(tri-o-tolylphosphine)Palladium(0)Hartwig, John F.; Paul, FredericJournal of the American Chemical Society (1995), 117 (19), 5373-4CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The 14-electron Pd(0) complex [(o-tol)3P]2Pd (1) undergoes oxidative addn. reactions with aryl bromides that occur by a dissociative mechanism. The products of aryl bromide addn. are dimeric with only one phosphine bound to each palladium. In contrast to typical oxidative addns. to Pt(0), Pd(0), and Ni(0) phosphine compds. that occur to intermediates with two or three phosphine ligands, the addn. to 1 occurs after phosphine dissocn. and generation of a monophosphine intermediate. Clean first order kinetics were obsd. when monitoring the disappearance of 1. Inverse first order rate behavior in phosphine demonstrated that the reaction is dissociative. Reaction rates in benzene, toluene, and xylene were essentially identical, suggesting that solvent coordination does not stabilize the unsatd. intermediate. Rather, displacement of a phosphine by coordinated aryl bromide before irreversible oxidative addn., or coordination of a ligand C-H bond appears to stabilize the monophosphine complex. This mechanism is unusual, but falls in accord with those for the opposite reaction, reductive elimination from Pd(II). Many of these elimination reactions occur after phosphine dissocn. and, therefore, lead to initial products that contain only one phosphine ligand. This initial product from reductive elimination is analogous to the intermediate generated from 1 that undergoes oxidative addn.(f) Paul, F.; Patt, J.; Hartwig, J. F. Structural Characterization and Simple Synthesis of {Pd[P(o-Tol)3]2}. Spectroscopic Study and Structural Characterization of the Dimeric Palladium(II) Complexes Obtained by Oxidative Addition of Aryl Bromides and Their Reactivity with Amines. Organometallics 1995, 14, 3030– 3039, DOI: 10.1021/om00006a0535fStructural Characterization and Simple Synthesis of {Pd[P(o-Tol)3]2}. Spectroscopic Study and Structural Characterization of the Dimeric Palladium(II) Complexes Obtained by Oxidative Addition of Aryl Bromides and Their Reactivity with AminesPaul, Frederic; Patt, Joe; Hartwig, John F.Organometallics (1995), 14 (6), 3030-9CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)The Pd(0) complex {Pd[P(o-Tol)3]2} was prepd. by addn. of P(o-Tol)3 to crude [Pd(DBA)2], which is an approx. equimolar mixt. of Pd2(DBA)3 and Pd(DBA)3, followed by crystn. from the reaction medium by addn. of ether. The formation of {Pd[P(o-Tol)3]2} appeared to be driven by its insoly. in the benzene/ether solvent mixt. Benzene solns. of [Pd(DBA)2] and P(o-Tol)3 did not contain amts. of the L2Pd compd. that could be detected by 31P NMR spectroscopy. {Pd[P(o-Tol)3]2} was characterized crystallog. and showed an exactly linear geometry. Similar Pd(0) compds. {Pd[P(2,4-dimethylphenyl)3]2}, {Pd[P(2-methyl-4-fluorophenyl)3]2}, and the low-coordinate trialkylphosphine complex {Pd[P(t-Bu)3]2} were also prepd. by this method, but [Pd(PCy3)2(DBA)] was produced from reactions involving PCy3 and [Pd(DBA)2], and [Pd(TMPP)(DBA)2] was isolated after addn. of tris(1,3,5-trimethoxyphenyl)phosphine (TMPP) to [Pd(DBA)2]. The oxidative addn. of aryl halides to {Pd[P(o-Tol)3]2} at room temp. led to dimeric {Pd[P(o-Tol)3](Ar)(Br)}2. An example of these compds. was characterized crystallog. as well as by soln. mol. wt. anal. This aryl halide complex is dimeric in the solid state as well as in soln. The NMR spectra of the large triarylphosphine complexes showed temp. dependent behavior, presumably due to isomerizations and ligand rotations that occurred on the NMR time scale. The aryl halide compds. did not form four-coordinate monometallic species in the presence of excess P(o-Tol)3, but they did undergo cleavage to four-coordinate monometallic complexes upon addn. of primary and secondary amines.(g) Amatore, C.; Jutand, A.; Suarez, A. Intimate mechanism of oxidative addition to zerovalent palladium complexes in the presence of halide ions and its relevance to the mechanism of palladium-catalyzed nucleophilic substitutions. J. Am. Chem. Soc. 1993, 115, 9531– 9541, DOI: 10.1021/ja00074a0185gIntimate mechanism of oxidative addition to zerovalent palladium complexes in the presence of halide ions and its relevance to the mechanism of palladium-catalyzed nucleophilic substitutionsAmatore, Christian; Jutand, Anny; Suarez, AlejandraJournal of the American Chemical Society (1993), 115 (21), 9531-41CODEN: JACSAT; ISSN:0002-7863.The mechanism of oxidative addn. of aryl halides to low-ligated zerovalent palladium species obtained by redn. of their divalent precursor complexes, PdIICl2(PR3)2, was investigated on the basis of 31P NMR and electrochem. In strong contrast to usual expectations, the reaction proceeds via a complex sequence of steps: (i) fast addn. of the aryl halide to a halide ligated zerovalent palladium center, so as to afford a pentacoordinated anionic arylpalladium(II) center; (ii) a fast but reversible uphill elimination from the pentacoordinated anionic arylpalladium(II) center of a halide ion ligand, possibly through its substitution by a solvent ligand; (iii) rearrangement of this second short-lived intermediate into the thermodynamically stable trans-arylpalladium(II) product of the reaction. On the basis of this detailed mechanism, a new mechanism is proposed for the catalysis of nucleophilic substitutions by palladium complexes. In contrast to the catalytic cycles that are usually considered, this new catalytic cycle accounts for the well-used effects of halide ions as well as that of small metal cations and rationalizes their role in the overall efficiency of palladium-catalyzed nucleophilic substitutions.(h) Amatore, C.; Pflüger, F. Mechanism of oxidative addition of palladium(0) with aromatic iodides in toluene, monitored at ultramicroelectrodes. Organometallics 1990, 9, 2276– 2282, DOI: 10.1021/om00158a0265hMechanism of oxidative addition of palladium(0) with aromatic iodides in toluene, monitored at ultramicroelectrodesAmatore, Christian; Pfluger, FernandoOrganometallics (1990), 9 (8), 2276-82CODEN: ORGND7; ISSN:0276-7333.The mechanism and rates of oxidative addn. of substituted iodobenzenes to tetrakis(triphenylphosphine)palladium(0) is investigated in toluene by means of electrochem. at ultramicroelectrodes. The mechanism is identical with that previously reported in THF (F. F. Fauvarque et al., 1981) and to correspond to a similar slope for Hammett's correlation (ρ = +2 in THF vs +2.3 ± 0.2 in toluene). Moreover, the important change in polarity (εs ≈ 7.58 in THF and 2.38 in toluene) does not affect the enthalpy of activation and leads to similar activation entropies in both solvents. This suggests that the transition state of addn. of aryl halides to the coordinately unsatd. zerovalent palladium complex Pd0(PPh3)2 has no significant ionic character.(i) Fauvarque, J.-F.; Pflüger, F. Kinetics of oxidative addition of zerovalent palladium to aromatic iodides. J. Organomet. Chem. 1981, 208, 419– 427, DOI: 10.1016/S0022-328X(00)86726-15iKinetics of oxidative addition of zerovalent palladium to aromatic iodidesFauvarque, Jean Francois; Pfluger, Fernando; Troupel, MichelJournal of Organometallic Chemistry (1981), 208 (3), 419-27CODEN: JORCAI; ISSN:0022-328X.The kinetics of oxidative addn. of Pd(0) to arom. iodides in THF contg. the ligand Ph3P were studied by amperometry on the Pd oxidn. wave. The reaction is 1st order each in arom. iodide and Pd(0), and the rate const. is inversely proportional to the free-ligand concn. The reactive Pd(0) intermediate is Pd(PPh3)2. With substituted arom. iodides, the rate consts. give a linear Hammett relationship with ρ = +2. The mechanism is related to an arom. nucleophilic substitution with some assistance from halogen-Pd interaction.(j) Stille, J. K.; Lau, K. S. Y. Acc. Chem. Res. 1977, 10, 434, DOI: 10.1021/ar50120a002There is no corresponding record for this reference.(k) Fitton, P.; Rick, E. A. The addition of aryl halides to tetrakis(triphenylphosphine)palladium(0). J. Organomet. Chem. 1971, 28, 287– 291, DOI: 10.1016/S0022-328X(00)84578-75kAddition of aryl halides to tetrakis(triphenylphosphine)palladium(0)Fitton, P.; Rick, Edward A.Journal of Organometallic Chemistry (1971), 28 (2), 287-91CODEN: JORCAI; ISSN:0022-328X.The order of reactivity of halobenzenes towards oxidative addn. to Pd(PPh3)4 was PhI > PhBr > PhCl. Iodobenzene will react with Pd(PPh3)4 in benzene at room temp., and bromobenzene will react at 80°. Chlorobenzene, however, is unreactive even at 135°. Aryl chlorides substituted with electron-donating groups similarly are unreactive, but aryl chlorides substituted with electron-withdrawing groups are reactive. This increased reactivity, and order of reactivity 4-O2NC6H4Cl > 4-CNC6H4Cl > 4-PhCOC6H4Cl > PhCl suggests that the mechanism of the oxidative addn. is similar to that of a bimol. nucleophilic aromatic displacement reaction in which breaking of the bond to the leaving group is involved in the rate detg. step. - 6Organotransition Metal Chemistry From Bonding to Catalysis; Hartwig, J. F., Ed.; University Science Books: Sausalito, CA, 2010; Chapter 19.There is no corresponding record for this reference.
- 7Pullen, R.; Olding, A.; Smith, J. A.; Bissember, A. C. Manuscript in preparation.There is no corresponding record for this reference.
- 8Wallow, T. I.; Novak, B. M. Highly Efficient and Accelerated Suzuki Aryl Couplings Mediated by Phosphine-Free Palladium Sources. J. Org. Chem. 1994, 59, 5034– 5037, DOI: 10.1021/jo00096a0568Highly efficient and accelerated Suzuki aryl couplings mediated by phosphine-free palladium sourcesWallow, Thomas I.; Novak, Bruce M.Journal of Organic Chemistry (1994), 59 (17), 5034-7CODEN: JOCEAH; ISSN:0022-3263.Suzuki aryl cross-couplings employing aryl bromides and aryl iodides proceed under mild conditions (65°) with high efficiency (substrate-to-catalyst ratios >500) in the presence of phosphine-free Pd catalysts derived from Pd acetate, Pd2(dba)3·C6H6 (dba = dibenzylideneacetone), and [(η3-C3H5)PdCl]2. Phosphine inhibition is shown to play a key role in limiting catalytic efficiency; qual. comparison studies show that the phosphine-free systems are 1-2 orders of magnitude more active than phosphine-supported catalytic systems. Pd[P(Ph)3]4 proved to be the least active of the catalytic species screened. The phosphine-free methodol. appears to be generally applicable; cross-couplings of aryl iodides yielding biaryls p-MeOC6H4C6H4CF3-p 6 and p-MeOC6H4C6H4Me-o 7 proceed without noticeable steric or electronic effects. Cross-couplings employing aryl bromides are insensitive to electronic effects in the synthesis of 6 but are slowed by steric hindrance in the synthesis of 7. Acceleration of cross-coupling is obsd. in the presence of polar cosolvents and at high pH.
- 9
The poor reactivity of aryl iodides with vinyltributyltin in Stille couplings employing Pd/PPh3 at ∼50 °C has been reported; see:
Farina, V.; Krishnan, B. Large rate accelerations in the stille reaction with tri-2-furylphosphine and triphenylarsine as palladium ligands: mechanistic and synthetic implications. J. Am. Chem. Soc. 1991, 113, 9585– 9595, DOI: 10.1021/ja00025a0259Large rate accelerations in the stille reaction with tri-2-furylphosphine and triphenylarsine as palladium ligands: mechanistic and synthetic implicationsFarina, Vittorio; Krishnan, BalaJournal of the American Chemical Society (1991), 113 (25), 9585-95CODEN: JACSAT; ISSN:0002-7863.The effect of changing the palladium ligands on the rates of typical Stille cross-coupling reactions was studied. Large rate enhancements (typically 102-103 over triphenylphosphine-based catalyst) were obsd. with tri-2-furylphosphine and triphenylarsine, which are recommended as the new ligands of choice in the palladium-catalyzed coupling between olefinic stannanes and electrophiles. On the basis of the evidence presented, the transmetalation, which is the rate-detg. step in the catalytic cycle, is postulated to involve a π-complex between the metal and the stannane double bond. In general, ligands that readily dissoc. from Pd(II) and allow ready formation of this π-complex are the ones that produce the fastest coupling rates. The utility of the new ligands is demonstrated with several synthetic examples. - 10Huff, B. E.; Koenig, T. M.; Mitchell, D.; Staszak, M. A. Synthesis of unsymmetrical biaryls using a modified Suzuki cross-coupling: 4-biphenylcarboxaldehyde. Org. Synth. 1998, 75, 53, DOI: 10.1002/0471264180.os075.07There is no corresponding record for this reference.
- 11
In general, iodide consumption was comparable to product yield.
There is no corresponding record for this reference. - 12
Analogous results were obtained in experiments employing higher [Pd(PPh3)4] loadings (2.5% and 5%) (see the Supporting Information).
There is no corresponding record for this reference. - 13
We also investigated the effect of incrementally varying the Na2CO3 loading on the outcome of the reaction at 70 °C. Our results suggest that the Na2CO3 loading (1–5 equiv) has little effect on the efficiency of reactions employing p-iodotoluene (see the Supporting Information). Substituting either Cs2CO3 or NMe4OH for Na2CO3 also provided inefficient couplings of p-iodotoluene (see the Supporting Information).
There is no corresponding record for this reference. - 14
We also performed experiments employing higher PPh3 loadings (see the Supporting Information).
There is no corresponding record for this reference. - 15
It has been established that PPh3 can effect the reduction of Pd(OAc)2 to generate catalytically active Pd(0) species. See, for example: ref (5c).
(a) Carole, W. A.; Colacot, T. J. Understanding Palladium Acetate from a User Perspective. Chem. - Eur. J. 2016, 22, 7686– 7695, DOI: 10.1002/chem.20160145015aUnderstanding Palladium Acetate from a User PerspectiveCarole, William A.; Colacot, Thomas J.Chemistry - A European Journal (2016), 22 (23), 7686-7695CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The behavior of palladium acetate is reviewed with respect to its synthesis, characterization, structure (in both soln. and solid state), and activation pathways. In addn., comparisons of catalytic activities between pure palladium acetate and two common byproducts, Pd3(OAc)5(NO2) and polymeric [Pd(OAc)2]n, typically present in com. available material are reviewed. Hence, this minireview serves as a concise guide for the users of palladium acetate from both academia and industry.(b) Amatore, C.; Carré, E.; Jutand, A.; M’Barki, M. Rates and Mechanism of the Formation of Zerovalent Palladium Complexes from Mixtures of Pd(OAc)2 and Tertiary Phosphines and Their Reactivity in Oxidative Additions. Organometallics 1995, 14, 1818– 1826, DOI: 10.1021/om00004a03915bRates and Mechanism of the Formation of Zerovalent Palladium Complexes from Mixtures of Pd(OAc)2 and Tertiary Phosphines and Their Reactivity in Oxidative AdditionsAmatore, Christian; Carre, Emmanuelle; Jutand, Anny; M'Barki, Mohamed AmineOrganometallics (1995), 14 (4), 1818-26CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)Mixts. of Pd(OAc)2 and tertiary phosphines spontaneously afford Pd(0) complexes. Kinetic studies demonstrate that this reaction proceeds from the complex Pd(OAc)2(PR3)2 via an inner-sphere redn. which is the rate-detg. step of the overall reaction. The phosphine is thus oxidized to the corresponding phosphine oxide. The formation of the Pd(0) complex is sensitive to electronic and steric factors. The more the triarylphosphine is substituted by electron-withdrawing groups, the faster the reaction. The Pd(0) complex thus formed reacts with Ph iodide via an oxidn. addn., and this reaction is faster when the phosphine is more electron-rich.(c) Amatore, C.; Jutand, A.; M’Barki, M. Evidence of the formation of zerovalent palladium from Pd(OAc)2 and triphenylphosphine. Organometallics 1992, 11, 3009– 3013, DOI: 10.1021/om00045a01215cEvidence of the formation of zerovalent palladium from Pd(OAc)2 and triphenylphosphineAmatore, Christian; Jutand, Anny; M'Barki, Mohamed AmineOrganometallics (1992), 11 (9), 3009-13CODEN: ORGND7; ISSN:0276-7333.The mixt. Pd(OAc)2 + nPPh3 (n ≥2) and Pd(OAc)2(PPh3)2, commonly used as catalysts in reactions involving aryl and vinyl halides, aryl triflates or allylic acetates, spontaneously generate a zerovalent Pd complex that reacts with PhI. PPh3 reduces the PdII from Pd(OAc)2(PPh3)2 by an intramol. reaction and is oxidized to OPPh3. In the presence of an excess of PPh3, the zerovalent Pd complex generated in situ has the same 31P NMR and cyclic voltammetry properties as those of Pd0(PPh3)4. - 16
Pd(OAc)2 was used in preference to Pdx(dba)y, as challenges in establishing the exact speciation and purity of the latter have been identified. See, for example:
(a) Amatore, C.; Jutand, A. Role of dba in the reactivity of palladium(0) complexes generated in situ from mixtures of Pd(dba)2 and phosphines. Coord. Chem. Rev. 1998, 178–180, 511– 528, DOI: 10.1016/S0010-8545(98)00073-316aRole of dba in the reactivity of palladium(0) complexes generated in situ from mixtures of Pd(dba)2 and phosphinesAmatore, Christian; Jutand, AnnyCoordination Chemistry Reviews (1998), 178-180 (Pt. 1), 511-528CODEN: CCHRAM; ISSN:0010-8545. (Elsevier Science S.A.)A review with 98 refs. The ligand dba plays a crucial role both on the structure and on the reactivity of Pd(0) complexes generated in situ in mixts. of Pd(dba)2 and phosphine ligands. Whatever the ligand, the major complex is always Pd(dba)L2 where L is a monodentate phosphine ligand or Pd(dba)(L-L) where L-L is a bidentate phosphine ligand. In all cases, the most reactive species in the oxidative addn. with Ph iodide is the lowest ligated complex PdL2 or Pd(L-L) in equil. with the major complex and dba. However, Pd(dba)(L-L) also reacts with Ph iodide. The presence of the major complexes ligated by dba diminishes the concn. of the more reactive species PdL2 or Pd(L-L) and consequently controls the rate of the overall reaction. The overall reactivity is governed both by the intrinsic reactivity of the reactive species and its concn., two factors which can be antagonistic and a nonlinear Hammett correlation of the oxidative addn. with the basicity of the phosphine is obsd. for monodentate ligands.(b) Zalesskiy, S. S.; Ananikov, V. P. Pd2(dba)3 as a Precursor of Soluble Metal Complexes and Nanoparticles: Determination of Palladium Active Species for Catalysis and Synthesis. Organometallics 2012, 31, 2302– 2309, DOI: 10.1021/om201217r16bPd2(dba)3 as a Precursor of Soluble Metal Complexes and Nanoparticles: Determination of Palladium Active Species for Catalysis and SynthesisZalesskiy, Sergey S.; Ananikov, Valentine P.Organometallics (2012), 31 (6), 2302-2309CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)Tris(dibenzylideneacetone)dipalladium (Pd2(dba)3) is ubiquitously used as a source of sol. Pd species for catalysis and as a precursor in the synthesis of more complex Pd structures. In spite of the massive usage of this convenient Pd complex, its nature in soln. was not revealed in detail and the applications rely on the assumed state and purity of the compd. A convenient NMR procedure is developed to reveal the nature of Pd2(dba)3 and to det. the purity of the complex. Surprisingly, com. available samples of Pd2(dba)3 may readily contain up to 40% of Pd nanoparticles in a wide range of sizes (10-200 nm). The routinely accepted practice of use of Pd2(dba)3 without anal. of the purity (both com. available and prepd. by common procedures) can introduce significant errors in the estn. of catalyst efficiency and lead to incorrect values of TON, TOF, and reported mol % values in the catalytic procedures. The presence of Pd nanoparticles in the catalyst precursor provides an opportunity for heterogeneous catalytic systems of different nature to be directly accessible from Pd2(dba)3. A modified procedure for the synthesis of Pd2(dba)3·CHCl3 with 99% purity is reported. - 17
See, for example:
Bumagin, N. A.; Bykov, V. V.; Beletskaya, I. P. Synthesis of diaryls from phenylboric acid and aryl iodides in an aqueous medium. Bull. Acad. Sci. USSR, Div. Chem. Sci. 1989, 38, 2206, DOI: 10.1007/BF00962156There is no corresponding record for this reference.Bumagin, N. A.; Bykov, V. V.; Beletskaya, I. P. Bull. Acad. Sci. USSR, Div. Chem. Sci. (Engl. Transl.) 1989, 38, 2206, DOI: 10.1007/BF00962156There is no corresponding record for this reference. - 18
We also performed experiments investigating the effect of halide additives (20% NH4I, NH4Br, NH4F, NMe4I, NMe4Br, NMe4Cl, LiBr, LiCl, KBr, KCl, CsBr, or CsCl) on Suzuki–Miyaura couplings (see the Supporting Information).
There is no corresponding record for this reference. - 19
It is generally acknowledged that defining key transmetalation intermediates is even more complicated with aryltrifluoroborate nucleophiles. See, for example:
Lennox, A. J. J.; Lloyd-Jones, G. C. The Slow-Release Strategy in Suzuki–Miyaura Coupling. Isr. J. Chem. 2010, 50, 664– 674, and references cited therein DOI: 10.1002/ijch.20100007419The slow-release strategy in Suzuki-Miyaura couplingLennox, Alastair J. J.; Lloyd-Jones, Guy C.Israel Journal of Chemistry (2010), 50 (5-6), 664-674CODEN: ISJCAT; ISSN:0021-2148. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Despite great advances in metal-catalyzed cross-coupling reactions, their efficacy is often compromised by side reactions, reducing the yield, or requiring a large excess of one component. Suzuki-Miyaura cross-coupling is no exception, as the boronic acid functionality can be susceptible to a range of undesired processes. A no. of methods were developed to mitigate these side reactions, and herein we focus on the slow-release strategy. These conditions involve deployment of a masking reagent that protects the vulnerable boronic acid functionality from degrdn., particularly protodeboronation, while simultaneously facilitating controlled liberation of the active reagent into the catalytic milieu. Under suitably tailored conditions, this dual-action approach ensures that the concn. of the free boronic acid is minimized, thus attenuating its degrdn. but still facilitating transmetallation of the organoboron species with the key organopalladium intermediate. - 20(a) Andersen, N. G.; Keay, B. A. 2-Furyl Phosphines as Ligands for Transition-Metal-Mediated Organic Synthesis. Chem. Rev. 2001, 101, 997– 1030, DOI: 10.1021/cr000024o20a2-Furyl Phosphines as Ligands for Transition-Metal-Mediated Organic SynthesisAndersen, Neil G.; Keay, Brian A.Chemical Reviews (Washington, D. C.) (2001), 101 (4), 997-1030CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review with 177 refs.(b) Niemeyer, Z. L.; Milo, A.; Hickey, D. P.; Sigman, M. S. Parameterization of phosphine ligands reveals mechanistic pathways and predicts reaction outcomes. Nat. Chem. 2016, 8, 610– 617, DOI: 10.1038/nchem.250120bParameterization of phosphine ligands reveals mechanistic pathways and predicts reaction outcomesNiemeyer, Zachary L.; Milo, Anat; Hickey, David P.; Sigman, Matthew S.Nature Chemistry (2016), 8 (6), 610-617CODEN: NCAHBB; ISSN:1755-4330. (Nature Publishing Group)The mechanistic foundation behind the identity of a phosphine ligand that best promotes a desired reaction outcome is often nonintuitive, and thus was addressed in numerous exptl. and theor. studies. Multivariate correlations of reaction outcomes using 38 different phosphine ligands were combined with classic potentiometric analyses to study a Suzuki reaction, for which the site selectivity of oxidative addn. is highly dependent on the nature of the phosphine. These studies shed light on the generality of hypotheses regarding the structural influence of different classes of phosphine ligands on the reaction mechanism(s), and deliver a methodol. that should prove useful in future studies of phosphine ligands.
- 21(a) Tolman, C. A. Electron donor-acceptor properties of phosphorus ligands. Substituent additivity. J. Am. Chem. Soc. 1970, 92, 2953– 2956, DOI: 10.1021/ja00713a00621aElectron donor-acceptor properties of phosphorus ligands. Substituent additivityTolman, Chadwick A.Journal of the American Chemical Society (1970), 92 (10), 2953-6CODEN: JACSAT; ISSN:0002-7863.A rapid method is described for detg. electron donor-acceptor properties of triply connected P ligands based on the A1 carbonyl stretching frequency of Ni(CO)3L in CH2Cl2. Data are given for 70 ligands and a substituent additivity rule is proposed. Forty-seven substituent parameters χi are derived and found to correlate well with Kabachnik's σ parameters, based on ionization consts. of P acids.(b) Otto, S.; Roodt, A. Quantifying the electronic cis effect of phosphine, arsine and stibine ligands by use of rhodium(I) Vaska-type complexes. Inorg. Chim. Acta 2004, 357, 1– 10, DOI: 10.1016/S0020-1693(03)00436-521bQuantifying the electronic cis effect of phosphine, arsine and stibine ligands by use of rhodium(I) Vaska-type complexesOtto, Stefanus; Roodt, AndreasInorganica Chimica Acta (2004), 357 (1), 1-10CODEN: ICHAA3; ISSN:0020-1693. (Elsevier Science B.V.)The cis effects of phosphine, arsine and stibine ligands were evaluated by measuring the IR stretching frequency in dichloromethane of the carbonyl ligand in Rh(I) Vaska-type complexes, trans-[RhCl(CO)(L)2]. These data were correlated with those obtained by Tolman for the electronic trans influences in [Ni(L)(CO)3]. The electronic contribution, χFc, of ferrocenyl was detd. as 0.8 from these plots by evaluating PPh2Fc as ligand. To accommodate arsine and stibine ligands an addnl. correction term, to compensate for differences in the donor atom, was added to Tolman's equation for calcn. of the Tolman electronic parameter of phosphine ligands. In the resulting equation: v(CONi) = 2056.1 + Σ3i=1 χi + CL values for CL of CP = 0, CAs = -1.5 and CSb = -3.1 are suggested for phosphine, arsine and stibine ligands, resp. The crystal and mol. structures of trans-[RhCl(CO)(PPh2Fc)2]·2C6H6, trans-[RhCl(CO){P(NMe2)3}2] and trans-[RhCl(CO)(AsPh3)2] are reported. The Tolman cone angles for PPh2Fc and P(NMe2)3 were detd. as 169° and 166°, while the effective cone angles for PPh2Fc, P(NMe2)3 and AsPh3 were detd. as 171°, 168° and 147°, resp.(c) Ackermann, M.; Pascariu, A.; Höcher, T.; Siehl, H.-U.; Berge, S. Electronic Properties of Furyl Substituents at Phosphorus and Their Influence on 31P NMR Chemical Shifts. J. Am. Chem. Soc. 2006, 128, 8434– 8440, DOI: 10.1021/ja057085u21cElectronic Properties of Furyl Substituents at Phosphorus and Their Influence on 31P NMR Chemical ShiftsAckermann, Marco; Pascariu, Aurelia; Hoecher, Thomas; Siehl, Hans-Ullrich; Berger, StefanJournal of the American Chemical Society (2006), 128 (26), 8434-8440CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The electronic properties of 2-furyl and 3-furyl substituents attached to phosphines and phosphonium salts were studied by IR spectroscopy and exptl. and computational 31P NMR spectroscopy. The heteroarom. systems proved to be electron withdrawing with respect to Ph substituents. However, phosphorus atoms with attached furyl substituents are strongly shielded in NMR. The reason for this phenomenon was studied by solid state 31P MAS NMR expts. The chem. shift tensor was extd., and the orientation within the mols. was detd. The tensor component σ33, which is effected the most by furyl systems, is oriented perpendicular to the P-C bonds of the substituents. P-furyl bonds are shorter than P-Ph bonds. We assume therefore a lower ground-state energy of the mols., because of the electron withdrawing properties of the 2-furyl systems. The σpara component of the 31P NMR magnetic shielding is therefore smaller, which results in an overall increase of the magnetic shielding.
- 22
The observation that oxidative addition is not the turnover-limiting step in this reaction is consistent with previous studies. See, for example:
Amatore, C.; Jutand, A.; Le Duc, G. Kinetic Data for the Transmetalation/Reductive Elimination in Palladium-Catalyzed Suzuki–Miyaura Reactions: Unexpected Triple Role of Hydroxide Ions Used as Base. Chem. - Eur. J. 2011, 17, 2492– 2503, and references cited therein DOI: 10.1002/chem.20100191122Kinetic Data for the Transmetalation/Reductive Elimination in Palladium-Catalyzed Suzuki-Miyaura Reactions: Unexpected Triple Role of Hydroxide Ions Used as BaseAmatore, Christian; Jutand, Anny; Le Duc, GaetanChemistry - A European Journal (2011), 17 (8), 2492-2503CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The mechanism of the reaction of trans-[ArPdX(PPh3)2] (Ar = p-Z-C6H4; Z = CN, F, H; X = I, Br, Cl) with Ar'B(OH)2 (Ar' = p-Z'-C6H4; Z' = CN, H, OMe) has been established in DMF in the presence of the base OH- in the context of real palladium-catalyzed Suzuki-Miyaura reactions. The formation of the cross-coupling product ArAr' and [Pd0(PPh3)3] has been followed through the application of electrochem. techniques. Kinetic data have been obtained for the first time, with detn. of the obsd. rate const., kobs, of the overall reaction. trans-[ArPdX(PPh3)2] is not reactive in the absence of the base. The base OH- plays three roles. It favors the reaction: (1) by formation of trans-[ArPd(OH)(PPh3)2], a key complex which, in contrast to trans-[ArPdX(PPh3)2], reacts with Ar'B(OH)2 (rate-detg. transmetalation), and (2) by unexpected promotion of the reductive elimination from the intermediate trans-[ArPdAr'(PPh3)2], which generates ArAr' and a Pd0 species. Conversely, the base OH- disfavors the reaction by formation of the unreactive anionic Ar'B(OH)3-. As a consequence of these antagonistic effects of OH-, the overall reactivity is controlled by the concn. of OH- and passes through a max. as the concn. of OH- is increased. Therefore, the base favors the rate-detg. transmetalation and unexpectedly also the reductive elimination. - 23
It was reported that trans-[Pd(PPh3)2I2] was formed in reactions with [Pd(PPh3)4] and (iodoethynyl)benzene (in addition to the expected oxidative addition product). See:
(a) Weigelt, M.; Becher, D.; Poetsch, E.; Bruhn, C.; Steinborn, D. Zur oxidativen Addition von 1-Halogenalk-1-inen – Synthese und Struktur von Phenylalkinylpalladium-Komplexen. Z. Anorg. Allg. Chem. 1999, 625, 1542– 1547, DOI: 10.1002/(SICI)1521-3749(199909)625:9<1542::AID-ZAAC1542>3.0.CO;2-9There is no corresponding record for this reference.The formation of trans-[Pd(PPh3)2I2] from trans-[Pd(PPh3)2(R)(I)] has also been reported:
(b) Gulia, N.; Pigulski, B.; Szafert, S. Palladium End-Capped Polyynes via Oxidative Addition of 1-Haloalkynes to Pd(PPh3)4. Organometallics 2015, 34, 673– 682, DOI: 10.1021/om501105w23bPalladium End-Capped Polyynes via Oxidative Addition of 1-Haloalkynes to Pd(PPh3)4Gulia, Nurbey; Pigulski, Bartlomiej; Szafert, SlawomirOrganometallics (2015), 34 (4), 673-682CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)Reported here is the use of 1-haloacetylenes and 1-halopolyynes as synthons for the prepn. of new palladium(II) end-capped polyynes. The 1-haloalkynes were obtained in a series of transformations from para-substituted bromoarenes that included Sonogashira coupling followed by halogenation and chain elongation via Cadiot-Chodkiewicz protocol. The key step for the synthesis of metal complexes was oxidative addn. of 1-haloalkynes to Pd(PPh3)4, which allowed obtaining a series of metal compds. 1-5-CnX with carbon chains up to hexatriyne in 75-100% yield. All the compds. were characterized by NMR and HRMS or elemental anal. The 13C spectra of the 1-haloalkynes showed interesting, although expected, shifts of the carbon chain atoms close to the halogen termini. X-ray crystal structures were obtained for three polyynes-two butadiynes (2-C4[Pd]Br and 3-C4[Pd]Br) and one hexatriyne (1-C6[Pd]Br)-and the latter is the first reported X-ray crystal structure of palladium end-capped hexatriyne.trans-[Pd(PPh3)2I2] was also formed in reactions with [Pd(PPh3)4] and 9-iodo-m-carborane. See:
(c) Marshall, W. J.; Young, R. J., Jr.; Grushin, V. V. Mechanistic Features of Boron–Iodine Bond Activation of B–Iodocarboranes. Organometallics 2001, 20, 523– 533, DOI: 10.1021/om0008575There is no corresponding record for this reference. - 24
We were also able to isolate and characterize trans-[Pd(PPh3)2I2]. The spectroscopic data obtained on this compound were consistent both with equivalent data reported in the literature and also with an authentic sample of trans-[Pd(PPh3)2I2] that we prepared via a literature procedure:
Hahn, F. E.; Lugger, T.; Beinhoff, M. Palladium(II) Complexes with Benzoxazol-2-ylidene Ligands: Crystal Structures of trans-Chloro(benzoxazol-2-ylidene)bis(triphenylphosphine)palladium(II) Chloride and cis-Diiodo(benzoxazol-2-ylidene)(triphenylphosphine)palladium(II). Z. Naturforsch. B Chem. Sci. 2004, 59, 196– 20124Palladium(II) complexes with benzoxazol-2-ylidene ligands: crystal structures of trans-chloro(benzoxazol-2-ylidene)bis(triphenylphosphine)palladium(II) chloride and cis-diiodo(benzoxazol-2-ylidene)(triphenylphosphine)palladium(II)Hahn, F. Ekkehardt; Luegger, Thomas; Beinhoff, MatthiasZeitschrift fuer Naturforschung, B: Chemical Sciences (2004), 59 (2), 196-201CODEN: ZNBSEN; ISSN:0932-0776. (Verlag der Zeitschrift fuer Naturforschung)The Pd(II) complexes trans-[PdCl(L)(PPh3)2]Cl, 5, and cis-[PdI2(L)PPh3], 7, (L = benzoxazol-2-ylidene) were synthesized by treatment of trans-[PdX2(PPh3)2] (4: X = Cl, 6: X = iodo) with 2-(trimethylsiloxy)phenyl isocyanide 1, and subsequent hydrolysis of the Si-O bond. The crystal structures of 5 and 7·CH2Cl2 were established by x-ray diffraction. NMR and IR studies indicate that the unexpected cis-configuration of 7 obtained from trans-[PdI2(PPh3)2] is not the result of a soln. equil. between the cis and the trans isomers. - 25
When the study shown in eq 2 was performed in DMF/H2O instead of n-PrOH/H2O, similar results were obtained. When the experiment shown in eq 2 was performed using p-bromotoluene instead of p-iodotoluene, 31P NMR spectroscopy indicated that trans-[Pd(PPh3)2Br2] was not present in the reaction mixture. trans-[Pd(PPh3)2Br2] was also not observed in equivalent experiments performed in DMF/H2O.
There is no corresponding record for this reference. - 26Carrow, B. P.; Hartwig, J. F. Distinguishing Between Pathways for Transmetalation in Suzuki–Miyaura Reactions. J. Am. Chem. Soc. 2011, 133, 2116– 2119, DOI: 10.1021/ja110832626Distinguishing between pathways for transmetalation in Suzuki-Miyaura reactionsCarrow, Brad P.; Hartwig, John F.Journal of the American Chemical Society (2011), 133 (7), 2116-2119CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Transmetalation of arylpalladium complexes proceeds rather via palladium hydroxo complex and arylboronic acid pathway, than via arylpalladium halide and aryl(trihydroxo)borate. A systematic study of the stoichiometric reactions of isolated arylpalladium hydroxo and halide complexes with arylboronic acids and aryltrihydroxyborates is performed to evaluate the relative rates of the two reaction pathways commonly proposed to account for transmetalation in the Suzuki-Miyaura reaction. On the basis of the relative populations of the palladium and organoboron species generated under conditions common for the catalytic process and the obsd. rate consts. for the stoichiometric reactions between the two classes of reaction components, we conclude that the reaction of a palladium hydroxo complex with boronic acid, not the reaction of a palladium halide complex with trihydroxyborate, accounts for transmetalation in catalytic Suzuki-Miyaura reactions conducted with weak base and aq. solvent mixts.
- 27
For a recent review on transmetalation in the Suzuki–Miyaura cross-coupling, see:
Lennox, A. J. J.; Lloyd-Jones, G. C. Transmetalation in the Suzuki-Miyaura coupling: the fork in the trail. Angew. Chem., Int. Ed. 2013, 52, 7362– 7370, DOI: 10.1002/anie.20130173727Transmetalation in the Suzuki-Miyaura Coupling: The Fork in the TrailLennox, Alastair J. J.; Lloyd-Jones, Guy C.Angewandte Chemie, International Edition (2013), 52 (29), 7362-7370CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The Suzuki-Miyaura coupling is one of the few transition-metal-catalyzed C-C bond-forming reactions that were used in applications ranging from discovery chem. to manufg. processes. Although coupling proceeds through the generic three-stage oxidative addn., transmetalation, reductive elimination' sequence, there are a no. of features that differentiate the Suzuki-Miyaura process from other transition-metal-catalyzed cross-couplings. Most of these features are centered around, or are a consequence of, activation of the boron reagent for transmetalation through one or both of two distinct pathways. This review focuses on the evidence that was presented for this fork in the trail, and the potential to apply such mechanistic insight to the design of reaction conditions. - 28
For selected studies on transmetalation in the Suzuki–Miyaura cross-coupling, see: refs (22and26).
(a) Thomas, A. A.; Wang, H.; Zahrt, A. F.; Denmark, S. E. Structural, Kinetic, and Computational Characterization of the Elusive Arylpalladium(II)boronate Complexes in the Suzuki–Miyaura Reaction. J. Am. Chem. Soc. 2017, 139, 3805– 3821, DOI: 10.1021/jacs.6b1338428aStructural, Kinetic, and Computational Characterization of the Elusive Arylpalladium(II)boronate Complexes in the Suzuki-Miyaura ReactionThomas, Andy A.; Wang, Hao; Zahrt, Andrew F.; Denmark, Scott E.Journal of the American Chemical Society (2017), 139 (10), 3805-3821CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The existence of the oft-invoked intermediates contg. the crucial Pd-O-B subunit, the "missing link", has been established in the Suzuki-Miyaura cross-coupling reaction. The use of low-temp., rapid injection NMR spectroscopy (RI-NMR), kinetic studies, and computational anal. has enabled the generation, observation, and characterization of these highly elusive species. The ability to confirm the intermediacy of Pd-O-B-contg. species provided the opportunity to clarify mechanistic aspects of the transfer of the org. moiety from boron to palladium in the key transmetalation step. Specifically, these studies establish the identity of two different intermediates contg. Pd-O-B linkages, a tri-coordinate (6-B-3) boronic acid complex and a tetra-coordinate (8-B-4) boronate complex, both of which undergo transmetalation leading to the cross-coupling product. Two distinct mechanistic pathways have been elucidated for stoichiometric reactions of these complexes: (1) transmetalation via an unactivated 6-B-3 intermediate that dominates in the presence of an excess of ligand, and (2) transmetalation via an activated 8-B-4 intermediate that takes place with a deficiency of ligand.(b) Thomas, A. A.; Denmark, S. E. Pre-transmetalation intermediates in the Suzuki-Miyaura reaction revealed: The missing link. Science 2016, 352, 329– 332, DOI: 10.1126/science.aad698128bPre-transmetalation intermediates in the Suzuki-Miyaura reaction revealed: The missing linkThomas, Andy A.; Denmark, Scott E.Science (Washington, DC, United States) (2016), 352 (6283), 329-332CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Despite the widespread application of Suzuki-Miyaura cross-coupling to forge carbon-carbon bonds, the structure of the reactive intermediates underlying the key transmetalation step from the boron reagent to the palladium catalyst remains uncertain. Here we report the use of low-temp. rapid injection NMR spectroscopy and kinetic studies to generate, observe, and characterize these previously elusive complexes. Specifically, this work establishes the identity of three different species contg. palladium-oxygen-boron linkages, a tricoordinate boronic acid complex, and two tetracoordinate boronate complexes with 2:1 and 1:1 stoichiometry with respect to palladium. All of these species transfer their boron-bearing aryl groups to a coordinatively unsatd. palladium center in the crit. transmetalation event.(c) Ortuño, M. A.; Lledós, A.; Maseras, F.; Ujaque, G. The Transmetalation Process in Suzuki–Miyaura Reactions: Calculations Indicate Lower Barrier via Boronate Intermediate. ChemCatChem 2014, 6, 3132– 3138, DOI: 10.1002/cctc.20140232628cThe transmetalation process in Suzuki-Miyaura reactions: Calculations indicate lower barrier via boronate IntermediateOrtuno, Manuel A.; Lledos, Agusti; Maseras, Feliu; Ujaque, GregoriChemCatChem (2014), 6 (11), 3132-3138CODEN: CHEMK3; ISSN:1867-3880. (Wiley-VCH Verlag GmbH & Co. KGaA)Recent exptl. reports have called into question the validity of the boronate mechanism (through a R-B(OH)3- intermediate) for the transmetalation step in the Suzuki-Miyaura cross-coupling, favoring instead the palladium hydroxo pathway (through an [LnPd(R')(OH)] intermediate). Herein we report DFT calcns. with the M06 functional performed on realistic model systems, including a combination of explicit solvent mols. along with a continuum method. These computational results support the boronate mechanism. The mechanistic proposal is shown to be compatible with the available exptl. evidence.(d) Lima, C. F. R. A. C.; Rodrigues, A. S. M. C.; Silva, V. L. M.; Silva, A. M. S.; Santos, L. M. N. B. F. Role of the Base and Control of Selectivity in the Suzuki–Miyaura Cross-Coupling Reaction. ChemCatChem 2014, 6, 1291– 1302, DOI: 10.1002/cctc.20130108028dRole of the Base and Control of Selectivity in the Suzuki-Miyaura Cross-Coupling ReactionLima, Carlos F. R. A. C.; Rodrigues, Ana S. M. C.; Silva, Vera L. M.; Silva, Artur M. S.; Santos, Luis M. N. B. F.ChemCatChem (2014), 6 (5), 1291-1302CODEN: CHEMK3; ISSN:1867-3880. (Wiley-VCH Verlag GmbH & Co. KGaA)The outcome of the Suzuki-Miyaura cross-coupling for the direct competition reaction between two boronic acids was evaluated under routine synthesis conditions. The reaction selectivity depends on the amt. of the base used, with fewer bases favoring the reactivity of the boronic acid with lower pKa (stronger acid). The dependence of the reaction selectivity on base stoichiometry increases with the increase in the difference in the pKa values of the competing boronic acids. These results confirm a relation between acid-base chem. and the Suzuki-Miyaura reaction catalytic cycle. Also, under these specific conditions, the most reactive organoboron species toward transmetalation is the borate anion R-B(OH)3- instead of the neutral boronic acid R-B(OH)2. Hence, the main role of the base in the reaction mechanism is to increase the reactivity of the boronic acid toward the Pd-halide complex by converting it into the resp. organoborate. Boric acid, an important reaction byproduct, affects the selectivity in the Suzuki reaction because its gradual formation in the reaction medium disturbs the acid-base equil.(e) Amatore, C.; Jutand, A.; Le Duc, G. Mechanistic Origin of Antagonist Effects of Usual Anionic Bases (OH–, CO32–) as Modulated by their Countercations (Na+, Cs+, K+) in Palladium-Catalyzed Suzuki–Miyaura Reactions. Chem. - Eur. J. 2012, 18, 6616– 6625, DOI: 10.1002/chem.20120051628eMechanistic Origin of Antagonist Effects of Usual Anionic Bases (OH-, CO32-) as Modulated by their Countercations (Na+, Cs+, K+) in Palladium-Catalyzed Suzuki-Miyaura ReactionsAmatore, Christian; Jutand, Anny; Le Duc, GaetanChemistry - A European Journal (2012), 18 (21), 6616-6625, S6616/1-S6616/8CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The mechanism of the reaction of trans-ArPdBrL2 (Ar = p-Z-C6H4, Z = CN, H; L = PPh3) with Ar'B(OH)2 (Ar' = p-Z'-C6H4, Z' = H, CN, MeO), which is a key step in the Suzuki-Miyaura process, was established in DMF with two bases, acetate (Bu4NOAc) or carbonate (Cs2CO3) and compared with that of hydroxide (Bu4NOH). As anionic bases are inevitably introduced with a countercation M+ (e.g., M+OH-), the role of cations in the transmetalation/reductive elimination was 1st studied. Cations M+ (Na+, Cs+, K+) are not innocent since they induce an unexpected decelerating effect in the transmetalation via their complexation to the OH ligand in the reactive ArPd(OH)L2, partly inhibiting its transmetalation with Ar'B(OH)2. A decreasing reactivity order is obsd. when M+ is assocd. with OH-: Bu4N+> K+> Cs+> Na+. Acetates give trans-ArPd(OAc)L2, which does not undergo transmetalation with Ar'B(OH)2. This explains why acetates are not used as bases in Suzuki-Miyaura reactions that involve Ar'B(OH)2. Carbonates (Cs2CO3) give rise to slower reactions than those performed from Bu4NOH at the same concn., even if the reactions are accelerated in the presence of H2O due to the generation of OH-. The mechanism of the reaction with carbonates is then similar to that established for Bu4NOH, involving ArPd(OH)L2 in the transmetalation with Ar'B(OH)2. Due to the low concn. of OH- generated from CO32- in H2O, both transmetalation and reductive elimination result slower than those performed from Bu4NOH at equal concns. as Cs2CO3. Therefore, the overall reactivity is finely tuned by the concn. of the common base OH- and the ratio [OH-]/[Ar'B(OH)2]. Hence, the anionic base (pure OH- or OH- generated from CO32-) assocd. with its countercation (Na+, Cs+, K+) plays four antagonist kinetic roles: acceleration of the transmetalation by formation of the reactive ArPd(OH)L2, acceleration of the reductive elimination, deceleration of the transmetalation by formation of unreactive Ar'B(OH)3- and by complexation of ArPd(OH)L2 by M+.(f) Schmidt, A. F.; Kurokhtina, A. A.; Larina, E. V. Role of a base in Suzuki-Miyaura reaction. Russ. J. Gen. Chem. 2011, 81, 1573– 1574, DOI: 10.1134/S107036321107033428fRole of a base in Suzuki-Miyaura reactionSchmidt, A. F.; Kurokhtina, A. A.; Larina, E. V.Russian Journal of General Chemistry (2011), 81 (7), 1573-1574CODEN: RJGCEK; ISSN:1070-3632. (MAIK Nauka/Interperiodica)There is no expanded citation for this reference.(g) Butters, M.; Harvey, J.; Jover, J.; Lennox, A.; Lloyd-Jones, G.; Murray, P. Aryl Trifluoroborates in Suzuki–Miyaura Coupling: The Roles of Endogenous Aryl Boronic Acid and Fluoride. Angew. Chem., Int. Ed. 2010, 49, 5156– 5160, DOI: 10.1002/anie.20100152228gAryl Trifluoroborates in Suzuki-Miyaura Coupling: The Roles of Endogenous Aryl Boronic Acid and FluorideButters, Mike; Harvey, Jeremy N.; Jover, Jesus; Lennox, Alastair J. J.; Lloyd-Jones, Guy C.; Murray, Paul M.Angewandte Chemie, International Edition (2010), 49 (30), 5156-5160, S5156/1-S5156/68CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)The Suzuki-Miyaura coupling (SM) of 4-FC6H4BF3- K+ (I) with 3,5-(CF3)2C6H3Br was studied in a toluene/ water biphasic soln., and in a THF/ water soln.; both systems are commonly employed for the SM of trifluoroborates. The reactions in toluene/water, failed to go to completion: turnover ceased after 6 h, affording 55% of the base-catalyzed protodeboronation product PhF and ≤ 32% of coupling product 4-[3,5-(CF3)2C6H3]C6H4F (II). In aq. THF the reaction proceeded much more efficiently (5.5 h; > 95% yield of II), with few side products (< 0.1-2% ), even when the reaction was performed in air. But reaction of 4-C6H4B(OH)2 under identical conditions, gave II in variable yield, and afforded substantially more of side products 4-(4-FC6H4)C6H44-F and 4-FC6H4OH (2-40% ), compared to trifluoroborate substrate I. The reaction intermediates were established and a mechanism, supported by B3LYP calcns. was proposed for the SM.(h) Huang, Y.-L.; Weng, C.-M.; Hong, F.-E. Density Functional Studies on Palladium-Catalyzed Suzuki–Miyaura Cross-Coupling Reactions Assisted by N- or P-Chelating Ligands. Chem. - Eur. J. 2008, 14, 4426– 4434, DOI: 10.1002/chem.20080001128hDensity functional studies on palladium-catalyzed Suzuki-Miyaura cross-coupling reactions assisted by N- or P-chelating ligandsHuang, Yi-Luen; Weng, Chia-Ming; Hong, Fung-E.Chemistry - A European Journal (2008), 14 (14), 4426-4434CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)DFT studies with the B3LYP functional have been carried out on the Suzuki-Miyaura cross-coupling reactions of Ph chloride and phenylboronic acid catalyzed by palladium complexes with N- or P-chelating ligands. The full catalytic cycle, from the addn. of reactants to the catalyst to the release of the cross-coupled product from the complexed intermediate, has been examd. The stages within the cycle, such as oxidative addn., transmetalation, and reductive elimination, were validated by linking the mechanistically relevant intermediates and transition states. Various derivs. of diimine, diphosphine, and diamine were considered as potential model ligands. The catalytic reaction employing diimine as the chelating ligand has been verified as the one with the most energetically feasible route.(i) Braga, A. A. C.; Ujaque, G.; Maseras, F. A DFT Study of the Full Catalytic Cycle of the Suzuki–Miyaura Cross-Coupling on a Model System. Organometallics 2006, 25, 3647– 3658, DOI: 10.1021/om060380i28iA DFT Study of the Full Catalytic Cycle of the Suzuki-Miyaura Cross-Coupling on a Model SystemBraga, Ataualpa A. C.; Ujaque, Gregori; Maseras, FeliuOrganometallics (2006), 25 (15), 3647-3658CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)A computational study with the Becke3LYP DFT functional is carried out on the cross-coupling reaction of vinyl bromide, H2C:CHBr and vinylboronic acid, H2C:CHB(OH)2 catalyzed by Pd diphosphine [Pd(PH3)2] in the presence of an excess of OH- base. The full catalytic cycle is computed, starting from the sepd. reactants and the catalyst and finishing with the cross-coupled product and the regeneration of the catalyst. The different stages in the cycle (oxidative addn., isomerization, transmetalation, reductive elimination) were characterized through calcn. of the corresponding intermediates and transition states. Different alternative mechanisms are considered, depending on the no. of phosphine ligands at Pd, and on the cis or trans isomery around the metal center. The results indicate the existence of a no. of competitive pathways of reasonably low energy.(j) Braga, A. A. C.; Morgon, N. H.; Ujaque, G.; Lledós, A.; Maseras, F. Computational study of the transmetalation process in the Suzuki–Miyaura cross-coupling of aryls. J. Organomet. Chem. 2006, 691, 4459– 4466, DOI: 10.1016/j.jorganchem.2006.02.01528jComputational study of the transmetalation process in the Suzuki-Miyaura cross-coupling of arylsBraga, Ataualpa A. C.; Morgon, Nelson H.; Ujaque, Gregori; Lledos, Agusti; Maseras, FeliuJournal of Organometallic Chemistry (2006), 691 (21), 4459-4466CODEN: JORCAI; ISSN:0022-328X. (Elsevier Ltd.)The transmetalation step of the Suzuki-Miyaura cross-coupling between aryl groups is analyzed by DFT calcns. with the Becke3LYP functional. The halide considered is PhBr, and the organoboronic acid is PhB(OH)2. The model catalyst is Pd(PH3)2, and the base, OH-. The transmetalation is considered to start from the Pd(Ph)(PH3)2Br complex, the product of the oxidative addn. The results are compared with those of a previous study on the analogous reaction with vinyl groups, and it is shown that the reaction mechanism is very similar.(k) Braga, A. A. C.; Morgon, N. H.; Ujaque, G.; Maseras, F. Computational Characterization of the Role of the Base in the Suzuki–Miyaura Cross-Coupling Reaction. J. Am. Chem. Soc. 2005, 127, 9298– 9307, DOI: 10.1021/ja050583i28kComputational characterization of the role of the base in the Suzuki-Miyaura cross-coupling reactionBraga, Ataualpa A. C.; Morgon, Nelson H.; Ujaque, Gregori; Maseras, FeliuJournal of the American Chemical Society (2005), 127 (25), 9298-9307CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The role of the base in the transmetalation step of the Suzuki-Miyaura cross-coupling reaction is analyzed computationally by means of DFT calcns. with the Becke3LYP functional. The model system studied consists of Pd(CH:CH2)(PH3)2Br as the starting catalyst complex, CH2:CHB(OH)2 as the organoboronic acid, and OH- as the base. The two main mechanistic proposals, consisting of the base attacking first either the palladium complex or the organoboronic acid, are evaluated through geometry optimization of the corresponding intermediates and transition states. Supplementary calcns. are carried out on the uncatalyzed reaction and on a process where the starting complex is Pd(CH:CH2)(PH3)2(OH). These calcns., considered together with available exptl. data, strongly suggest that the main mechanism of transmetalation in the catalytic cycle starts with the reaction of the base and the organoboronic acid.(l) Miyaura, N. Cross-coupling reaction of organoboron compounds via base-assisted transmetalation to palladium(II) complexes. J. Organomet. Chem. 2002, 653, 54– 57, DOI: 10.1016/S0022-328X(02)01264-028lCross-coupling reaction of organoboron compounds via base-assisted transmetalation to palladium(II) complexesMiyaura, NorioJournal of Organometallic Chemistry (2002), 653 (1-2), 54-57CODEN: JORCAI; ISSN:0022-328X. (Elsevier Science B.V.)A review. Until recently, organoboronic acids have been of limited use in org. synthesis due to their low reactivity for ionic reactions. During the past two decades, however, it has become increasingly clear that they are valuable reagents capable of undergoing many carbon-carbon bond formations in org. syntheses. In 1979, catalytic carbon-carbon bond formation via transmetalation between organoboron compds. and palladium(II) halides was found to proceed in the presence of a base.(m) Matos, K.; Soderquist, J. A. Alkylboranes in the Suzuki–Miyaura Coupling: Stereochemical and Mechanistic Studies. J. Org. Chem. 1998, 63, 461– 470, DOI: 10.1021/jo971681sThere is no corresponding record for this reference.(n) Aliprantis, A. O.; Canary, J. W. Observation of Catalytic Intermediates in the Suzuki Reaction by Electrospray Mass Spectrometry. J. Am. Chem. Soc. 1994, 116, 6985– 6986, DOI: 10.1021/ja00094a08328nObservation of Catalytic Intermediates in the Suzuki Reaction by Electrospray Mass SpectrometryAliprantis, Antonios O.; Canary, James W.Journal of the American Chemical Society (1994), 116 (15), 6985-6CODEN: JACSAT; ISSN:0002-7863.Pyridyl bromides, e.g., I (R = H, CHO, X = 2, 3-Br), couple with phenylboronic acids II (R1 = H, Me, R2 = H, Me) in the presence of Pd[PPh3]4 catalyst to give (2- and 3-aryl-substituted)pyridines III. Intermdiate formation of palladium complexes for this coupling reaction, e.g., [(pyrH)(R1R2C6H3)Pd(PPh3)2]+, are obsd. by electrospray Mass Spectrometry.(o) Smith, G. B.; Dezeny, G. C.; Hughes, D. L.; King, A. O.; Verhoeven, T. Mechanistic Studies of the Suzuki Cross-Coupling Reaction. J. Org. Chem. 1994, 59, 8151– 8156, DOI: 10.1021/jo00105a03628oMechanistic Studies of the Suzuki Cross-Coupling ReactionSmith, George B.; Dezeny, George C.; Hughes, David L.; King, Anthony O.; Verhoeven, Thomas R.Journal of Organic Chemistry (1994), 59 (26), 8151-6CODEN: JOCEAH; ISSN:0022-3263. (American Chemical Society)The key step in the synthesis of the drug losartan is a palladium-catalyzed cross-coupling reaction of an aryl bromide and a boronic acid. The reaction scheme was defined in kinetic studies using HPLC, and computer simulation served to depict the time dependence of the concns. of palladium species, which were not obsd. exptl. Two catalyst poisons were identified and characterized. One was an isomeric impurity of the aryl bromide; the other was formed in the reaction mixt. upon hydrolysis of the boronic acid and two of its impurities.(p) Thomas, A. A.; Zahrt, A. F.; Delaney, C. P.; Denmark, S. E. Elucidating the Role of the Boronic Esters in the Suzuki–Miyaura Reaction: Structural, Kinetic, and Computational Investigations. J. Am. Chem. Soc. 2018, 140, 4401– 4416, DOI: 10.1021/jacs.8b0040028pElucidating the Role of the Boronic Esters in the Suzuki-Miyaura Reaction: Structural, Kinetic, and Computational InvestigationsThomas, Andy A.; Zahrt, Andrew F.; Delaney, Connor P.; Denmark, Scott E.Journal of the American Chemical Society (2018), 140 (12), 4401-4416CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The Suzuki-Miyaura reaction is the most practiced palladium-catalyzed, cross-coupling reaction because of its broad applicability, low toxicity of the metal (B), and the wide variety of com. available boron substrates. A wide variety of boronic acids and esters, each with different properties, have been developed for this process. Despite the popularity of the Suzuki-Miyaura reaction, the precise manner in which the org. fragment is transferred from boron to palladium has remained elusive for these reagents. Herein, we report the observation and characterization of pretransmetalation intermediates generated from a variety of commonly employed boronic esters. The ability to confirm the intermediacy of pretransmetalation intermediates provided the opportunity to clarify mechanistic aspects of the transfer of the org. moiety from boron to palladium in the key transmetalation step. A series of structural, kinetic, and computational investigations revealed that boronic esters can transmetalate directly without prior hydrolysis. Furthermore, depending on the boronic ester employed, significant rate enhancements for the transfer of the B-aryl groups were obsd. Overall, two crit. features were identified that enable the transfer of the org. fragment from boron to palladium: (1) the ability to create an empty coordination site on the palladium atom and (2) the nucleophilic character of the ipso carbon bound to boron. Both of these features ultimately relate to the electron d. of the oxygen atoms in the boronic ester. - 29
Lloyd-Jones and Lennox note that “Elucidation of the dominant pathway to transmetalation is not at all straightforward: one must establish the kinetically active boron and palladium intermediates, and they may not necessarily be the most abundant species present in the medium.”. (27)
There is no corresponding record for this reference. - 30
Detailed studies exploring the transmetalation of trans-[Pd(PPh3)2(Ph)(I)] with p-tol-B(OH)3K (and 18-crown-6) and Pd(PPh3)(Ph)(μ-OH)]2 with p-tol-B(OH)2 under stoichiometric conditions were performed in THF/H2O solvent mixtures. (26) Extensive studies employing electrochemical methods to investigate the transmetalation of trans-[Pd(PPh3)2(Ar)(X)] with Ar-B(OH)2 (n-Bu4NOH as the base) under catalytic conditions were performed in a DMF/MeOH solvent mixture. (22)
There is no corresponding record for this reference. - 31
Germane observations regarding the adverse effect of PPh3 have also been reported in Pd-catalyzed Stille cross-couplings; see ref (9).
(a) Farina, V.; Baker, S. R.; Benigni, D. A.; Hauck, S. I.; Sapino, C., Jr Palladium catalysis in cephalosporin chemistry: general methodology for the synthesis of cephem side chains. J. Org. Chem. 1990, 55, 5833– 5847, DOI: 10.1021/jo00310a01431aPalladium catalysis in cephalosporin chemistry: general methodology for the synthesis of cephem side chainsFarina, Vittorio; Baker, Stephen R.; Benigni, Daniel A.; Hauck, Sheila I.; Sapino, Chester, Jr.Journal of Organic Chemistry (1990), 55 (23), 5833-47CODEN: JOCEAH; ISSN:0022-3263.The Pd-catalyzed coupling of 3-(triflyloxy)cephems with organotin compds.leads to 3-alkenyl-, 3-alkynyl-, and 3-arylcephems under exceptionally mild conditions. While this approach was not satisfactory for 3-allylcephems, the related coupling of easily available 3-(chloromethyl)cephems with stannanes provided a high-yielding route to such 3-allylcephems and also to 3-benzyl- and 3-homoallylcephems. The choice of the catalyst was crucial in both cases. PPh3-based catalysts were quite unsatisfactory. A much better ligand in this respect was tri(2-furyl)phosphine, which is introduced as a useful new ligand in organopalladium chem. The effect of this ligand on coupling rates is discussed. It is suggested that reduced electron d. at Pd enhances the rate of the transmetalation, considered to be the rate-detg. step in these coupling reactions. Cases of unexpected transfer order among unsym. substituted stannanes were obsd., as well as a complex coupling reaction between (chloromethyl)cephems and alkynylstannanes. These new facets of the Stille coupling, as well as the potential of this chem. for the development of new cephalosporin antibiotics, are discussed.(b) Scott, W. J.; Stille, J. K. Palladium-catalyzed coupling of vinyl triflates with organostannanes. Synthetic and mechanistic studies. J. Am. Chem. Soc. 1986, 108, 3033– 3040, DOI: 10.1021/ja00271a03731bPalladium-catalyzed coupling of vinyl triflates with organostannanes. Synthetic and mechanistic studiesScott, William J.; Stille, J. K.Journal of the American Chemical Society (1986), 108 (11), 3033-40CODEN: JACSAT; ISSN:0002-7863.The palladium-catalyzed coupling reaction of vinyl triflates, e.g., (I), with acetylenic, vinyl, allyl, and alkyl tin reagents in the presence of LiCl or another suitable salt takes place in high yields under mild reaction conditions; however, benzyl and Ph tin reagents give poor yield of coupled product. The utilization of a tin or silicon hydride reagent in place of the organotin partner yields the alkene by reductive cleavage of the triflate group. The palladium-catalyzed reaction of vinyl triflates with Me3SnSnMe3 gives vinyl stannanes in high yields. Regioselectively formed vinyl triflates can be used to produce 1,3-dienes as the regioisomeric coupled products. - 33
For a study focused on the reactivity of Pd(0) complexes derived from mixtures of Pd(dba)2 and PPh3 and P(2-furyl)3 ligands, see for example:
Amatore, C.; Jutand, A.; Meyer, G.; Atmani, H.; Khalil, F.; Chahdi, F. O. Comparative Reactivity of Palladium(0) Complexes Generated in Situ in Mixtures of Triphenylphosphine or Tri-2-furylphosphine and Pd(dba)2. Organometallics 1998, 17, 2958– 2964, DOI: 10.1021/om971064u33Comparative Reactivity of Palladium(0) Complexes Generated in Situ in Mixtures of Triphenylphosphine or Tri-2-furylphosphine and Pd(dba)2Amatore, Christian; Jutand, Anny; Meyer, Gilbert; Atmani, Hamid; Khalil, Fouad; Chahdi, Fouad OuazzaniOrganometallics (1998), 17 (14), 2958-2964CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)For PPh3, mixts. of Pd(dba)2 and nTFP (TFP = tri-2-furylphosphine, n ≥ 2) in DMF and THF (S) lead to the formation of Pd(dba)(TFP)2, SPd(TFP)3 in equil. with SPd(TFP)2. The substitution of dba by the phosphine in Pd(dba)L2 to form SPdL3 is easier for L = TPF than for L = PPh3. The less ligated complex SPd(TFP)2 is the reactive species in the oxidative addn. with Ph iodide. In THF, {Pd(dba)2 + nTFP}, a mixt. often used as a catalyst promoter in several synthetic org. reactions, is found to be less reactive than {Pd(dba)2 + nPPh3} for small values of n (n = 2 or 4) whereas it is more reactive for higher values of n (n > 6). Conversely, in DMF, {Pd(dba)2 + nTFP} is always found to be more reactive than {Pd(dba)2 + nPPh3} whatever n (n ≥ 2). - 34
Farina and Krishnan also state “When a large excess of P(2-furyl)3 was added to a solution of trans-[Pd(PPh3)2(Ph)(I)] and PPh3, no trace of trans-[Pd{P(2-furyl)3}2(Ph)(I)] was observed, the corresponding signal for P(2-furyl)3 being the only new peak in the spectrum.”. (9)
There is no corresponding record for this reference.
Supporting Information
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.organomet.8b00189.
Experimental procedures and compound characterization data (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.