ACS Publications. Most Trusted. Most Cited. Most Read
A Fluorinated Ligand Enables Room-Temperature and Regioselective Pd-Catalyzed Fluorination of Aryl Triflates and Bromides
My Activity

Figure 1Loading Img
  • Open Access
Article

A Fluorinated Ligand Enables Room-Temperature and Regioselective Pd-Catalyzed Fluorination of Aryl Triflates and Bromides
Click to copy article linkArticle link copied!

View Author Information
Department of Chemistry, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United States
Department of Chemistry, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, United States
Open PDFSupporting Information (2)

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2015, 137, 41, 13433–13438
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.5b09308
Published September 28, 2015

Copyright © 2015 American Chemical Society. This publication is licensed under these Terms of Use.

Abstract

Click to copy section linkSection link copied!

A new biaryl monophosphine ligand (AlPhos, L1) allows for the room-temperature Pd-catalyzed fluorination of a variety of activated (hetero)aryl triflates. Furthermore, aryl triflates and bromides that are prone to give mixtures of regioisomeric aryl fluorides with Pd-catalysis can now be converted to the desired aryl fluorides with high regioselectivity. Analysis of the solid-state structures of several Pd(II) complexes, as well as density functional theory (DFT) calculations, shed light on the origin of the enhanced reactivity observed with L1.

Copyright © 2015 American Chemical Society

Introduction

Click to copy section linkSection link copied!

The preparation of fluorine-containing aromatic compounds (ArF) has been a long-standing challenge in organic synthesis. (1) While these compounds are highly prized in the pharmaceutical (2) and agrochemical industries, (3) methods to access aryl fluorides selectively, under mild and general reaction conditions, remain elusive. Since the advent of the Balz–Schiemann (4) reaction and the Halex (5) process, significant advances have been made toward this end, (6) especially with regard to transition metal-catalyzed methods. (7) In analogy to the practicality, simplicity, and generality of Pd(0)/Pd(II)-catalyzed aryl carbon–heteroatom bond-forming processes, the coupling of aryl (pseudo)halides with simple metal fluoride salts (“F”) (Figure 1a) would be ideal. However, several challenges associated with developing such a process have been revealed both experimentally (8) and theoretically. (9) Among these difficulties are the formation of stable [LPd(II)F]2 dimers and competitive P–F and P–C bond formation, which suggests a high barrier for C–F reductive elimination (Figure 1a). To circumvent the difficult reductive elimination from Pd(II) complexes, electrophilic fluorine sources (“F+”) have been employed to oxidize the metal center to Pd(III) or Pd(IV) intermediates from which reductive elimination is more facile. (10)
The desired C–F reductive elimination from [LPdAr(F)] is unique to biaryl monophosphine-ligated complexes, and has been observed stoichiometrically, (11) albeit at high temperatures and with substrates that contain activated (12) aryl groups. Subsequently, it was discovered that incorporating a substituent at the C3′ position of the biaryl monophosphine ligand, for example, as in HGPhos ((L2), Figure 1b), gives rise to a more active catalyst system. (13)

Figure 1

Figure 1. (a) The proposed catalytic cycle for Pd-catalyzed fluorination. (b) A Pd(II) complex with L2 highlighted in red. The interaction between ligand-bound Pd(II) and C1′ is shown by a dashed line. (c) Biaryl monophosphine ligand (Alphos, L1).

While these discoveries have shown that Pd(II)-catalyzed fluorination is indeed viable, several problems remain. Most importantly, the generation of regioisomeric byproducts (14) (presumably through a Pd–benzyne intermediate) lowers the yield of the desired product and renders purification difficult or even impossible. Additionally, the use of elevated reaction temperatures is required to achieve full conversion of the starting materials. To address these issues, which we ascribe to the difficult C–F reductive elimination from Pd(II) complexes, we sought to design a ligand to improve this elementary step as well as the overall efficiency of the catalytic transformation. As shown in Figure 1b, biaryl monophosphine ligands coordinate to the Pd(II) metal center in a bidentate fashion, making contacts through both the phosphine and the C1′ carbon of the adjacent aromatic ring. (15) Because reductive elimination from Pd(II) complexes is typically favored from a three-coordinate intermediate, (16) we hypothesized that modification at C3′ with an electron-withdrawing group would diminish the donation of electron density from C1′ to the Pd(II) metal center, providing an intermediate with more three-coordinate character and thereby generating a more active catalyst.

Results and Discussion

Click to copy section linkSection link copied!

To test our hypothesis, we designed and synthesized a ligand that was less electron-rich than the previously reported L2. This was accomplished by removing the methoxy group adjacent to the dialkylphosphino group on the “upper” aromatic ring and replacing the hydrogen atoms of the pendant 4-(n-Bu)Ph group at C3′ with fluorines (17) (Figure 1c). The synthesis of L1 is described in Scheme 1. Lithium–halogen exchange of 2,4,6-triisopropylbromobenzene with n-BuLi followed by nucleophilic addition into hexafluorobenzene provides perfluoro biaryl L1a. To improve solubility of the final ligand and prevent subsequent nucleophilic additions into the activated perfluoro aromatic, n-BuLi was added at −78 °C to give the alkylated product via SNAr, which was then halogenated using N-iodosuccinimide and H2SO4 to yield L1b. Terphenyl L1c was prepared by nucleophilic addition of L1b into a benzyne intermediate, generated from 3-fluoroanisole, followed by quenching the resulting aryl anion with bromine at −30 °C. Treatment of L1c with t-BuLi at −78 °C followed by the addition of CuCl and Ad2PCl and then heating to 140 °C overnight furnished L1 in 82% yield (25% yield over five steps). The synthesis was amenable to scale up as more than 10 g of L1 was prepared in a single batch.

Scheme 1

Scheme 1. Synthesis of L1 and Pd(0) Precatalyst 1a

Scheme aReagents and conditions: (a) THF, −78 °C, n-BuLi, 1 h; added to C6F6 in THF, 0 °C to rt, 2 h, 87%. (b) THF, −78 °C, n-BuLi, 30 min, 99%. (c) AcOH, EtOAc, H2SO4, N-iodosuccinimide, 50 °C, 17 h, 80%. (d) THF, −78 °C, t-BuLi, 1 h; added to 3-fluoroanisole and n-BuLi in THF, −78 to −25 °C, 44%. (e) THF, −78 °C, t-BuLi, 1 h; CuCl −78 °C to rt; Ad2PCl, toluene, 140 °C, 18 h, 82%. (f) [COD·Pd(CH2TMS)2], pentane, rt, 48 h, 78%.

The highly active Pd(0) precatalyst [(L1Pd)2·COD] (1) was prepared in 78% yield by treating L1 with an equivalent of [COD·Pd(CH2TMS)2] in pentane, which could be easily isolated and stored under an inert atmosphere (Scheme 1). (18, 19) Because of the electron-rich nature and high reactivity of biaryl monophosphine-ligated Pd(0) complexes, there exist only a few structurally characterized examples. (20) We have previously described several Pd(0) precatalysts, of which indirect evidence was obtained suggesting an empirical formula of [(LPd)n·COD] (n = 1–2); however, the insolubility of these complexes precluded structural characterization by single-crystal X-ray diffraction. (13c, 18) Single crystals of 1 suitable for X-ray diffraction were obtained providing the first structural evidence for this class of precatalyst (Figure 2). Complex 1 crystallizes in the triclinic centrosymmetric space group P1̅ with half a molecule of 1 and three molecules of benzene per asymmetric unit. The second half of 1 is generated by the crystallographic inversion center. The full molecule is a binuclear complex in which each Pd atom is chelated by one L1 ligand. Located between the two palladium centers is a cyclooctadiene molecule (COD), which coordinates through its two double bonds to the metal atoms in a side-on fashion that can be described as η2. This results in a distorted trigonal planar coordination environment for the palladium atoms with the following angles: P–Pd–C1′, 88°; P–Pd–COD, 135°; and C1′–Pd–COD, 137°.

Figure 2

Figure 2. Crystal structure of [(L1Pd)2·COD] (1). Thermal ellipsoids are shown at 50% probability; hydrogen atoms and residual benzene molecules are omitted for clarity.

The reactivity of L1 and L2 was compared by preparing analogous [LPdAr(F)] complexes and examining the formation of the ArF resulting from the stoichiometric C–F reductive elimination from the Pd(II) intermediates at room temperature (Scheme 2). These reactions were performed in the presence and absence of added aryl bromide, which serves as a trapping agent for the resulting Pd(0) that is formed after reductive elimination. (21, 13a) After a period of 12 h, the L1-supported complex provides approximately twice the amount of the desired product than that bearing L2. Both L1- and L2-ligated complexes provided the ArF product and are the first examples of room-temperature C–F reductive elimination from isolated [LPd(Ar)F] complexes.

Scheme 2

Scheme 2. Stoichiometric C–F Reductive Elimination from Pd(II) Complexes at Room Temperaturea

Scheme aYields determined by 19F NMR. bNo ArBr added. cArBr (R = C(O)n-Bu). dArBr (R = n-Bu).

The enhanced reactivity in the stoichiometric process exhibited by the L1-supported Pd(Ar)F complex was successfully extended to the catalytic reaction. By using 1, a variety of activated (hetero)aromatic triflates (Table 1) were transformed to the fluoride derivatives at room temperature. (22) A range of functional groups were tolerated in this transformation, including nitro (2), formyl (3), methyl ketone (4), redox active anthraquinone (5), and sulfonamide/tertiary amine (6). Heteroaryl triflates were also competent substrates as flavone 7, quinolines 8 and 9, and quinazoline 10 derivatives were prepared in high yields. Notably, XAV939, (23) a tankyrase inhibitor and potential cancer therapeutic, could be fluorinated to give 11, demonstrating the applicability of this method to fluorinate medicinally relevant compounds. Additionally, aryl triflates can be selectively transformed in the presence of aryl chlorides, leaving a reactive handle for downstream modifications (1214). While the long reaction times at room temperature speak to the stability of the catalyst, the transformation can be completed in 24 h by increasing the temperature even while lowering the catalyst loading (Table 1, 7 and 9).
Table 1. Room-Temperature Fluorination of Aryl Triflatesa
Table a

Isolated yields are reported as an average of two runs. Reaction conditions unless otherwise noted: Aryl triflate (1 mmol), CsF (3 mmol), tol or 2-MeTHF (10 mL).

Table b

0.50 mmol scale.

Table c

19F NMR yield.

Table d

Aryl bromide (0.5 mmol), KF (0.25 mmol), AgF (1.0 mmol), tol (5 mL). tol = toluene, 2-MeTHF = 2-methyl tetrahydrofuran.

With our previously described catalyst systems, (11, 13, 14) mixtures of regioisomeric aryl fluorides are sometimes formed during the Pd-catalyzed nucleophilic fluorination reaction. (14) In general, substrates with para electron-donating groups or meta substitution are plagued by this undesired reaction pathway. Previously, by using cyclohexane as the solvent, the amount of regioisomer formation could be reduced. In some instances, however, the quantity of undesired regioisomer formed was still significant.
To probe the effectiveness of catalysts derived from 1 to further suppress the formation of regioisomeric byproducts, we examined 4-(n-Bu)PhOTf (X = OTf) and 4-(n-Bu)PhBr (X = Br) as model substrates (Table 2). With either aryl electrophile, the L2-supported precatalyst provides substantial amounts of undesired isomer B (entry 1). The enhanced reactivity of 1, however, allows for the use of lower reaction temperatures than were previously possible (entry 2, X = OTf), revealing a temperature dependence on the formation of B (entries 3–5, X = OTf). The same temperature dependence is not observed for aryl bromides, although using 1 diminishes the amount of B formed. Currently, we have no explanation for this dichotomy.
Table 2. Temperature Dependence on Regioisomer Formationa,b
Table a

Yields were determined by 19F NMR. Numbers in parentheses indicate % conversion of the starting material.

Table b

The reaction time was 48 h.

The success of 1 in suppressing the formation of B with 4-(n-Bu)phenyl triflate or bromide prompted us to study the reaction of a number of substrates that are prone to give regioisomeric mixtures of ArF products. As shown in Table 3, a variety of aryl triflates and bromides converted to the corresponding aryl fluorides in high yield with >100:1 selectivity for the desired regioisomer (1521, 2325). Previously, substrates with strongly electron-donating groups in the para position, such as −OMe, have been the most challenging substrates for this methodology. For example, 4-methoxyphenyl triflate (15) and 4-bromoanisole (13c) have previously been shown to form the undesired meta isomer as the major product. Using 1, the observed selectivity is reversed and the desired product dominates (22). Several fluorinated analogues of biologically active compounds were also prepared to give (±)-naproxen 23, diethylstilbesterol 24, and tyrosine 25 derivatives with exceptional selectivity. No loss in stereochemical integrity was observed with 25 under the reaction conditions, and reduction (ArH) was only observed for 16 (0.75%), 19 (0.64%), and 22 (0.52%) (see the Supporting Information). (24)
Table 3. Regioselective Fluorinationa,b,c
Table a

Isolated yields are reported as an average of two runs. Reaction conditions unless otherwise noted: Aryl triflate (1 mmol), CsF (3 mmol), cy (10 mL).

Table b

Aryl bromide (1 mmol), KF (0.5 mmol), AgF (2.0 mmol), cy (10 mL).

Table c

0.5 mmol scale.

Table d

19F NMR yield.

Table e

No regioisomer detected by 19F NMR. cy = cyclohexane.

To gain further insight into the enhanced reactivity of 1, we prepared the [LPd(Ar)Br] complexes, from the reaction of 4-(n-Bu)PhBr and CODPd(CH2TMS)2 using L1 and L2 as supporting ligands (Figure 3) (see the Supporting Information). The respective complexes were characterized in the solid state using single-crystal X-ray diffraction. The initial hypothesis from which the design of L1 derived predicts that significant lengthening of the C1′–Pd bond would be observed with the more electron-deficient supporting ligand (L1). In fact, only a slight increase in this bond length in D was apparent, as compared to that seen with C (see the Supporting Information).
By overlaying the crystal structures of C and D, a more subtle difference was, however, revealed (Figure 3c). With C, the pendant 4-(n-Bu)Ph group of the ligand is shifted toward the Pd-bound bromide to form an aromatic CH′···Br contact of 3.874(3) Å (measured from C to Br), which is in agreement with typical aromatic CH···Br interaction distances found in the solid state (3.7–4.2 Å; measured from C to Br). (25) The perfluorinated aryl group of D lacks the required hydrogen atom to achieve the same attractive interaction, giving a C···Br distance of 4.370(2) Å. The analogous [LPd(Ar)F] complexes were synthesized (E and F), and single-crystal X-ray analysis of E shows that this behavior is conserved, with E exhibiting an aromatic CH···F interaction distance of 3.387(3) Å (measured from C to F) (Figure 3d). Typical aromatic CH···F interaction distances found in the solid state range from 3.3 to 3.7 Å (measured from C to F). (25) Unfortunately, all attempts to obtain single crystals of F suitable for X-ray diffraction yielded decomposition and/or pseudocrystalline material.

Figure 3

Figure 3. (a) Schematic of [L2Pd(Ar)X] complexes C and E where L2 is highlighted in red. (b) Schematic of [L1Pd(Ar)X] complexes D and F. (c) Crystal structure overlay of C and D. C is shown in red. Hydrogen atoms and an isopropyl group are omitted for clarity. (d) Crystal structure of E highlighting the intramolecular CH···F contact. Hydrogen atoms and the n-butyl group on the Pd-bound aryl omitted for clarity. Thermal ellipsoids are shown at 50% probability.

Short CH···F contacts are frequently observed in structurally characterized Pd–F complexes. They serve to stabilize the Pd–F bond by alleviating electronic repulsions between the lone pairs on the fluoride atom and the filled d-orbitals of Pd. (8) As shown in the solid state, E achieves this Pd–F bond stabilization intramolecularly while F cannot. The lack of an intramolecular CH···F interaction in F results in ground-state destabilization relative to E, which may account for the observed enhancement in reactivity.
Density functional theory (DFT) calculations were carried out to compare complexes E and F and gain further insight into the stabilizing CH···F interaction that was observed in the solid-state structure of E. To reduce conformational complexity, the Pd-bound 4-(n-Bu)Ph of complexes E and F was replaced with a phenyl group in our computational study, giving complexes E′ and F′, respectively. The calculated ground-state structure of E′ is in good agreement with the crystallographic data of E, exhibiting an attractive CH···F contact (3.24 Å, measured from C to F) (Figure 4a). Furthermore, NPA (natural population analysis) charge calculations of complex E′ showed that the Pd–bound fluorine atom bears a partial negative charge of −0.700 and the hydrogen atom at the 2 position of the pendant 4-(n-Bu)Ph group carries a partial positive charge of +0.251. Taken together, these analyses clearly indicate the presence of an intramoleuclar stabilizing interaction. As anticipated, this interaction is absent in complex F′, which lacks the necessary hydrogen atom required for this attractive interaction (Figure 4b). In addition, NPA charge analysis of complex F′ reveals that both the Pd-Bound fluorine atom and the fluorine atom at the 2 position of the pendant 4-(n-Bu)Ph group carry a partial negative charge (−0.690 and −0.318, respectively), further demonstrating the lack of an attractive interaction in complex F′.

Figure 4

Figure 4. Computed ground-state structures of complexes E′ and F′. Geometries were optimized at the B3LYP/6-31G(d)-SDD(Pd) level of theory. Hydrogen atoms are omitted for clarity. (a) The CH···F interaction of complex E′ is 3.24 Å, measured from C to F. (b) Complex F′ showing the lack of an intermolecular stabilizing interaction.

Additionally, we performed DFT calculations of the C–F bond-forming reductive elimination step for complexes E′ and F′ (Figure 5). This process involves a three-membered transition state where the C–F bond is being formed while the Pd–F and the Pd–C bonds are being broken in a concerted fashion. The calculations showed that the C–F reductive elimination of complex F′ (Figure 5b) has an activation barrier (ΔG) 0.7 kcal/mol lower than that of complex E′ (Figure 5a), which is in agreement with the experimentally observed enhanced reactivity of the L1-based catalyst.

Figure 5

Figure 5. Energies were computed at the M06/6-311+G(d,p)-SDD(Pd)/SMD(toluene) level of theory with geometries optimized at the B3LYP/6-31G(d)-SDD(Pd) level. Hydrogen atoms are omitted for clarity. (a) Transition state minimization of complex E′. (b) Transition state minimization of complex F′.

Conclusion

Click to copy section linkSection link copied!

A new ligand (L1) that allows for room-temperature C–F reductive elimination from [LPd(Ar)F] has been developed. The enhanced reactivity was exploited to provide aryl fluorides from aryl triflates and bromides with greater than 100:1 selectivity for the desired regioisomer. Crystallographic analysis of several Pd complexes revealed that L2 provides a stabilizing intramolecular aromatic CH···F interaction, which is not accessible with L1, giving rise to a more active catalyst system. Finally, DFT calculations were performed comparing complexes E′ and F′ corroborating the observations made in the solid state. This study provides insights into ligand structure, which should aid in the discovery and design of more effective catalysts for C–F bond formation.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.5b09308.

  • Additional procedural, crystallographic, and spectral data (PDF)

  • X-ray data for six complexes (CIF)

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.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • Stephen L. Buchwald - Department of Chemistry, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United States Email: [email protected]
  • Authors
    • Aaron C. Sather - Department of Chemistry, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United States
    • Hong Geun Lee - Department of Chemistry, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United States
    • Valentina Y. De La Rosa - Department of Chemistry, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United States
    • Yang Yang - Department of Chemistry, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United StatesDepartment of Chemistry, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, United States
    • Peter Müller - Department of Chemistry, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United States
  • Notes
    The authors declare the following competing financial interest(s): MIT has obtained patents for the ligands/precatalysts that are described in this Article from which S.L.B. and former/current co-workers receive royalty payments.

Acknowledgment

Click to copy section linkSection link copied!

This work is dedicated to the grandfather of A.C.S. (Albert B. Sather; AlPhos, L1). We thank Drs. Yiming Wang, Michael Pirnot, and John Nguyen for assistance with the preparation of the manuscript. We thank the National Institutes of Health (NIH) for financial support of this research (R01GM46059). A.C.S. thanks the NIH for a postdoctoral fellowship (1F32GM108092-01A1). The content is solely the responsibility of the authors and does not necessarily represent the official views of the NIH. V.Y.D.L.R. thanks the MIT UROP program. We thank Prof. Peng Liu (University of Pittsburgh) for help with computational studies. Calculations were performed at the Center for Simulation and Modeling at the University of Pittsburgh. The X-ray diffractometer was purchased with the help of funding from the National Science Foundation (CHE 0946721).

References

Click to copy section linkSection link copied!

This article references 25 other publications.

  1. 1
    (a) Campbell, M. G.; Ritter, T. Chem. Rev. 2015, 115, 612 DOI: 10.1021/cr500366b
    (b) Purser, S.; Moore, P. R.; Swallow, S.; Gouverneur, V. Chem. Soc. Rev. 2008, 37, 320 DOI: 10.1039/B610213C
    (c) Neumann, C. N.; Ritter, T. Angew. Chem., Int. Ed. 2015, 54, 3216 DOI: 10.1002/anie.201410288
  2. 2
    (a) Kirk, K. L. Org. Process Res. Dev. 2008, 12, 305 DOI: 10.1021/op700134j
    (b) Müller, K.; Faeh, C.; Diederich, F. Science 2007, 317, 1881 DOI: 10.1126/science.1131943
    (c) Smart, B. E. J. Fluorine Chem. 2001, 109, 3 DOI: 10.1016/S0022-1139(01)00375-X
    (d) Kirsch, P. Modern Fluoroorganic Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: New York, 2013; p 299.
  3. 3
    (a) Jeschke, P. ChemBioChem 2004, 5, 570 DOI: 10.1002/cbic.200300833
    (b) Fujiwara, T.; O’Hagan, D. J. Fluorine Chem. 2014, 167, 16 DOI: 10.1016/j.jfluchem.2014.06.014
  4. 4
    (a) Balz, G.; Schiemann, G. Ber. Dtsch. Chem. Ges. B 1927, 60, 1186 DOI: 10.1002/cber.19270600539
    (b) Kirk, K. L.; Cohen, L. A. J. Am. Chem. Soc. 1973, 95, 4619 DOI: 10.1021/ja00795a026
    (c) Cresswell, A. J.; Davies, S. G.; Roberts, P. M.; Thomson, J. E. Chem. Rev. 2015, 115, 566 DOI: 10.1021/cr5001805
  5. 5
    Finger, G. C.; Kruse, C. W. J. Am. Chem. Soc. 1956, 78, 6034 DOI: 10.1021/ja01604a022
  6. 6
    (a) Fujimoto, T.; Becker, F.; Ritter, T. Org. Process Res. Dev. 2014, 18, 1041 DOI: 10.1021/op500121w
    (b) Sun, H.; DiMagno, S. G. Angew. Chem., Int. Ed. 2006, 45, 2720 DOI: 10.1002/anie.200504555
    (c) Allen, L. J.; Muhuhi, J. M.; Bland, D. C.; Merzel, R.; Sanford, M. S. J. Org. Chem. 2014, 79, 5827 DOI: 10.1021/jo5003054
    (d) Anbarasan, P.; Neumann, H.; Beller, M. Angew. Chem., Int. Ed. 2010, 49, 2219 DOI: 10.1002/anie.200905855
  7. 7
    (a) Tang, P.; Furuya, T.; Ritter, T. J. Am. Chem. Soc. 2010, 132, 12150 DOI: 10.1021/ja105834t
    (b) Furuya, T.; Kaiser, H. M.; Ritter, T. Angew. Chem., Int. Ed. 2008, 47, 5993 DOI: 10.1002/anie.200802164
    (c) Hollingworth, C.; Gouverneur, V. Chem. Commun. 2012, 48, 2929 DOI: 10.1039/c2cc16158c
    (d) Fier, P. S.; Hartwig, J. F. J. Am. Chem. Soc. 2012, 134, 10795 DOI: 10.1021/ja304410x
    (e) Ichiishi, N.; Brooks, A. F.; Topczewski, J. J.; Rodnick, M. E.; Sanford, M. S.; Scott, P. J. H. Org. Lett. 2014, 16, 3224 DOI: 10.1021/ol501243g
    (f) Casitas, A.; Canta, M.; Solà, M.; Costas, M.; Ribas, X. J. Am. Chem. Soc. 2011, 133, 19386 DOI: 10.1021/ja2058567
    (g) Mu, X.; Zhang, H.; Chen, P.; Liu, G. Chem. Sci. 2014, 5, 275 DOI: 10.1039/C3SC51876K
    (h) Truong, T.; Klimovica, K.; Daugulis, O. J. Am. Chem. Soc. 2013, 135, 9342 DOI: 10.1021/ja4047125
    (i) Ye, Y.; Schimler, S. D.; Hanley, P. S.; Sanford, M. S. J. Am. Chem. Soc. 2013, 135, 16292 DOI: 10.1021/ja408607r
    (j) Wannberg, J.; Wallinder, C.; Ünlüsoy, M.; Sköld, C.; Larhed, M. J. Org. Chem. 2013, 78, 4184 DOI: 10.1021/jo400255m
    (k) Noël, T.; Maimone, T. J.; Buchwald, S. L. Angew. Chem., Int. Ed. 2011, 50, 8900 DOI: 10.1002/anie.201104652
  8. 8
    (a) Grushin, V. V. Acc. Chem. Res. 2010, 43, 160 DOI: 10.1021/ar9001763
    (b) Grushin, V. V.; Marshall, W. J. Organometallics 2007, 26, 4997 DOI: 10.1021/om700469k
    (c) Grushin, V. V. Chem. - Eur. J. 2002, 8, 1006 DOI: 10.1002/1521-3765(20020301)8:5<1006::AID-CHEM1006>3.0.CO;2-M
  9. 9
    Yandulov, D. V.; Tran, N. T. J. Am. Chem. Soc. 2007, 129, 1342 DOI: 10.1021/ja066930l
  10. 10
    (a) Wang, X.; Mei, T.-S.; Yu, J.-Q. J. Am. Chem. Soc. 2009, 131, 7520 DOI: 10.1021/ja901352k
    (b) Mazzotti, A. R.; Campbell, M. G.; Tang, P.; Murphy, J. M.; Ritter, T. J. Am. Chem. Soc. 2013, 135, 14012 DOI: 10.1021/ja405919z
    (c) Hull, K. L.; Anani, W. Q.; Sanford, M. S. J. Am. Chem. Soc. 2006, 128, 7134 DOI: 10.1021/ja061943k
    (d) Chan, K. S. L.; Wasa, M.; Wang, X.; Yu, J.-Q. Angew. Chem., Int. Ed. 2011, 50, 9081 DOI: 10.1002/anie.201102985
    (e) Pérez-Temprano, M. H.; Racowski, J. M.; Kampf, J. W.; Sanford, M. S. J. Am. Chem. Soc. 2014, 136, 4097 DOI: 10.1021/ja411433f
    (f) Ball, N. D.; Sanford, M. S. J. Am. Chem. Soc. 2009, 131, 3796 DOI: 10.1021/ja8054595
    (g) Furuya, T.; Benitez, D.; Tkatchouk, E.; Strom, A. E.; Tang, P.; Goddard, W. A.; Ritter, T. J. Am. Chem. Soc. 2010, 132, 3793 DOI: 10.1021/ja909371t
    (h) Ding, Q.; Ye, C.; Pu, S.; Cao, B. Tetrahedron 2014, 70, 409 DOI: 10.1016/j.tet.2013.11.034
    (i) Lou, S.-J.; Xu, D.-Q.; Xia, A.-B.; Wang, Y.-F.; Liu, Y.-K.; Du, X.-H.; Xu, Z.-Y. Chem. Commun. 2013, 49, 6218 DOI: 10.1039/c3cc42220h
  11. 11
    Watson, D. A.; Su, M.; Teverovskiy, G.; Zhang, Y.; García-Fortanet, J.; Kinzel, T.; Buchwald, S. L. Science 2009, 325, 1661 DOI: 10.1126/science.1178239
  12. 12
    Roy, A. H.; Hartwig, J. F. J. Am. Chem. Soc. 2001, 123, 1232 DOI: 10.1021/ja0034592
  13. 13
    (a) Maimone, T. J.; Milner, P. J.; Kinzel, T.; Zhang, Y.; Takase, M. K.; Buchwald, S. L. J. Am. Chem. Soc. 2011, 133, 18106 DOI: 10.1021/ja208461k
    (b) Milner, P. J.; Maimone, T. J.; Su, M.; Chen, J.; Müller, P.; Buchwald, S. L. J. Am. Chem. Soc. 2012, 134, 19922 DOI: 10.1021/ja310351e
    (c) Lee, H. G.; Milner, P. J.; Buchwald, S. L. J. Am. Chem. Soc. 2014, 136, 3792 DOI: 10.1021/ja5009739
  14. 14
    Milner, P. J.; Kinzel, T.; Zhang, Y.; Buchwald, S. L. J. Am. Chem. Soc. 2014, 136, 15757 DOI: 10.1021/ja509144r
  15. 15
    Barder, T. E.; Biscoe, M. R.; Buchwald, S. L. Organometallics 2007, 26, 2183 DOI: 10.1021/om0701017
  16. 16
    (a) Hoffmann, R. In IUPAC. Frontiers of Chemistry; Laidler, K. J., Ed.; Pergamon Press: Oxford, 1982; p 247.
    (b) Tatsumi, K.; Hoffmann, R.; Yamamoto, A.; Stille, J. K. Bull. Chem. Soc. Jpn. 1981, 54, 1857 DOI: 10.1246/bcsj.54.1857
  17. 17
    (a) Yagupol’skii, L. M.; Ya Il’chenko, A.; Kondratenko, N. V. Russ. Chem. Rev. 1974, 43, 32 DOI: 10.1070/RC1974v043n01ABEH001787
    (b) Hansch, C.; Leo, A.; Taft, R. W. Chem. Rev. 1991, 91, 165 DOI: 10.1021/cr00002a004
  18. 18
    (a) Lee, H. G.; Milner, P. J.; Colvin, M. T.; Andreas, L.; Buchwald, S. L. Inorg. Chim. Acta 2014, 422, 188 DOI: 10.1016/j.ica.2014.06.008
    (b) Lee, H. G.; Milner, P. J.; Buchwald, S. L. Org. Lett. 2013, 15, 5602 DOI: 10.1021/ol402859k
  19. 19

    The unreacted ligand could be reisolated from the reaction mixture, allowing 1 to be prepared on the gram scale without losing appreciable amounts of L1 (see the Supporting Information).

  20. 20
    (a) Tschan, M. J. L.; García-Suárez, E. J.; Freixa, Z.; Launay, H.; Hagen, H.; Benet-Buchholz, J.; van Leeuwen, P. W. N. M. J. Am. Chem. Soc. 2010, 132, 6463 DOI: 10.1021/ja100521m
    (b) Barder, T. E.; Walker, S. D.; Martinelli, J. R.; Buchwald, S. L. J. Am. Chem. Soc. 2005, 127, 4685 DOI: 10.1021/ja042491j
    (c) Walker, S. D.; Barder, T. E.; Martinelli, J. R.; Buchwald, S. L. Angew. Chem., Int. Ed. 2004, 43, 1871 DOI: 10.1002/anie.200353615
    (d) Andreu, M. G.; Zapf, A.; Beller, M. Chem. Commun. 2000, 2475 DOI: 10.1039/b006791l
  21. 21
    Baranano, D.; Hartwig, J. F. J. Am. Chem. Soc. 1995, 117, 2937 DOI: 10.1021/ja00115a033
  22. 22

    [(L1Pd)2·COD] (1) and [(L2Pd)2·COD] were compared as precatalysts for the conversion of an aryl bromide (4-bromovalerophenone) and an aryl triflate (4-pentanoylphenyl trifluoromethanesulfonate) to the corresponding aryl fluoride (4-fluorovalerophenone) at room temperature. The reactions were analyzed after 24 h, and these experiments revealed that the use of 1 catalyzes the transformation approximately 3 times faster than the use of [(L2Pd)2·COD] (see the Supporting Information).

  23. 23
    Huang, S.-M. Nature 2009, 461, 614 DOI: 10.1038/nature08356
  24. 24

    The reduction content (ArH) of 2, 21, and 10 was not determined.

  25. 25
    Bissantz, C.; Kuhn, B.; Stahl, M. J. Med. Chem. 2010, 53, 5061 DOI: 10.1021/jm100112j

Cited By

Click to copy section linkSection link copied!

This article is cited by 102 publications.

  1. Tian Tian, Myuto Kashihara, Weidan Yan, Yasushi Nishihara. Palladium-Catalyzed Decarbonylative Nucleophilic Halogenation of Acyl Fluorides and Chlorides: Synthesis of Aryl Halides via Reductive Elimination of the C–X (X = I, Br, and Cl) Bond and Mechanistic Implications. ACS Catalysis 2024, 14 (16) , 11905-11917. https://doi.org/10.1021/acscatal.4c03731
  2. Tengxiao Ji, Yang Cao, Daniel Lin, Monika Stolar, Curtis P. Berlinguette. Molecular Interlayer with Large Cations Supports Efficient, Stable Perovskite Solar Cells. ACS Applied Energy Materials 2024, 7 (13) , 5371-5378. https://doi.org/10.1021/acsaem.4c00344
  3. Alejandra Pita-Milleiro, Macarena G. Alférez, Juan J. Moreno, María F. Espada, Celia Maya, Jesús Campos. Unveiling the Latent Reactivity of Cp* Ligands (C5Me5–) toward Carbon Nucleophiles on an Iridium Complex. Inorganic Chemistry 2023, 62 (15) , 5961-5971. https://doi.org/10.1021/acs.inorgchem.2c04381
  4. Jun-Qi Zhang, Dandan Hu, Jiali Wang, Bukuo Ni, Hongjun Ren. Bimetallic Metal–Organic Coordination Polymers Facilitated the Selective C–F Cleavage of Polyfluoroarenes. Organic Letters 2022, 24 (43) , 7905-7911. https://doi.org/10.1021/acs.orglett.2c02918
  5. Wen-Jun Yue, Ruben Martin. Ni-Catalyzed Site-Selective Hydrofluoroalkylation of Terminal and Internal Olefins. ACS Catalysis 2022, 12 (19) , 12132-12137. https://doi.org/10.1021/acscatal.2c04412
  6. Oriol Planas, Vytautas Peciukenas, Markus Leutzsch, Nils Nöthling, Dimitrios A. Pantazis, Josep Cornella. Mechanism of the Aryl–F Bond-Forming Step from Bi(V) Fluorides. Journal of the American Chemical Society 2022, 144 (32) , 14489-14504. https://doi.org/10.1021/jacs.2c01072
  7. Su Yong Go, Hyunho Chung, Samuel Jaeho Shin, Sohee An, Ju Hyun Youn, Tae Yeong Im, Ji Yong Kim, Taek Dong Chung, Hong Geun Lee. A Unified Synthetic Strategy to Introduce Heteroatoms via Electrochemical Functionalization of Alkyl Organoboron Reagents. Journal of the American Chemical Society 2022, 144 (20) , 9149-9160. https://doi.org/10.1021/jacs.2c03213
  8. Zhuyong Zhang, Junfei Luo, Hongyin Gao. Rapid Access to Fluorinated Anilides via DAST-Mediated Deoxyfluorination of Arylhydroxylamines. Organic Letters 2021, 23 (23) , 9332-9336. https://doi.org/10.1021/acs.orglett.1c03779
  9. Ryan P. King, Shane W. Krska, Stephen L. Buchwald. A Neophyl Palladacycle as an Air- and Thermally Stable Precursor to Oxidative Addition Complexes. Organic Letters 2021, 23 (20) , 7927-7932. https://doi.org/10.1021/acs.orglett.1c02307
  10. Zihang Qiu, Chao-Jun Li. Transformations of Less-Activated Phenols and Phenol Derivatives via C–O Cleavage. Chemical Reviews 2020, 120 (18) , 10454-10515. https://doi.org/10.1021/acs.chemrev.0c00088
  11. Stéphane Caron. Where Does the Fluorine Come From? A Review on the Challenges Associated with the Synthesis of Organofluorine Compounds. Organic Process Research & Development 2020, 24 (4) , 470-480. https://doi.org/10.1021/acs.oprd.0c00030
  12. Huaizhi Li, Bhoomireddy Rajendra Prasad Reddy, Xihe Bi. Transformation of Alkynes into α- or β-Difluorinated Alkyl Azides by an Efficient One-Pot Two-Step Procedure. Organic Letters 2019, 21 (23) , 9358-9362. https://doi.org/10.1021/acs.orglett.9b03593
  13. Elizabeth A. Meucci, Alireza Ariafard, Allan J. Canty, Jeff W. Kampf, Melanie S. Sanford. Aryl–Fluoride Bond-Forming Reductive Elimination from Nickel(IV) Centers. Journal of the American Chemical Society 2019, 141 (33) , 13261-13267. https://doi.org/10.1021/jacs.9b06896
  14. Aaron C. Sather, Theodore A. Martinot. Data-Rich Experimentation Enables Palladium-Catalyzed Couplings of Piperidines and Five-Membered (Hetero)aromatic Electrophiles. Organic Process Research & Development 2019, 23 (8) , 1725-1739. https://doi.org/10.1021/acs.oprd.9b00233
  15. Lorenz M. Baumgartner, Joseph M. Dennis, Nicholas A. White, Stephen L. Buchwald, Klavs F. Jensen. Use of a Droplet Platform To Optimize Pd-Catalyzed C–N Coupling Reactions Promoted by Organic Bases. Organic Process Research & Development 2019, 23 (8) , 1594-1601. https://doi.org/10.1021/acs.oprd.9b00236
  16. Juan J. Moreno, María F. Espada, Jesús Campos, Joaquín López-Serrano, Stuart A. Macgregor, Ernesto Carmona. Base-Promoted, Remote C–H Activation at a Cationic (η5-C5Me5)Ir(III) Center Involving Reversible C–C Bond Formation of Bound C5Me5. Journal of the American Chemical Society 2019, 141 (6) , 2205-2210. https://doi.org/10.1021/jacs.8b11752
  17. Sydonie D. Schimler, Robert D. J. Froese, Douglas C. Bland, Melanie S. Sanford. Reactions of Arylsulfonate Electrophiles with NMe4F: Mechanistic Insight, Reactivity, and Scope. The Journal of Organic Chemistry 2018, 83 (18) , 11178-11190. https://doi.org/10.1021/acs.joc.8b01762
  18. Joseph M. Dennis, Nicholas A. White, Richard Y. Liu, Stephen L. Buchwald. Breaking the Base Barrier: An Electron-Deficient Palladium Catalyst Enables the Use of a Common Soluble Base in C–N Coupling. Journal of the American Chemical Society 2018, 140 (13) , 4721-4725. https://doi.org/10.1021/jacs.8b01696
  19. Jason C. Mixdorf, Alexandre M. Sorlin, Qi Zhang, and Hien M. Nguyen . Asymmetric Synthesis of Allylic Fluorides via Fluorination of Racemic Allylic Trichloroacetimidates Catalyzed by a Chiral Diene-Iridium Complex. ACS Catalysis 2018, 8 (2) , 790-801. https://doi.org/10.1021/acscatal.7b03786
  20. Constanze N. Neumann and Tobias Ritter . Facile C–F Bond Formation through a Concerted Nucleophilic Aromatic Substitution Mediated by the PhenoFluor Reagent. Accounts of Chemical Research 2017, 50 (11) , 2822-2833. https://doi.org/10.1021/acs.accounts.7b00413
  21. Stefan Kraft, Kristen Ryan, and Robert B. Kargbo . Recent Advances in Asymmetric Hydrogenation of Tetrasubstituted Olefins. Journal of the American Chemical Society 2017, 139 (34) , 11630-11641. https://doi.org/10.1021/jacs.7b07188
  22. Bryan T. Ingoglia and Stephen L. Buchwald . Oxidative Addition Complexes as Precatalysts for Cross-Coupling Reactions Requiring Extremely Bulky Biarylphosphine Ligands. Organic Letters 2017, 19 (11) , 2853-2856. https://doi.org/10.1021/acs.orglett.7b01082
  23. Jaeok Park, Chun Yuen Leung, Alexios N. Matralis, Cyrus M. Lacbay, Michail Tsakos, Guillermo Fernandez De Troconiz, Albert M. Berghuis, and Youla S. Tsantrizos . Pharmacophore Mapping of Thienopyrimidine-Based Monophosphonate (ThP-MP) Inhibitors of the Human Farnesyl Pyrophosphate Synthase. Journal of Medicinal Chemistry 2017, 60 (5) , 2119-2134. https://doi.org/10.1021/acs.jmedchem.6b01888
  24. Sydonie D. Schimler, Megan A. Cismesia, Patrick S. Hanley, Robert D. J. Froese, Matthew J. Jansma, Douglas C. Bland, and Melanie S. Sanford . Nucleophilic Deoxyfluorination of Phenols via Aryl Fluorosulfonate Intermediates. Journal of the American Chemical Society 2017, 139 (4) , 1452-1455. https://doi.org/10.1021/jacs.6b12911
  25. Aaron C. Sather and Stephen L. Buchwald . The Evolution of Pd0/PdII-Catalyzed Aromatic Fluorination. Accounts of Chemical Research 2016, 49 (10) , 2146-2157. https://doi.org/10.1021/acs.accounts.6b00247
  26. Nathália F. Carvalho and Josefredo R. Pliego, Jr. . Theoretical Design and Calculation of a Crown Ether Phase-Transfer-Catalyst Scaffold for Nucleophilic Fluorination Merging Two Catalytic Concepts. The Journal of Organic Chemistry 2016, 81 (18) , 8455-8463. https://doi.org/10.1021/acs.joc.6b01624
  27. David A. Petrone, Juntao Ye, and Mark Lautens . Modern Transition-Metal-Catalyzed Carbon–Halogen Bond Formation. Chemical Reviews 2016, 116 (14) , 8003-8104. https://doi.org/10.1021/acs.chemrev.6b00089
  28. Greta Utecht‐Jarzyńska, Shicheng Shi, Pengcheng Gao, Szymon Jarzyński, Md. Mahbubur Rahman, Roger Lalancette, Roman Szostak, Michal Szostak. IPr* F – Highly Hindered, Fluorinated N‐Heterocyclic Carbenes. Chemistry – A European Journal 2024, 30 (64) https://doi.org/10.1002/chem.202402847
  29. Josefredo R. Pliego. Theoretical design of new ligands to boost reaction rate and selectivity in palladium‐catalyzed aromatic fluorination. Journal of Computational Chemistry 2024, 28 https://doi.org/10.1002/jcc.27513
  30. V. A. Litvinova, A. S. Tikhomirov, A. E. Shchekotikhin. Methods for functionalization of anthraquinones. Russian Chemical Reviews 2024, 93 (10) , RCR5141. https://doi.org/10.59761/RCR5141
  31. Jie Han, Zhan‐Ming Zhang. (Radio)Fluorination Reactions for the Synthesis of Aryl Fluorides: Recent Advances and Perspectives. European Journal of Organic Chemistry 2024, 27 (32) https://doi.org/10.1002/ejoc.202400502
  32. Josefredo R. Pliego. Elucidating the Origin of Regioselectivity in Palladium-Catalyzed Aromatic Fluorination: Mechanistic Investigation and Microkinetic Analysis. Catalysis Letters 2024, 154 (8) , 4546-4555. https://doi.org/10.1007/s10562-024-04706-x
  33. . Diamondoids in Catalysis. 2024, 213-237. https://doi.org/10.1002/9783527812950.ch8
  34. Benita Kapuku, D. Scott Bohle. Synthesis and Photolysis Properties of a New Chloroquine Photoaffinity Probe. Molecules 2024, 29 (5) , 1084. https://doi.org/10.3390/molecules29051084
  35. Skyler Markham, Debbie C. Crans, Bruce Atwater. Using Speciation to Gain Insight into Sustainable Coupling Reactions and Their Catalysts. 2024, 559-576. https://doi.org/10.1002/9781119870647.ch26
  36. Su Yong Go. Introduction of Heteroatoms to Alkyl Carbocations Generated from Alkylboron Reagents via Electrochemical Activation. 2024, 67-169. https://doi.org/10.1007/978-981-99-8994-2_4
  37. Su Yong Go. Recent Achievements of C(sp3)‒Heteroatom Bond Formation in Electroorganic Synthesis and History of C(sp3)‒B Bond Activation. 2024, 55-65. https://doi.org/10.1007/978-981-99-8994-2_3
  38. Amardeep Singh, Haritha Sindhe, Akshay Kamble, Karthikeyan Rajkumar, Vanshika Agrawal, Satyasheel Sharma. Recent advances in late-stage monofluorination of natural products and their derivatives. Organic Chemistry Frontiers 2024, 1830 https://doi.org/10.1039/D4QO01111B
  39. Tongyu Huo, Xinyi Zhao, Zengrui Cheng, Jialiang Wei, Minghui Zhu, Xiaodong Dou, Ning Jiao. Late-stage modification of bioactive compounds: Improving druggability through efficient molecular editing. Acta Pharmaceutica Sinica B 2023, 356 https://doi.org/10.1016/j.apsb.2023.11.021
  40. Annelise S. Melo, Marcelo S. Valle, Josefredo R. Pliego. Accelerating SN2 nucleophilic fluorination of primary alkyl bromide using an 18-crown-6 and fluorinated bulky alcohol phase transfer system. Journal of Fluorine Chemistry 2023, 269 , 110146. https://doi.org/10.1016/j.jfluchem.2023.110146
  41. Wei‐Chieh Chang, Helena Randel, Thomas Weyhermüller, Alexander A. Auer, Christophe Farès, Christophe Werlé. Ein Kooperatives Rhodium/Sekundäres Phosphinoxid [Rh/P(O) n Bu 2 ]‐Template zur Katalytischen Hydrodefluorierung von Perfluoroarenen. Angewandte Chemie 2023, 135 (23) https://doi.org/10.1002/ange.202219127
  42. Wei‐Chieh Chang, Helena Randel, Thomas Weyhermüller, Alexander A. Auer, Christophe Farès, Christophe Werlé. A Cooperative Rhodium/Secondary Phosphine Oxide [Rh/P(O) n Bu 2 ] Template for Catalytic Hydrodefluorination of Perfluoroarenes. Angewandte Chemie International Edition 2023, 62 (23) https://doi.org/10.1002/anie.202219127
  43. Yuxuan Ye, Seoung‐Tae Kim, Ryan P. King, Mu‐Hyun Baik, Stephen L. Buchwald. Studying Regioisomer Formation in the Pd‐Catalyzed Fluorination of Cyclic Vinyl Triflates: Evidence for in situ Ligand Modification**. Angewandte Chemie 2023, 135 (15) https://doi.org/10.1002/ange.202300109
  44. Yuxuan Ye, Seoung‐Tae Kim, Ryan P. King, Mu‐Hyun Baik, Stephen L. Buchwald. Studying Regioisomer Formation in the Pd‐Catalyzed Fluorination of Cyclic Vinyl Triflates: Evidence for in situ Ligand Modification**. Angewandte Chemie International Edition 2023, 62 (15) https://doi.org/10.1002/anie.202300109
  45. Simon L. Rössler, Nathalie M. Grob, Stephen L. Buchwald, Bradley L. Pentelute. Abiotic peptides as carriers of information for the encoding of small-molecule library synthesis. Science 2023, 379 (6635) , 939-945. https://doi.org/10.1126/science.adf1354
  46. Akifumi Sumida, Kenta Ogawa, Hiroaki Imoto, Kensuke Naka. Steric and electronic effects of arsa-Buchwald ligands on Suzuki–Miyaura coupling reaction. Dalton Transactions 2023, 52 (9) , 2838-2844. https://doi.org/10.1039/D2DT04139A
  47. Alexander Düfert. Übergangsmetallkatalysierte Kupplungsreaktionen. 2023, 615-751. https://doi.org/10.1007/978-3-662-65244-2_6
  48. Y. Nishimoto, M. Yasuda. Synthesis of Fluorides. 2023https://doi.org/10.1016/B978-0-323-96025-0.00002-8
  49. Bing‐Bing Wu, Jie Xu, Qian Gao, Kang‐Jie Bian, Guo‐Kai Liu, Xi‐Sheng Wang. A General and Efficient Solution to Monofluoroalkylation: Divergent Synthesis of Aliphatic Monofluorides with Modular Synthetic Scaffolds. Angewandte Chemie 2022, 134 (36) https://doi.org/10.1002/ange.202208938
  50. Bing‐Bing Wu, Jie Xu, Qian Gao, Kang‐Jie Bian, Guo‐Kai Liu, Xi‐Sheng Wang. A General and Efficient Solution to Monofluoroalkylation: Divergent Synthesis of Aliphatic Monofluorides with Modular Synthetic Scaffolds. Angewandte Chemie International Edition 2022, 61 (36) https://doi.org/10.1002/anie.202208938
  51. Renyu Guo, Brittany P. Witherspoon, M. Kevin Brown. Stereoselective [2+2]-Cycloadditions of chiral allenic ketones and alkenes: Applications towards the synthesis of benzocyclobutenes and endiandric acids. Tetrahedron 2022, 122 , 132932. https://doi.org/10.1016/j.tet.2022.132932
  52. Jiali Cai, Minna Zhi, Junyuan Hu, Tingting Pu, Kai Guo, Lili Zhao. Mechanistic study of the bismuth mediated fluorination of arylboronic esters and further rational design. RSC Advances 2022, 12 (37) , 24208-24216. https://doi.org/10.1039/D2RA04296G
  53. Diogo L. Poeira, João Macara, M. Manuel B. Marques. Pyrans and Their Benzo Derivatives: Structure and Reactivity. 2022, 243-328. https://doi.org/10.1016/B978-0-12-818655-8.00055-X
  54. Juan José Moreno, María F. Espada, Celia Maya, Jesús Campos, Joaquín López-Serrano, Stuart A. Macgregor, Ernesto Carmona. Isomerization of a cationic (η5-C5Me5)Ir(III) complex involving remote C–C and C–H bond formation. Polyhedron 2021, 207 , 115363. https://doi.org/10.1016/j.poly.2021.115363
  55. Ayane Suto, Junichiro Yamaguchi. Fluorination -A Decade of Progress (2010-2020). Journal of Synthetic Organic Chemistry, Japan 2021, 79 (10) , 910-967. https://doi.org/10.5059/yukigoseikyokaishi.79.910
  56. Tien Tan Bui, Hee‐Kwon Kim. Recent Advances in Photo‐mediated Radiofluorination. Chemistry – An Asian Journal 2021, 16 (16) , 2155-2167. https://doi.org/10.1002/asia.202100399
  57. Jie Sheng, Hui‐Qi Ni, Shan‐Xiu Ni, Yan He, Ru Cui, Guang‐Xu Liao, Kang‐Jie Bian, Bing‐Bing Wu, Xi‐Sheng Wang. Diversity‐Oriented Synthesis of Aliphatic Fluorides via Reductive C(sp 3 )−C(sp 3 ) Cross‐Coupling Fluoroalkylation. Angewandte Chemie 2021, 133 (27) , 15147-15154. https://doi.org/10.1002/ange.202102481
  58. Jie Sheng, Hui‐Qi Ni, Shan‐Xiu Ni, Yan He, Ru Cui, Guang‐Xu Liao, Kang‐Jie Bian, Bing‐Bing Wu, Xi‐Sheng Wang. Diversity‐Oriented Synthesis of Aliphatic Fluorides via Reductive C(sp 3 )−C(sp 3 ) Cross‐Coupling Fluoroalkylation. Angewandte Chemie International Edition 2021, 60 (27) , 15020-15027. https://doi.org/10.1002/anie.202102481
  59. Liang‐Neng Wang, Pan‐Ting Tang, Ming Li, Jia‐Wei Li, Yue‐Jin Liu, Ming‐Hua Zeng. Double Ligands Enabled Ruthenium Catalyzed ortho ‐C−H Arylation of Dialkyl Biarylphosphines: Straight and Economic Synthesis of Highly Steric and Electron‐Rich Aryl‐Substituted Buchwald‐Type Phosphines. Advanced Synthesis & Catalysis 2021, 363 (11) , 2843-2849. https://doi.org/10.1002/adsc.202100283
  60. Josefredo R. Pliego. Catalytic cycle and off-cycle steps in the palladium-catalyzed fluorination of aryl bromide with biaryl monophosphine ligands: Theoretical free energy profile. Molecular Catalysis 2021, 506 , 111540. https://doi.org/10.1016/j.mcat.2021.111540
  61. Thomas R. Puleo, Stephen J. Sujansky, Shawn E. Wright, Jeffrey S. Bandar. Organic Superbases in Recent Synthetic Methodology Research. Chemistry – A European Journal 2021, 27 (13) , 4216-4229. https://doi.org/10.1002/chem.202003580
  62. Yueqian Sang, Biying Zhou, Meng‐Meng Zheng, Xiao‐Song Xue, Jin‐Pei Cheng. DFT Modeling of Catalytic Fluorination Reactions: Mechanisms, Reactivities, and Selectivities. 2021, 307-362. https://doi.org/10.1002/9783527825158.ch10
  63. Constanze N. Neumann, Tobias Ritter, . Transition‐Metal‐Mediated and Transition‐Metal‐Catalyzed Carbon–Fluorine Bond Formation. 2020, 1-181. https://doi.org/10.1002/0471264180.or104.03
  64. Nicholas E. S. Tay, Wei Chen, Alison Levens, Vincent A. Pistritto, Zeng Huang, Zhanhong Wu, Zibo Li, David A. Nicewicz. 19F- and 18F-arene deoxyfluorination via organic photoredox-catalysed polarity-reversed nucleophilic aromatic substitution. Nature Catalysis 2020, 3 (9) , 734-742. https://doi.org/10.1038/s41929-020-0495-0
  65. Alessandro Castrogiovanni, Dominik Lotter, Fabian R. Bissegger, Christof Sparr. JoyaPhos: An Atropisomeric Teraryl Monophosphine Ligand. Chemistry – A European Journal 2020, 26 (44) , 9864-9868. https://doi.org/10.1002/chem.202001269
  66. V.S. Subramaniam. AlPHOS, alphos. 2020https://doi.org/10.1002/9783527809080.cataz00645
  67. Ádám Sinai, Dániel Cs. Simkó, Fruzsina Szabó, Attila Paczal, Tamás Gáti, Attila Bényei, Zoltán Novák, András Kotschy. Aryl‐Diadamantyl Phosphine Ligands in Palladium‐Catalyzed Cross‐Coupling Reactions: Synthesis, Structural Analysis, and Application. European Journal of Organic Chemistry 2020, 2020 (9) , 1122-1128. https://doi.org/10.1002/ejoc.201901834
  68. Oriol Planas, Feng Wang, Markus Leutzsch, Josep Cornella. Fluorination of arylboronic esters enabled by bismuth redox catalysis. Science 2020, 367 (6475) , 313-317. https://doi.org/10.1126/science.aaz2258
  69. Gregory B. Boursalian, Tobias Ritter. Palladium-Mediated Fluorination for Preparing Aryl Fluorides. 2020, 497-513. https://doi.org/10.1007/978-981-10-3896-9_23
  70. Usisipho Feleni, Unathi Sidwaba, Nomaphelo Ntshongontshi, Lindsay Wilson, Emmanuel Iwuoha. Biocompatible Palladium Telluride Quantum Dot-Amplified Biosensor for HIV Drug. Electrocatalysis 2020, 11 (1) , 68-76. https://doi.org/10.1007/s12678-019-00563-0
  71. Yang-Jie Mao, Gen Luo, Hong-Yan Hao, Zhen-Yuan Xu, Shao-Jie Lou, Dan-Qian Xu. Anion ligand promoted selective C–F bond reductive elimination enables C(sp 2 )–H fluorination. Chemical Communications 2019, 55 (96) , 14458-14461. https://doi.org/10.1039/C9CC07726J
  72. David J. Jones, Mark Lautens, Gerard P. McGlacken. The emergence of Pd-mediated reversible oxidative addition in cross coupling, carbohalogenation and carbonylation reactions. Nature Catalysis 2019, 2 (10) , 843-851. https://doi.org/10.1038/s41929-019-0361-0
  73. Koji Kubota, Rina Takahashi, Hajime Ito. Mechanochemistry allows carrying out sensitive organometallic reactions in air: glove-box-and-Schlenk-line-free synthesis of oxidative addition complexes from aryl halides and palladium(0). Chemical Science 2019, 10 (22) , 5837-5842. https://doi.org/10.1039/C9SC01711A
  74. Kevin D. Dykstra, Naoko Ichiishi, Shane W. Krska, Paul F. Richardson. Emerging fluorination methods in organic chemistry relevant for life science application. 2019, 1-90. https://doi.org/10.1016/B978-0-12-812733-9.00001-5
  75. Xiaowei Li, Xiaolin Shi, Xiangqian Li, Dayong Shi. Recent advances in transition-metal-catalyzed incorporation of fluorine-containing groups. Beilstein Journal of Organic Chemistry 2019, 15 , 2213-2270. https://doi.org/10.3762/bjoc.15.218
  76. Tharwat Mohy El Dine, Omar Sadek, Emmanuel Gras, David M. Perrin. Expanding the Balz–Schiemann Reaction: Organotrifluoroborates Serve as Competent Sources of Fluoride Ion for Fluoro‐Dediazoniation. Chemistry – A European Journal 2018, 24 (56) , 14933-14937. https://doi.org/10.1002/chem.201803575
  77. Thorsten Scherpf, Christopher Schwarz, Lennart T. Scharf, Jana‐Alina Zur, Andreas Helbig, Viktoria H. Gessner. Ylide‐Functionalized Phosphines: Strong Donor Ligands for Homogeneous Catalysis. Angewandte Chemie 2018, 130 (39) , 13041-13046. https://doi.org/10.1002/ange.201805372
  78. Thorsten Scherpf, Christopher Schwarz, Lennart T. Scharf, Jana‐Alina Zur, Andreas Helbig, Viktoria H. Gessner. Ylide‐Functionalized Phosphines: Strong Donor Ligands for Homogeneous Catalysis. Angewandte Chemie International Edition 2018, 57 (39) , 12859-12864. https://doi.org/10.1002/anie.201805372
  79. Keisuke Ishita, Ahmed Khalil, Rohit Tiwari, Judith Gallucci, Werner Tjarks. Bis(tri‐ tert ‐butylphosphine)palladium(0)‐Catalyzed Iodine–Fluorine Exchange at closo ‐Carboranes. European Journal of Inorganic Chemistry 2018, 2018 (24) , 2821-2825. https://doi.org/10.1002/ejic.201800486
  80. Gwilherm Evano, Antoine Nitelet, Pierre Thilmany, Damien F. Dewez. Metal-Mediated Halogen Exchange in Aryl and Vinyl Halides: A Review. Frontiers in Chemistry 2018, 6 https://doi.org/10.3389/fchem.2018.00114
  81. Syo Ishida, Tom Sheppard, Takashi Nishikata. Site selectivities in fluorination. Tetrahedron Letters 2018, 59 (9) , 789-798. https://doi.org/10.1016/j.tetlet.2018.01.044
  82. Gregory B. Boursalian, Tobias Ritter. Palladium-Mediated Fluorination for Preparing Aryl Fluorides. 2018, 1-17. https://doi.org/10.1007/978-981-10-1855-8_23-1
  83. Keith G. Andrews, Radmila Faizova, Ross M. Denton. A practical and catalyst-free trifluoroethylation reaction of amines using trifluoroacetic acid. Nature Communications 2017, 8 (1) https://doi.org/10.1038/ncomms15913
  84. Xiang Zhang. Mechanistic study on silver(I)-catalyzed aminofluorination of unactivated alkenes. Journal of Physical Organic Chemistry 2017, 30 (7) , e3655. https://doi.org/10.1002/poc.3655
  85. Xiaodong Qiu, Minyan Wang, Yue Zhao, Zhuangzhi Shi. Rhodium(I)‐Catalyzed Tertiary Phosphine Directed C−H Arylation: Rapid Construction of Ligand Libraries. Angewandte Chemie 2017, 129 (25) , 7339-7343. https://doi.org/10.1002/ange.201703354
  86. Xiaodong Qiu, Minyan Wang, Yue Zhao, Zhuangzhi Shi. Rhodium(I)‐Catalyzed Tertiary Phosphine Directed C−H Arylation: Rapid Construction of Ligand Libraries. Angewandte Chemie International Edition 2017, 56 (25) , 7233-7237. https://doi.org/10.1002/anie.201703354
  87. Fui Fong Yong, Adrian M. Mak, Wenqin Wu, Michael B. Sullivan, Edward G. Robins, Charles W. Johannes, Howard Jong, Yee Hwee Lim. Empirical and Computational Insights into N‐Arylation Reactions Catalyzed by Palladium meta ‐Terarylphosphine Catalyst. ChemPlusChem 2017, 82 (5) , 750-757. https://doi.org/10.1002/cplu.201700042
  88. Michael N. Balfour, Caio H. Franco, Carolina B. Moraes, Lucio H. Freitas-Junior, Hélio A. Stefani. Synthesis and trypanocidal activity of a library of 4-substituted 2-(1H-pyrrolo[3,2-c]pyridin-2-yl)propan-2-ols. European Journal of Medicinal Chemistry 2017, 128 , 202-212. https://doi.org/10.1016/j.ejmech.2017.01.040
  89. Nilay Hazari, Patrick R. Melvin, Megan Mohadjer Beromi. Well-defined nickel and palladium precatalysts for cross-coupling. Nature Reviews Chemistry 2017, 1 (3) https://doi.org/10.1038/s41570-017-0025
  90. Julien Roger, Sylviane Royer, Hélène Cattey, Aleksandr Savateev, Radomyr V. Smaliy, Aleksandr N. Kostyuk, Jean‐Cyrille Hierso. Diastereoselective Synthesis of Dialkylated Bis(phosphino)ferrocenes: Their Use in Promoting Silver‐Mediated Nucleophilic Fluorination of Chloroquinolines. European Journal of Inorganic Chemistry 2017, 2017 (2) , 330-339. https://doi.org/10.1002/ejic.201600502
  91. Tharun T. Ponduru, H. V. Rasika Dias. Dinuclear zinc( ii ) pyrazolates with different degrees of ring-fluorination and their use in zinc( ii ) mediated olefin aziridination. Dalton Transactions 2017, 46 (43) , 14733-14737. https://doi.org/10.1039/C7DT02951A
  92. Yuxuan Ye, Takashi Takada, Stephen L. Buchwald. Palladium‐Catalyzed Fluorination of Cyclic Vinyl Triflates: Effect of TESCF 3 as an Additive. Angewandte Chemie 2016, 128 (50) , 15788-15792. https://doi.org/10.1002/ange.201608927
  93. Yuxuan Ye, Takashi Takada, Stephen L. Buchwald. Palladium‐Catalyzed Fluorination of Cyclic Vinyl Triflates: Effect of TESCF 3 as an Additive. Angewandte Chemie International Edition 2016, 55 (50) , 15559-15563. https://doi.org/10.1002/anie.201608927
  94. Hanna Baumgarth, Gregor Meier, Thomas Braun, Beatrice Braun‐Cula. Rhodium and Iridium Fluorido and Bifluorido Complexes Derived from Peroxido Precursors. European Journal of Inorganic Chemistry 2016, 2016 (28) , 4565-4572. https://doi.org/10.1002/ejic.201600682
  95. Nathaniel H. Park, Timothy J. Senter, Stephen L. Buchwald. Rapid Synthesis of Aryl Fluorides in Continuous Flow through the Balz–Schiemann Reaction. Angewandte Chemie 2016, 128 (39) , 12086-12090. https://doi.org/10.1002/ange.201606601
  96. Nathaniel H. Park, Timothy J. Senter, Stephen L. Buchwald. Rapid Synthesis of Aryl Fluorides in Continuous Flow through the Balz–Schiemann Reaction. Angewandte Chemie International Edition 2016, 55 (39) , 11907-11911. https://doi.org/10.1002/anie.201606601
  97. Takashi Nishikata, Syo Ishida, Ryo Fujimoto. Site‐Selective Tertiary Alkyl–Fluorine Bond Formation from α‐Bromoamides Using a Copper/CsF Catalyst System. Angewandte Chemie 2016, 128 (34) , 10162-10166. https://doi.org/10.1002/ange.201603426
  98. Takashi Nishikata, Syo Ishida, Ryo Fujimoto. Site‐Selective Tertiary Alkyl–Fluorine Bond Formation from α‐Bromoamides Using a Copper/CsF Catalyst System. Angewandte Chemie International Edition 2016, 55 (34) , 10008-10012. https://doi.org/10.1002/anie.201603426
  99. Xiang Zhang. Insights into the mechanism of silver-catalyzed decarboxylative fluorination. Computational and Theoretical Chemistry 2016, 1082 , 11-20. https://doi.org/10.1016/j.comptc.2016.02.016
  100. Aaron C. Sather, Hong Geun Lee, Valentina Y. De La Rosa, Yang Yang, Peter Mueller, Stephen L. Buchwald. ChemInform Abstract: A Fluorinated Ligand Enables Room‐Temperature and Regioselective Pd‐Catalyzed Fluorination of Aryl Triflates and Bromides.. ChemInform 2016, 47 (12) https://doi.org/10.1002/chin.201612051
Load all citations

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2015, 137, 41, 13433–13438
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.5b09308
Published September 28, 2015

Copyright © 2015 American Chemical Society. This publication is licensed under these Terms of Use.

Article Views

20k

Altmetric

-

Citations

Learn about these metrics

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.

  • Abstract

    Figure 1

    Figure 1. (a) The proposed catalytic cycle for Pd-catalyzed fluorination. (b) A Pd(II) complex with L2 highlighted in red. The interaction between ligand-bound Pd(II) and C1′ is shown by a dashed line. (c) Biaryl monophosphine ligand (Alphos, L1).

    Scheme 1

    Scheme 1. Synthesis of L1 and Pd(0) Precatalyst 1a

    Scheme aReagents and conditions: (a) THF, −78 °C, n-BuLi, 1 h; added to C6F6 in THF, 0 °C to rt, 2 h, 87%. (b) THF, −78 °C, n-BuLi, 30 min, 99%. (c) AcOH, EtOAc, H2SO4, N-iodosuccinimide, 50 °C, 17 h, 80%. (d) THF, −78 °C, t-BuLi, 1 h; added to 3-fluoroanisole and n-BuLi in THF, −78 to −25 °C, 44%. (e) THF, −78 °C, t-BuLi, 1 h; CuCl −78 °C to rt; Ad2PCl, toluene, 140 °C, 18 h, 82%. (f) [COD·Pd(CH2TMS)2], pentane, rt, 48 h, 78%.

    Figure 2

    Figure 2. Crystal structure of [(L1Pd)2·COD] (1). Thermal ellipsoids are shown at 50% probability; hydrogen atoms and residual benzene molecules are omitted for clarity.

    Scheme 2

    Scheme 2. Stoichiometric C–F Reductive Elimination from Pd(II) Complexes at Room Temperaturea

    Scheme aYields determined by 19F NMR. bNo ArBr added. cArBr (R = C(O)n-Bu). dArBr (R = n-Bu).

    Figure 3

    Figure 3. (a) Schematic of [L2Pd(Ar)X] complexes C and E where L2 is highlighted in red. (b) Schematic of [L1Pd(Ar)X] complexes D and F. (c) Crystal structure overlay of C and D. C is shown in red. Hydrogen atoms and an isopropyl group are omitted for clarity. (d) Crystal structure of E highlighting the intramolecular CH···F contact. Hydrogen atoms and the n-butyl group on the Pd-bound aryl omitted for clarity. Thermal ellipsoids are shown at 50% probability.

    Figure 4

    Figure 4. Computed ground-state structures of complexes E′ and F′. Geometries were optimized at the B3LYP/6-31G(d)-SDD(Pd) level of theory. Hydrogen atoms are omitted for clarity. (a) The CH···F interaction of complex E′ is 3.24 Å, measured from C to F. (b) Complex F′ showing the lack of an intermolecular stabilizing interaction.

    Figure 5

    Figure 5. Energies were computed at the M06/6-311+G(d,p)-SDD(Pd)/SMD(toluene) level of theory with geometries optimized at the B3LYP/6-31G(d)-SDD(Pd) level. Hydrogen atoms are omitted for clarity. (a) Transition state minimization of complex E′. (b) Transition state minimization of complex F′.

  • References


    This article references 25 other publications.

    1. 1
      (a) Campbell, M. G.; Ritter, T. Chem. Rev. 2015, 115, 612 DOI: 10.1021/cr500366b
      (b) Purser, S.; Moore, P. R.; Swallow, S.; Gouverneur, V. Chem. Soc. Rev. 2008, 37, 320 DOI: 10.1039/B610213C
      (c) Neumann, C. N.; Ritter, T. Angew. Chem., Int. Ed. 2015, 54, 3216 DOI: 10.1002/anie.201410288
    2. 2
      (a) Kirk, K. L. Org. Process Res. Dev. 2008, 12, 305 DOI: 10.1021/op700134j
      (b) Müller, K.; Faeh, C.; Diederich, F. Science 2007, 317, 1881 DOI: 10.1126/science.1131943
      (c) Smart, B. E. J. Fluorine Chem. 2001, 109, 3 DOI: 10.1016/S0022-1139(01)00375-X
      (d) Kirsch, P. Modern Fluoroorganic Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: New York, 2013; p 299.
    3. 3
      (a) Jeschke, P. ChemBioChem 2004, 5, 570 DOI: 10.1002/cbic.200300833
      (b) Fujiwara, T.; O’Hagan, D. J. Fluorine Chem. 2014, 167, 16 DOI: 10.1016/j.jfluchem.2014.06.014
    4. 4
      (a) Balz, G.; Schiemann, G. Ber. Dtsch. Chem. Ges. B 1927, 60, 1186 DOI: 10.1002/cber.19270600539
      (b) Kirk, K. L.; Cohen, L. A. J. Am. Chem. Soc. 1973, 95, 4619 DOI: 10.1021/ja00795a026
      (c) Cresswell, A. J.; Davies, S. G.; Roberts, P. M.; Thomson, J. E. Chem. Rev. 2015, 115, 566 DOI: 10.1021/cr5001805
    5. 5
      Finger, G. C.; Kruse, C. W. J. Am. Chem. Soc. 1956, 78, 6034 DOI: 10.1021/ja01604a022
    6. 6
      (a) Fujimoto, T.; Becker, F.; Ritter, T. Org. Process Res. Dev. 2014, 18, 1041 DOI: 10.1021/op500121w
      (b) Sun, H.; DiMagno, S. G. Angew. Chem., Int. Ed. 2006, 45, 2720 DOI: 10.1002/anie.200504555
      (c) Allen, L. J.; Muhuhi, J. M.; Bland, D. C.; Merzel, R.; Sanford, M. S. J. Org. Chem. 2014, 79, 5827 DOI: 10.1021/jo5003054
      (d) Anbarasan, P.; Neumann, H.; Beller, M. Angew. Chem., Int. Ed. 2010, 49, 2219 DOI: 10.1002/anie.200905855
    7. 7
      (a) Tang, P.; Furuya, T.; Ritter, T. J. Am. Chem. Soc. 2010, 132, 12150 DOI: 10.1021/ja105834t
      (b) Furuya, T.; Kaiser, H. M.; Ritter, T. Angew. Chem., Int. Ed. 2008, 47, 5993 DOI: 10.1002/anie.200802164
      (c) Hollingworth, C.; Gouverneur, V. Chem. Commun. 2012, 48, 2929 DOI: 10.1039/c2cc16158c
      (d) Fier, P. S.; Hartwig, J. F. J. Am. Chem. Soc. 2012, 134, 10795 DOI: 10.1021/ja304410x
      (e) Ichiishi, N.; Brooks, A. F.; Topczewski, J. J.; Rodnick, M. E.; Sanford, M. S.; Scott, P. J. H. Org. Lett. 2014, 16, 3224 DOI: 10.1021/ol501243g
      (f) Casitas, A.; Canta, M.; Solà, M.; Costas, M.; Ribas, X. J. Am. Chem. Soc. 2011, 133, 19386 DOI: 10.1021/ja2058567
      (g) Mu, X.; Zhang, H.; Chen, P.; Liu, G. Chem. Sci. 2014, 5, 275 DOI: 10.1039/C3SC51876K
      (h) Truong, T.; Klimovica, K.; Daugulis, O. J. Am. Chem. Soc. 2013, 135, 9342 DOI: 10.1021/ja4047125
      (i) Ye, Y.; Schimler, S. D.; Hanley, P. S.; Sanford, M. S. J. Am. Chem. Soc. 2013, 135, 16292 DOI: 10.1021/ja408607r
      (j) Wannberg, J.; Wallinder, C.; Ünlüsoy, M.; Sköld, C.; Larhed, M. J. Org. Chem. 2013, 78, 4184 DOI: 10.1021/jo400255m
      (k) Noël, T.; Maimone, T. J.; Buchwald, S. L. Angew. Chem., Int. Ed. 2011, 50, 8900 DOI: 10.1002/anie.201104652
    8. 8
      (a) Grushin, V. V. Acc. Chem. Res. 2010, 43, 160 DOI: 10.1021/ar9001763
      (b) Grushin, V. V.; Marshall, W. J. Organometallics 2007, 26, 4997 DOI: 10.1021/om700469k
      (c) Grushin, V. V. Chem. - Eur. J. 2002, 8, 1006 DOI: 10.1002/1521-3765(20020301)8:5<1006::AID-CHEM1006>3.0.CO;2-M
    9. 9
      Yandulov, D. V.; Tran, N. T. J. Am. Chem. Soc. 2007, 129, 1342 DOI: 10.1021/ja066930l
    10. 10
      (a) Wang, X.; Mei, T.-S.; Yu, J.-Q. J. Am. Chem. Soc. 2009, 131, 7520 DOI: 10.1021/ja901352k
      (b) Mazzotti, A. R.; Campbell, M. G.; Tang, P.; Murphy, J. M.; Ritter, T. J. Am. Chem. Soc. 2013, 135, 14012 DOI: 10.1021/ja405919z
      (c) Hull, K. L.; Anani, W. Q.; Sanford, M. S. J. Am. Chem. Soc. 2006, 128, 7134 DOI: 10.1021/ja061943k
      (d) Chan, K. S. L.; Wasa, M.; Wang, X.; Yu, J.-Q. Angew. Chem., Int. Ed. 2011, 50, 9081 DOI: 10.1002/anie.201102985
      (e) Pérez-Temprano, M. H.; Racowski, J. M.; Kampf, J. W.; Sanford, M. S. J. Am. Chem. Soc. 2014, 136, 4097 DOI: 10.1021/ja411433f
      (f) Ball, N. D.; Sanford, M. S. J. Am. Chem. Soc. 2009, 131, 3796 DOI: 10.1021/ja8054595
      (g) Furuya, T.; Benitez, D.; Tkatchouk, E.; Strom, A. E.; Tang, P.; Goddard, W. A.; Ritter, T. J. Am. Chem. Soc. 2010, 132, 3793 DOI: 10.1021/ja909371t
      (h) Ding, Q.; Ye, C.; Pu, S.; Cao, B. Tetrahedron 2014, 70, 409 DOI: 10.1016/j.tet.2013.11.034
      (i) Lou, S.-J.; Xu, D.-Q.; Xia, A.-B.; Wang, Y.-F.; Liu, Y.-K.; Du, X.-H.; Xu, Z.-Y. Chem. Commun. 2013, 49, 6218 DOI: 10.1039/c3cc42220h
    11. 11
      Watson, D. A.; Su, M.; Teverovskiy, G.; Zhang, Y.; García-Fortanet, J.; Kinzel, T.; Buchwald, S. L. Science 2009, 325, 1661 DOI: 10.1126/science.1178239
    12. 12
      Roy, A. H.; Hartwig, J. F. J. Am. Chem. Soc. 2001, 123, 1232 DOI: 10.1021/ja0034592
    13. 13
      (a) Maimone, T. J.; Milner, P. J.; Kinzel, T.; Zhang, Y.; Takase, M. K.; Buchwald, S. L. J. Am. Chem. Soc. 2011, 133, 18106 DOI: 10.1021/ja208461k
      (b) Milner, P. J.; Maimone, T. J.; Su, M.; Chen, J.; Müller, P.; Buchwald, S. L. J. Am. Chem. Soc. 2012, 134, 19922 DOI: 10.1021/ja310351e
      (c) Lee, H. G.; Milner, P. J.; Buchwald, S. L. J. Am. Chem. Soc. 2014, 136, 3792 DOI: 10.1021/ja5009739
    14. 14
      Milner, P. J.; Kinzel, T.; Zhang, Y.; Buchwald, S. L. J. Am. Chem. Soc. 2014, 136, 15757 DOI: 10.1021/ja509144r
    15. 15
      Barder, T. E.; Biscoe, M. R.; Buchwald, S. L. Organometallics 2007, 26, 2183 DOI: 10.1021/om0701017
    16. 16
      (a) Hoffmann, R. In IUPAC. Frontiers of Chemistry; Laidler, K. J., Ed.; Pergamon Press: Oxford, 1982; p 247.
      (b) Tatsumi, K.; Hoffmann, R.; Yamamoto, A.; Stille, J. K. Bull. Chem. Soc. Jpn. 1981, 54, 1857 DOI: 10.1246/bcsj.54.1857
    17. 17
      (a) Yagupol’skii, L. M.; Ya Il’chenko, A.; Kondratenko, N. V. Russ. Chem. Rev. 1974, 43, 32 DOI: 10.1070/RC1974v043n01ABEH001787
      (b) Hansch, C.; Leo, A.; Taft, R. W. Chem. Rev. 1991, 91, 165 DOI: 10.1021/cr00002a004
    18. 18
      (a) Lee, H. G.; Milner, P. J.; Colvin, M. T.; Andreas, L.; Buchwald, S. L. Inorg. Chim. Acta 2014, 422, 188 DOI: 10.1016/j.ica.2014.06.008
      (b) Lee, H. G.; Milner, P. J.; Buchwald, S. L. Org. Lett. 2013, 15, 5602 DOI: 10.1021/ol402859k
    19. 19

      The unreacted ligand could be reisolated from the reaction mixture, allowing 1 to be prepared on the gram scale without losing appreciable amounts of L1 (see the Supporting Information).

    20. 20
      (a) Tschan, M. J. L.; García-Suárez, E. J.; Freixa, Z.; Launay, H.; Hagen, H.; Benet-Buchholz, J.; van Leeuwen, P. W. N. M. J. Am. Chem. Soc. 2010, 132, 6463 DOI: 10.1021/ja100521m
      (b) Barder, T. E.; Walker, S. D.; Martinelli, J. R.; Buchwald, S. L. J. Am. Chem. Soc. 2005, 127, 4685 DOI: 10.1021/ja042491j
      (c) Walker, S. D.; Barder, T. E.; Martinelli, J. R.; Buchwald, S. L. Angew. Chem., Int. Ed. 2004, 43, 1871 DOI: 10.1002/anie.200353615
      (d) Andreu, M. G.; Zapf, A.; Beller, M. Chem. Commun. 2000, 2475 DOI: 10.1039/b006791l
    21. 21
      Baranano, D.; Hartwig, J. F. J. Am. Chem. Soc. 1995, 117, 2937 DOI: 10.1021/ja00115a033
    22. 22

      [(L1Pd)2·COD] (1) and [(L2Pd)2·COD] were compared as precatalysts for the conversion of an aryl bromide (4-bromovalerophenone) and an aryl triflate (4-pentanoylphenyl trifluoromethanesulfonate) to the corresponding aryl fluoride (4-fluorovalerophenone) at room temperature. The reactions were analyzed after 24 h, and these experiments revealed that the use of 1 catalyzes the transformation approximately 3 times faster than the use of [(L2Pd)2·COD] (see the Supporting Information).

    23. 23
      Huang, S.-M. Nature 2009, 461, 614 DOI: 10.1038/nature08356
    24. 24

      The reduction content (ArH) of 2, 21, and 10 was not determined.

    25. 25
      Bissantz, C.; Kuhn, B.; Stahl, M. J. Med. Chem. 2010, 53, 5061 DOI: 10.1021/jm100112j
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.5b09308.

    • Additional procedural, crystallographic, and spectral data (PDF)

    • X-ray data for six complexes (CIF)


    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.