Mechanistic Studies on the Copper-Catalyzed N-Arylation of

May 23, 2016 - Sameeran Kumar Das , Pangkita Deka , Monikha Chetia , Ramesh C. Deka , Pankaj Bharali , Utpal Bora. Catalysis Letters 2018 148 (2), 547...
0 downloads 0 Views 3MB Size
Research Article pubs.acs.org/acscatalysis

Mechanistic Studies on the Copper-Catalyzed N‑Arylation of Alkylamines Promoted by Organic Soluble Ionic Bases Simon Sung,† David Sale,‡ D. Christopher Braddock,† Alan Armstrong,† Colin Brennan,‡ and Robert P. Davies*,† †

Department of Chemistry, Imperial College London, South Kensington, London SW7 2AZ, United Kingdom Process Studies Group, Syngenta, Jealott’s Hill Research Centre, Bracknell, Berkshire RG42 6EY, United Kingdom



S Supporting Information *

ABSTRACT: Experimental studies on the mechanism of copper-catalyzed amination of aryl halides have been undertaken for the coupling of piperidine with iodobenzene using a Cu(I) catalyst and the organic base tetrabutylphosphonium malonate (TBPM). The use of TBPM led to high reactivity and high conversion rates in the coupling reaction, as well as obviating any mass transfer effects. The often commonly employed O,O-chelating ligand 2-acetylcyclohexanone was surprisingly found to have a negligible effect on the reaction rate, and on the basis of NMR, calorimetric, and kinetic modeling studies, the malonate dianion in TBPM is instead postulated to act as an ancillary ligand in this system. Kinetic profiling using reaction progress kinetic analysis (RPKA) methods show the reaction rate to have a dependence on all of the reaction components in the concentration range studied, with firstorder kinetics with respect to [amine], [aryl halide], and [Cu]total. Unexpectedly, negative first-order kinetics in [TBPM] was observed. This negative rate dependence in [TBPM] can be explained by the formation of an off-cycle copper(I) dimalonate species, which is also argued to undergo disproportionation and is thus responsible for catalyst deactivation. The key role of the amine in minimizing catalyst deactivation is also highlighted by the kinetic studies. An examination of the aryl halide activation mechanism using radical probes was undertaken, which is consistent with an oxidative addition pathway. On the basis of these findings, a more detailed mechanistic cycle for the C−N coupling is proposed, including catalyst deactivation pathways. KEYWORDS: copper, Ullmann Reaction, C−N bond coupling, amination, RPKA, catalyst deactivation, organic bases



INTRODUCTION

and complementary method to palladium-catalyzed Buchwald− Hartwig amination.6 As highlighted in recent reports, these advantages have allowed the modified Ullmann reaction to have a key role in the total synthesis of a number of natural products and other compounds with biological activity.7 Although computationally derived mechanisms for the copper-catalyzed C−N cross coupling between aryl halides and amines have been proposed in the literature, 8,9 experimentally obtained mechanistic information is still scarce. Nevertheless a number of experimental studies have probed the structures and behavior of intermediates in the catalytic cycle, and the three-coordinate copper(I) amide catalyst resting state (Complex I, Scheme 2) has been proposed as a key catalytic intermediate in a number of literature reports.9,10 In addition, reactivity studies have shown aryl halide activation likely proceeds via direct oxidative addition of the aryl halide with this catalyst resting state.9,10b Kinetic studies on the coppercatalyzed C−N cross coupling reaction have also been used

The classic copper-mediated cross-coupling reaction between aryl halides and amines to form C−N bonds, as first reported by Ullmann, has many drawbacks including high (often stoichiometric) copper loadings, high reaction temperatures, and long reaction times.1 Recently, protocols that employ bidentate nitrogen- and oxygen-based ancillary ligands, such as 1,10-phenanthroline,2 1,3-diketones,3 and amino acids,4 have overcome some of these issues and enabled the cross-coupling reactions to be carried out under milder reaction temperatures (≤110 °C) with lower copper loadings (≤10%) (Scheme 1).5 The relatively low cost and low toxicity of copper, the use of cheap and low molecular weight ancillary ligands, and good functional group tolerance have made this reaction an attractive Scheme 1. Modified Ullmann Reaction

Received: February 19, 2016 Revised: April 28, 2016

© XXXX American Chemical Society

3965

DOI: 10.1021/acscatal.6b00504 ACS Catal. 2016, 6, 3965−3974

ACS Catalysis



Research Article

RESULTS AND DISCUSSION 1. Modified Ullmann Amination Reaction. The C−N cross-coupling of piperidine (1) and iodobenzene (2) to give N-phenylpiperidine (3) using 10.0 mol % catalytic loadings of CuI was chosen as the model reaction in this work, following on previously published results from our group involving similar systems.16 First the catalytic reactivity of this reaction using either fully soluble TBPM base or poorly soluble K3PO4 base, both in dimethyl sulfoxide solvent at room temperature and with 20.0 mol % 2-acetylcyclohexanone (L) as the ancillary ligand, were compared (Figure 1). In agreement with Liu’s

Scheme 2. Proposed Catalytic Cycle for the Modified Ullmann Amination Reaction

to confirm the intermediacy of this copper(I) amide species.11,12 However, the exact roles played by the ligand, solvent, and base, and the presence and impact of off-cycle equilibria remain a matter of debate. A more detailed kinetic profiling would allow the currently proposed general catalytic cycle, as shown in Scheme 2, to be further developed to include other potential reaction steps and solution equilibria such as catalyst deactivation and inhibition. In addition, it would improve fundamental understanding of the mechanism and in turn allow a more rational approach to the development of new catalysts as well as the improvement of existing systems. However, the attainment of accurate kinetic data for coppercatalyzed amination reactions is challenging. Commonly employed inorganic bases, such as K3PO4, NaOtBu, or Cs2CO3, exhibit low solubility in the reaction media.5 As a result, mass transfer effects can arise. These significantly affect the rate of deprotonation and therefore also the overall rate of reaction.13 This complicates the attainment of an accurate chemical kinetic profile. As a result, detailed kinetic studies of copper-catalyzed N-arylation reactions to date have been limited to low pKa substrates such as amide nucleophiles (as reported by Buchwald and co-workers).14 In this work, experimental studies on the mechanism of copper-catalyzed C−N cross coupling between aryl iodides and alkylamine substrates are reported using tetra(n-butyl)phosphonium malonate (TBPM) as the base. This organic ionic base has been previously shown by Liu and co-workers15 to be highly effective in copper-catalyzed protocols, capable of operating at reaction temperatures as low as 0 °C for the amination of aryl iodides or room temperature for aryl bromide substrates. In addition to the benefits associated with its high reactivity in modified-Ullmann protocols, TBPM is fully soluble in various polar solvents.15 Thus, the homogeneous nature of the reaction solutions eliminates undesired mass transfer effects that are present using alternative inorganic bases. As a result, kinetic analysis of the cross coupling reaction can be simplified, which allows full determination of the rate dependences in reactants and provides additional insights into possible catalytic intermediates and catalyst deactivation pathways. Kinetic modeling has also been used to support the experimental data.

Figure 1. Profiles of [3] over time produced using TBPM or K3PO4 as the base in the presence or absence of 2-acetylcyclohexanone (L) for the reaction shown. Product concentrations were determined by GC using naphthalene as an internal standard.

work on soluble organic bases,15 the reaction rate was observed to be significantly faster with TBPM compared to K3PO4 (red and green curves, respectively, Figure 1). Furthermore, 97% yield of product 3 was obtained after only 3 h using TBPM, whereas the yield of 3 using K3PO4 plateaued at the lower value of 77% after 22 h. The influence of the commonly used ancillary diketone (L) on the reaction rate was also assessed. Surprisingly, the [3] over time profiles produced in the presence and absence of L with TBPM are very similar (red and blue curves, Figure 1). In marked contrast, only 7% yield of 3 was observed in the absence of L with K3PO4 after 18 h (not shown in Figure 1). Hence, while the diketone significantly improves the overall reactivity with K3PO4, it does not seem to influence the reactivity at all with TBPM as the base. One possible explanation for this observation is that TBPM, unlike K3PO4, is not basic enough to deprotonate L to give the active 1,3diketonate ligand. In order to test this theory, a 1H NMR 3966

DOI: 10.1021/acscatal.6b00504 ACS Catal. 2016, 6, 3965−3974

Research Article

ACS Catalysis spectrum of a mixture of L and TBPM together in DMSO-d6 was obtained.17 However, this revealed that TBPM is indeed sufficiently basic in DMSO to deprotonate L (which is present predominantly in its enol tautomer form), as evidenced by the loss of its acidic enol O−H proton resonance at 15.92 ppm. Alternative explanations are that the excess of malonate anion (present in TBPM) coordinates more strongly to the copper(I) catalyst center than the 1,3-diketonate ligand, or that the accelerating effect of L is solely due to its role in improving mass transfer of the base with K3PO4, which is not relevant when dealing with soluble bases such as TBPM. Reaction profiles using the ancillary ligands N,N-dimethylglycine, 2-picolinic acid, or 1,10-phenanthroline15 (all with TBPM base) showed similar reactivity to the 1,3-diketone and ligand free systems.17 Hence, in order to simplify the system as much as possible, while still maintaining high reactivity and conversion rates, it was decided to perform the reactions in the kinetic study using ligand-free conditions. In order to further verify the suitability of the model reaction for kinetic profiling, the system was further analyzed for mass transfer effects and for the presence of any competing sidereactions. In contrast to the heterogeneous reaction mixture obtained with K3PO4 as the base, employing soluble TBPM led to an entirely homogeneous reaction solution. Moreover, very similar [2] over time profiles for unstirred and stirred reactions verify that mass transfer effects are not present.17 Although the C−N coupling reaction proceeds to give 97% yield of the target arylamine with TPBM as the base (Figure 1), GC monitoring of the reaction revealed 100% conversion of phenyl iodide (2). The missing mass balance was investigated using 1H NMR analysis of the completed reaction mixtures. These studies revealed a 3% yield of benzene was also present, presumably arising due to hydrodehalogenation of 2. Hydrodehalogenation side reactions of aryl halides have been reported in other copper-mediated reactions in the literature.11,18,19 However, given the very small quantity of benzene generated during the reaction, it was thought unlikely to significantly affect the interpretation of any kinetic data. Further 1H NMR studies show the amine proton to be the most likely source of the hydrogen atom in the hydrodehalogenation reaction.17 2. Kinetic Analysis Using RPKA. Reaction progress kinetic analysis (RPKA), as described by Blackmond,20,21 was used to explore the rate dependence in each reactant and identify possible catalyst deactivation and inhibition pathways. RPKA involves analyzing experimental data collected by continuously monitoring reactions performed under synthetically relevant conditions. By utilizing a parameter called the “excess” ([e]), which is defined as the difference in the initial substrate concentrations (eq 1), kinetic information can be obtained from same excess and different excess experiments.20 The experimental data can then be plotted as reaction rate functions against substrate concentration to extract kinetic information. These plots are commonly referred to as graphical rate equations. RPKA has been demonstrated by numerous reports in the literature to be an effective method for elucidating reaction mechanisms for catalytic processes.22 [e] = [amine]0 − [aryl halide]0

Figure 2. (a) Graphical rate equation for the reaction shown. (b) Comparison between [2] determined from calorimetry heat flow and 1 H NMR spectroscopy measurements.

equations to exclude any rate effects caused by heat transfer effects and also to limit inaccuracies as the limiting substrate concentration approaches zero (see the Supporting Information for the full plots).21 Validation of the reaction calorimetry data was confirmed by the excellent correlation between the conversion calculated from the detected heat flow and the conversion determined by 1H NMR analysis (Figure 2b). In addition, the calculated enthalpy of reaction, ΔHrxn, for all calorimetric experiments obtained in this study over all the various initial reactant concentrations was consistent within a 2% margin (ΔHrxn = 172.5 ± 3.5 kJ mol−1). Monitoring of the reaction rate against [2] with CuI loadings of either 2.50 or 5.00 mol % gives overlaying plots when the reaction rate is normalized by [Cu]total0.9 (Figure 3). This indicates that the copper-catalyzed coupling reaction between 1 and 2 is approximately first order in [Cu]total.23 For the same excess experiments, the reaction rates at any given value of [2] would be expected to be equal in each experiment when the concentration of the active catalyst (or on-cycle catalyst) is constant (i.e., no catalyst activation or deactivation processes are present). However, at any given [2], the rate in an experiment with lower initial reactant concentrations (red curve, Figure 4) was found to be higher than that in the experiment with higher initial concentrations (blue curve). This indicates a drop in [Cu]active in the higher concentration experiment, which can be due to either catalyst inhibition or deactivation. To determine if catalyst inhibition by reaction products or byproducts was the cause of the decreased

(1)

The model copper-catalyzed C−N cross-coupling reaction, as defined in the previous section and shown in Figure 2, was monitored in situ by reaction calorimetry to produce the graphical rate equation also shown in Figure 2. Only the data between 20−80% conversion is displayed in the graphical rate 3967

DOI: 10.1021/acscatal.6b00504 ACS Catal. 2016, 6, 3965−3974

Research Article

ACS Catalysis

The rate vs [2] curve produced with these added components overlays with the previous same excess experiment and therefore indicates that catalyst inhibition by processes such as product coordination, protonation of 1 by 4, or solution conductivity changes caused by ionic species 5 can all be discounted. Because catalyst inhibition can be ruled out, catalyst deactivation is implied (see discussion later for analysis of catalyst decomposition routes). The rate dependences in [TBPM], [1], and [2] were next determined using the different excess protocol. Two reactions were conducted using either 0.1500 or 0.2000 M [1]0, so that at any given time, the concentrations of all the reactants except 1 are the same in both experiments. An overlay of the curves produced by normalizing the reactions rates by [1]1 (Figure 5)

Figure 3. Normalized graphical rate equation (from dividing rate by [Cu]total0.9) of experiments using concentrations listed in the table.

Figure 5. Normalized graphical rate equations (from dividing rate by [1]1) of experiments using concentrations listed in the table.

shows the reaction is first-order in [1]. Similarly, the graphical rate equations of the two different excess experiments shown in Figure 6 reveal a first-order dependence in [2] on the reaction rate. The graphical rate equations produced from different excess experiments using varied [TBPM]0 unexpectedly revealed that

Figure 4. Graphical rate equations from standard and same excess experiments using concentrations stated in the table.

reactivity, the same excess experiment was repeated but with the added reaction products/byproducts N-phenylpiperidine (3), tetra(n-butyl)phosphonium monoanionic malonic acid (4), and tetra(n-butyl)phosphonium iodide (5), each at 0.0500 M concentration (green curve, Figure 4).

Figure 6. Normalized graphical rate equations (from dividing rate by [2]1) of experiments using concentrations listed in the table. 3968

DOI: 10.1021/acscatal.6b00504 ACS Catal. 2016, 6, 3965−3974

Research Article

ACS Catalysis higher [TBPM]0 led to slower reaction rates. This indicates a negative rate dependence in [TBPM]. When the reaction rates are normalized by dividing [TBPM]−1, the curves overlay moderately well (Figure 7), suggesting that the order in

Figure 7. Normalized graphical rate equations (from dividing rate by [TBPM]−1) using concentrations listed in the table.

[TBPM] is approximately negative first-order in this concentration regime. As far as we are aware, this is the first time a negative order in base has been implied in copper-catalyzed amination reactions. Thus, overall under these conditions, the reaction rate was found to have a first order dependence in [1], [2], and [Cu]total and negative order dependence in [TBPM]. In addition, the same excess experiments revealed the catalyst to deactivate over the course of the reaction. As the reaction rate can be influenced by a large number of factors, the initial visual fit to a straightforward system exhibiting overall-second order kinetics, as suggested by the linearity and overlay of the normalized rate curves in Figure 5 and Figure 6, is likely misleading. The reduction in rate over time caused by decreasing [Cu]total as a result of catalyst deactivation, is likely offset by an approximately equal increase in rate caused by the continually decreasing [TBPM] as it is consumed in the reaction. These findings highlight the important role kinetic profiling using RPKA can play in uncovering the many different influences on reaction rate which might otherwise remain hidden or obscured. 3. Studies on Catalyst Deactivation. Decreases in rate between sequential reactions using the same catalyst solution were monitored using calorimetry in order to confirm catalyst deactivation was occurring. A reaction between 1.000 mmol piperidine (1) and 0.500 mmol iodobenzene (2) in DMSO and 2.000 mmol TBPM was initiated by addition of 0.050 mmol CuI with a total solution volume of 10 mL (Figure 8a and b, blue curve). After the heat flow from the first reaction became negligible, indicating the reaction had completed, a mixture of piperidine (1) and iodobenzene (2) (0.500 mmol each in 200 μL DMSO solution) was injected to start another reaction cycle using the same catalyst. Thus, the approximately same initial concentrations of 1 and 2 were regenerated, but now with a lower 0.1500 M [TBPM]. It was expected that the second

Figure 8. (a) Profiles of heat flow over time after sequential additions of 1 and 2 (0.0500 M each); (b) Graphical rate equations for each sequential reaction.

reaction would be faster due to the lower [TBPM]. However, the peak heat flow was lower in the second reaction (Figure 8a), and the reaction rate was slower (red curve, Figure 8b) (reaction rate at [2] = 0.040 M decreased by 16% between runs). Furthermore, after a second 0.500 mmol portion of 1 and 2 was added, the reaction rate decreased even further (by an additional 21%) and only 85% conversion was achieved (green curve, Figure 8b). The total solvent volume increased by only 2% with each portion of additional 1 and 2, and as a result, [Cu]total was diluted by the same percentage. As the decrease in reaction rate between the sequential reactions is significantly greater than 2%, catalyst deactivation can be shown to be occurring. To identify the catalyst deactivation pathway, a series of experiments was carried out in which all but one of the reactants were stirred together for 60 min before addition of the final reactant to initiate the reaction. The conversion of iodobenzene (2) over time was monitored for each of these experiments, as shown in Figure 9. The reaction rate remained approximately unchanged when either CuI (blue curve, Figure 9), TBPM (red curve), or 2 (orange curve) were used to initiate the reaction. However, when the amine 1 was used to initiate the reaction (purple curve), the observed rate was much slower than the other experiments, and moreover, the conversion of 2 was significantly curtailed at just 14%. In addition, the reaction solution in this experiment visibly changed appearance from colorless to blue within 15 min, 3969

DOI: 10.1021/acscatal.6b00504 ACS Catal. 2016, 6, 3965−3974

Research Article

ACS Catalysis

Scheme 3. Proposed Catalyst Deactivation Pathway and the Competitive Coordination of Piperidine (1) to [CuIMal]− (6)a

Figure 9. Conversion of 2 over time where the reagent(s) shown was used to initiate the reaction after a delay of 60 min. Conversions were determined by GC using naphthalene as an internal standard. a

indicative of the formation of copper(II) species. When both 1 and TBPM were added together to initiate the reaction (green curve), the rate and overall conversion of 2 once more resembled the values seen with CuI, TBPM, or 2 initiation. These results suggest that the catalyst is stable in the absence of TBPM, or when TBPM and amine 1 are both present, but becomes unstable when TBPM is present in the absence of the amine. A simple qualitative experiment in which CuI in DMSO was treated with TBPM further verified this finding, with the solution rapidly changing from colorless to blue in appearance. A similar color change is observed in the final reaction solutions produced in the kinetic experiments discussed previously and has also been reported in the literature for Ullmann couplings using inorganic bases.24,25 In the later cases, this was attributed to disproportionation of a (as yet unidentified) copper(I) species.24,25 These results now pinpoint a copper(I)−base complex as the most likely candidate to undergo this disproportionation reaction. It is noteworthy that isolated solutions of copper(I) carboxylate complexes without stabilizing coligands, such as phosphines or alkenes, have previously been reported to be unstable and prone to disproportionation at room temperature even under inert atmospheres.26 In addition, all isolated copper(I) malonate complexes to date contain PPh 3 coligands, presumably to prevent such disproportionation.27 The results also demonstrate how amine 1 can act as a stabilizing coligand to reduce the degree of disproportionation of the copper(I) malonate intermediate. Scheme 3 shows this in the context of the reaction mechanism where 1 competitively coordinates to the copper(I) malonate complex 6 to give heteroleptic complex 7, whereas coordination of another malonate dianion would give the homoleptic complex 9. Complex 7 can subsequently undergo N−H deprotonation to give the malonate-ligated copper(I) piperidine complex 8, which is a commonly accepted intermediate in the catalytic cycle (I, Scheme 2). However, complex 9 would be susceptible to disproportionation to give Cu(0) and Cu(II) species. In the literature, copper(I) disproportionation has been shown to occur from copper(I) complexes ligated by two 1,3-diketonate,

Tetra(n-butyl)phosphonium counter cations are not shown.

1,3-keto ester enolate, or 1,3-keto amide enolate ancillary ligands.24,28 In agreement with these other studies, disproportionation is proposed to occur from the copper(I) dimalonate species 9 in Scheme 3. In order to confirm the proposed stabilizing effect of the amine on the catalyst, a multiple addition experiment was performed in which the reaction rates of sequential reactions were monitored over successive additions of aryl iodide 2 to a reaction solution that contained a large excess of amine 1 (Figure 10). In accordance with the mechanism proposed in Scheme 3, a large excess of 1 would favor the formation of the piperidine-ligated copper(I) intermediate 7 over the off-cycle dimalonate complex 9. This in turn would effectively remove the negative rate dependence in [TBPM] and prevent catalyst deactivation. In addition, the negligible effect of addition of monoanionic malonic acid (4) on the reaction rate (see Figure 4) indicates that the N−H deprotonation of 7 is highly favored, thus leading to saturation of the catalyst as the malonate-ligated copper(I) piperidide complex 8. As a result, the rate equation in the presence of a large excess of 1 can be estimated as being dependent only on [2] and [Cu]total (eq 2).

rate = k[Cu]total [2]

(2)

Figure 10 shows that the rate vs [2] profiles from this experiment are linear and overlay, thus confirming that [Cu]total is constant and catalyst deactivation has been prevented at high [1]0. The amine has therefore been shown to have an important role in minimizing catalyst deactivation. As the graphical rate equations were in agreement with eq 2, the rate constant for iodobenzene activation could be calculated using the gradients of the linear least-squares fits to the plots in Figure 10 to be 18.0 ± 0.7 M−1 min−1. 4. Identification of Aryl Halide Activation Mechanism. In order to investigate the aryl halide activation mechanism, the overall catalytic reaction between piperidine (1) and iodobenzene (2) with TBPM was performed in the dark. Product 3 was still obtained in excellent yield (97%) after 3 h, similar to when the reaction was performed under ambient 3970

DOI: 10.1021/acscatal.6b00504 ACS Catal. 2016, 6, 3965−3974

Research Article

ACS Catalysis

compound with the nucleophile bonded to the pendant methyl moiety. The copper-catalyzed reaction of piperidine (1) and radical probe 10 produced the arylamine product in 45% yield and hydrodehalogenation side product allyl phenyl ether in 2% yield, while no detectable amount of either cyclized side products were observed by 1H NMR spectroscopy or GC-MS (Scheme 4). Therefore, it is likely that the copper-catalyzed aryl amination and hydrodehalogenation reactions do not proceed via aryl radical intermediates under these conditions. This result is in agreement with the absence of biphenyl formation from phenyl radical homocoupling when iodobenzene (2) was used as the substrate in the kinetic study.19 Furthermore, it is also in agreement with a previous report by Hartwig and co-workers who found that the copper-catalyzed N-arylation of diarylamines was unlikely to involve free aryl radicals and that the aryl halide activation involved oxidative addition to copper(I).10b The aryl halide activation in the copper-catalyzed Narylation of alkylamines studied in this work can therefore be considered to progress via an oxidative addition mechanism as opposed to a single-electron transfer process. 5. Proposed Mechanistic Cycle and Kinetic Modeling. Based on the results from the kinetic and other mechanistic studies herein, Scheme 5 shows our proposed reaction

Figure 10. Graphical rate equations using sequential additions of 2 (0.585 mmol) to the same solution after reaction completion. Linear least-squares fit to y = mx + c gave the following parameters to three significant figures: 1st reaction: m = 0.0348 ± 1.66 × 10−4 min−1, c = 9.51 × 10−5 ± 3.94 × 10−6 M min−1, r2 = 0.984; 2nd reaction: m = 0.0360 ± 1.59 × 10−4 min−1, c = 1.61 × 10−4 ± 4.32 × 10−6 M min−1, r2 = 0.988; 3rd reaction: m = 0.0374 ± 1.01 × 10−4 min−1, c = 1.02 × 10−4 ± 2.72 × 10−6 M min−1, r2 = 0.995.

Scheme 5. Proposed Reaction Mechanism, Including Catalyst Deactivation Pathway, For the Copper-Catalyzed Cross-Coupling Reaction between Piperidine and Iodobenzene with TBPM as the Basea

light. This control experiment confirmed that the aryl halide activation is not photochemically induced; however, it does not omit aryl radicals as potential reaction intermediates. To determine if aryl radicals are formed during the reaction, the radical probe 2-(allyloxy)iodobenzene (10) was used as a substrate (Scheme 4). If the reaction proceeds via a nonradical Scheme 4. Employment of the Radical Probe 10 Reveals Aryl Halide Activation Is Unlikely To Occur via a Radical-Based Mechanisma

a a

Yields were determined by 1H NMR spectroscopy using naphthalene as an internal standard.

Tetra(n-butyl)phosphonium counter cations are not shown.

mechanism for the copper-catalyzed C−N cross coupling reaction. This mechanism is also supported by detailed 1H NMR spectroscopic studies.17 As dianionic malonate can adopt a large number of coordination modes to copper(I),30 the mode of malonate coordination is not defined in the proposed mechanism. Malonate coordination to the initially formed copper(I) malonate (6) is proposed to give an off-cycle copper(I) di(malonate) species (9). Catalyst deactivation via copper(I) disproportionation to give copper(II) and copper(0) is then

based mechanism, the expected C(sp2)−N bond formation would occur. In comparison, if the mechanism involves the formation of aryl radicals, the allyloxy aryl radical intermediate would rapidly ring-close via a 5-exo-trig process (k = 9.6 × 109 s−1 at 25 °C in DMSO).29 The subsequent methyl radical could then react to give 3-methyl-2,3-dihydrobenzofuran or a 3971

DOI: 10.1021/acscatal.6b00504 ACS Catal. 2016, 6, 3965−3974

Research Article

ACS Catalysis suggested to occur from this off-cycle complex 9. As piperidine (1) was found to stabilize the copper catalyst from deactivation, it is proposed to competitively coordinate against malonate to complex 6 to give the malonate and piperidine heteroleptic copper(I) intermediate 7. The next step involves N−H deprotonation of the coordinated piperidine ligand by malonate to give the malonate-ligated copper(I) piperidide complex 8. Complex 8 has a 1:1:1 ratio of ancillary ligand (malonate), copper(I), and deprotonated nucleophile, which is in accordance with the general structure of proposed catalytic resting state species in other copper-catalyzed C−N crosscoupling reactions.9a,10a,b,14b,25b As the radical probe 10 had no effect on the aryl amination reaction, aryl halide activation of iodobenzene (2) is proposed to occur via an oxidative addition/reductive elimination mechanism to give arylamine product 3 and regenerate complex 6. On account of the dianionic malonate from TBPM acting as both the ancillary ligand and base, the general structure of the on-cycle species in the proposed mechanism in Scheme 5 are similar to those in the general catalytic cycle shown in Scheme 2. The experimental studies herein therefore provide further support for the identity of the on-cycle intermediates. The proposed mechanism in Scheme 5 also incorporates for the first time a catalyst deactivation pathway which proceeds via an offcycle copper(I)-base complex (9). The steady-state rate law for the proposed catalytic cycle shown in Scheme 5 (including all steps proposed as reversibly connected to the cycle, but neglecting the irreversible deactivation step) is shown in eq 3 and derived in the Supporting Information. rate =

k4K 2K3[1][2][mal][Cu]total 1 + K 2[1] + K 2K3[1][mal] + K1[mal]

Figure 11. Simulated concentrations of complexes 6, 7, 8, 9, and [Cu(II)] + [Cu(0)] (a) over the course of the reaction in Figure 2 and (b) under the same conditions but using an excess of amine 1 (2.000 M). Produced in COPASI using optimized kinetic parameters.17

(3)

In order to obtain the first order kinetics in both [1] and [2] (as observed in the experimental RPKA studies), either the free catalyst (the “1” term) or the K1[mal] term must dominate the denominator; the latter being the most likely (as backed up by the calculations below). This therefore implies that the intrinsic steady-state dependence of [mal] is zero-order. It is the irreversible draining of catalyst out of the system (which is not considered in the steady-state rate law) which allows the order in [mal] to manifest itself as negative in the RPKA experiments. In order to further access the viability of our proposed mechanistic cycle, the reaction simulation and data-fitting program COPASI31 was employed to model the experimental calorimetric data.17 A number of different mechanistic scenarios were investigated including changes in the order of piperidine and malonate coordination, disproportionation from copper(I) malonate (6) instead of copper(I) di(malonate) (9), disproportionation from an equivalent of 6 and 9 each, and formation of a di(piperidido)cuprate off-cycle complex. However, the best-fit model, being both consistent with the experimental reaction calorimetry data and displaying low standard deviations on the rate constants, was the mechanism shown in Scheme 5. The simulated concentrations of all copper containing species over the course of the reaction under standard conditions and also in the presence of a large excess of amine are shown in Figure 11. Since K1 is predicted to be large, malonate coordination to copper(I) malonate (6) to give copper(I) di(malonate) (9) occurs readily leading to a high [9] over the course of the reaction, and in particular in the early stages (Figure 11a).

Hence a significant quantity of catalyst is predicted to reside (under these conditions) in an inactive off-cycle form (9). However, increasing the initial concentration of amine 1 (Figure 11b) drives the amine coordination and deprotonation steps forward to the catalyst resting state 8, which is now predicted to be the predominate copper containing species. This suppresses formation of 9 and inhibits catalyst deactivation, and is in agreement with the experimental kinetic studies (vide supra). A consequence of an unfavorable equilibrium in the reverse direction from catalyst resting state 8, through intermediates 7 and 6, back to unstable off-cycle copper(I) dimalonate complex 9 is that solutions of CuI, amine 1, and TBPM together are relatively stable against immediate complete catalyst deactivation. This is also in agreement with the experimental observations presented in this manuscript that showed solutions of CuI and TBPM are stable to disproportionation in the presence of 1 and is also supported by the multiple addition experiments (Figure 8 and 10). The modeling also predicts kdeact to be lower but of a similar magnitude to k4. Hence, the relative magnitudes of the rate and equilibrium constants in the other equilibrium reaction steps (both on- and off-cycle) can significantly influence the reaction outcome. A better understanding of these positive and negative influences on the catalytic cycle will undoubtedly benefit further optimization of the copper(I) amination process.



CONCLUSION In order to build a better understanding of the mechanism of copper catalyzed amination reactions, experimental mechanistic 3972

DOI: 10.1021/acscatal.6b00504 ACS Catal. 2016, 6, 3965−3974

ACS Catalysis



ACKNOWLEDGMENTS This work was supported by Syngenta and the UK EPSRC (Grant EP/J500239/1). We thank the reviewers for their suggested improvements to the modelling studies, which have been implemented in the final version of the manuscript.

studies have been undertaken on the representative C−N crosscoupling reaction between piperidine (1) and iodobenzene (2) using a Cu(I) catalyst. On the basis of initial screenings, tetrabutylphosphonium malonate (TBPM) was employed as the base with no additional ligands present. In contrast to poorly soluble inorganic bases, such as K3PO4, the high solubility of TBPM in the reaction media was shown to obviate all base mass transfer effects, thus permitting the collection of more reliable and precise kinetic data. Kinetic profiling reveals the reaction rate to have a dependence on all of the reaction components in the concentration range studied. First-order kinetics are observed with respect to [1], [2], and [Cu]total, and unexpectedly negative-order kinetics in [Base]. The negative rate dependence in [Base] can been attributed to the formation of an off-cycle copper(I) dimalonate species (9), in which the malonate dianion is acting as a ligand in addition to its role as the base. This off-cycle species is then postulated to undergo disproportionation and is thus responsible for catalyst deactivation. A new catalytic scheme incorporating this offcycle catalytic deactivation step has been proposed (Scheme 5), with similar deactivation processes likely common to all copper(I)-catalyzed Ullmann reactions. Modeling of the kinetic experimental data using COPASI is shown to be congruent with the proposed scheme. The key role of the amine 1 in minimizing catalyst deactivation has also been highlighted by the kinetic studies, and it is demonstrated how a high concentration of amine can ensure excellent catalyst stability and improve the observed TON. In addition, an examination of the aryl halide activation mechanism using radical probes is consistent with an oxidative addition pathway. Existing studies on the Ullmann amination reaction rarely acknowledge the negative effect ligands (ancillary or base derived) can have on the reaction rate, for example, by causing the introduction of off-cycle equilibria and reduced catalyst stability. The results presented in this work suggest that ligand (and base) choice for these catalytic systems should not only focus on improving the rate of the on-cycle steps but also aim to minimize or eliminate these detrimental off-cycle processes. To this end, we are currently investigating a range of existing and new ancillary ligand systems using similar experimental mechanistic studies. It is anticipated that building a better understanding of the various positive and negative influences on these complex reaction kinetics will lead to the development of better performing and more robust copper-catalyzed amination protocols.





REFERENCES

(1) (a) Ullmann, F. Ber. Dtsch. Chem. Ges. 1903, 36, 2382−2384. (b) Yamamoto, T.; Kurata, Y. Can. J. Chem. 1983, 61, 86−91. (c) Sperotto, E.; van Klink, G. P. M.; van Koten, G.; de Vries, J. G. Dalton Trans. 2010, 39, 10338−10351. (2) (a) Goodbrand, H. B.; Hu, N.-X. J. Org. Chem. 1999, 64, 670− 674. (b) Gujadhur, R. K.; Bates, C. G.; Venkataraman, D. Org. Lett. 2001, 3, 4315−4317. (c) Moriwaki, K.; Satoh, K.; Takada, M.; Ishino, Y.; Ohno, T. Tetrahedron Lett. 2005, 46, 7559−7562. (d) Tseng, C.-K.; Lee, C.-R.; Tseng, M.-C.; Han, C.-C.; Shyu, S.-G. Dalton Trans. 2014, 43, 7020−7027. (e) Tseng, C.-K.; Tseng, M.-C.; Han, C.-C.; Shyu, S.G. Chem. Commun. 2011, 47, 6686−6688. (3) (a) de Lange, B.; Lambers-Verstappen, M.; Schmieder-van de Vondervoort, L.; Sereinig, N.; de Rijk, R.; de Vries, A.; de Vries, J. Synlett 2006, 2006, 3105−3109. (b) Shafir, A.; Buchwald, S. L. J. Am. Chem. Soc. 2006, 128, 8742−8743. (c) Shafir, A.; Lichtor, P. A.; Buchwald, S. L. J. Am. Chem. Soc. 2007, 129, 3490−3491. (4) (a) Ma, D.; Cai, Q.; Zhang, H. Org. Lett. 2003, 5, 2453−2455. (b) Zhang, H.; Cai, Q.; Ma, D. J. Org. Chem. 2005, 70, 5164−5173. (5) (a) Evano, G.; Blanchard, N.; Toumi, M. Chem. Rev. 2008, 108, 3054−3131. (b) Ley, S. V.; Thomas, A. W. Angew. Chem., Int. Ed. 2003, 42, 5400−5449. (c) Monnier, F.; Taillefer, M. Angew. Chem., Int. Ed. 2009, 48, 6954−6971. (6) (a) Beletskaya, I. P.; Cheprakov, A. V. Organometallics 2012, 31, 7753−7808. (b) Evano, G.; Theunissen, C.; Pradal, A. Nat. Prod. Rep. 2013, 30, 1467−1489. (7) (a) Clement, J.-B.; Hayes, J. F.; Sheldrake, H. M.; Sheldrake, P. W.; Wells, A. S. Synlett 2001, 2001, 1423−1427. (b) Ma, D.; Xia, C. Org. Lett. 2001, 3, 2583−2586. (c) Ma, D.; Xia, C.; Jiang, J.; Zhang, J. Org. Lett. 2001, 3, 2189−2191. (d) Ma, D.; Zhang, Y.; Yao, J.; Wu, S.; Tao, F. J. Am. Chem. Soc. 1998, 120, 12459−12467. (e) Okano, K.; Tokuyama, H.; Fukuyama, T. Chem. Commun. 2014, 50, 13650− 13663. (8) (a) Jones, G. O.; Liu, P.; Houk, K. N.; Buchwald, S. L. J. Am. Chem. Soc. 2010, 132, 6205−6213. (b) Zhang, S.; Ding, Y. Organometallics 2011, 30, 633−641. (9) (a) Franc, G.; Cacciuttolo, Q.; Lefèvre, G.; Adamo, C.; Ciofini, I.; Jutand, A. ChemCatChem 2011, 3, 305−309. (b) Lefèvre, G.; Franc, G.; Adamo, C.; Jutand, A.; Ciofini, I. Organometallics 2012, 31, 914− 920. (c) Lefèvre, G.; Franc, G.; Tlili, A.; Adamo, C.; Taillefer, M.; Ciofini, I.; Jutand, A. Organometallics 2012, 31, 7694−7707. (d) Yu, H. Z.; Jiang, Y. Y.; Fu, Y.; Liu, L. J. Am. Chem. Soc. 2010, 132, 18078− 18091. (10) (a) Creutz, S. E.; Lotito, K. J.; Fu, G. C.; Peters, J. C. Science 2012, 338, 647−651. (b) Giri, R.; Hartwig, J. F. J. Am. Chem. Soc. 2010, 132, 15860−15863. (c) Huffman, L. M.; Stahl, S. S. J. Am. Chem. Soc. 2008, 130, 9196−9197. (d) Ribas, X.; Güell, I. Pure Appl. Chem. 2014, 86, 345−360. (e) Tseng, C.-K.; Lee, C.-R.; Han, C.-C.; Shyu, S.G. Chem. - Eur. J. 2011, 17, 2716−2723. (f) Liu, X.; Zhang, S.; Ding, Y. Dalton Trans 2012, 41, 5897−5902. (11) Arai, S.; Hida, M.; Yamagishi, T. Bull. Chem. Soc. Jpn. 1978, 51, 277−282. (12) (a) Bethell, D.; Jenkins, I. L.; Quan, P. M. J. Chem. Soc., Perkin Trans. 2 1985, 1789−1795. (b) Paine, A. J. J. Am. Chem. Soc. 1987, 109, 1496−1502. (13) (a) Larsson, P.-F.; Bolm, C.; Norrby, P.-O. Chem. - Eur. J. 2010, 16, 13613−13616. (b) Larsson, P.-F.; Wallentin, C.-J.; Norrby, P.-O. ChemCatChem 2014, 6, 1277−1282. (c) Meyers, C.; Maes, B. U. W.; Loones, K. T. J.; Bal, G.; Lemière, G. L. F.; Dommisse, R. A. J. Org. Chem. 2004, 69, 6010−6017.

ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acscatal.6b00504. Experimental procedures, experimental and spectroscopic data, full graphs related to the kinetic studies, and details of the kinetic modeling using COPASI (PDF)



Research Article

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel: +44 (0)207 5945754. Notes

The authors declare no competing financial interest. 3973

DOI: 10.1021/acscatal.6b00504 ACS Catal. 2016, 6, 3965−3974

Research Article

ACS Catalysis (14) (a) Strieter, E. R.; Blackmond, D. G.; Buchwald, S. L. J. Am. Chem. Soc. 2005, 127, 4120−4121. (b) Strieter, E. R.; Bhayana, B.; Buchwald, S. L. J. Am. Chem. Soc. 2009, 131, 78−88. (15) Yang, C.-T.; Fu, Y.; Huang, Y.-B.; Yi, J.; Guo, Q.-X.; Liu, L. Angew. Chem., Int. Ed. 2009, 48, 7398−7401. (16) Sung, S.; Braddock, D. C.; Armstrong, A.; Brennan, C.; Sale, D.; White, A. J. P.; Davies, R. P. Chem. - Eur. J. 2015, 21, 7179−7192. (17) For full details, see Supporting Information. (18) (a) Cannon, K. A.; Geuther, M. E.; Kelly, C. K.; Lin, S.; MacArthur, A. H. R. Organometallics 2011, 30, 4067−4073. (b) Fier, P. S.; Hartwig, J. F. J. Am. Chem. Soc. 2012, 134, 5524−5527. (c) Fier, P. S.; Hartwig, J. F. J. Am. Chem. Soc. 2012, 134, 10795−10798. (19) (a) Keegstra, M. A.; Peters, T. H. A.; Brandsma, L. Tetrahedron 1992, 48, 3633−3652. (b) Moroz, A. A.; Shvartsberg, M. S. Russ. Chem. Rev. 1974, 43, 679−689. (20) Blackmond, D. G. Angew. Chem., Int. Ed. 2005, 44, 4302−4320. (21) Mathew, J. S.; Klussmann, M.; Iwamura, H.; Valera, F.; Futran, A.; Emanuelsson, E. A. C.; Blackmond, D. G. J. Org. Chem. 2006, 71, 4711−4722. (22) (a) Baxter, R. D.; Sale, D.; Engle, K. M.; Yu, J.-q.; Blackmond, D. G. J. Am. Chem. Soc. 2012, 134, 4600−4606. (b) Colladon, M.; Scarso, A.; Sgarbossa, P.; Michelin, R. A.; Strukul, G. J. Am. Chem. Soc. 2007, 129, 7680−7689. (c) Ji, Y.; Benkovics, T.; Beutner, G. L.; Sfouggatakis, C.; Eastgate, M. D.; Blackmond, D. G. J. Org. Chem. 2015, 80, 1696− 1702. (d) Kina, A.; Iwamura, H.; Hayashi, T. J. Am. Chem. Soc. 2006, 128, 3904−3905. (e) Ruiz-Castillo, P.; Blackmond, D. G.; Buchwald, S. L. J. Am. Chem. Soc. 2015, 137, 3085−3092. (f) Scott, M.; Sud, A.; Boess, E.; Klussmann, M. J. Org. Chem. 2014, 79, 12033−12040. (g) Shekhar, S.; Ryberg, P.; Hartwig, J. F.; Mathew, J. S.; Blackmond, D. G.; Strieter, E. R.; Buchwald, S. L. J. Am. Chem. Soc. 2006, 128, 3584−3591. (23) The slightly lower order in [Cu]total from exactly 1.0 could be the result of another process that is also dependent on [Cu]total, such as the catalyst deactivation highlighted in this report. For similar rate orders in catalyst, see refs 10b and 14a. (24) He, C.; Zhang, G.; Ke, J.; Zhang, H.; Miller, J. T.; Kropf, A. J.; Lei, A. J. Am. Chem. Soc. 2013, 135, 488−493. (25) (a) Sambiagio, C.; Munday, R. H.; Marsden, S. P.; Blacker, A. J.; McGowan, P. C. Chem. - Eur. J. 2014, 20, 17606−17615. (b) Tye, J. W.; Weng, Z.; Johns, A. M.; Incarvito, C. D.; Hartwig, J. F. J. Am. Chem. Soc. 2008, 130, 9971−9983. (26) (a) Aalten, H. L.; Van Koten, G.; Goubitz, K.; Stam, C. H. Organometallics 1989, 8, 2293−2299. (b) Cotton, F. A.; Dikarev, E. V.; Petrukhina, M. A. Inorg. Chem. 2000, 39, 6072−6079. (c) Drew, M. G. B.; Edwards, D. A.; Richards, R. J. Chem. Soc., Chem. Commun. 1973, 124−125. (d) Grodzicki, A.; Łakomska, I.; Piszczek, P.; Szymańska, I.; Szłyk, E. Coord. Chem. Rev. 2005, 249, 2232−2258. (e) Ogura, T.; Mounts, R. D.; Fernando, Q. J. Am. Chem. Soc. 1973, 95, 949−951. (27) (a) Darensbourg, D. J.; Holtcamp, M. W.; Longridge, E. M.; Khandelwal, B.; Klausmeyer, K. K.; Reibenspies, J. H. J. Am. Chem. Soc. 1995, 117, 318−328. (b) Oldham, C.; Sandford, W. F. J. Chem. Soc., Dalton Trans. 1977, 2068−2070. (28) Cheng, B.; Yi, H.; He, C.; Liu, C.; Lei, A. Organometallics 2015, 34, 206−211. (29) Annunziata, A.; Galli, C.; Marinelli, M.; Pau, T. Eur. J. Org. Chem. 2001, 2001, 1323−1329. (30) Chawla, S. K.; Arora, M.; Nättinen, K.; Rissanen, K.; Yakhmi, J. V. Polyhedron 2004, 23, 3007−3019. (31) Hoops, S.; Sahle, S.; Gauges, R.; Lee, C.; Pahle, J.; Simus, N.; Singhal, M.; Xu, L.; Mendes, P.; Kummer, U. Bioinformatics 2006, 22, 3067−3074.

3974

DOI: 10.1021/acscatal.6b00504 ACS Catal. 2016, 6, 3965−3974