Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Sep 3.
Published in final edited form as: J Org Chem. 2021 Aug 4;86(17):11378–11387. doi: 10.1021/acs.joc.1c00877

Pd-Catalyzed Alkene Diamination Reactions with O-Benzoylhydroxylamine Electrophiles. Evidence Supporting a Pd(II/IV) Catalytic Cycle, the Role of 2,4-Pentanedione Derivatives as Ligands, and Expanded Substrate Scope

Janelle K Kirsch 1, Gabriel A Gonzalez 1, Mason S Faculak 1, John P Wolfe 1,*
PMCID: PMC9289953  NIHMSID: NIHMS1823162  PMID: 34344155

Abstract

This article describes continued studies on Pd-catalyzed alkene diamination reactions between N-allylguanidines or ureas and O-benzoylhydroxylamine derivatives, which serve as N-centered electrophiles. The transformations generate cyclic guanidines and ureas bearing dialkylaminomethyl groups in moderate to good yield. We describe new mechanistic experiments that have led to a revised mechanistic hypothesis that involves a key oxidative addition of the electrophile to a PdII complex, followed by reductive elimination from PdIV to form the alkyl carbon-nitrogen bond. In addition, we demonstrate that acac, not a phosphine, serves as a key ligand for palladium. Moreover, simple acac derivatives bearing substituted aryl groups outperform acac in the catalytic reactions, and phosphines inhibit catalysis in many cases. These discoveries have led to a significant expansion in the scope of this chemistry, which now allows for the coupling of a variety of cyclic amines, acyclic secondary amines, and primary amines. In addition, we also demonstrate these new conditions allow for the use of amide nucleophiles, in addition to guanidines and ureas.

Graphical Abstract

graphic file with name nihms-1823162-f0001.jpg

INTRODUCTION

Metal-catalyzed alkene diamination reactions are an important sub-class of alkene difunctionalization reactions that have attracted considerable attention in recent years.1,2,3,4,5 One strategy that we found particularly interesting due to both its potential utility, as well as its current limitations, involves the coupling of readily available nitrogen-centered electrophiles with alkenes bearing tethered nitrogen nucleophiles. In 2009 Michael,6 reported the first metal-catalyzed alkene diamination reactions of this type, which utilized NFBS (N-fluorobenzene-sulfonimide) as a source of electrophilic nitrogen. More recently, Wang7 has described the Cu-catalyzed coupling of similar alkene substrates with O-acylated hydroxylamine electrophiles. However, despite the utility of these methods, they have limitations with respect to either scope,8 or control of stereochemistry in reactions involving 1,2-disubstituted alkenes.9

In 2018 we reported the Pd-catalyzed alkene diamination of guanidines and ureas bearing tethered alkenes (e.g., 1a-b) with O-acylated hydroxylamine derivatives such as morpholino benzoate 2a (eq 1).10 These transformations provide cyclic guanidines or ureas bearing appended aminoalkyl groups (e.g., 3a-b) in good to excellent chemical yield. In addition, reactions involving deuterated alkene substrates (e.g., 1c-d) proceed with modest diastereoselectivity favoring products resulting from anti-addition of the two nitrogen atoms to the alkene (e.g., 3c-d).

graphic file with name nihms-1823162-f0008.jpg (1)

We initially hypothesized that the mechanism of these reactions is similar to that of related alkene carboamination reactions between 1a-d and aryl halide or triflate electrophiles and likely involves a typical Pd0/PdII catalytic cycle,10 (Scheme 1, Path A) that is initiated by oxidative addition of 2a to the metal center to afford 4.11 Subsequent coordination of the alkene group of 1d to the metal followed by alkene aminopalladation (5 to 6),12 and then reductive elimination from 6, would afford the product 3d.13 However, a PdII/PdIV catalytic cycle in which 6 was oxidized to 7 by additional 2a,14 followed by reductive elimination to generate 3d and 4 also seemed plausible.13

Scheme 1. Initial mechanistic hypothesis.

Scheme 1.

In this Article we describe new experiments that have led us to revise our mechanistic hypothesis, along with the discovery that acac (2,4-pentanedione) and simple acac derivatives bearing substituted arenes serve as highly effective ligands for this transformation. These discoveries have also allowed for significant expansion in scope with respect to electrophile structure.

RESULTS AND DISCUSSION

Control experiments and precatalyst effects.

Following our preliminary studies, we set out to further examine this reaction and its mechanism (Scheme 2). We elected to first explore the putative oxidative addition of the morpholino benzoate electrophile 2a to either Pd0 or PdII. However, in initial scouting experiments we were surprised to find that the phosphine ligand JackiePhos did not bind to Pd(acac)2 as judged by a lack of change of the 31P NMR chemical shift of JackiePhos vs the mixture of JackiePhos and Pd(acac)2 (eq 2). In order to explore the possibility of oxidative addition to a Pd0/JackiePhos complex, we carried out NMR studies in which the Buchwald PdII(JackiePhos) G3 precatalyst was reduced in situ, as judged by 31P NMR analysis,15 and then treated with 2a (eq 3). Surprisingly, we saw no evidence for oxidative addition of 2a to this Pd0 complex. In addition, we did not observe any evidence for oxidative addition of 2a to the PdII complex Pd(acac)2 at 100 °C, either with or without JackiePhos present (eq 4).16 The results of these experiments immediately suggested two things: (i) since JackiePhos does not appear to bind to Pd(acac)2, it may not be the actual ligand for Pd in these alkene diamination reactions; and (ii) our original mechanistic hypothesis is likely incorrect, as the proposed initial oxidative addition step does not appear to be viable with this catalyst system.

Scheme 2. Exploration of JackiePhos ligation and the viability of initial oxidative addition.

Scheme 2.

Given these results, we decided to revisit the influence of precatalyst and ligand on the outcome of the coupling reactions between 2a and N-allylurea substrate 1b. We had previously reported that the coupling of 2a with 1b in the presence of Cs2CO3, Pd(acac)2, and JackiePhos, afforded a 90% yield of the desired product 3b (Table 1, entry 1). We first carried out the key control experiment in which JackiePhos was omitted from these conditions,17 and obtained essentially the same result as when JackiePhos was included (entry 2). This result suggests that the actual ligand for Pd is acac, not JackiePhos, and is further supported by the fact that Pd(tfa)2 or other simple palladium pre-catalysts alone gave poor results (entries 3–5), but use of Pd(tfa)2 and acac provided 3b in 95% yield (entry 6).

Table 1:

Precatalyst effects.a

graphic file with name nihms-1823162-t0009.jpg
entry [Pd] Ligand (mol %) yield (%)b

1 Pd(acac)2 JackiePhos (16 mol %) 90c
2 Pd(acac)2 none 95
3 Pd2(dba)3 none 43
4 Pd(OAc)2 none 24
5 Pd(tfa)2 none 35
6 Pd(tfa)2 acac (8 mol %) 95
a

Conditions: 1.0 equiv 1b, 3.0 equiv 2a, 4 mol % [Pd], 2 equiv CS2CO3, dioxane (0.1 M), 100 °C, 16 h.

b

Yields were determined by 1H NMR analysis using phenanthrene as an internal standard.

c

Isolated yield reported in reference 10.

Revised mechanistic hypothesis and stereochemical details.

Given that the experiments described in Scheme 2 indicated our initial mechanistic hypothesis (Scheme 1) is not correct, we sought to gain additional information about the mechanism of these transformations. Michael’s previously reported Pd-catalyzed alkene diamination reactions of protected aminoalkenes with N-fluorobenzenesulfonimide have been shown to proceed via initial anti-aminopalladation of the alkene, followed by oxidative addition of the electrophile to an alkyl PdII complex, and subsequent reductive elimination with inversion of configuration from the resulting PdIV intermediate.6 In order to probe whether our reactions may proceed via a similar mechanism, we first sought to confirm that alkene aminopalladation could occur in the absence of the electrophile. As such, 1b was treated with Cs2CO3 and a stoichiometric amount of Pd(acac)2. These conditions afforded a roughly 1:1:4 mixture of products 8-10a in 62% combined NMR yield (Scheme 3). All three of these products likely arise from alkene aminopalladation (1b to 11a), with 8 likely derived from β-hydride elimination of 11a to 8a followed by alkene isomerization. The β-hydride elimination generates an equivalent of H-Pd(acac), which can react with 1b to afford 11b, or undergo disproportionation with 11a to form 11b and Pd(acac)2. Subsequent reductive elimination from 11b then produces 9. Finally 10a derives from aryl C–H functionalization of 11a. As such, the presence of the electrophile is not required for alkene aminopalladation to occur, and Pd(acac)2 is effective at inducing alkene aminopalladation of 1b under our otherwise standard reaction conditions.

Scheme 3. Evidence in support of initial alkene aminopalladation.

Scheme 3.

Since product 3d results from net anti-addition to the alkene, we sought to gain additional information about the stereochemistry of alkene aminopalladation and reductive elimination. The overall anti-addition could result from either anti-aminopalladation followed by reductive elimination with retention of configuration, or may derive from syn-aminopalladation and subsequent reductive elimination with inversion of configuration.18 In order to determine the alkene addition stereochemistry, we elected to take advantage of the relatively facile aminopalladation/C–H functionalization sequence described above (1b to 10a). As shown in Scheme 4, treatment of deuterated alkene substrate 1d with Pd(acac)2 and Cs2CO3 in the presence or absence of JackiePhos afforded 10b in modest yield (37% and 20%, respectively), with ca 1.5:1 diastereoselectivity favoring the product resulting from anti-aminopalladation.19

Scheme 4. Stereochemistry of alkene aminopalladationa.

Scheme 4.

aConditions: 1.0 equiv 1d or 1e, 3.0 equiv 2a, 1 equiv Pd(acac)2, 2 equiv ligand, 2 equiv Cs2CO3, dioxane (0.1 M), 100 °C, 1 h. bYields for entries 1 and 2 are isolated yields, whereas the yield for entry 3 was determined by 1H NMR analysis using phenanthrene as an internal standard.

Given the low diastereoselectivity of the conversion of 1d to 10b (ca 1.5:1 dr) relative to that for the conversion of 1d to 3d (Table 2, 6:1 dr), it seemed possible that the aminopalladation stereoselectivity was being eroded by reversible β-hydride elimination from complex 11d.20 However, the cyclization/C–H functionalization of di-deuterated substrate 1e, which should undergo much slower β–deuterium elimination from complex 11e than the rate for β–hydride elimination from complex 11d,21 proceeded with comparable (1.6:1) diastereoselectivity (Scheme 4, entry 3). Although the selectivity for anti- vs. syn aminopalladation is modest, this result suggests that since the overall conversion of 1d to 3d proceeds with net anti-addition, the C–N bond-forming reductive elimination must occur with retention of configuration. Thus, although these transformations are mechanistically similar to the reactions previously described by Michael,6 the nature of the nitrogen nucleophile has an impact on the stereochemistry of reductive elimination. Transformations of the more electron-rich O-benzoylhydroxylamine-derived electrophiles appear to involve stereoretentive reductive elimination, rather than stereoinvertive reductive elimination as observed with N-fluorobenzenesulfonimide by Michael.6

Table 2.

Effect of electrophile concentration and ligand on Anti-addition stereoselectivitya

graphic file with name nihms-1823162-t0010.jpg
entry equiv 2a ligand dr yield (%)b

1 3.0 JackiePhos 6:1 75c
2 3.0 none 6:1 74
3 1.5 JackiePhos 2.2:1 79
4 1.5 None 2.8:1 76
a

Conditions: 1.0 equiv 1d, 1.5 or 3.0 equiv 2a, 4 mol % Pd(acac)2, 8 mol% ligand, 2 equiv CS2CO3, dioxane (0.1 M), 100 °C, 16 h.

b

isolated yields.

c

Isolated yield reported in reference 10.

Collectively, the results of the experiments shown above suggest the mechanism of this transformation involves initial, and likely reversible, anti-aminopalladation of the alkene to generate intermediate 11d from 1d (Scheme 5).22 Oxidative addition of the amine electrophile 2a to 11d provides PdIV intermediate 13d, and then reductive elimination affords the observed product 3d with regeneration of the PdII catalyst.23 Although our control experiments illustrate that 2a does not undergo oxidative addition to Pd(acac)2, the alkylpalladium complex 11d is considerably more electron-rich than Pd(acac)2, and should undergo much more facile oxidative addition.

Scheme 5. Revised mechanistic hypothesis.

Scheme 5.

Effect of electrophile concentration on stereoselectivity.

In our initial studies, we found that the Pd(acac)2/JackiePhos - catalyzed coupling of 1d with 2a afforded 3d in 75% yield and 6:1 dr (Table 2, entry 1).10 A similar outcome was obtained when Pd(acac)2 alone (no phosphine) was used as the catalyst for this reaction (entry 2). However, when the amount of 2a was decreased to only 1.5 equiv, 3d was generated in much lower diastereoselectivity (2.2:1 to 2.8:1 dr; entries 3–4), regardless of whether or not JackiePhos was present. This suggests that the stereoselectivity of this transformation is dependent at least to some extent on the rate of the oxidative addition step in the catalytic cycle.

The influence of electrophile concentration on diastereoselectivity may be due to the relative rates of anti-aminopalladation/oxidative addition vs. syn-aminopalladation, coupled with kinetic vs thermodynamic control of selectivity. As illustrated in Scheme 6, it possible for 12d and 12d’ to equilibrate through exchange of the coordinated alkene for the anionic urea nitrogen, with displacement of an acac ligand. If anti-aminopalladation of 12d to 11d is faster than syn-aminopalladation of 12d’ to 11d’ (k1 > k2), and if the oxidative addition sequence from 11d to 13d is faster than retro-aminopalladation of 11d to 12d (k3 > k−1), then the selectivity for the anti-addition stereoisomer 3d should be relatively high.24

Scheme 6. Competing stereochemical pathways.

Scheme 6.

On the other hand, anti-aminopalladation complex 11d and its syn-addition diastereomer 11d’ are expected to be very close in energy, since they differ only in the configuration of an H/D stereocenter, and if oxidative addition is slow enough that 11d and 11d’ equilibrate through reversible aminopalladation (k−1 and k2 > k3), stereoselectivity should be relatively low, as it would reflect the thermodynamic ratio of 11d and 11d’. The data shown in Table 2 are consistent with the former scenario (k1 > k2 and k3 > k−1) when the concentration of 2a is relatively high. However, if the concentration of 2a is relatively low, the rate k3 of the bimolecular oxidative addition should decrease, and selectivity for the anti-addition product 3d should erode, as is observed.

Influence of acac structure on reactivity.

The use of acac as a “standalone” (no added phosphine) ligand in Pd-catalyzed cross-coupling reactions is extremely rare, and the influence of acac ligand structure on reactivity in Pd-catalyzed reactions has not thoroughly been explored.25 In order to determine whether acac structure has an impact on reactivity or yield, we quickly screened Pd(acac-F6)2 as a catalyst for the coupling of 1b with 2a to afford 3b (Table 3). Interestingly, this complex bearing an electron-deficient acac derivative gave very poor results (Table 3, entry 2, 10% yield of 3b) in comparison to the results obtained with Pd(acac)2 (entry 1, 90% yield). This result indicated the structure of the acac ligand does influence reactivity, and in a profound manner. Accordingly, we explored the coupling of more challenging urea substrates 1f and 1g with several different acac-derived ligands (Table 3). With our original catalyst system of Pd(acac)2/JackiePhos, the coupling of 1f with 2a afforded 3e in only 45% isolated yield (entry 3).10 By omitting JackiePhos and using only Pd(acac)2 as catalyst we observed 74% NMR yield of 3e (entry 4). Changing the ligand to 14 or 15a (entries 5 and 7) resulted in significantly reduced yields. However, improved results were obtained with 15c-15e (83–88%; entries 8–10). Finally, the best result was obtained with electron-rich analog 15f (84% isolated yield, entry 11). This ligand also provided considerably improved results in the reaction of 1g with 2a. Under our original conditions with JackiePhos as ligand, only a trace amount of desired product 3f was obtained (entry 12). However, use of 15f as a ligand in this reaction provided 3f in 42% isolated yield (entry 14). Thus, it appears that JackiePhos inhibits catalysis for some substrate combinations.

Table 3:

Acac ligand effects.a

graphic file with name nihms-1823162-t0011.jpg
entry R [Pd] ligand yield (%/)b

1 NO2 Pd(acac)2 none >95 (69)c
2 NO2 Pd(acac-F6)2 none 11d
3 Cl Pd(acac)2 JackiePhos 87(45)e
4 Cl Pd(acac)2 none 74
5 Cl Pd(tfa)2 14 16
6 Cl Pd(tfa)2 acac (15b) 77
7 Cl Pd(tfa)2 15a 24
8 Cl Pd(tfa)2 15c 83
9 Cl Pd(tfa)2 15d 86
10 Cl Pd(tfa)2 15e 88
11 Cl Pd(tfa)2 15f 94 (84)
12 OMe Pd(acac)2 JackiePhos <5
13 OMe Pd(acac)2 none 30
14 OMe Pd(tfa)2 15f 75 (42)
a

Conditions: 1.0 equiv 1b, 1.5 equiv 2a, 4 mol % [Pd], 8 mol % ligand, 2.0 equiv Cs2CO3, dioxane (0.1 M), 100 °C, 16 h.

b

Yields were determined by 1H NMR analysis using phenanthrene as an internal standard. Yields in parentheses are isolated yields.

c

The reaction was conducted for one hour.

d

The reaction was conducted with 3.0 equiv 2a.

e

Isolated yield from reference 10.

Further exploration of substrate scope.

Given the observation that JackiePhos may inhibit catalysis in some systems, combined with the fact that in our preliminary studies with Pd(acac)2/JackiePhos we only successfully coupled electrophiles derived from cyclic amines,10 we sought to further explore the scope of Pd/15f-catalyzed alkene diamination reactions. As shown in Table 4, this catalyst was effective for the preparation of cyclic guanidines (3a, 3g-i), ureas (3b, 3j-n), and amides (3o-p). However, the presence of a gem-dimethyl group adjacent to the amide carbonyl was necessary to obtain satisfactory results; a substrate lacking the gem-dimethyl group failed to react. This is presumably the result of a Thorpe-Ingold effect that facilitates the alkene aminopalladation step. The reactions were effective with both cyclic and acyclic secondary amine-derived electrophiles, and the coupling of the primary cyclohexylaminobenzoate electrophile with 1b afforded 3k in 45% yield. This expansion in scope is significant, as reactions of primary or acyclic secondary amine-derived electrophiles gave little or no desired product using our original conditions. Interestingly, although phosphine ligands appear to inhibit catalysis, an electrophile with a potentially coordinating heteroaryl group reacted smoothly with 1h and 1b to afford 3i and 3l in good yield. The presence of an ester functional group was also tolerated (3n).

Table 4:

Alkene diamination reactions.a,b

graphic file with name nihms-1823162-t0012.jpg
a

Conditions: 1.0 equiv 1a-j, 3.0 equiv 2a-e, 4 mol % Pd(tfa)2, 8 mol% 15f, 2 equiv CS2CO3, dioxane (0.1 M), 100 °C, 16 h, 0.2 mmol scale.

b

Isolated yields (average of two experiments).

c

The reaction was conducted on a 1.0 mmol scale

d

The reaction was conducted using benzotrifluoride as solvent (0.4 M).

SUMMARY AND CONCLUSION

In conclusion, through continued studies on Pd-catalyzed alkene diamination reactions involving O-benzoyl hydroxylamine electrophiles, we have demonstrated that phosphine ligands are not required, and in some cases inhibit catalysis. Instead, acac or acac-derivatives serve as the actual ligand for palladium that promotes reactivity in these transformations. In addition, we have illustrated that the structure of the acac ligand has a significant influence on reactivity, and this observation will likely find utility in other Pd-catalyzed reactions. We have also found that these transformations do not proceed via our originally postulated Pd(0/II) catalytic cycle that was initiated by oxidative addition of the N-electrophile to Pd(0). Instead, catalysis cycles through Pd(II)/Pd(IV) oxidation states, and is initiated by alkene aminopalladation. In comparison of our results with those of Michael, it appears that the stereochemical outcome of the reductive elimination from Pd(IV) is dependent on the nucleophilicity/basicity of the N-electrophile, with basic amines favoring reductive elimination with retention of configuration. Finally, this new catalyst system has allowed for a significant expansion of the scope of electrophiles that can be employed as coupling partners. Future studies will be directed towards further exploration of the classes of nucleophiles (in addition to guanidines, ureas, and amides) that can undergo diamination with O-benzoyl hydroxylamines.

EXPERIMENTAL SECTION

General:

All reactions were carried out under a nitrogen atmosphere using oven or flame-dried glassware and reactions conducted above room temperature were heated in an oil bath. All reagents, palladium sources, phosphine ligands, dicarbonyl ligands 14 and 15a-e, were obtained from commercial sources and used without further purification unless otherwise noted. N-Allylureas and guanidines 1a26, 1b,d,e,f,g,h,10 1i,7a 1j,27 ligand 15f,28 and electrophiles 2a,29,30 2b,30 2c,31 2d,32 2e,33 and 2f34 were prepared according to previously reported procedures. Dioxane was purified by distillation from Na metal and benzophenone under a nitrogen atmosphere. All yields refer to isolated compounds that are estimated to be ≥ 95% pure as judged by 1H NMR analysis unless otherwise noted. The yields reported in the experimental section describe the result of a single experiment, whereas yields reported in Table 4 are average yields of two or more experiments. Thus, the yields reported in the experimental section may differ from those shown in Table 4.

Preparation and Characterization of Products

General Procedure for Pd-Catalyzed Alkene Difunctionalization Reactions.

A flame-dried Schlenk tube equipped with a stirbar was cooled under a stream of N2 and charged with the appropriate palladium source (4 mol %), the appropriate ligand (8–16 mol %), the appropriate urea, guanidine, or amide substrate (0.2 mmol, 1 equiv), the appropriate electrophile (0.3–0.6 mmol, 1.5–3.0 equiv) and cesium carbonate (0.4 mmol, 2 equiv). The tube was purged with N2, dioxane (2 mL, 0.1 M) was added, and the reaction mixture was headed to 100 °C for 16 h. The mixture was then cooled to rt and filtered through cotton, which was then eluted with a small amount diethyl ether. The resulting solution was concentrated in vacuo, and the crude product was purified by flash chromatography on silica gel.

N-(1,3-Dibenzyl-4-(morpholinomethyl)imidazolidin-2-ylidene)cyanamide (3a).7a

The general procedure was used for the coupling of 1a (30.4 mg, 0.1 mmol, 1 equiv) with morpholino benzoate (2a) (103.6 mg, 0.5 mmol, 5 equiv) using a catalyst composed of Pd(TFA)2 (1.3 mg, 0.004 mmol, 4 mol %) and 1,3-bis(4-methoxyphenyl)propane-1,3-dione (15f) (2.3 mg, 0.008 mmol, 8 mol %) except the reaction was conducted on a 0.1 mmol scale with 1 mL of dioxane. The crude product was purified by flash chromatography on silica gel (10% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 62.0 mg (80%) of the title compound as a yellow oil. 1H NMR (500 MHz, CDCl3) δ 7.29–6.95 (m, 10 H), 5.24 (d, J = 15.5 Hz, 1 H), 4.56–4.46 (m, 2 H) 4.05 (d, J = 15.5 Hz, 1 H), 3.33–3.24 (m, 4 H), 3.01 (m, 1 H), 2.64 (app t, J = 9.5 Hz, 1 H), 2.52 (dd, J = 9.6, 7.1 Hz, 1 H), 1.87 (dd, J = 12.8, 5.6 Hz, 1 H), 1.77–1.66 (m, 4 H), 1.53 (dd, J = 12.8. 6.9 Hz, 1 H); 13C{1H} NMR (126 MHz, CDCl3) δ 158.4, 135.9, 135.4, 128.9, 128.8, 128.2, 128.1, 128.0, 127.9, 116.5, 66.7, 61.0, 54.1, 51.9, 49.5, 49.3, 47.8; IR (film) 2919, 2171, 1596 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C23H27N5O: 390.2288; found: 390.2292.

1-Benzyl-4-(morpholinomethyl)-3-(4-nitrophenyl)imidazolidin-2-one (3b).7a

The general procedure was used for the coupling of 1b (62.3 mg, 0.2 mmol, 1 equiv) with morpholino benzoate (2a) (62.2 mg, 0.3 mmol, 1.5 equiv) using a catalyst composed of Pd(acac)2 (2.4 mg, 0.008 mmol, 4 mol %) except the reaction was run for 1 h. The crude product was purified by flash chromatography on silica gel (10% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 56.5 mg (71%) of the title compound as a yellow solid, m.p. 108–110 °C. 1H NMR (400 MHz, CDCl3) δ 8.04 (d, J = 9.3 Hz, 2 H), 7.59 (d, J = 9.3 Hz, 2 H), 7.15–7.02 (m, 5 H), 4.30–4.19 (m, 2 H), 3.44–3.34 (m, 4 H), 2.85 (dd, J = 8.9, 2.8 Hz, 1 H), 2.75 (t, J = 8.7 Hz, 1 H), 1.98 (dd, J = 13.0, 3.1 Hz, 1 H), 1.97–1.88 (m, 2 H), 1.83–1.75 (m, 2 H), 1.71 (dd, J = 13.0, 9.3 Hz, 1 H); 13C{1H} NMR (126 MHz, C6D6) δ 155.8, 145.0, 142.0, 136.6, 128.6, 128.2, 127.9, 124.6, 116.8, 66.4, 58.7, 53.8, 50.0, 47.5, 45.3; IR (film) 2921, 1709 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C21H24N4O4: 397.1870; found: 397.1868.

1-Benzyl-4-(morpholinomethyl)-3-(4-nitrophenyl)imidazolidin-2-one (3b).7a

The general procedure was used for the coupling of 1b (31.1mg, 1.0 mmol, 1 equiv) with morpholino benzoate (2a) (62.2 mg, 3.0 mmol, 1.5 equiv) using a catalyst composed of Pd(acac)2 (1.33 mg, 0.004 mmol, 4 mol %) except the reaction was run for 1 h. The crude product was purified by flash chromatography on silica gel (10% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 35.0 mg (88%) of the title compound as a yellow solid, m.p. 108–110 °C. Spectroscopic data were identical to those tabulated above.

(4S*,4’R*)-1-Benzyl-4-(morpholinomethyl-d)-3-(4-nitrophenyl)imidazolidin-2-one (3d).7a

The general procedure was used for the coupling of 1d (62.3 mg, 0.2 mmol, 1 equiv) with morpholino benzoate (2a) (62.2 mg, 0.3 mmol, 1.5 equiv or 124.3 mg, 0.6 mmol, 3 equiv) using a catalyst composed of Pd(acac)2 (2.4 mg, 0.008 mmol, 4 mol %) and either no ligand or JackiePhos (12.7 mg, 0.016 mmol, 8 mol %) The crude product was purified by flash chromatography on silica gel (10% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 58.6 mg (6:1 dr, 74%, entry 2), 62.8 mg (2.2:1 dr, 79%, entry 3), 60.3 mg (2.8:1 dr, 76%, entry 4), and 59.0 mg (6.7:1, 74%, ref 6) of the title compound as a yellow foam. Characterization data for entry 2: This compound was obtained as a ca 6:1 mixture of diastereomers as judged by 1H NMR analysis; data are for the mixture. 1H NMR (500 MHz, CDCl3) δ 8.04 (d, J = 9.3 Hz, 2 H), 7.61 (d, J = 9.3 Hz, 2 H), 7.14–7.02 (m, 5 H), 4.31–4.19 (m, 2 H), 3.52–3.24 (m, 5 H), 2.85 (dd, J = 8.9, 2.8 Hz, 1 H), 2.75 (t, J = 8.7 Hz, 1 H), 1.95–1.88 (m, 3 H), 1.83–1.79 (m, 2 H), 1.72 (dd, J = 13.7, 9.3 Hz, 1 H); 13C{1H} NMR (126 MHz, CDCl3) δ 156.4, 145.2, 142.0, 136.1, 128.8, 128.3, 127.9, 124.9, 117.4, 66.7, 58.9 (t, J = 19.0 Hz), 54.2, 50.8, 47.8, 45.9; IR (film) 2922, 1710 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C21H23DN4O4: 398.1933 found: 398.1929.

1-Benzyl-3-(4-chlorophenyl)-4-(morpholinomethyl)imidazolidin-2-one (3e).7a

The general procedure was used for the coupling of 1f (60.2 mg, 0.2 mmol, 1 equiv) with morpholino benzoate (2a) (62.2 mg, 0.3 mmol, 1.5 equiv) using a catalyst composed of Pd(TFA)2 (2.7 mg, 0.008 mmol, 4 mol %) and 1,3-bis(4-methoxyphenyl)propane-1,3-dione (15f) (4.5 mg, 0.016 mmol, 8 mol %). The crude product was purified by flash chromatography on silica gel (10% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 65.4 mg (85%) of the title compound as a yellow oil. 1H NMR (400 MHz, CDCl3) δ 7.46 (d, J = 8.8 Hz, 2 H), 7.38–7.27 (m, 7 H), 4.46 (s, 2 H), 4.27 (tt, J = 8.6, 3.8 Hz, 1 H), 3.66–3.54 (m, 4 H), 3.45 (t, J = 8.9 Hz, 1 H), 3.29 (dd, J = 9.0, 4.1 Hz, 1 H), 2.55 (dd, J = 12.9, 3.3 Hz, 1 H), 2.50–2.23 (m, 5 H); 13C{1H} NMR (126 MHz, CDCl3) δ 157.4, 137.7, 136.7, 128.9, 128.7, 128.3, 128.2, 127.6, 121.1, 66.8, 59.6, 54.2, 51.1, 47.9, 46.4; IR (film) 2917, 1700 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C21H24ClN3O2: 386.1635; found: 386.163.

1-Benzyl-3-(4-methoxyphenyl)-4-(morpholinomethyl)imidazolidin-2-one (3f).

The general procedure was used for the coupling of 1g (59.3 mg, 0.2 mmol, 1 equiv) with morpholino benzoate (2a) (124.4 mg, 0.6 mmol, 3 equiv) using a catalyst composed of Pd(TFA)2 (2.7 mg, 0.008 mmol, 4 mol %) and 1,3-bis(4-methoxyphenyl)propane-1,3-dione (15f) (4.5 mg, 0.016 mmol, 8 mol %). The crude product was purified by flash chromatography on silica gel (10% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 34.5 mg (45%) of the title compound as a clear oil. 1H NMR (500 MHz, CDCl3) δ 7.40 – 7.27 (m, 7H), 6.93 – 6.83 (m, 2H), 4.56 – 4.36 (m, 2H), 4.23 (s, br, 1H), 3.80 (s, 3H), 3.60 (s, br, 4H), 3.45 (s, br, 1H), 3.26 (s, br, 1H), 2.54 (s, br, 1H), 2.47 – 2.25 (m, 5H); 13C{1H} NMR (126 MHz, CDCl3) at 50°C δ 158.5, 156.7, 137.4, 132.4, 128.8, 128.5, 127.7, 123.6, 114.6, 67.0, 60.5, 55.7, 54.4, 52.4, 48.4, 47.2. IR (film) 2915, 2852, 1693, 1512 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C22H28N3O3: 382.2125; found: 382.2123.

N-{1,3-Dibenzyl-4-[(diethylamino)methyl]imidazolidin-2-ylidene}cyanamide (3g).

The general procedure was used for the coupling of 1a (60.9 mg, 0.2 mmol, 1 equiv) with N,N-diethylamino benzoate (2b) (116.0 mg, 0.6 mmol, 3 equiv) using a catalyst composed of Pd(TFA)2 (2.7 mg, 0.008 mmol, 4 mol %) and 1,3-bis(4-methoxyphenyl)propane-1,3-dione (15f) (4.5 mg, 0.016 mmol, 8 mol %). The crude product was purified by flash chromatography on silica gel (30% 3:1 (ethyl acetate:ethanol) in hexanes) to give the desired product along with minor impurities. The compound was dissolved in ethyl acetate and extracted with 1M HCl (3 × ca. 3 mL). The combined aqueous layers were basified to a pH of 14 using 3M NaOH (ca. 5 mL) and then extracted with EtOAc (3 × ca. 3 mL). The combined organic layers were dried with anhydrous Na2SO4, filtered, and concentrated in vacuo to give to afford 41.5 mg (55%) of the title compound as a clear oil. 1H NMR (500 MHz, CDCl3) δ 7.36 – 7.33 (m, 4H), 7.30 – 7.27 (m, 6H), 5.29 – 5.26 (d, J = 15.5, 1H), 4.81 – 4.78 (d, J = 15.1 Hz, 1H), 4.65 – 4.62 (d, J = 15.2 Hz, 1H), 4.33 – 4.29 (d, J = 15.5, 1H), 3.56 – 3.50 (m, 1H), 3.38 – 3.35 (t, 1H), 3.14 – 3.11 (m, 1H), 2.55 – 2.51 (dd, J = 13.0 Hz, 4.9 Hz, 1H), 2.39 – 2.30 (m, 4H), 2.28 – 2.24 (m, 1H), 0.85 – 0.82 (t, 6H).13C{1H} NMR (126 MHz, CDCl3) δ 158.2, 136.0, 135.5, 128.8, 128.3, 128.2, 128.1, 128.0, 116.7, 55.3, 53.2, 49.5, 49.4, 47.7, 47.6, 11.6. IR (film) 3030, 2967, 2169, 1593, 1507 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C23H30N5: 376.2496; found: 376.2487.

N-{1,3-Dibenzyl-4-[(diethylamino)methyl]imidazolidin-2-ylidene}-4-methylbenzenesulfonamide (3h).

The general procedure was used for the coupling of N-{[allyl(benzyl)amino][benzylamino]methylene}-4-methylbenzenesulfonamide 1h (86.7 mg, 0.2 mmol) with N,N-diethylamino benzoate (2b) (116.0 mg, 0.6 mmol) using a catalyst composed of Pd(TFA)2 (2.66 mg, 0.008 mmol, 4 mol %) and 1,3-bis(4-methoxyphenyl)propane-1,3-dione (15f) (4.55 mg, 0.016 mmol, 8 mol %). The crude product was purified by flash chromatography on silica gel (30% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 83.4 mg (83%) of the title compound as a yellow oil. 1H NMR (500 MHz, C6D6) δ 8.29 (d, J = 8.2 Hz, 2H), 7.32 (d, J = 7.3 Hz, 2H), 7.27 (d, J = 7.2 Hz, 2H), 7.15 – 7.08 (m, 4H), 7.07 – 7.02 (m, 2H), 6.83 (d, J = 8.0 Hz, 2H), 5.51 (d, J = 15.2 Hz, 1H), 4.94 (d, J = 14.9 Hz, 1H), 4.51 (d, J = 14.9 Hz, 1H), 4.20 (d, J = 15.2 Hz, 1H), 3.23 – 3.13 (m, 1H), 2.86 – 2.75 (m, 2H), 2.15 (dd, J = 13.1, 4.9 Hz, 1H), 2.06 – 1.95 (m, 4H), 1.90 (s, 3H), 1.89 – 1.85 (m, 1H), 0.60 (t, J = 7.1 Hz, 6H); 13C{1H} NMR (126 MHz, C6D6) δ 156.1, 144.3, 140.4, 137.0, 136.5, 128.8, 128.6, 128.5, 128.5, 128.0, 127.6, 126.2, 55.1, 52.8, 50.6, 49.0, 48.6, 47.2, 20.7, 11.4; IR (film) 2968, 1563, 1497, 1453 cm-1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C29H37N4O2S: 505.2632; found:505.2629.

N-(1,3-Dibenzyl-4-{[4-(pyrimidin-2-yl)piperazin-1-yl]methyl}imidazolidin-2-ylidene)-4-methylbenzenesulfonamide (3i).

The general procedure was used for the coupling of 1h (86.7 mg, 0.2 mmol) and 4-(pyrimidin-2-yl)piperazin-1-yl benzoate (2c) (170.6 mg, 0.6 mmol) using a catalyst composed of Pd(TFA)2 (2.66 mg, 0.008 mmol, 4 mol %) and 1,3-bis(4-methoxyphenyl)propane-1,3-dione (15f) (4.55 mg, 0.016 mmol, 8 mol %). The crude product was purified by flash chromatography on silica gel (30% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 89.1 mg (75%) of the title compound as a clear oil. 1H NMR (500 MHz, C6D6) δ 8.30 (d, 2H), 8.21 (d, J = 9.3 Hz, 2H), 7.76 (d, J = 9.2 Hz, 2H), 7.40 – 7.29 (m, 5H), 6.49 (dd, J = 6.6, 2.9 Hz, 1H), 4.51 (q, J = 14.9 Hz, 2H), 4.40 (m, J = 8.6 Hz, 1H), 3.75 (m, 4H), 3.53 (t, J = 8.8 Hz, 1H), 3.41 (d, J = 9.2 Hz, 1H), 2.64 (d, J = 13.0 Hz, 1H), 2.59 – 2.53 (m, 2H), 2.49 – 2.35 (m, 3H); 13C{1H} NMR (126 MHz, C6D6) δ 161.8, 157.4, 156.2, 144.2, 140.5, 137.0, 136.4, 128.8, 128.6, 128.5, 128.5, 128.4, 127.5, 126.2, 109.7, 60.2, 53.2, 51.6, 50.8, 48.8, 48.5, 43.4, 20.7; IR (film) 3028, 2924, 1585, 1564, 1497 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C33H38N7O2S: 596.2802; found: 596.2801.

1-Benzyl-4-{[benzyl(methyl)amino]methyl}-3-(4-nitrophenyl)imidazolidin-2-one (3j).

The general procedure was used for the coupling of 1b (62.3 mg, 0.2 mmol, 1 equiv) with N-methylbenzylamino benzoate (2d) (97.0 mg, 0.4 mmol, 2 equiv) using a catalyst composed of Pd(TFA)2 (2.7 mg, 0.008 mmol, 4 mol %) and 1,3-bis(4-methoxyphenyl)propane-1,3-dione (15f) (4.5 mg, 0.016 mmol, 8 mol %) except the crude reaction mixture was filtered through Celite. The crude product was purified by flash chromatography on silica gel (30% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 70.3 mg (82%) of the title compound as a yellow oil. 1H NMR (500 MHz, CDCl3) δ 8.16 (d, J = 9.3 Hz, 2H), 7.64 (d, J = 9.3 Hz, 2H), 7.38 – 7.25 (m, 8H), 7.21 (d, J = 6.6 Hz, 2H), 4.41 (q, J = 10.8 Hz, 2H), 4.21 (m, 1H), 3.52 (d, J = 12.9 Hz, 1H), 3.47 – 3.39 (m, 2H), 3.30 (dd, J = 9.1, 2.5 Hz, 1H), 2.60 (d, J = 2.7 Hz, 1H), 2.48 – 2.41 (m, 1H), 2.29 (s, 3H); 13C{1H} NMR (126 MHz, CDCl3) δ 156.5, 145.5, 141.9, 138.5, 136.3, 129.2, 128.9, 128.6, 128.4, 128.0, 127.7, 125.1, 117.3, 63.5, 57.6, 52.0, 47.9, 45.8, 44.0; IR (film) 2921, 2850, 1712, 1595, 1503 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C25H27N4O3: 431.2085; found: 431.2063.

1-Benzyl-4-[(cyclohexylamino)methyl]-3-(4-nitrophenyl)imidazolidin-2-one (3k).

The general procedure was used for the coupling of 1b (62.3 mg, 0.2 mmol, 1 equiv) with O-benzoyl-N-cyclohexylhydroxylamine (2e) (88.0 mg, 0.4 mmol, 2 equiv) using a catalyst composed of Pd(TFA)2 (2.7 mg, 0.008 mmol, 4 mol %) and 1,3-bis(4-methoxyphenyl)propane-1,3-dione (15f) (4.5 mg, 0.016 mmol, 8 mol %) except the crude reaction mixture was filtered through Celite. The crude product was purified by flash chromatography on silica gel (30% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 46.2 mg (52%) of the title compound as a yellow oil. 1H NMR (500 MHz, CDCl3) δ 8.21 (d, J = 9.3 Hz, 2H), 7.77 (d, J = 9.2 Hz, 2H), 7.38 – 7.28 (m, 5H), 4.57 (d, J = 15.0 Hz, 1H), 4.42 (d, J = 15.0 Hz, 1H), 4.33 – 4.25 (m, 1H), 3.49 (t, J = 9.1 Hz, 1H), 3.37 (dd, J = 8.9, 3.2 Hz, 1H), 2.86 (dd, J = 12.5, 2.6 Hz, 1H), 2.76 (dd, J = 12.5, 7.7 Hz, 1H), 2.29 (m, 1H), 1.78 – 1.62 (m, 4H), 1.60 – 1.54 (m, 1H), 1.24 – 1.08 (m, 4H), 1.01 – 0.87 (m, 2H); 13C{1H} NMR (126 MHz, CDCl3) δ 156.5, 145.2, 142.0, 136.2, 128.7, 128.2, 127.8, 124.9, 117.6, 56.8, 53.3, 47.8, 46.9, 45.5, 33.7, 26.0, 24.8; IR (film) 2926, 2852, 1698, 1596 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C23H29N4O3: 409.2234; found: 409.2230.

1-Benzyl-3-(4-nitrophenyl)-4-{[4-(pyrimidin-2-yl)piperazin-1-yl]methyl}imidazolidin-2-one (3l).

The general procedure was used for the coupling of 1b (62.3 mg, 0.2 mmol, 1 equiv) with 4-(pyrimidin-2-yl)piperazin-1-yl benzoate (2c) (113.7 mg, 0.4 mmol, 2 equiv) using a catalyst composed of Pd(TFA)2 (2.7 mg, 0.008 mmol, 4 mol %) and 1,3-bis(4-methoxyphenyl)propane-1,3-dione (15f) (4.5 mg, 0.016 mmol, 8 mol %) except the crude reaction mixture was filtered through Celite. The crude product was purified by flash chromatography on silica gel (30% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 85.9 mg (91%) of the title compound as a yellow oil. 1H NMR (500 MHz, CDCl3) δ 8.30 (d, 2H), 8.21 (d, J = 9.3 Hz, 2H), 7.76 (d, J = 9.2 Hz, 2H), 7.40 – 7.29 (m, 5H), 6.49 (dd, J = 6.6, 2.9 Hz, 1H), 4.51 (q, J = 14.9 Hz, 2H), 4.40 (m, 1H), 3.75 (m, 4H), 3.53 (t, J = 8.8 Hz, 1H), 3.41 (d, J = 9.2 Hz, 1H), 2.64 (d, J = 13.0 Hz, 1H), 2.59 – 2.53 (m, 2H), 2.49 – 2.35 (m, 3H); 13C{1H} NMR (126 MHz, CDCl3) δ 157.9, 129.0, 128.4, 128.1, 125.2, 117.6, 110.3, 59.2, 53.9, 51.3, 48.0, 46.1, 43.7; IR (film) 2925, 2851, 1709, 1585, 1548, 1501 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C25H28N7O3: 474.2255; found: 474.2251.

1-Benzyl-4-[(diethylamino)methyl]-3-(4-nitrophenyl)imidazolidin-2-one (3m).

The general procedure was used for the coupling of 1b (62.3 mg, 0.2 mmol, 1 equiv) with N,N-diethylamino benzoate (2b) (97.0 mg, 0.4 mmol, 2 equiv) using a catalyst composed of Pd(TFA)2 (2.7 mg, 0.008 mmol, 4 mol %) and 1,3-bis(4-methoxyphenyl)propane-1,3-dione (15f) (4.5 mg, 0.016 mmol, 8 mol %) except the crude reaction mixture was filtered through Celite. The crude product was purified by flash chromatography on silica gel (30% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 56.6 mg (74%) of the title compound as a yellow oil. 1H NMR (500 MHz, CDCl3) δ 8.20 (d, 2H), 7.77 (d, 2H), 7.39 – 7.27 (m, 5H), 4.60 (d, J = 14.8 Hz, 1H), 4.37 (d, J = 14.8 Hz, 1H), 4.32 – 4.23 (m, 1H), 3.45 (t, J = 8.8 Hz, 1H), 3.37 (d, J = 9.1 Hz, 1H), 2.62 (d, J = 13.1 Hz, 1H), 2.58 – 2.49 (m, 2H), 2.46 – 2.36 (m, 3H), 0.91 (t, J = 7.0 Hz, 6H); 13C{1H} NMR (126 MHz, CDCl3) δ 156.4, 145.5, 141.8, 136.3, 128.8, 128.3, 127.8, 125.0, 117.2, 53.9, 52.0, 47.8, 47.7, 45.7, 11.8; IR (film) 2969, 1712, 1595, 1503 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C21H27N4O3: 383.2085; found: 383.2078.

Ethyl 1-{[1-benzyl-3-(4-nitrophenyl)-2-oxoimidazolidin-4-yl]methyl}piperidine-4-carboxylate (3n).

The general procedure was used for the coupling of 1b (62.3 mg, 0.2 mmol, 1 equiv) with ethyl 1-(benzoyloxy)piperidine-4-carboxylate (2f) (111.0 mg, 0.4 mmol, 2 equiv) using a catalyst composed of Pd(TFA)2 (2.7 mg, 0.008 mmol, 4 mol %) and 1,3-bis(4-methoxyphenyl)propane-1,3-dione (15f) (4.5 mg, 0.016 mmol, 8 mol %) except the crude reaction mixture was filtered through Celite. The crude product was purified by flash chromatography on silica gel (30% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 75.0 mg (81%) of the title compound as a yellow oil. 1H NMR (500 MHz, CDCl3) δ 8.21 (d, J = 9.3 Hz, 2H), 7.75 (d, J = 9.3 Hz, 2H), 7.39 – 7.29 (m, 5H), 4.49 (q, J = 14.9 Hz, 2H), 4.36 – 4.28 (m, 1H), 4.12 (q, J = 7.1 Hz, 2H), 3.48 (t, J = 8.9 Hz, 1H), 3.34 (dd, J = 9.2, 2.6 Hz, 1H), 2.86 (d, J = 11.1 Hz, 1H), 2.64 – 2.51 (m, 2H), 2.36 (dd, J = 12.9, 9.4 Hz, 1H), 2.28 – 2.05 (m, 3H), 1.84 (t, J = 14.8 Hz, 2H), 1.73 – 1.61 (m, 2H), 1.25 (t, J = 7.1 Hz, 3H); 13C{1H} NMR (126 MHz, CDCl3) δ 174.8, 156.4, 145.4, 141.9, 136.2, 128.9, 128.6, 128.3, 128.1, 124.9, 117.4, 117.2, 60.4, 59.0, 54.5, 52.9, 51.4, 51.3, 47.8, 46.0, 40.7, 28.2, 28.1, 14.3, 14.1; IR (film) 2927, 1714, 1595, 1503 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C25H31N4O5: 467.2296; found: 467.2297.

1-Methoxy-3,3-dimethyl-5-(morpholinomethyl)pyrrolidin-2-one (3o).

The general procedure was used for the coupling of 1i (31.4 mg, 0.2 mmol, 1 equiv) with morpholino benzoate (2a) (124.4 mg, 0.6 mmol, 3 equiv) and a catalyst composed of Pd(TFA)2 (2.7 mg, 0.008 mmol, 4 mol %) and 1,3-bis(4-methoxyphenyl)propane-1,3-dione (15f) (4.5 mg, 0.016 mmol, 8 mol %). The crude product was purified by flash chromatography on silica gel (20% 3:1 (ethyl acetate:ethanol) in hexanes) to afford 33.4 mg (69%) of the title compound as a pale-yellow oil. 1H NMR (500 MHz, C6D6) δ 3.59 (s, 3H), 3.56 – 3.51 (m, 4H), 3.41 (m, 1H), 2.42 (dd, J = 12.5, 4.4 Hz, 1H), 2.21 – 2.07 (m, 4H), 2.03 (dd, J = 12.6, 7.3 Hz, 1H), 1.56 (dd, J = 12.7, 7.2 Hz, 1H), 1.31 (dd, J = 12.7, 7.9 Hz, 1H), 1.13 (s, 3H), 0.97 (s, 3H); 13C{1H} NMR (126 MHz, CDCl3) δ 176.3, 66.7, 62.6, 61.5, 54.4, 52.7, 38.1, 37.5, 25.9, 25.2; IR (film) 2961, 2930, 2853, 2812, 1706 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C12H23N2O3: 243.1703; found: 243.1707.

3,3-Dimethyl-5-(morpholinomethyl)-1-(4-nitrophenyl)pyrrolidin-2-one (3p).

The general procedure was used for the coupling of 1j (74.4 mg, 0.3 mmol, 1 equiv) with morpholino benzoate (2a) (93.3 mg, 0.45 mmol, 1.5 equiv) and a catalyst composed of Pd(TFA)2 (3.9 mg, 0.012 mmol, 4 mol %) and 1,3-bis(4-methoxyphenyl)propane-1,3-dione (15f) (4.5 mg, 0.024 mmol, 8 mol %), except the reaction was conducted on a 0.3 mmol scale with α,α,α-trifluorotoluene (0.75 mL, 0.4M) instead of dioxane as solvent. The crude product was purified by flash chromatography on silica gel (20% 3:1 (ethyl acetate:ethanol) in hexanes) to afford the desired product along with minor impurities. The compound was dissolved in ethyl acetate and extracted with 1M HCl (3 × ca. 3 mL). The combined aqueous layers were basified to a pH of 14 using 3M NaOH (ca. 5 mL) and then extracted with EtOAc (3 × ca. 3 mL). The combined organic layers were dried with anhydrous Na2SO4, filtered, and concentrated in vacuo to give 55.0 mg (55%) of the title compound as a pale-yellow oil. 1H NMR (500 MHz, CDCl3) δ 8.23 (d, J = 9.2 Hz, 2H), 7.68 (d, J = 8.7 Hz, 2H), 4.45 – 4.26 (m, 1H), 3.71 – 3.51 (m, 4H), 2.65 – 2.58 (m, 1H), 2.49 – 2.30 (m, 5H), 2.26 (dd, J = 13.0, 7.9 Hz, 1H), 1.94 (dd, J = 13.1, 6.1 Hz, 1H), 1.33 (s, 3H), 1.22 (s, 3H); 13C{1H} NMR (126 MHz, CDCl3) δ 179.9, 144.3, 124.6, 122.6, 66.9, 62.3, 54.3, 54.0, 41.3, 39.3, 26.4, 26.1, one carbon signal is missing due to accidental equivalence; IR (film) 2964, 2855, 2247, 1698, 1592 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C17H24N3O4: 334.1761; found: 334.1755.

Experiment shown in Scheme 3.

The general procedure was used except 2a was not included in the reaction. Instead 1a (15.6 mg, 0.05 mmol, 1 equiv), Pd(acac)2 (15.2 mg, 0.05 mmol, 1 equiv), Cs2CO3 (32.6 mg, 0.1 mmol, 2 equiv), and dioxane (0.5 mL, 0.1M) were heated to 100 °C for 16 h. This reaction afforded a mixture of 8 (10%), 9 (12%), and 10a (40%) as judged by 1H NMR analysis using phenanthrene as an internal standard. The crude products were not isolated from this experiment. Instead, a separate reaction was carried out using NaOtBu as the base (this led to higher NMR yields of the three products) as described below.

1-Benzyl-4-methyl-3-(4-nitrophenyl)-1,3-dihydro-2H-imidazol-2-one (8).

The general procedure was used except 2a was not included in the reaction, and NaOtBu was used as the base. Substrate 1b (31.1 mg, 0.1 mmol, 1 equiv), Pd(acac)2 (30.5 mg, 0.1 mmol, 1 equiv), and NaOtBu (19.2 mg, 0.2 mmol, 2 equiv), and dioxane (1 mL, 0.1M) were heated to 100 °C for 16 h. The crude product was purified by flash chromatography on silica gel (20% ethyl acetate in hexanes) and then further purified by preparative TLC (40% EtOAc in hexanes) to afford 5.1 mg (17%) of the title compound as a yellow oil. 1H NMR (500 MHz, CDCl3) δ 8.33 (d, J = 8.9 Hz, 2H), 7.56 (d, J = 8.9 Hz, 2H), 7.41 – 7.29 (m, 5H), 6.05 (s, 1H), 4.81 (s, 2H), 2.01 (s, 3H); 13C{1H} NMR (126 MHz, CDCl3) δ 152.6, 146.3, 141.1, 136.7, 129.0, 128.3, 128.2, 127.2, 124.7, 118.2, 109.1, 47.3, 11.8; IR (film) 3116, 2926, 1690, 1641, 1594 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C17H16N3O3: 310.1186; found: 310.1185. In addition, this procedure also afforded 1-benzyl-4-methyl-3-(4-nitrophenyl)imidazolidin-2-one 9 (4.9 mg, 16%, yellow oil), and 10a (ca 5 mg, ca 16%, yellow oil). Data for 9 and 10a are provided below.

1-Benzyl-4-methyl-3-(4-nitrophenyl)imidazolidin-2-one (9).

1H NMR (500 MHz, CDCl3) δ 8.22 (d, J = 9.4 Hz, 2H), 7.73 (d, J = 9.3 Hz, 2H), 7.41 – 7.28 (m, 5H), 4.60 – 4.44 (m, 2H), 4.44 – 4.31 (m, 1H), 3.57 (app. t, J = 8.8 Hz, 1H), 3.00 (dd, J = 8.9, 3.8 Hz, 1H), 1.35 (d, J = 6.2 Hz, 3H); 13C{1H} NMR (126 MHz, CDCl3) δ 156.5, 145.2, 142.2 136.3, 129.0, 128.4, 128.0, 125.1, 118.0, 48.9, 48.1, 29.9, 19.2; IR (film) 2917, 2849, 1710, 1596 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C17H18N3O3: 312.1343; found: 312.1342.

2-Benzyl-7-nitro-1,2,9,9a-tetrahydro-3H-imidazo[1,5-a]indol-3-one (10a).

1H NMR (500 MHz, CDCl3) δ 8.18 (d, J = 8.7 Hz, 1H), 8.05 (s, 1H), 7.57 (d, J = 8.7 Hz, 1H), 7.40 – 7.27 (m, 5H), 4.71 (p, J = 9.1 Hz, 1H), 4.59 (d, J = 14.9 Hz, 1H), 4.32 (d, J = 15.0 Hz, 1H), 3.73 (t, J = 9.1 Hz, 1H), 3.38 (dd, J = 16.2, 9.4 Hz, 1H), 3.27 (t, J = 8.7 Hz, 1H), 3.01 (dd, J = 16.2, 9.2 Hz, 1H); 13C{1H} NMR (126 MHz, CDCl3) δ 157.3, 148.7, 144.0, 136.1, 133.2, 129.0, 128.4, 128.1, 125.3, 121.1, 114.4, 56.4, 51.2, 48.1, 36.0; IR (film) 2922, 1712, 1596 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C17H16N3O3 310.1186; found: 310.1195.

(9R*,9aR*)-2-Benzyl-7-nitro-1,2,9,9a-tetrahydro-3H-imidazo[1,5-a]indol-3-one-9-d (10b).

The general procedure was used except 2a was not included in the reaction. Instead 1d (31.2 mg, 0.1 mmol, 1 equiv), Pd(acac)2 (30.4 mg, 0.1 mmol, 1 equiv), JackiePhos, if appropriate, (159.3 mg, 0.2 mmol, 2 equiv), Cs2CO3 (65.2 mg, 0.2 mmol, 2 equiv), and dioxane (0.1 mL, 0.1M) were heated to 100 °C for 16 h. The crude product was isolated via preparative TLC (40% EtOAc in hexanes) to afford the title compound as a yellow oil 11.6 mg (entry 1, 1.7:1 dr, 37%) and 6.3 mg (entry 2, 1.5:1 dr, 20%). 1H NMR for entry 1: (500 MHz, CDCl3) δ 8.18 (dd, J = 8.7, 2.3 Hz, 1H), 8.08 – 7.99 (m, 1H), 7.57 (d, J = 8.7 Hz, 1H), 7.40 – 7.25 (m, 5H), 4.70 (q, J = 8.7 Hz, 1H), 4.59 (d, J = 14.9 Hz, 1H), 4.32 (d, J = 14.9 Hz, 1H), 3.73 (t, J = 9.1 Hz, 1H), 3.36 (d, J = 9.5 Hz, 0.40H), 3.27 (t, J = 9.3, 8.0 Hz, 1H), 2.99 (d, J = 9.2 Hz, 0.65H); 1H NMR for entry 2: (500 MHz, CDCl3) δ 8.18 (dd, J = 8.6, 2.3 Hz, 1H), 8.10 – 8.01 (m, 1H), 7.57 (d, J = 8.7 Hz, 1H), 7.40 – 7.25 (m, 5H), 4.71 (q, J = 8.7 Hz, 1H), 4.59 (d, J = 14.9 Hz, 1H), 4.32 (d, J = 14.9 Hz, 1H), 3.73 (t, J = 9.1 Hz, 1H), 3.36 (d, J = 9.4 Hz, 0.43H), 3.27 (t, J = 8.6 Hz, 1H), 2.99 (d, J = 9.3 Hz, 0.64H); 13C NMR for entry 2 (176 MHz, CDCl3) δ 157.2, 148.5, 143.8, 135.9, 132.93, 128.9, 128.2, 127.9, 125.1, 121.0, 114.2, 56.1, 51.0, 47.9; IR (film) 2926, 1712, 1595 cm−1, HRMS (ESI TOF) m/z: [M + H]+ calcd for C17H14DN3O3: 311.1256; found: 311.1249.

(9R*,9aR*)-2-benzyl-7-nitro-1,2,9,9a-tetrahydro-3H-imidazo[1,5-a]indol-3-one-9,9a-d2 (10c).

The general procedure was used except 2a was not included in the reaction. Instead 1e (15.9 mg, 0.05 mmol, 1 equiv), Pd(acac)2 (15.3 mg, 0.05 mmol, 1 equiv), Cs2CO3 (32.6 mg, 0.1 mmol, 2 equiv), and dioxane (0.05 mL, 0.1M) were heated to 100 °C for 16 h. The crude product (1.6:1 dr, 38% crude 1H NMR yield). was purified via a pipet column (2% methanol in CH2Cl2) followed by preparative TLC (40% EtOAc in hexanes). However, in the midst of COVID-19 shutdown confusion, one of the authors neglected to record the mass of this product. As such, we report only the NMR yield, but provide data for the product. 1H NMR (500 MHz, CDCl3) δ 8.18 (d, J = 8.4 Hz, 1H), 8.05 (s, 1H), 7.57 (dd, J = 8.7, 2.4 Hz, 1H), 7.39 – 7.27 (m, 5H), 4.63 – 4.52 (m, 1H), 4.35 – 4.29 (m, 1H), 3.72 (d, J = 9.2 Hz, 1H), 3.36 (s, 0.45H), 3.26 (d, J = 9.3 Hz, 1H), 2.99 (s, 0.70H); 13C NMR (126 MHz, CDCl3) δ 157.2, 148.6, 143.8, 135.9, 132.9, 128.9, 128.2, 127.9, 125.1, 121.0, 121.0, 114.2, 50.9, 50.9, 48.0, 35.7, 35.4, 31.9; IR (film) 2923, 2852, 1713, 1595 cm−1; HRMS (ESI TOF) m/z: [M + H]+ calcd for C17H13D2N3O3: 312.1312; found: 312.1305.

Supplementary Material

Supporting Information

ACKNOWLEDGMENT

The authors thank the NIH-NIGMS (GM 124040) for financial support of this work. JKK is the recipient of a University of Michigan Rackham Predoctoral Fellowship, which provided partial support for these studies. We also thank Mr. Michael Gatazka for synthesizing ligand 15f.

Footnotes

The authors declare no competing financial interests.

ASSOCIATED CONTENT

Supporting Information

Copies of 1H and 13C NMR spectra for new compounds. The Supporting Information is available free of charge on the ACS Publications website at http://pubs.acs.org.

REFERENCES

  • (1).For general reviews on metal-catalyzed alkene diamination reactions, see: [Google Scholar]; (a) de Figruiredo RM Transition-metal-catalyzed diamination of olefins. Angew. Chem., Int. Ed 2009, 48, 1190. [DOI] [PubMed] [Google Scholar]; (b) De Jong S; Nosal DG; Wardrop D Methods for direct alkene diamination, new & old. Tetrahedron 2012, 68, 4067. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Chemler SR; Copeland DA Synthesis of saturated heterocycles via metal-catalyzed alkene diamination, aminoalkoxylation, or dialkoxylation reactions. Top. Heterocycl. Chem 2013, 32, 39. [Google Scholar]; (d) Muñiz K; Martinez C Development of intramolecular vicinal diamination of alkenes: from palladium to bromine catalysis. J. Org. Chem 2013, 78, 2168. [DOI] [PubMed] [Google Scholar]; (e) Broggini G; Borelli T; Giofre S; Mazza A Intramolecular oxidative palladium-catalyzed amination involving double C-H functionalization of unactivated olefins. Synthesis 2017, 49, 2803. [Google Scholar]
  • (2).(a) Streuff J. Hövelmann CH; Nieger M; Muñiz K. Palladium(II)-catalyzed intramolecular diamination of unfunctionalized alkenes. J. Am. Chem. Soc 2005, 127, 14586. [DOI] [PubMed] [Google Scholar]; (b) Muñiz K; Hövelmann CH; Streuff J; Campos-Gomez E First palladium- and nickel-catalyzed oxidative diamination of alkenes: cyclic urea, sulfamide, and guanidine building blocks. Pure Appl. Chem 2008, 80, 1089. [Google Scholar]; (c) Muñiz K; Hövelmann CH; Streuff J Oxidative diamination of alkenes with ureas as nitrogen sources: mechanistic pathways in the presence of a high oxidation state palladium catalyst. J. Am. Chem. Soc 2008, 130, 763. [DOI] [PubMed] [Google Scholar]
  • (3).For examples of Cu-catalyzed diamination reactions in the presence of external oxidants, see: [Google Scholar]; (a) Khoder ZM; Wong CE; Chemler SR Stereoselective synthesis of isoxazolidines via copper-catalyzed alkene diamination. ACS. Catal 2017, 7, 4775. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Chen M; Wang L-J; Ren P-X; Hou X-Y; Fang Z; Han M-N; Li W. Copper-catalyzed diamination of alkenes of unsaturated ketohydrazones with amines. Org. Lett 2018, 20, 510. [DOI] [PubMed] [Google Scholar]; (c) Sequeira FC; Turnpenny BW; Chemler SR Copper-promoted and copper-catalyzed intermolecular alkene diamination. Angew. Chem. Int. Ed 2010, 49, 6365. Thus far these reactions have not been demonstrated to be stereospecific. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (4).For examples of Pd-catalyzed alkene diamination reactions that involve allylpalladium intermediates, see: [Google Scholar]; (a) Du H; Yuan W; Zhao B; Shi Y A Pd(0)-catalyzed diamination of terminal olefins at allylic and homoallylic carbons via formal C-H activation under solvent-free conditions. J. Am. Chem. Soc 2007, 129, 7496. [DOI] [PubMed] [Google Scholar]; (b) Wu Z; Wen K; Zhang J; Zhang W Pd(II)-catalyzed aerobic intermolecular 1,2-diamination of conjugated dienes: a regio- and chemoselective [4+2] annulation for the synthesis of tetrahydroquinoxalines. Org. Lett 2017, 19, 2813. [DOI] [PubMed] [Google Scholar]
  • (5).Conway JH Jr., Rovis T Regiodivergent Iridium(III)-Catalyzed Diamination of Alkenyl Amides with Secondary Amines: Complementary Access to γ- or δ-Lactams. J. Am. Chem. Soc 2018, 140, 135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (6).(a) Ingalls EL; Sibbald PA; Kaminsky W; Michael FE Enantioselective palladium-catalyzed diamination of alkenes using N-fluorobenzenesulfonimide. J. Am. Chem. Soc 2013, 135, 8854. [DOI] [PubMed] [Google Scholar]; (b) Sibbald PA; Rosewall CF; Swartz RD; Michael FE Mechanism of N-fluorobenzenesulfonimide promoted diamination and carboamination reactions: divergent reactivity of a Pd(IV) species. J. Am. Chem. Soc 2009, 131, 15945. [DOI] [PubMed] [Google Scholar]; (c) Sibbald PA; Michael FE Palladium-catalyzed diamination of unactivated alkenes using N-fluorobenzenesulfonimide as a source of electrophilic nitrogen. Org. Lett 2009, 11, 1147. [DOI] [PubMed] [Google Scholar]
  • (7).(a) Shen K; Wang Q Copper-catalyzed diamination of unactivated alkenes with hydroxylamines. Chem. Sci 2015, 6, 4279.For related alkene oxyamination reactions, see: [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Hemric BN; Chen AW; Wang Q Copper-catalyzed modular amino oxygenation of alkenes: access to diverse 1,2-amino oxygen-containing skeletons. J. Org. Chem 2019, 84, 1468. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Hemric BN; Shen K; Wang Q Copper-catalyzed amino lactonization and amino oxygenation of alkenes using O-benzoylhydroxylamines. J. Am. Chem. Soc 2016, 138, 5813. [DOI] [PubMed] [Google Scholar]
  • (8).The Michael diamination method is currently limited to only the NFBS electrophile.
  • (9).Wang has demonstrated that the Cu-catalyzed alkene diamination reactions between O-acylated hydroxylamines and substrates bearing 1,2-disubstituted alkenes are not stereospecific or stereoselective. This appears to be due to the configurational instability/radical character of the C-Cu bond in the intermediate generated upon amino cupration of the alkene. See reference 7a.
  • (10).Peterson LJ; Kirsch JK; Wolfe JP Pd-catalyzed alkene diamination reactions of nitrogen electrophiles: synthesis of cyclic guanidines and ureas bearing dialkylaminomethyl groups. Org. Lett 2018, 20, 3513. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (11).Hartwig has previously demonstrated the oxidative addition of analogous O-acyl oximes to Pd(0). Tan Y; Hartwig JF Palladium-catalyzed amination of aromatic C-H bonds with oxime esters. J. Am. Chem. Soc 2010, 132, 3676. [DOI] [PubMed] [Google Scholar]
  • (12).(a) McDonald RI; Liu G; Stahl SS Palladium(II)-catalyzed alkene functionalization via nucleopalladation: stereochemical pathways and enantioselective catalytic applications. Chem. Rev 2011, 111, 2981. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Fornwald RM; Fritz JA; Wolfe JP Influence of catalyst structure and reaction conditions on anti- vs. syn-aminopalladation pathways in Pd-catalyzed alkene carboamination reactions. Chem. Eur. J 2014, 20, 8782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (13).Yin G; Mu X; Liu G Palladium(II)-catalyzed oxidative difunctionalization of alkenes: bond-forming at a high-valent palladium center. Acc. Chem. Res 2016, 49, 2413. [DOI] [PubMed] [Google Scholar]
  • (14).O-acylated hydroxylamines have been shown to be capable of oxidizing PdII to PdIV in some systems. For representative examples, see: [Google Scholar]; (a) Dong Z; Dong G Ortho vs. ipso: site-selective Pd and norbornene-catalyzed arene C-H amination using aryl halides. J. Am. Chem. Soc 2013, 135, 18350. [DOI] [PubMed] [Google Scholar]; (b) Luo B; Gao J -M.; Lautens, M. Palladium-catalyzed norbornene-mediated tandem amination/cyanation reaction: a method for the synthesis of ortho-aminated benzonitriles. Org. Lett 2016, 18, 4166. [DOI] [PubMed] [Google Scholar]; (c) Zhou Y; Bao X DFT study of Pd(0)-promoted intermolecular C-H amination with O-benzoyl hydroxylamines. Org. Lett 2016, 18, 4506. [DOI] [PubMed] [Google Scholar]
  • (15).(a) Ingoglia BT; Wagen CC; Buchwald SL Biaryl monophosphine ligands in palladium-catalyzed C–N coupling: an updated user’s guide. Tetrahedron, 2019, 75, 4199. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Friis SD; Skrydstrup T; Buchwald SL Mild Pd-catalyzed aminocarbonylation of (hetero)aryl bromides with a palladacycle precatalyst. Org. Lett 2014, 16, 4296. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (16).The experiments shown in eq 34 were carried out in deuterated benzene or toluene due to the high cost of deuterated 1,4-dioxane. Although dioxane provided the best results in catalytic reactions, toluene was a suitable solvent, and the coupling of 1b with morpholino benzoate afforded 3b in 82% NMR yield when conducted in toluene solvent rather than dioxane.
  • (17).In our initial optimization studies (reference 10), although we did note the significance of using Pd(acac)2 as precatalyst, we also observed a clear ligand-effect trend where chemical yields steadily improved as phosphine ligand structures increased in size and electron-deficiency. Unfortunately, we incorrectly interpreted this trend as suggesting that a bulky, electron-poor phosphine was required, and failed to carry out the key control experiment in which phosphine was omitted. We now believe the observed phosphine ligand effects simply reflect inhibition of catalysis by phosphine ligands. The bulky, electron-poor ligands do not bind with high affinity to Pd(acac)2, and inhibit the catalytic reaction to a lesser degree than the more electron-rich phosphines.
  • (18).The related transformation reported by Michael has been demonstrated to proceed via anti-aminopalladation, followed by reductive elimination with inversion of configuration. See reference 6b.
  • (19).Carbon-carbon bond-forming reductive elimination occurs with retention of configuration. See: Milstein D; Stille JK Mechanism of reductive elimination. Reaction of alkylpalladium(II) complexes with tetraorganotin, organolithium, and Grignard reagents. Evidence for palladium(IV) intermediacy. J. Am. Chem. Soc 1979, 101, 4981. [Google Scholar]
  • (20).We have previously shown that alkylpalladium intermediates related to 11 can undergo erosion of stereochemical purity through reversible β-hydride elimination/re-insertion/σ-bond rotation events. See: Hay MB; Wolfe JP Palladium-catalyzed synthesis of 2,1’-disubstituted tetrahydrofurans from γ-hydroxy internal alkenes. Evidence for alkene insertion into a Pd-O bond and stereochemical scrambling via β -hydride elimination. J. Am. Chem. Soc 2005, 127, 16468. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (21).β-Hydride elimination reactions from Pd exhibit large primary kinetic isotope effects (>5). See: Mueller JA; Goller CP; Sigman MS Elucidating the significance of β-hydride elimination and the dynamic role of acid/base chemistry in a palladium-catalyzed aerobic oxidation of alcohols. J. Am. Chem. Soc 2004, 126, 9724. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (22).We have illustrated a mechanism where deprotonation precedes anti-aminopalladation, given the relatively low nucleophilicity of the neutral urea nitrogen atom, and the fact that the measured pKa of 4-nitrophenyl urea is 14.0, which is close to that of the conjugate acid of carbonate (pKa 10.33). See: Kavalek J; Sterba V; El Bahaie S Dissociation Constants of m- and p-substituted phenythioureas in water and in methanol. Coll. Czech. Chem. Commun 1983, 48, 1430. However, we cannot rule out an alternate pathway in which deprotonation occurs after aminopalladation. [Google Scholar]
  • (23).The fact that the coupling of 1b with 2a provided comparable results when conducted open to air (82% NMR yield) is also more consistent with Pd(II)/(IV) catalysis than a Pd(0)/(II) catalytic cycle given the air-sensitivity of most Pd(0) complexes.
  • (24).Due to the high energy of Pd(IV) intermediates, oxidative addition from Pd(II) to Pd(IV) tends to be relatively slow, and reductive elimination from Pd(IV) to Pd(II) tends to be relatively fast. For a general review on Pd(IV) catalysis, see: Sehnal P; Taylor RJK; Fairlamb IS Emergence of Palladium(IV) Chemistry in Synthesis and Catalysis. Chem. Rev 2010, 110, 824. [DOI] [PubMed] [Google Scholar]
  • (25).(a) Cui X; Li J; Liu L; Guo QX 1,3-Dicarbonyl compounds as phosphine-free ligands for Pd-catalyzed Heck and Suzuki reactions. Chin. Chem. Lett 2007, 18, 625. [Google Scholar]; (b) Jones RC Controlling reactivity in the Fujiwara-Moritani reaction: examining solvent effects and the addition of 1,3-dicarbonyl ligands on the oxidative coupling of electron rich arenes and acrylates. Tetrahedron Lett 2020, 61, 15147. [Google Scholar]; (c) Harada S; Yano H; Obora Y Palladium-catalyzed, ligand-controlled chemoselective oxidative coupling reactions of benzene derivatives with acrylamides under an oxygen atmosphere. ChemCatChem 2013, 5, 121. [Google Scholar]; (d) For studies on the reactivity of Pd complexes that contain both an NHC ligand and an acac-derived ligand in Pd-catalyzed N-arylation reactions, see: Marion N; Navarro O; Stevens ED; Ecarnot EC; Bell A; Amoroso D; Nolan SP Modified [(IPr)Pd(R-acac)Cl] complexes: influence of the acac substitution on the catalytic activity in aryl amination. Chem. Asian. J 2010, 5, 841. [DOI] [PubMed] [Google Scholar]
  • (26).Peterson LJ; Luo J; Wolfe JP Synthesis of Cyclic Guanidines Bearing N-Arylsulfonyl and N-Cyano Protecting Groups via Pd-Catalyzed Alkene Carboamination Reactions. Org. Lett 2017, 19, 2817. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (27).Deng X; Zhang L; Liu H; Bai Y; He W mCPBA-Mediated Dioxygenation of Unactivated Alkenes for the Synthesis of 5-Imino-2-Tetrahydrofuranyl Methanol Derivatives. Tetrahedron Lett 2020, 61, 152620. [Google Scholar]
  • (28).Howard JL; Sagatov Y; Repusseau L; Schotten C; Browne DL Controlling Reactivity Through Liquid Assisted Grinding: the Curious Case of Mechanochemical Fluorination. Green Chem, 2017, 19, 2798. [Google Scholar]
  • (29).Biloski AJ; Ganem B Improved Oxidation of Amines with Dibenzoyl Peroxide. Synthesis 1983, 537–538. [Google Scholar]
  • (30).Berman AM; Johnson JS Copper-Catalyzed Electrophilic Amination of Organozinc Nucleophiles: Documentation of O-Benzoyl Hydroxylamines as broadly Useful R2N(+) and RHN(+) Synthons. J. Org. Chem 2006, 71, 219. [DOI] [PubMed] [Google Scholar]
  • (31).Niljianskul N; Zhu S; Buchwald SL Enantioselective Synthesis of α-Aminosilanes by Copper Catalyzed Hydroamination of Vinylsilanes. Angew. Chem. Int. Ed 2015, 54, 1638. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (32).Luo B; Gao J-M; Lautens M Palladium-Catalyzed Norbornene-Mediated Tandem Amination/Cyanation Reaction: A Method for the Synthesis of ortho-Aminated Benzonitriles. Org. Lett 2016, 18, 4166. [DOI] [PubMed] [Google Scholar]
  • (33).Phanstiel IV O; Wang QX; Powell DH; Ospina MP; Leeson BA Synthesis of Secondary Amines via N-(Benzoyloxy)amines and Organoboranes J. Org. Chem 1999, 64, 803. [DOI] [PubMed] [Google Scholar]
  • (34).Fan L; Liu J; Bai L; Wang Y; Luan X Rapid Assembly of Diversely Functionalized Spiroindenes by a Three-Component Palladium-Catalyzed C–H Amination/Phenol Dearomatization Domino Reaction. Angew. Chem. Int. Ed 2017, 56, 14257. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information

RESOURCES