Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2024 May 17.
Published in final edited form as: J Am Chem Soc. 2023 May 3;145(19):10743–10755. doi: 10.1021/jacs.2c13047

Breaking the tert-Butyllithium Contact Ion Pair: A Gateway to Alternate Selectivity in Lithiation Reactions

Michael P Crockett 1, Jeanette Piña 1, Achyut Ranjan Gogoi 1,, Remy F Lalisse 1,, Andrew V Nguyen 1, Osvaldo Gutierrez 1, Andy A Thomas 1
PMCID: PMC10245630  NIHMSID: NIHMS1901528  PMID: 37133911

Abstract

The effects of Lewis basic phosphoramides on the aggregate structure of t-BuLi have been investigated in detail by NMR and DFT methods. It was determined that hexamethylphosphoramide (HMPA) can shift the equilibrium of t-BuLi to include the triple ion pair (t-Bu–Li–t-Bu)/HMPA4Li+ which serves as a reservoir for the highly reactive separated ion pair t-Bu/HMPA4Li+. Because the Li-atom’s valences are saturated in this ion pair, the Lewis acidity is significantly decreased; in turn, the basicity is maximized which allowed for the typical directing effects within oxygen heterocycles to be overridden and for remote sp3 C–H bonds to be deprotonated. Furthermore, these newly accessed lithium aggregation states were leveraged to develop a simple γ-lithiation and capture protocol of chromane heterocycles with a variety of alkyl halide electrophiles in good yields.

Graphical Abstract

graphic file with name nihms-1901528-f0001.jpg

1. INTRODUCTION

The widespread application of organolithium reagents can be ascribed to their innate ability to predictably transform organic substrates into organolithium compounds, priming them to be further functionalized.1,2 Throughout the evolution of these lithiation-substitution sequences, the generality of this approach has been well demonstrated for substrates that incorporate Lewis basic (directing) or anion stabilizing subunits.3 For example, directed α-lithiations of heterocycles emerged as a critical strategy for the synthesis of building blocks in drug discovery (Figure 1A).4Despite these achievements, the strong coordinating ability of the Li-ion with Lewis basic directing groups has precluded remote positions in heterocycles from being functionalized, which is a significant limitation.

Figure 1.

Figure 1.

(A) Directed ortho lithiation strategies and applications in synthesis. (B) Traditional lithiation site selectivity of chromane heterocycles. (C) This work-breaking the carbon lithium bond to alter lithiation site selectivity in oxygen heterocycles.

Although there are numerous illustrations of employing Lewis basic ligands to switch from one directing group to another, very few lithiation reactions are able to suppress embedded heteroatoms altogether and activate weakly acidic positions.5 As part of our efforts to identify new avenues to synthesize organometallics, we sought to develop a lithiation protocol that could bypass precoordination of the Li-ion with oxygen heterocycles and allow for remote positions to be deprotonated through a direct bimolecular acid base reaction. In particular, chromanes and tetrahydrobenzoxepines, which are popular cores in pharmaceuticals, remain challenging to selectively functionalize at the benzylic position because the Lewis basic O atoms direct and control the site of functionalization (Figure 1B).6,7

In this article, we describe a new scheme for lithiation reactions based on the concept of ion-pairing (Figure 1C). Specifically, we have found that Lewis basic phosphoramides can shift the equilibrium of strong organolithium reagents, such as t-BuLi, to include triple ion pairs (TIP) and separated ion pairs (SIP).8 In this system the Li-atom’s valences are saturated (L4Li+//R) which hinders coordination to heteroatoms and, in turn, allows for remote benzylic C–H bonds to be selectively lithiated. For example, the benzylic site of chromane can now be lithiated and trapped in preference to traditional ortho-aromatic sites. Moreover, the decreased Lewis acidity of the saturated Li-atom (L4Li+) combined with the enhanced nucleophilicity of the newly formed benzylchromane anion allows for primary and secondary alkyl halide electrophiles bearing functional groups, such as epoxides, to be well tolerated.

2. BACKGROUND

A growing body of experimental and theoretical investigations on organolithium reagents have shown that Lewis bases, such as tetramethylethylenediamine (TMEDA) or hexamethylphosphoramide (HMPA),9 can lead to drastic changes in reactivity,10 regioselectivity,11 and stereoselectivity12 by shifting the aggregate–aggregate exchange equilibrium to include lower aggregation states. At the extreme end, contact ion pairs (CIP) have been shown to convert into triple ion pairs (TIP) and even into separated ion pairs (SIP) (Figure 2A).13 For instance, Reich demonstrated that additions of HMPA to various organolithium reagents bearing anion stabilizing groups such as dithiosubstituted 1,14 2,15 or 39 allowed for the conversion of CIP into TIP as well as SIP. Importantly, these mechanistic insights revealed that the regioselectivity of lithium dithiane additions to enones was a function of the ion pair structure, where the CIP (absence of HMPA) with an intact C–Li bond give 1,2-addition through a four-centered transition state, i.e., “induced proximity”, and the SIP (presence of HMPA) give predominantly 1,4-addition via an open transition state (Figure 2B).9c Although this model system clearly indicated that CIP and SIP exhibit unique reactivity patterns and is likely to apply to similar organolithium compounds, it is difficult to predict if strong organolithium bases, such as t-BuLi, have access to appreciable concentrations of TIP and SIP under standard reaction conditions. In this context the ability to perform a similar aggregate structure–activity relationship study with t-BuLi to determine if the SIP t-Bu/L4Li+ can be accessed and used to bypass directing groups for the lithiation of chromane heterocycles at the benzylic position would be ideal; however, the high reactivity of the t-BuLi CIP and predicted higher reactivity of TIP and SIP make their observation difficult in practice.

Figure 2.

Figure 2.

(A) Generation of triple ion pairs and separated ion pairs from alkyllithium reagents using Lewis bases. (B) Altered reactivity observed between CIP and SIP in enone additions.

3. GOALS OF STUDY

The central goal of this project is to determine if Lewis bases, such as phosphoramides, can shift the equilibrium of strong organolithium reagents to include separated ion pairs and in turn allow for common directing effects within oxygen heterocycles to be overridden. While there are numerous examples of stabilized organolithium reagents converting into SIP as described above, there are no known literature reports that describe TIP or SIP with unstabilized organolithium reagents such as t-BuLi.16 We suspect that this is because TIP and SIP are likely present in only trace amounts and are too reactive for detection by routine spectroscopic methods. Our laboratory specializes in rapid-injection NMR (RI-NMR) spectroscopy which provides us the unique avenue to investigate these highly reactive chemical processes.17 The major advantage of RI-NMR over other physical techniques is that the reaction is performed inside the NMR spectrometer, allowing for structural and kinetic data to be directly linked. Specific goals of this study include (1) investigating if t-BuLi/Lewis base combinations can override the directing effects in common oxygen heterocycles allowing for remote sp3 C–H bonds to be selectively lithiated and functionalized, (2) harnessing any new found reactivity by developing a general method to functionalize remote positions in oxygen heterocycles, (3) fully characterizing the active aggregate structures of t-BuLi in the presence of Lewis bases, and (4) quantum mechanical simulation of the metalation process by computational modeling.

4. RESULTS

4.1. Lithiation of Oxygen Heterocycles: Effect of Phosphoramides on Chemoselectivity.

4.1.1. Preliminary Reactions of Tetrahydrofuran with Organolithiums and Phosphoramides.

Tetrahydrofuran (THF) has proven to be a particularly useful solvent for lithiation reactions due to its ability to deaggregate organolithiums which often leads to extreme enhancements (and on occasion decreases) in the reactivity.10 However, THF is also susceptible to lithiation with organolithiums, especially at higher temperatures.18 Typically, THF is lithiated alpha to oxygen which is then followed by a rapid reverse [3 + 2] cycloaddition yielding lithium enolate 5 and ethylene 6 (Figure 3A).19 Recent investigations of substituted tetrahydrofurans have revealed a dependence on the solvent medium and metalating agent for the reverse [3 + 2] rate.20 In contrast, Clayden found that in the presence of HMPA, an alternate fragmentation product 8 could be obtained (Figure 3B).21 We wondered if these results indicated that the site selectivity for the lithiation of THF had been altered to favor a β-lithiation pathway (Figure 3C).22 While it is possible that this alternative lithiation product 7 could be formed through lithiation alpha to oxygen followed by carbene formation and a 1,2 hydride shift, another likely mechanism would proceed through a direct β-metalation and subsequent anti-Baldwin elimination pathway.23

Figure 3.

Figure 3.

Altered site selectivity for THF lithiation. (A) α-Lithiation and fragmentation of THF. (B) Clayden observed alternate fragmentation products using HMPA. (C) Proposed mechanism for the formation of 7. (D) Preliminary results for lithiation and trapping of THF-d4.

To determine if lithiation was occurring directly at the β-position, THF isotopomer THF-d4 was subjected to t-BuLi and 6.0 equiv of trispyrrolidinephosphoramide (TPPA); TPPA is a nonmutagenic alternative to HMPA and exhibits nearly identical reactivity patterns to HMPA (Figure 3D).24,25 The only product that could be detected by high resolution electrospray ionization mass spectrometry (HR-ESI-MS) after trapping with benzoyl chloride was 11-d3. This result strongly suggests that lithiation is taking place directly at the β-position and that the heteroatom within the THF substrate was no longer binding the Li-ion. Importantly, this represents a rare example of selectively removing a stronger C–H bond in the presence of weaker ones.26

4.1.2. Exploring Site Selectivity of Lithiation with Oxygen Heterocycles and t-BuLi/Phosphoramides.

Given the ability to override the typical directing effects imparted by the oxygen atom within THF, other saturated oxygen heterocycles were evaluated to determine if the unique site selectivity for lithiation could be generalized (Figure 4 and Tables S3S9). To investigate this possibility, oxygen heterocycles 12–18 were subjected independently to t-BuLi in the presence and absence of phosphoramides as well as to other standard metalation conditions for comparison (Figure 4 and Tables S3S9).27 It became apparent that the product distribution derived by the addition of phosphoramides was selective for the benzylic position whereas the parent alkyl lithium reagent resulted in exclusive ortho-methylation in all cases. For example, addition of t-BuLi to a mixture of chromane 12 and TPPA followed by the addition of methyl iodide was cleanly methylated at the benzylic position. Furthermore, when compared to Schlosser’s base (t-BuLi/KO-t-Bu),5b the t-BuLi/phosphoramide combination proved to be much more selective for the benzylic position. To further highlight the high degree of selectivity, methyl chromane 13 was evaluated under the same conditions indicating that similar functional groups could be differentiated.

Figure 4.

Figure 4.

Site selectivity in the metalation of oxygen heterocycles. aYields determined by NMR relative to a trimethoxybenzene internal standard. bOxygen heterocycle (1.0 equiv) treated with t-BuLi (1.0 equiv) at −78 °C and then warmed to rt for 1 h before cooling to −78 °C and adding MeI (1.1 equiv). cOxygen heterocycle (1.0 equiv) treated with ligand and t-BuLi (1.0 equiv) at −78 °C for 10 min before adding MeI (1.1 equiv). dOxygen heterocycle (1.0 equiv) treated with ligand and n-BuLi (1.0) at at −78 °C for 1 h before adding MeI (1.1 equiv). eOxygen heterocycle (1.0 equiv) treated with ligand and n-BuLi (1.0) at at −78 °C for 10 min before adding MeI (1.1 equiv).

Although there are many examples of switching from one directing group to another under metalation conditions by employing various Lewis basic ligands, few protocols are able to selectively suppress the directing effects of multiple heteroatoms.5a Therefore, we sought to evaluate the selectivity on a variety of chromane cores bearing additional heteroatomic groups to determine if they could also be overridden. Piperidine (14), fluoro (15), and chloro (16) substituted chromanes were selected and evaluated under the same conditions; vide infra (Figure 4). It became apparent that the product distribution derived by the addition of phosphoramides was complementary to the parent alkyllithium in all cases. Although Schlosser’s base exhibited very little site selectivity with 14 or 15, it was selective for ortho-metalation in 16.5b

We next turned our attention toward other oxygen heterocycles to determine if the selectivity was unique to the chromane core. Benzoxepin 17 provided the 5-methyl product in excellent selectivity with the t-BuLi/phosphoramide combination. Interestingly, Schlosser’s base becomes more selective for ortho-metalation with 17 in comparison to 12. Moreover, oxane 18, a nonbenzofused system, was cleanly methylated at the benzylic site whereas neither t-BuLi or Schlosser’s base reacted.

Significantly, these results combined demonstrate that heteroatomic directing groups can be overridden and benzyl anions can be generated and captured within oxygen heterocycles. More explicitly, the direct synthesis of benzyl anions within oxygen heterocycles provides a powerful synthon in creating pharmaceutically relevant structures.28 Accordingly, the focus of our investigations switched to determine the generality of this protocol by investigating the functional group compatibilities of the chromane substructure and alkyl halide coupling partners.

4.1.3. Lithiation and Alkylation of Chromanes at the Benzylic Position with Alkyl Halides.

The development of new methods for the construction of sp3–sp3 carbon–carbon bonds is of paramount importance. Lithiation-substitution sequences mirror contemporary transition metal catalyzed cross-coupling methodologies but are not as susceptible to detrimental isomerization or β-hydride elimination pathways (Figure 5).29 After a brief optimization, chromane 12 was found to undergo alkylation at the 4-position with 1-bromooctane using 2.0 equiv of HMPA and 1.2 equiv of t-BuLi (Figure 5; Table S10). In addition, TPPA and HMPA were found to behave similarly in all cases (Table S10). This is particularly attractive in a practical sense as HMPA can then be replaced with nonmutagenic alternatives in applications that require it on larger scales.25 Interestingly, t-BuLi/DMPU was found to provide the same site-selectivity as the t-BuLi/phosphoramide combination though the yields were inferior.30

Figure 5.

Figure 5.

Scope of the chromane γ-alkylation. Standard reactions conditions were carried out on 0.5 mmol scale employing 2.0 equiv of (a) HMPA or (b) TPPA. Yields reported are isolated yields of spectroscopically pure product and are an average of two runs. (c) DMPU was used in place of HMPA or TPPA, and NMR yields are provided. (d) Single run experiment. (e) Product contains ∼5% of an inseparable aromatic alkylated isomer. (f) 6.0 equiv of HMPA employed in the reaction with no annealing period.

With optimized reaction conditions in hand, primary alkyl halides were explored first; 1-bromooctane was found to furnish the product 19 in excellent yields. Likewise, methyl iodide was found to be similarly as effective to 19, providing yet another avenue to install methyl groups at unique locations which have great pharmaceutical significance.31 Furthermore, protected oxygen functionality was also found to provide the alkylated products 21, 22, and 23 in moderate to good yield. Of note, epichlorohydrin was also found to provide the product 24 in synthetically useful yields.

To further demonstrate the functional group compatibility, a selection of heterocycle containing electrophiles were then evaluated. Indole (25) and piperazine (26) substrates provided the product in good to excellent yields showcasing the ability to incorporate nitrogen heterocycles of broad pharmaceutical relevance.32 Quinone (27) and thiophene (28) substrates provided the corresponding trapped products in slightly reduced but synthetically useful yields presumably due to competitive lithiation of the more acidic protons in these scaffolds.

Next, a set of secondary alkyl halides were evaluated as shown in Figure 5. Five (29), six (30), and seven (31) membered bromocarbocycles all provided the alkylated products in moderate to good yields. Cyclopropyl bromide (32) resulted in no product formation presumably due to the large amount of I-strain associated with the transition state during substitution33 Furthermore, oxacycles (33) and (34) also provided good yields of the desired alkylation products. Finally, acyclic secondary bromide (35) provided the product in suitable yields albeit with low diastereoselectivity. The ability to couple two secondary sp3 hybridized carbons is particularly attractive for the formation of three-dimensional products that are required for many pharmaceutical applications.34

To complement the electrophile scope, substituted chromane heterocycles were also evaluated (Figure 5). Methyl chromane 36 was found to be alkylated exclusively at the benzylic position within the ring, and no competing side chain alkylation was observed.35 Fluorinated and chlorinated chromanes 37 and 38 were found to provide the trapped products in good to excellent yields. Interestingly, brominated chromane 39 could also be trapped, albeit in low yields as a result of competition with lithium halogen exchange. Lithium–halogen exchange is known to be one of the fastest chemical processes for organolithium reagents and can be competitive with proton transfer from an acidic substrate.36 The ability to detect any of the desired product 39 from the aryl bromide further highlights the speed of the lithiation (see section 4.2.1). Evaluation of the trapping of biaryl 40 resulted in slightly reduced yield but with exclusive selectivity for the benzylic position which highlights the ability to maintain selectivity in the presence of other ortho-directing aryl ethers. Piperidine functionalized chromane 41 also provided the product in excellent yields, showcasing the ability to override the influence of nitrogen directing groups as well. Furthermore, a fused pyridine heterocycle 42 was also tolerated under these conditions demonstrating how this method could be amenable to more complex product synthesis. The ability of phosphoramides to completely override the directing group present in each of these substrates clearly demonstrates the level of control the active lithium aggregate has on site selectivity. Therefore, the next phase of our investigations focused on determining the active lithiation species.

4.2. Effect of Phosphoramides on the Organolithium Aggregate Structure: An NMR and DFT Study.

Before embarking on an in-depth mechanistic analysis into the effects of phosphoramides on the t-BuLi aggregate structure, a critical aspect of the mechanism needed to be clarified. To rule out the possibility of a lithiated phosphoramide serving as the base or an anionic intermediate, HMPA isotopomer HMPA-d18 was subjected under the standard lithiation conditions with chromane 12.37 Under these conditions, only protic isobutene (44) could be detected by 1H and 13C NMR spectroscopy, i.e., no deuterium exchange (Figures 6 and S34). This result strongly suggests that HMPA is coordinating the Li+ cation and not serving as a proton shuttle. Therefore, to better understand the origins for the observed site selectivity changes as described above, an in-depth NMR investigation into the aggregate structure of the active lithiation species was undertaken.

Figure 6.

Figure 6.

No deuterium incorporation was observed when employing HMPA-d18 and THF-d8 under the standard reaction conditions.

Based on Reich’s previous studies, we hypothesized that the change in site selectivity in our system could be due to the formation of TIP or SIP. Such intermediates have never been observed with unstabilized alkyllithium reagents which is presumably due to their high reactivity.9a Armed with our rapid-injection NMR (RI-NMR) apparatus, we set out to identify the active aggregation state at low temperature. Although the natural abundance 7Li isotope is observable by NMR spectroscopy, it possesses a large quadrupole moment as well as a 3/2 nuclear spin which often leads to line broadening and precludes the observation of J-couplings between nuclei (Figure S4). Therefore, to detect and gain structural information, the carbon–lithium bond requires the enrichment of both the quaternary carbon and Li-atom within the t-BuLi [((CH3)313C–6Li] CIP-46.

The t-BuLi isotopomer 46 was effectively synthesized through reductive metalation using our newly developed 6Li-dendrites and 13C enriched (CH3)313CCl 45 (Figure 7).38To provide insight into the aggregate structure and composition of the active lithiation species formed by combining phosphoramides with t-BuLi, a series of 6Li, 31P, and 13C NMR spectroscopic experiments were performed by titrating HMPA (Reich’s method)9a into solutions of doubly isotopically enriched CIP-46.

Figure 7.

Figure 7.

(A) Preparation of 6Li enriched dendrites. (B) Synthesis of (CH3)313C6Li using 6Li-dendrites and (CH3)313CCl.

4.2.1. NMR Investigations on the Effect of HMPA on the t-BuLi Aggregate Structure.

Initial attempts to prepare samples of t-BuLi in the presence of HMPA at −80 °C resulted only in the formation of the THF elimination product 7 as described earlier (see section 4.1.1).9a To overcome this limitation, t-BuLi samples for HMPA titrations were prepared at lower temperatures, −115 °C (Figure 8).39 The 6Li spectrum of the t-BuLi isotopomer 46 consisted of a single doublet (JLi–C = 12.9 Hz) at 1.03 ppm and was assigned as the monomeric contact ion pair structure CIP-46a.9a Furthermore, only one signal at 17.20 ppm was observed in 13C NMR spectrum bearing the characteristic 1:1:1 triplet (JC–Li = 13.1 Hz). Upon addition of 1.0 equiv of HMPA, CIP-46a was partially converted to a new monomeric contact ion pair t-BuLi-HMPA1 CIP-46b that was observable in the 6Li and 13C NMR spectra. This structural assignment for CIP-46b was supported by the observation of small (JLi–P = 3 Hz) scalar couplings between the 6Li and HMPA at −130 °C (Figure S13).40

Figure 8.

Figure 8.

HMPA-d18 titration of (CH3)313C6Li in 3:2 THF-d8:Et2O at −125 ° C at 0.1 M.

Upon addition of 2.0 equiv of HMPA, two new species appeared in the 6Li NMR spectrum in a 1:1 ratio (Figure 8). One species, at 2.7 ppm, is clearly resolved into a 1:2:1 triplet (JLi–C = 12.1 Hz) which is consistent with the formation of (R–Li–R) as triple ion pair TIP-46. The second species appeared at −0.36 ppm and is resolved into a quintet (JLi–P = 2.8 Hz) and is consistent with the formation of the (HMPA)4Li+ counterion within the TIP-46. Likewise, in the 13C NMR spectrum, a new C–Li contact ion is formed as indicated by the appearance of a second 1:1:1 triplet (JC–Li = 12.1 Hz). Combining the observations from the two spectra together is most consistent with the formation of the (t-Bu)2Li/L4Li+ triple ion pair TIP-46.41 Furthermore, increasing the concentration of HMPA to 3.0 equiv increases the fraction of the t-BuLi CIP-46 that is sequestered in this triple ion pair TIP-46 from 9% to 15% further indicating that HMPA has a role in the formation of this aggregate.42,43 Of note, upon increasing the amount of HMPA from 1 to 2 (or 3) equiv, the observed monomeric contact ion pair could not be detected as a distinct species t-BuLi-HMPAX CIP-46a-d (X = a, 0; b, 1; c, 2; d, 3) even at −125 °C. This is presumably due to increased rates of exchange between bound and free HMPA. Results from our computations described below indicate that each addition of HMPA decreases the Lewis acidity of the Li-cation within CIP-46 and likely leads to the observed increased rates of exchange.

4.2.2. Thermochemical Calculations of the Effect of HMPA on the t-BuLi Aggregate Structure.

To gain further insight into the various aggregates formed during the titration of t-BuLi with HMPA, vide infra, dispersion-corrected DFT in THF using the CPCM implicit solvation model was used to calculate the ground-state energies (Figure 9). Starting from CIP-46a, the loss of a THF solvent molecule from CIP-46a and concomitant coordination of an HMPA molecule to form CIP-46b was predicted to be exergonic. Interestingly, upon each subsequent coordination of HMPA to CIP-46b resulted in a highly exergonic downward cascade leading to TIP-46. Upon inspection of complexation interaction energies, it became clear that each addition of HMPA was weakening the t-Bu···Li interaction and lowering the C–Li bond order (see Supporting Information Figure S91). Importantly, the Lewis acidity of the (HMPA)4Li+ counterion within TIP-46 is greatly increased, and as a result, this decreases the ability for ligand dissociation to occur and serves as a thermodynamic sink for TIP-46.

Figure 9.

Figure 9.

Calculated equilibria for t-BuLi titration with HMPA at −125 °C. See Supporting Information for equilibria calculations at −80 °C.

The goal of the NMR studies described above was to identify the aggregate responsible for the observed change in metalation site selectivity within oxygen heterocycles. Although several newfound aggregates, such as CIP-46b and TIP-46, were observable at low temperatures (−125 °C) where interaggregate exchange is slow, the high reactivity of these samples precluded investigation at higher temperatures (>−110 °C) where interaggregate exchange is presumably rapid. As a result, to gain further insight into the unique selectivity observed during the deprotonation step, a thorough computational investigation was undertaken to determine the active metalation aggregate and origins of selectivity (section 4.3).

4.2.3. Effect of HMPA on the Benzyl Anion in Chromane.

To probe the aggregate structure of the lithiated chromane 43, 1H, 13C, 7Li, and 31P NMR experiments were carried out. See Supporting Information for full details. Following Reich’s titration protocols, the 7Li NMR spectrum displayed a single signal that is resolved into a quintet (JLi–P = 7.8 Hz) (Figure 10A). Furthermore, the P NMR spectrum displayed at least two new signals at (26.88 ppm, JP–Li = 7.8 Hz) and (26.74 ppm, JP–Li = 9.1 Hz) with resolved J-couplings (Figure 10A). The most consistent interpretation of these spectra is the formation of the (HMPA)4Li+ and (HMPA)3Li+ cations. Interestingly, the 13C NMR indicates a significant loss in aromaticity of the lithiated chromane intermediate 43 as seen by the significant upfield shifts of nearly all aromatic signals. The benzylic carbon moved from 27.9 to 58.9 ppm (Δ(13C) – 31 ppm) suggesting significant delocalization of anionic character into the aromatic ring. The combined NMR spectra are most consistent with the assignment of the lithiated chromane as SIP-43 in the presence of 6.0 equiv of HMPA (Figure 10A).

Figure 10.

Figure 10.

(A) 7Li and 31P NMR spectra of lithiated chromane in THF-d8 at −115 °C. (B) 1H NMR spectra (RI-NMR) array of the injection of 1-bromooctane into the lithiated chromane in THF-d8 at −80 °C. The injection of substrate was performed at the 2 s time point.

Using our rapid-injection NMR (RI-NMR) apparatus,17d the nucleophilicity under standard reaction conditions of the SIP-43 was evaluated by injecting 1-bromooctane at −80 °C (Figure 10B). Although it was not possible to obtain an exact rate for the alkylation of SIP-43, it was fully consumed in less than a single scan (<1 s at −80 °C), highlighting the tremendous reactivity of the SIP in solution.

4.3. Computational Analysis of the Reaction Profile for Deprotonation with Phosphoramides: Benzylic Site Selectivity Explained.

Based on our empirical and NMR studies, there are three plausible scenarios that explain the role of phosphoramides in alternating the site selectivity (ortho vs benzyl) during the deprotonation event with the substrates described in section 4.1.1 (Figure 11). Therefore, to gain further insight into the deprotonation step, computational investigations were undertaken as detailed below. To facilitate our discussions, chromane 12 was chosen as a representative substrate in the calculations. For completeness, energy profiles (ground states, transition states, and products) for the metalation event for both ortho (TS-O) and benzyl (TS-B) positions of 12 were calculated using the dispersion-corrected DFT in tetrahydrofuran as solvent using the CPCM solvent model (see Supporting Information for full computational details; Figure 12).4447 This level of theory has been used previously in the characterization of methyllithium and tert-butyllithium oligomers.45d All structural figures were generated with CYLview.48

Figure 11.

Figure 11.

Possible lithiation scenarios using chromane as an example substrate.

Figure 12.

Figure 12.

(A) Comparison of energy profiles for the deprotonation step: ortho vs benzyl. (B) Transition-state structure CIP-46a-TS-O for directed ortho lithiation (no HMPA). (C) Transition-state structure SIP-46-TS-B for undirected benzyl lithiation (with HMPA).

We posit the following three scenarios: (1) a contact ion pair bearing HMPA ligands (CIP-46b, CIP-46c, or CIP-46d) is the active metalating species and is too sterically hindered to be influenced by directing effects, (2) TIP-46 has significantly reduced Lewis-acidity due to the steric bulk and donating nature of the ligands which both increases its basicity and prevents directing effects, or (3) TIP-46 serves as a reservoir for SIP-46 that, by lacking lithium, is not influenced by directing effects.

4.3.1. Scenario 1: Comparison of Energy Profiles for Metalation of Chromane with tert-Butyl Lithium Contact Ion Pairs.

The energy profiles (i.e., ortho vs benzyl selectivity) employing CIP-46a (no HMPA) are summarized in Figure 12. The activation energies for H-atom extraction for CIP-46a-TS-O and CIP-46a-TS-B are 16.4 and 18.6 kcal/mol, respectively, clearly reflecting a notable activation energy difference of 2.2 kcal/mol in favor of the ortho position. The next series employing a single HMPA molecule CIP-46b-TS-O (16.7 kcal/mol) vs CIP-46b-TS-B (20.7 kcal/mol) also favored the ortho pathway by 4.0 kcal/mol (Figure 12). The greater activation barrier for CIP-46b-TS-B over CIP-46b-TS-O can be attributed to higher distortion energy of the t-Bu fragment during the removal of the proton (see Supporting Information for further details). In addition, activation energies for proton removal for CIP-46c-TS-O and CIP-46c-TS-B (2 HMPA) are nearly identical at 12.6 and 13.2 kcal/mol, respectively, clearly predicting only a slight preference for the ortho position.

Upon inspection of energy profiles employing CIP-46d bearing three HMPA molecules, it becomes clear that the chromane 12 substrate can no longer bind the Li-cation, and as a result, the bimolecular proton extraction processes did not result in a realistic comparison to the other contact ion pairs described above. Because the three HMPA ligands are sterically demanding within CIP46d, the carbon–lithium bond distance became elongated (4.9 Å) during proton extraction and a result was reminiscent of the separated ion pair activation pathways described below (section 4.3.3). Combining our empirical, NMR and DFT mechanistic studies suggest that a contact ion pair ligated by HMPA is not responsible for the observed change in site selectivity.

4.3.2. Scenario 2: Comparison of Energy Profiles for Metalation of Chromane with Triple Ion TIP-53.

The second possibility also seems unlikely based on the energies of transition states TIP-46-TS-O and TIP-46-TS-B for proton extraction which are 17.2 and 18.0 kcal/mol, respectively. In this series the ortho position is slightly favored by 0.8 kcal/mol indicating that the process would favor the ortho position and likely to be unselective if TIP-46 were the dominant species. These combined observations lead us to investigate the possibility whereby TIP-46 dissociates into SIP-46 and monomeric CIP-46. Unfortunately, SIP-46 was completely undetectable during any of the HMPA NMR titrations carried out in this study. This is likely due to the high reactivity of this aggregate such that any SIP-46 generated reacts immediately, i.e., Curtin–Hammett conditions. Under such a scenario, the small amount of SIP-46 generated at any given time would be reacting under large excess of all possible substrates and provides a possible pathway for the observed site selectivity for the benzyl position.

4.3.3. Scenario 3: Comparison of Energy Profiles for Metalation of Chromane with Separated Ion Pair SIP-53.

In fact, the computed transition states for SIP-46-TS-O and SIP-46-TS-B were 11.3 and 6.9 kcal/mol, respectively, indicating that the benzylic position was predicted to be favored over the ortho position by 4.4 kcal/mol. The origin for the observed change in selectivity is the net result of the inability for the saturated (HMPA)4Li+ cation to bind the oxygen atom within chromane 12 which does not allow for the t-Bu anion to be directed. Combining our results with the fact that SIP are known to be 10 orders of magnitude more reactive than TIP, as described by Reich et al.,49 suggests that TIP-46 is the ground state complex and that SIP-46 is the active aggregate responsible for lithiation.

Of note, if an HMPA within the (HMPA)4Li+ cation in SIP-46 were exchanged with the chromane 12 substrate, it would open an avenue for a Lewis acid catalyzed pathway (Figure 13). This effect is 2-fold: (1) it sterically shields the ortho position and (2) acidifies the benzyl position. Importantly, this Lewis acid catalyzed pathway raised the SIP-46-TS-O-CAT and lowered the SIP-46-TS-B-CAT to 13.3 and 5.5, respectively (Figure 13). Although the Lewis acid catalyzed pathway is predicted to have a lower barrier, we suspect that HMPA is likely inhibiting this pathway to some extent. Thus, the active species is likely a separated ion pair that proceeds through a combination of these two pathways. Our combined observations indicate that separated ion pairs are accessible with strong organolithiums and can serve as reactive intermediates.

Figure 13.

Figure 13.

Comparison of energy profiles for the deprotonation step with separated ion pairs: Lewis acid vs non-Lewis acid catalyzed pathways.

5. SUMMARY

5.1. Observed Trends in Lithiation Site Selectivity.

It is clear from the earlier studies that the addition of phosphoramides changes the lithiation site selectivity for THF and other oxygen heterocycles. Because the oxygen atom within THF is electron withdrawing, the α hydrogen is known to be more acidic over the β hydrogen, indicating that pKa values alone cannot explain the observed site selectivities for all the substrates surveyed.20b,50 Thus, the selectivity for proton removal is likely the result of the summation of the localized steric, electronic, and stereoelectronic properties about each hydrogen atom. Since deprotonations of alkanes often exhibit late transition states, it is conceivable that small differences in the environments about the hydrogen atoms can lead to drastic changes in selectivity.51

This selectivity is best contextualized by revisiting the lithiation of some of the oxygen heterocycles presented earlier (Figure 14). In each of these substrates, there is at least one other site that is activated for alkylation. In chromane 12 for example, the ortho-aromatic site is activated for metalation in the absence of added ligands. Schlosser’s base is unable to overcome this activation and provides the methylated products as a mixture. By contrast, the t-BuLi/phosphoramide combination is able to completely override this directing effect. This is further demonstrated by examining chromanes that bear additional directing groups (14–16). Piperidine chromane 14 provides a product distribution that is similar to chromane under all examined conditions.52 Furthermore, the results with Schlosser’s base indicate that the site ortho to nitrogen is accessible though not favored, presumably due to the lower inductive withdrawing capability of nitrogen compared to oxygen. Metalation of fluorochromane 15 using the t-BuLi/phosphoramide combination demonstrates that the site selectivity is relatively insensitive to increased acidity in other positions on the substrate, whereas unassisted metalation results in selectivity ortho to oxygen, and Schlosser’s base results in metalation ortho to fluorine as it is more inductively withdrawing. Chlorochromane 16 on the other hand displays complete selectivity for the position ortho to oxygen with both t-BuLi and Schlosser’s base. This result falls directly in line with the trends summarized by Gschwend and Rodriguez53 where Cl is an inferior ortho-director to OMe but a superior meta-director. These two features combined favor the observed site selectivity in directed metalation systems. The t-BuLi/phosphoramide combination is able to overcome this double directing effect and shows excellent selectivity for benzylic metalation. With each additional directing group surveyed, it becomes clear how sensitive standard metalating conditions are to the precise substitution of the substrate. By comparison, the t-BuLi/phosphoramide combination is relatively insensitive to the functionalization on the substrate. This is further demonstrated by examining benzoxepine 17. Despite what would appear to be a simple ring expansion, a strong change in selectivity is observed for Schlosser’s base while the t-BuLi/phosphoramide combination maintains exclusive selectivity for the benzylic position. Finally, metalation/trapping of oxane 18 is only possible through the t-BuLi/phosphoramide combination which would suggest that the mechanism for activation of the alkyllithium reagent must be unique to this combination. These results combined demonstrate the unique capability of the t-BuLi/phosphoramide system to override directing effects and enable the remote functionalization of oxygen heterocycles.

Figure 14.

Figure 14.

Substrates that undergo selective benzylic lithiation.

5.2. Enhanced Functional Group Compatibility During Capturing Step.

While the change in site selectivity was the primary motivation of the study, it is worth commenting on the fact that the lithiated chromanes are far more nucleophilic than has been previously reported for similar organometallic reagents. Numerous reports throughout the past decades have detailed the direct alkylation of activated alkyl halide electrophiles with a variety of organolithium,54 organosodium,55 and organopotassium reagents.56 Many of these reactions use catalytic or stoichiometric copper additives to further expand the scope of the transformation to less activated electrophiles. In this context, the direct trapping of secondary alkyl bromides from the SIP presented here is a significant departure from the current synthetic usage of alkyllithium reagents. Typically, to enable the coupling with unactivated secondary electrophiles, transmetalation to another organometallic is required to be compatible with modern cross-coupling methods. Considering this synthetic advantage, it is worth considering the exact cause for the increased reactivity. To understand a possible origin of the reactivity, it is worth contemplating two possible scenarios: alkylation from a CIP and alkylation from a SIP (Figure 15). Alkylation that proceeds from a CIP reagent generally requires the formation of a product-separated ion pair as described by Cram et al. (Figure 15A).57 This separated ion pair must then reorganize to generate the CIP salt byproduct. Alternatively, if the lithium reagent starts as a separated ion pair, the lithium ion can be positioned such that during the capture step, a CIP can be formed immediately with minimal reorganization thus lowering the overall barrier of the transformation (Figure 15B).

Figure 15.

Figure 15.

(A) Reaction pathway from CIP species. (B) Reaction pathway from SIP species.

The ability to trap secondary electrophiles with the SIP was unexpected as the direct trapping of secondary lithium reagents with unactivated secondary halides has not previously been reported. This result is likely indicative of the highly nucleophilic nature of benzylic anions and has not been exploited previously simply because the preparation of benzyl lithium reagents is challenging. The methods that are reliable often generate a stoichiometric alkoxide or sulfide byproduct that interferes with electrophile trapping.6a,58 Regardless of the origin of the reactivity, the good yields observed with secondary electrophiles further emphasize how the enhanced nucleophilicity of separated anions can be leveraged for new reactivity to compliment other carbon–carbon bond forming reactions.59

6. CONCLUSIONS

Given the long history and success of lithiation reactions, these initial observations of TIP generated from highly basic and localized organolithium reagents have confirmed that such species exist in the presence of sufficiently donating ligands. The ability to override the directing ability of embedded heteroatoms using the alternate aggregate structure provides a powerful entry point for the alkylation of oxygen heterocycles. Furthermore, it has been established that the highly nucleophilic SIP generated display tremendous reactivity toward alkyl electrophiles which challenges the traditional notions of carbon–carbon bond formation from organometallic reagents. The investigation and elucidation of the active lithium aggregates were crucial to understanding the origins of this reactivity and have the potential to reshape our approach to the lithiation of weakly acidic C–H bonds.

Supplementary Material

SI
2

ACKNOWLEDGMENTS

A.A.T. is thankful for the generous financial support from the Welch Foundation (Grant A-2081-20210327) and Texas A&M University. O.G. gratefully acknowledges financial support from the National Institutes of Health (Grant R35GM137797), the Camille and Henry Dreyfus Foundation, and the Welch Foundation (Grant A-2102-20220331) for supporting this work. O.G. also acknowledges the Texas A&M University HPRC resources (https://hprc.tamu.edu) for computational resources.

Footnotes

The authors declare no competing financial interest.

REFERENCES

  • (1).(a) Inoue K; Okano K Trapping of Transient Organolithium Compounds. Asian J. Org. Chem 2020, 9, 1548–1561. [Google Scholar]; (b) Capriati V; Perna FM; Salomone A “The Great Beauty” of organolithium chemistry: a land still worth exploring. Dalton Trans 2014, 43, 14204–14210. [DOI] [PubMed] [Google Scholar]; (c) Gessner VH; Däschlein C; Strohmann C Structure formation principles and reactivity of organolithium compounds. Chem.—Eur. J 2009, 15, 3320–3334. [DOI] [PubMed] [Google Scholar]; (d) Chinchilla R; Nájera C; Yus M Metalated Heterocycles and Their Applications in Synthetic Organic Chemistry. Chem. Rev 2004, 104, 2667–2722. [DOI] [PubMed] [Google Scholar]; (e) Beak P; Basu A; Gallagher DJ; Park YS; Thayumanavan S Regioselective, Diastereoselective, and Enantioselective Lithiation-Substitution Sequences: Reaction Pathways and Synthetic Applications. Acc. Chem. Res 1996, 29, 552–560. [Google Scholar]
  • (2).Wietelmann U; Klett J 200 Years of Lithium and 100 Years of Organolithium Chemistry. Z. Anorg. Allg. Chem 2018, 644, 194–204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (3).Stegner P; Harder S Enantioselective Catalysis with s-Block Organometallics. In Early Main Group Metal Catalysis: Concepts and Reactions; Harder S, Ed.; Wiley-VCH: Weinheim, Germany, 2020; pp 251–277. [Google Scholar]
  • (4).(a) Snieckus V Directed ortho metalation. Tertiary amide and O-carbamate directors in synthetic strategies for polysubstituted aromatics. Chem. Rev 1990, 90, 879–933. [Google Scholar]; (b) Wu G; Huang M Organolithium Reagents in Pharmaceutical Asymmetric Processes. Chem. Rev 2006, 106, 2596–2616. [DOI] [PubMed] [Google Scholar]; (c) Kuo SC; Chen F; Hou D; Kim-Meade A; Bernard C; Liu J; Levy S; Wu GG A Novel Enantioselective Alkylation and Its Application to the Synthesis of an Anticancer Agent. J. Org. Chem 2003, 68, 4984–4987. [DOI] [PubMed] [Google Scholar]; (d) Tanoury GJ; Chen M; Dong Y; Forslund R; Jurkauskas V; Jones AD; Belmont D Stereoselective Lithiation and Carboxylation of Boc-Protected Bicyclopyrrolidine: Synthesis of a Key Building Block for HCV Protease Inhibitor Telaprevir. Org. Process Res. Dev 2014, 18, 1234–1244. [Google Scholar]
  • (5).(a) Shimano M; Meyers AI α-Ethoxyvinyllithium·HMPA. Further Studies on its Unusual Basic Properties. Tetrahedron Lett 1997, 38, 5415–5418. [Google Scholar]; (b) Schlosser: Schlosser M Zur aktivierung lithiumorganischer reagenzien. J. Organomet. Chem 1967, 8, 9–16. [Google Scholar]; (c) Schlosser M; Strunk S The “super-basic” butyllithium/potassium tert-butoxide mixture and other LiCKOR-reagents. Tetrahedron Lett 1984, 25, 741–744. [Google Scholar]; (d) Schlosser M The 2 × 3 Toolbox of Organometallic Methods for Regiochemically Exhaustive Functionalization. Angew. Chem., Int. Ed 2005, 44, 376–393. [DOI] [PubMed] [Google Scholar]; (e) Katsoulos G; Takagishi S; Schlosser M The Metalation of Fluoroanisoles: Optional Regioselectivity Due to Metal Mediated Control. Synlett 1991, 1991, 731–732. [Google Scholar]
  • (6).(a) Preparations of benzylithium have been achieved through a variety of methods though typically only the parent benzyllthium is reported. Cleavage of benzyl ethers: Gilman H; Schwebke GL Improved Method for the Preparation of Benzyllithium. J. Org. Chem 1962, 27, 4259–4261. [Google Scholar]; (b) Organometallic interchange from organotin and organomercury compounds: Gilman H; Moore FW; Jones RG Preferential Cleavage of Radicals in Some Organometallic Compounds. J. Am. Chem. Soc 1941, 63, 2482–2485. [Google Scholar]; (c) Gilman H; Rosenberg S Notes-Benzyllithium from Triphenylbenzyltin and Phenyllithium. J. Org. Chem 1959, 24, 2063–2064. [Google Scholar]; (d) Direct metalation: Broaddus CD Homogeneous Metalation of Alkylbenzenes. J. Org. Chem 1970, 35, 10–15. [Google Scholar]
  • (7).Delost MD; Smith DT; Anderson BJ; Njardarson JT From Oxiranes to Oligamers: Architectures of U.S. FDA Approved Pharmaceuticals Containing Oxygen Heterocycles. J. Med. Chem 2018, 61, 10996–11020. [DOI] [PubMed] [Google Scholar]
  • (8).The triple ion classification designates three ions bonded together. This nomenclature was standardized by Collum: Romesberg FE; Collum DB Determination of Structures of Solvated Lithium Dialkylamides by Semiempirical (MNDO) Methods. Comparison of Theory and Experiment. J. Am. Chem. Soc 1992, 114, 2112–2121. [Google Scholar]
  • (9).(a) Reich HJ; Borst JP; Dykstra RR; Green PD A Nuclear Magnetic Resonance Spectroscopic Technique for the Characterization of Lithium Ion Pair Structures in THF and THF/HMPA Solution. J. Am. Chem. Soc 1993, 115, 8728–8741. [Google Scholar]; (b) Reich HJ; Borst JP; Dykstra RR (Trimethylstannyl)- and (tributylstannyl)lithium: Solution Structure in Ether, THF, and HMPA. Organometallics 1994, 13, 1–3. [Google Scholar]; (c) Sikorski WH; Reich HJ The Regioselectivity of Addition of Organolithium Reagents to Enones and Enals: The Role of HMPA. J. Am. Chem. Soc 2001, 123, 6527–6535. [DOI] [PubMed] [Google Scholar]
  • (10).Reich HJ; Sanders AW; Fiedler AT; Bevan MJ The Effect of HMPA on the Reactivity of Epoxides, Aziridines, and Alkyl Halides with Organolithium Reagents. J. Am. Chem. Soc 2002, 124, 13386–13387. [DOI] [PubMed] [Google Scholar]
  • (11).Reich HJ; Sikorski WH Regioselectivity of Addition of Organolithium Reagents to Enones: The Role of HMPA. J. Org. Chem 1999, 64, 14–15. [DOI] [PubMed] [Google Scholar]
  • (12).Panek EJ; Rodgers TJ Role of Ion-Pair Solvation Complexes in the Protonation Stereospecificity of 10-tert-Butyl-9-alkyl-9-metallo-9,10-dihydroanthracenes. J. Am. Chem. Soc 1974, 96, 6921–6928. [Google Scholar]
  • (13).Reich HJ; Sikorski WH; Thompson JL; Sanders AW; Jones AC Interconversion of Contact and Separated Ion Pairs in Silyl- and Arylthio-Substituted Alkyllithium Reagents. Org. Lett 2006, 8, 4003–4006. [DOI] [PubMed] [Google Scholar]
  • (14).Reich HJ; Dykstra RR Structure Effects of Ion Pair Separation: Planar and Pyramidal Sulfur- and Silicon-Substituted Carbanions. J. Am. Chem. Soc 1993, 115, 7041–7042. [Google Scholar]
  • (15).Reich HJ; Sikorski WH; Gudmundsson BO; Dykstra RR Triple Ion Formation in Localized Organolithium Reagents. J. Am. Chem. Soc 1998, 120, 4035–4036. [Google Scholar]
  • (16).Harrison-Marchand A; Mongin F Mixed Aggregate (MAA): A Single Concept for All Dipolar Organometallic Aggregates. 1. Structural Data. Chem. Rev 2013, 113, 7470–7562. [DOI] [PubMed] [Google Scholar]
  • (17).(a) McGarrity JF; Prodolliet J; Smyth T Rapid injection NMR: A simple technique for the observation of reactive intermediates. Org. Magn. Reson 1981, 17, 59–65. [Google Scholar]; (b) Jones AC; Sanders AW; Bevan MJ; Reich HJ Reactivity of Individual Organolithium Aggregates: A RINMR Study of n-Butyllithium and 2-Methoxy-6-(methoxymethyl)phenyllithium. J. Am. Chem. Soc 2007, 129, 3492–3493. [DOI] [PubMed] [Google Scholar]; (c) Denmark SE; Williams BJ; Eklov BM; Pham SM; Beutner GL Design, Validation, and Implementation of a Rapid-Injection NMR System. J. Org. Chem 2010, 75, 5558–5572. [DOI] [PubMed] [Google Scholar]; (d) Thomas AA; Denmark SE; Ernest L Eliel, a Physical Organic Chemist with the Right Tool for the Job: Rapid Injection Nuclear Magnetic Resonance. In Stereochemistry and Global Connectivity: The Legacy of Ernest L. Eliel; Cheng HN, Ed.; American Chemical Society: Washington, DC, 2017; Vol. 2, pp 105–134. [Google Scholar]
  • (18).Gilman H; Gaj BJ Preparation and Stability of Some Organolithium Compounds in Tetrahydrofuran. J. Org. Chem 1957, 22, 1165–1168. [Google Scholar]
  • (19).Bates RB; Kroposki LM; Potter DE Cycloreversions of Anions from Tetrahydrofurans. Convenient Synthesis of Lithium Enolates of Aldehydes. J. Org. Chem 1972, 37, 560–562. [Google Scholar]
  • (20).(a) Mansueto R; Mallardo M; Perna FP; Salomone A; Capriati V Gated access to α-lithiated phenyltetrahydrofuran: functionalisation via direct lithiation of the parent oxygen heterocycle. Chem. Commun 2013, 49, 10160–10162. [DOI] [PubMed] [Google Scholar]; (b) Kennedy AR; Klett J; Mulvey RE; Wright DS Synergic Sedation of Sensitive Anions: Alkali-Mediated Zincation of Cyclic Ethers and Ethene. Science 2009, 326, 706–708. [DOI] [PubMed] [Google Scholar]; (c) Crosbie E; García-Alvarez P; Kennedy AR; Klett J; Mulvey RE; Robertson SD Structurally Engineered Deprotonation/Alumination of THF and THTP with Retention of Their Cycloanionic Structures. Angew. Chem., Int. Ed 2010, 49, 9388–9391. [DOI] [PubMed] [Google Scholar]
  • (21).Clayden J; Yasin SA Pathways for decomposition of THF by organolithiums: the role of HMPA. New J. Chem 2002, 26, 191–192. [Google Scholar]; (b) Fleming also observed the alternate fragmentation product as a side product. Fleming I; Mack SR; Fleming I; Mack SR; Clark BP α-Amino carbene or carbenoid formation in the reaction of a tertiary amide with PhMe2SiLi and its insertion into the Si–Li bond of a second equivalent. Chem. Commun 1998, 713–714. [Google Scholar]
  • (22). We reproduced the observations of Clayden and extended them with TPPA in addition to the HMPA used previously; see Table S1.
  • (23).(a) Baldwin JE Rules for Ring Closure. J. Chem. Soc., Chem. Commun 1976, 734–736. [Google Scholar]; (b) DePuy CH; Bierbaum VM Gas-phase elimination reactions of ethers induced by amide and hydroxide ions. J. Am. Chem. Soc 1981, 103, 5034–5038. [Google Scholar]
  • (24). DMPU and TMEDA were not found to be a suitable HMPA replacements under these conditions.
  • (25).(a) McDonald CE; Ramsey JD; Sampsell DG; Butler JA; Cecchini MR Tripyrrolidinophosphoric Acid Triamide as an Activator in Samarium Diiodide Reductions. Org. Lett 2010, 12, 5178–5181. [DOI] [PubMed] [Google Scholar]; (b) Berndt M; Hölemann A; Niermann A; Bentz C; Zimmer R; Reissig H-U Replacement of HMPA in Samarium Diiodide Promoted Cyclizations and Reactions of Organolithium Compounds. Eur. J. Org. Chem 2012, 2012, 1299–1302. [Google Scholar]; (c) Coste J; Le-Nguyen D; Castro B PyBOP®: A New Peptide Coupling Reagent Devoid of Toxic By-Product. Tetrahedron Lett 1990, 31, 205–208. [Google Scholar]
  • (26). While it is not possible at this time to determine if the observed selectivity is based on kinetic or thermodynamic control, it is clear that the stronger C–H bond has been removed.
  • (27). Further conditions are included in the Supporting Information. See Tables S3–S9.
  • (28).(a) Linden M; Roeters T; Harting R; Stokkingreef E; Gelpke AS; Kemperman G Debottlenecking the Synthesis Route of Asenapine. Org. Process Res. Dev 2008, 12, 196–201. [Google Scholar]; (b) Ray S; Grover PK; Kamboj VP; Setty BS; Kar AB; Anand N Antifertility Agents. 12. Structure-Activity Relation of 3,4-Diphenylchromenes and -Chromans. J. Med. Chem 1976, 19, 276–279. [DOI] [PubMed] [Google Scholar]
  • (29).Cherney AH; Hedley SJ; Mennen SM; Tedrow JS Xantphos as a Branch-Selective Ligand for the Acyclic sec-Alkyl Negishi Cross-Coupling of Heteroaryl Halides. Organometallics 2019, 38, 97–102. [Google Scholar]
  • (30).DMPU reacts with highly nucleophilic alkyllithium reagents; see Table S3 also: Mukhopadhyay T; Seebach D Substitution of HMPT by the Cyclic Urea DMPU as a Cosolvent for Highly Reactive Nucleophiles and Bases. Helv. Chim. Acta 1982, 65, 385–391. [Google Scholar]
  • (31).Schönherr H; Cernak T Profound Methyl Effects in Drug Discovery and a Call for New C–H Methylation Reactions. Angew. Chem., Int. Ed 2013, 52, 12256–12267. [DOI] [PubMed] [Google Scholar]
  • (32).Vitaku E; Smith DT; Njardarson JT Analysis of the Structural Diversity, Substitution Patterns, and Frequency of Nitrogen Heterocycles among U.S. FDA Approved Pharmaceuticals. J. Med. Chem 2014, 57, 10257–10274. [DOI] [PubMed] [Google Scholar]
  • (33).(a) Brown HC; Fletcher RS; Johannesen RB I-Strain as a Factor in the Chemistry of Ring Compounds. J. Am. Chem. Soc 1951, 73, 212–221. [Google Scholar]; (b) For a full list of substrates that were not amenable to the conditions listed here, see Figure S28 in the Supporting Information.
  • (34).Lovering F; Bikker J; Humblet C Escape from Flatland: Increasing Saturation as an Approach to Improving Clinical Success. J. Med. Chem 2009, 52, 6752–6756. [DOI] [PubMed] [Google Scholar]
  • (35).Maksić ZB; Eckert-Maksić M A correlation between the acidity and hybridization in non-conjugated hydrocarbons. Tetrahedron 1969, 25, 5113–5114. [Google Scholar]
  • (36).Bailey WF; Patricia JJ; Nurmi TT; Wang W Metal—halogen interchange between t-butyllithium and 1-iodo-5-hexenes provides no evidence for single-electron transfer. Tetrahedron Lett 1986, 27, 1861–1864. [Google Scholar]
  • (37).(a) Denmark SE; Chen C-T Alkylation of Chiral, Phosphorus-Stabilized Carbanions: Substituent Effects on the Alkylation Selectivity. J. Org. Chem 1994, 59, 2922–2924. [Google Scholar]; (b) Maekawa Y; Maruyama T; Murai T Sequential Deprotonation–Alkylation of Binaphthyloxy-Substituted Phosphonochalcogenoates: Chiral Tri- and Tetrasubstituted Carbon Centers Adjacent to a Phosphorus Atom. Org. Lett 2016, 18, 5264–5267. [DOI] [PubMed] [Google Scholar]
  • (38).Crockett MP; Aguirre LS; Jimenez LB; Hsu H-H; Thomas AA Preparation of Highly Reactive Lithium Metal Dendrites for the Synthesis of Organolithium Reagents. J. Am. Chem. Soc 2022, 144, 16631–16637. [DOI] [PubMed] [Google Scholar]
  • (39). See Supporting Information for full sample preparation details.
  • (40). 6Li–31P coupling constants are much smaller than the corresponding 7Li–31P coupling constants.
  • (41).Reich HJ; Sikorski WH; Sanders AW; Jones AC; Plessel KN Multinuclear NMR Study of the Solution Structure and Reactivity of Tris(trimethylsilyl)methyllithium and its Iodine Ate Complex. J. Org. Chem 2009, 74, 719–729. [DOI] [PubMed] [Google Scholar]
  • (42). Higher equivalents of HMPA lead to solution freezing at the temperatures required for titration.
  • (43).These data are nearly identical to spectroscopic data collected by Reich for the characterization of triple ion pairs with carbanions bearing stabilizing groups. See the following: Reich HJ; Borst JP Direct Nuclear Magnetic Resonance Spectroscopic Determination of Organolithium Ion Pair Structures in THF/HMPA Solution. J. Am. Chem. Soc 1991, 113, 1835–1837. [Google Scholar]
  • (44).(a) Lee C; Yang W; Parr RG Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789. [DOI] [PubMed] [Google Scholar]; (b) Becke AD Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys 1993, 98, 5648–5652. [Google Scholar]; (c) Grimme S Accurate description of van der Waals complexes by density functional theory including empirical corrections. J. Comput. Chem 2004, 25, 1463–1473. [DOI] [PubMed] [Google Scholar]; (d) Grimme S; Antony J; Ehrlich S; Krieg H A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys 2010, 132, 154104. [DOI] [PubMed] [Google Scholar]; (e) Grimme S Density functional theory with London dispersion corrections. Wiley Interdiscip. Rev.: Comput. Mol. Sci 2011, 1, 211–228. [Google Scholar]; (f) Ehrlich S; Moellmann J; Grimme S Dispersion-Corrected Density Functional Theory for Aromatic Interactions in Complex Systems. Acc. Chem. Res 2013, 46, 916–926. [DOI] [PubMed] [Google Scholar]
  • (45).(a) Ditchfield R; Hehre WJ; Pople JA Self-consistent Molecular-orbital Methods. IX. An Extended Gaussian-type Basis for Molecular-orbital Studies of Organic Molecules. J. Chem. Phys 1971, 54, 724–728. [Google Scholar]; (b) Hehre WJ; Ditchfield R; Pople JA Self—Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian—Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules. J. Chem. Phys 1972, 56, 2257–2261. [Google Scholar]; (c) Hariharan PC; Pople JA The Influence of Polarization Functions on Molecular Orbital Hydrogenation Energies. Theor. Chim. Acta 1973, 28, 213–222. [Google Scholar]; (d) Kwon O; Sevin F; McKee ML Density Functional Calculations of Methyllithium, t-Butyllithium, and Phenyllithium Oligomers: Effect of Hyperconjugation on Conformation. J. Phys. Chem. A 2001, 105, 913–922. [Google Scholar]
  • (46).(a) Klamt A; Schüürmann, G. COSMO: a new approach to dielectric screening in solvents with explicit expressions for the screening energy and its gradient. J. Chem. Soc. Perkin Trans. 2 1993, 0, 799–805. [Google Scholar]; (b) Tomasi J; Persico M Molecular Interactions in Solution: An Overview of Methods Based on Continuous Distributions of the Solvent. Chem. Rev 1994, 94, 2027–2094. [Google Scholar]; (c) Andzelm J; Kölmel C; Klamt A Incorporation of solvent effects into density functional calculations of molecular energies and geometries. J. Chem. Phys 1995, 103, 9312–9320. [Google Scholar]; (d) Barone V; Cossi M Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent Model. J. Phys. Chem. A 1998, 102, 1995–2001. [Google Scholar]; (e) Cossi M; Rega N; Scalmani G; Barone V Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation model. J. Comput. Chem 2003, 24, 669–681. [DOI] [PubMed] [Google Scholar]
  • (47).Frisch MJ; Trucks GW; Schlegel HB; Scuseria GE; Robb MA; Cheeseman JR; Scalmani G; Barone V; Petersson GA; Nakatsuji H; Li X; Caricato M; Marenich AV; Bloino J; Janesko BG; Gomperts R; Mennucci B; Hratchian HP; Ortiz JV; Izmaylov AF; Sonnenberg JL; Williams-Young D; Ding F; Lipparini F; Egidi F; Goings J; Peng B; Petrone A; Henderson T; Ranasinghe D; Zakrzewski VG; Gao J; Rega N; Zheng G; Liang W; Hada M; Ehara M; Toyota K; Fukuda R; Hasegawa J; Ishida M; Nakajima T; Honda Y; Kitao O; Nakai H; Vreven T; Throssell K; Montgomery JA Jr.; Peralta JE; Ogliaro F; Bearpark MJ; Heyd JJ; Brothers EN; Kudin KN; Staroverov VN; Keith TA; Kobayashi R; Normand J; Raghavachari K; Rendell AP; Burant JC; Iyengar SS; Tomasi J; Cossi M; Millam JM; Klene M; Adamo C; Cammi R; Ochterski JW; Martin RL; Morokuma K; Farkas O; Foresman JB; Fox DJ Gaussian 16, revision C.01; Gaussian, Inc.: Wallingford, CT, 2016. [Google Scholar]
  • (48).Legault CY (2009) CYLview, 1.0b, Université de Sherbrooke: Sherbrooke, Canada, http://www.cylview.org (accessed Mar 22, 2023). [Google Scholar]
  • (49).Jones AC; Sanders AW; Sikorski WH; Jansen KL; Reich HJ Reactivity of the Triple Ion and Separated Ion Pair of Tris(trimethylsilyl)methyllithium with Aldehydes: A RINMR Study. J. Am. Chem. Soc 2008, 130, 6060–6061. [DOI] [PubMed] [Google Scholar]
  • (50).Oeschger R; Su B; Yu I; Ehinger C; Romero E; He S; Hartwig J Diverse functionalization of strong alkyl C–H bonds by undirected borylation. Science 2020, 368, 736–741. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (51). See Supporting Information for further discussions on the regioselectivity for THF lithiation.
  • (52).Slocum DW; Jennings CA Directed Metalation Reactions. 6. Competition of Substituents for Ortho Direction of Metalation in Substituted Anisoles. J. Org. Chem 1976, 41, 3653–3664. [Google Scholar]
  • (53).Gschwend HW; Rodriguez HR Heteroatom-Facilitated Lithiations. Org. React 1979, 26, 1–360. [Google Scholar]
  • (54).Gladfelder JJ; Ghosh S; Podunavac M; Cook AW; Ma Y; Woltornist RA; Keresztes I; Hayton TW; Collum DB; Zakarian A Enantioselective Alkylation of 2-Alkylpyridines Controlled by Organolithium Aggregation. J. Am. Chem. Soc 2019, 141, 15024–15028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (55).Weidmann N; Ketels M; Knochel P Sodiation of Arenes and Heteroarenes in Continuous Flow. Angew. Chem., Int. Ed 2018, 57, 10748–10751. [DOI] [PubMed] [Google Scholar]
  • (56).Harenberg JH; Weidmann N; Knochel P Preparation of Functionalized Aryl, Heteroaryl, and Benzylic Potassium Organometallics Using Potassium Diisopropylamide in Continuous Flow. Angew. Chem., Int. Ed 2020, 59, 12321–12325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (57).Cram DJ; Gosser L Electrophilic Substitution at Saturated Carbon. XXI. Isoracemization Reactions Involving Ion-Pair Intermediates. J. Am. Chem. Soc 1964, 86, 5457–5465. [Google Scholar]
  • (58).Screttas CG; Heropoulos GA; Micha-Screttas M; Steele BR; Catsoulacos DP Reductive lithiation of alkyl phenyl sulfides in diethyl ether. A ready access to α,α-dialkylbenzyllithiums. Tetrahedron Lett 2003, 44, 5633–5635. [Google Scholar]
  • (59).Binder JT; Cordier CJ; Fu GC Catalytic Enantioselective Cross-Couplings of Secondary Alkyl Electrophiles with Secondary Alkylmetal Nucleophiles: Negishi Reactions of Racemic Benzylic Bromides with Achiral Alkylzinc Reagents. J. Am. Chem. Soc 2012, 134, 17003–17006. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

SI
2

RESOURCES