Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2024 Sep 4.
Published in final edited form as: J Am Chem Soc. 2023 Oct 19;145(43):23568–23584. doi: 10.1021/jacs.3c07317

Carbon-Nitrogen Bond Formation Using Sodium Hexamethyldisilazide: Solvent-Dependent Reactivities and Mechanisms

Qiulin You 1, David B Collum 1,*
PMCID: PMC11373886  NIHMSID: NIHMS2015774  PMID: 37857357

Abstract

The solvent-dependent reactivity of sodium hexamethyldisilazide (NaHMDS) toward carbon-centered electrophiles reveals reactions that are poorly represented or unrepresented in the literature including direct aminolysis of aromatic methyl esters to give carboxamides, nitriles, or amidines, depending on the choice of solvent. SNAr substitutions of aryl halides and opening of terminal epoxides are also examined. A combination of 1H and 29Si NMR spectroscopic studies using [15N]NaHMDS, kinetic studies, and computational studies reveal the complex mechanistic basis of the preferences for simple aryl carboxamides in toluene and dimethylethylamine and arylnitriles or amidines in THF. A prevalence of dimer- and mixed dimer-based chemistry even starting from the observable NaHMDS monomer in THF solution is notable.

Graphical Abstract:

graphic file with name nihms-2015774-f0001.jpg

Introduction

After decades of communal disinterest in organosodium chemistry, we began trying to shine a light on its potential by expanding the toolbox for generating and using organosodium species.13 This involved some minor adjustments to protocols for generating and handling sodium diisopropylamide (NaDA) 2,3 as well as developing a new reagent, sodium isopropyl(trimethyl)silyl amide (NaPTA), that manifests desirable solubilities and reactivities as a strong base.4 Our primary approach, however, is decidedly structural and mechanistic with the faith that understanding how solvation and aggregation influence reactivity and selectivity will propel applications through a combination of serendipity in our lab and need-driven progress by others. Even the most prominent organosodium reagent, sodium hexamethyldisilazide (NaHMDS),5 while garnering the attention of crystallographers,6 had evaded spectroscopic, mechanistic, and computational scrutiny until recent studies of its solvent-dependent structure and reactivity toward enolization.7 This brings us to the current work. Aminolyses illustrated with generic examples in Scheme 1 are legion. SNAr substitutions and peptide bond formations are of unquestioned importance in pharmaceutical chemistry.8,9 With that said, any reaction in which the electrophile is merely heated in ammonia or simple alkylamine will be difficult to improve upon. However, should these simple protocols fail, whether owing to low reactivity or poor selectivity, the experimentalist is left with few options. In these instances, metal amides could offer solutions through control of their coordination spheres.

Scheme 1.

Scheme 1.

Generic aminolyses

We describe herein a survey of the reactivity of NaHMDS toward carbon-centered electrophiles of potential interest in synthesis accompanied by detailed structural and mechanistic studies. NaHMDS plays the role of a highly reactive analog of ammonia and a preface to expanding investigations of sodium alkylsilazides.4 Some of the transformations surveyed as opportunities for mechanistic studies piqued our interest as synthetically promising and have surprisingly little or no presence in the NaHMDS literature. The more mechanistically inclined will find the prevalence of dimer- and mixed aggregate-based reactivity surprising.10 We also inadvertently stumbled into the complex world of organosilicon chemistry.

Results and Discussion

Aminolyses.

Substitutions of representative organic substrates by NaHMDS are summarized in Table 1. The yields are of isolated, purified products that have been desilylated during workup. 1H and 29Si NMR spectroscopic monitoring (vide infra) shows the reactions are often pristine and reveals unisolated silylated and sodiated intermediates. The times and temperatures listed in Table 1 are those required to achieve high conversion as confirmed by in situ monitoring using NMR or IR spectroscopies.

Table 1.

Reactions of NaHMDS with electrophiles in various solvents.

entry substrate conditions product yield
1 graphic file with name nihms-2015774-t0039.jpg
1
2.0 equiv NaHMDS 25 °C, 5 h toluene graphic file with name nihms-2015774-t0040.jpg
2
72%
2 graphic file with name nihms-2015774-t0041.jpg
3
3.0 equiv NaHMDS 25 °C, 0.3 h DMEA graphic file with name nihms-2015774-t0042.jpg
4
95%
3 graphic file with name nihms-2015774-t0043.jpg
5
3.0 equiv NaHMDS 25 °C, 0.3 h toluene graphic file with name nihms-2015774-t0044.jpg
4
90%
4 graphic file with name nihms-2015774-t0045.jpg
3
1.0 equiv NaHMDS 50 °C, 1.0 h THF graphic file with name nihms-2015774-t0046.jpg
6
86%
5 graphic file with name nihms-2015774-t0047.jpg
3
3.0 equiv NaHMDS 50 °C, 0.3 h THF graphic file with name nihms-2015774-t0048.jpg
7
92%
6 graphic file with name nihms-2015774-t0049.jpg
8
3.0 equiv NaHMDS 70 °C, 0.3 h toluenea graphic file with name nihms-2015774-t0050.jpg
9
85%
7 graphic file with name nihms-2015774-t0051.jpg
8
3.0 equiv NaHMDS 70 °C, 1.0 h THF graphic file with name nihms-2015774-t0052.jpg
10
96%
8 graphic file with name nihms-2015774-t0053.jpg
11
2.0 equiv NaHMDS 70 °C, 1.0 h toluene graphic file with name nihms-2015774-t0054.jpg
12
76%
9 graphic file with name nihms-2015774-t0055.jpg
11
3.0 equiv NaHMDS 70 °C, 1.0 h THF graphic file with name nihms-2015774-t0056.jpg
13
95%
10 graphic file with name nihms-2015774-t0057.jpg
14
3 equiv NaHMDS 50 °C, 1 h DMEA graphic file with name nihms-2015774-t0058.jpg
15
78%
11 graphic file with name nihms-2015774-t0059.jpg
14
3.0 equiv NaHMDS/THF 70 °C, 2.0 h 25 °C, 24 h graphic file with name nihms-2015774-t0060.jpg
16
92%
12 graphic file with name nihms-2015774-t0061.jpg
17
2.0 equiv NaHMDS 25 °C, 0.05 h DMEA graphic file with name nihms-2015774-t0062.jpg
16
95%
13 graphic file with name nihms-2015774-t0063.jpg
18
2.0 equiv NaHMDS 25 °C, 2 h toluene graphic file with name nihms-2015774-t0064.jpg
19
85%
14 graphic file with name nihms-2015774-t0065.jpg
20
2.0 equiv NaHMDS 25 °C, 1.0 h toluene graphic file with name nihms-2015774-t0066.jpg
21
77%
15 graphic file with name nihms-2015774-t0067.jpg
22
2.0 equiv NaHMDS 25 °C, 1.0 h toluene graphic file with name nihms-2015774-t0068.jpg
23
83%
16 graphic file with name nihms-2015774-t0069.jpg
24
2.0 equiv NaHMDS 110 °C, 2 h toluene graphic file with name nihms-2015774-t0070.jpg
25
55%
17 graphic file with name nihms-2015774-t0071.jpg
26
2.0 equiv NaHMDS 110 °C, 3 h toluene graphic file with name nihms-2015774-t0072.jpg
27
76%
18 graphic file with name nihms-2015774-t0073.jpg
28
2 equiv NaHMDS 60 °C, 24 h toluene graphic file with name nihms-2015774-t0074.jpg
(>50:1)
29
86%
19 graphic file with name nihms-2015774-t0075.jpg
30
2 equiv NaHMDS 25 °C, 24 h THF graphic file with name nihms-2015774-t0076.jpg
31
76%
a

Forms a gel-like mixture during the reaction.

For context, the aminolysis of methyl cinnamate to form carboxamide (entry 1) rather than alternatives such as 1,4-addition,11,12 enolization,5 or generalized destruction caught our attention. Ester aminolyses by LiHMDS and NaHMDS may populate pharmaceutical notebooks, but they are poorly represented in the published literature. In 1963, Kruger et al. aminolyzed an aryl ester with NaHMDS to form an O-methyloximino ether, an intermediate en route to carboxamides (as in entry 2). One is detected in mechanistic studies below.13 In 1998, Hwu and coworkers reported a NaHMDS-mediated conversion of aryl esters to nitriles with NaHMDS at 110–185 °C (as in entry 4), attributing a central importance to phenolic groups.14 One-pot conversions of aryl esters to amidines (entries 11 and 12) or carboxamides (such as entry 2) by NaHMDS are unreported. Additions to nitriles to form amidinates (as in entry 12) are well known for LiHMDS.15 Mechanistic studies below offer some thoughts on why analogous additions of NaHMDS to nitriles are rare.16

The potentially useful selectivities for the formation of carboxamides, nitriles, and amidines directly from esters have complex mechanistic underpinnings dependent on solvent, temperature, and equivalents of NaHMDS (vide infra). The 3- and 4-picoline methyl esters form gels in weak solvents (entries 9 and 11), presumably owing to head-to-tail oligomeric substrate-NaHMDS dimer complexes but without negative consequences.

SNAr reactions (entries 15–19) represent a reaction class of unquestioned importance.8 We find it somewhat confounding that SNAr reactions seem to be more effective with soft nucleophiles. Malonates, for example, are far superior nucleophiles17 than unstabilized enolates.18 We can find only one report of direct (uncatalyzed) SNAr substitution by LiHMDS19 and none for NaHMDS, while RNH2-based aminolyses facilitated by LiHMDS are legion albeit with an unknown role of the LiHMDS.20 We expended considerable effort trying to glean mechanistic insights (entry 15–17) only to be thwarted by deeply colored debris that appears to be diazo derivatives arising from NaHMDS reacting with the nitro moiety noted years ago (eq 1).21 Nitroarenes are spartan in the voluminous literature of lithium amide-mediated ortholithiations,22 possibly for similar reasons.

graphic file with name nihms-2015774-f0031.jpg (1)

A few problematic substrates are illustrated in Chart 1. Arenes 3234 gave complex products suggestive of orthometalations and possibly pyridyne intermediates. Halopyridines 35 and 36 and chloroisoquinoline 37 aminated but were prone to transfer a silyl group to the 3-position by a Fries-like process.23 Nitrobenzene 38 underwent a clean halogen dance,24 forming 2-fluoro-3-iodonitrobenzene.25 We replicated the highly regioselective epoxide openings reported by Withnall and coworkers in 200826 (entries 20 and 21) to evaluate them as candidates for mechanistic work and include these potentially useful results because they have gone largely unnoticed.27

Chart 1.

Chart 1.

Failed and troublesome substrates

To assist the reader, Scheme 2 represents a roadmap to what is forthcoming in the mechanistic studies. We explore the reaction coordinates to convert NaHMDS and methyl picolinate 3 to carboxamide 4 in DMEA (Part 1), to nitrile 6 in THF (Part 2), and conversion of nitrile 6 to amidine 7 in DMEA (Part 3) and THF (Part 4). Each part delves into the underlying organosodium and organosilicon chemistry by examining spectroscopically observable structural events, kinetic studies to evaluate solvation and aggregation in the rate-limiting transition structures,28 and computational probes of solvation, aggregation, critical transition structures, and elusive details along the reaction coordinates.29,30

Scheme 2.

Scheme 2.

Focus of the four studies examined spectroscopically, kinetically, and computationally.

Monitoring the reactions by 1H and 29Si NMR spectroscopies showed the reactions to be very clean and provided detailed structural assignments. [15N]NaHMDS7 distinguished O–Si and 15N–Si species owing to 15N–29Si coupling whereas combinations of [15N]NaHMDS and NaHMDS were exploited to examine reversible steps. DFT computations using Me3N in place of Me2NEt to reduce unnecessary degrees of freedom revealed an exothermic substitution of DMEA on dimer 39. We take the liberty of using the original numbers assigned to DMEA solvates in the schemes showing computed Me3N solvates. The shorthand of general form AmSn alludes to the hexamethyldisilazide fragment (A) and solvent (S).

Part 1: Mechanism of aminolysis of methyl-2-picolinate (3) with NaHMDS/DMEA.

We focused structure-reactivity studies on additions to methyl-2-picolinate in DMEA and THF. Toluene acts similarly to DMEA but affords broader resonances. The reaction coordinate for the addition of NaHMDS to picoline 3 in DMEA observable by NMR spectroscopy is summarized Scheme 3. Mono- and dichelated dimers 40 and 41 observable at −80 °C reflect the behavior of NaHMDS with other bifunctional ligands studied previously.7 Both dimers display coupling constants (1JN-Si = 8.5 and 8.7 Hz, respectively) and chemical shifts (−15.8 and −16.4 ppm, respectively) characteristic of [15N]NaHMDS dimers.7 Addition of excess 3 affords computationally viable doubly chelated monomer 42 with an upfield 29Si chemical shift and large coupling constant (−21.41 ppm, 1JN-Si = 13.1 Hz) characteristic of a NaHMDS monomer.7 Rate studies (below) place dimer 40 on the reaction coordinate. The stability of 43 allows carboxamide 431 to be isolated in high yield (Table 1, entry 5).32

Scheme 3.

Scheme 3.

Spectroscopically observed species on the reaction of methyl-2-picolinate with NaHMDS/DMEA.33

The computed structure of 40 (Figure 1) reveals a geometry that more closely approximates square planar than tetrahedral sodium. Serial substitution from 39 to 41 is computed to be mildly exothermic for both steps (ignoring translational entropy).34 DFT supports monomer stereoisomer 42 with the alternative stereoisomeric monomer (not drawn) to be 10.4 kcal/mol less stable.

Figure 1.

Figure 1.

Ball and stick depictions of DFT-computed dimer 40 displaying an approximate 10° rotation of the chelate and Na2N2 planes from coplanarity.

The only observable species during the aminolysis in DMEA are mixed aggregate 43 (Figure 2A) and Me3SiOMe with the latter confirmed by comparison with an authentic sample (29Si, δ 17.49 ppm). Two 29Si resonances of 43 at −16.0 ppm (1JN-Si = 9.7 Hz) and −16.6 ppm (1JN-Si = 10.3 Hz) are similar to [15N]NaHMDS dimer 39 (−15.2 ppm, JN-Si = 8.5 Hz) and coalesce into a single broad resonance at −20 °C. The imino 29Si resonance is markedly downfield (−9.70 ppm, JN-Si = 11.2 Hz). Addition of excess unlabelled NaHMDS to pre-formed [15N2]43 at −80 °C shows immediate incorporation of unlabelled silazide fragments (Figure 2B). DFT computations illustrate the magnetic inequivalence of the two silazide-derived Me3Si moieties in 43 (Figure 3).

Figure 2.

Figure 2.

Spectrum A is 29Si NMR spectrum of 0.05 M ester 3 with 0.10 M added [15N]NaHMDS in DMEA after reaction containing only mixed aggregate [15N]43 and residual NaHMDS dimer [15N]39 (Scheme 3). Spectrum B is the sample from spectrum A with 1.0 equiv (relative to ester 3) of unlabelled NaHMDS.

Figure 3.

Figure 3.

Ball and stick depictions of DFT-computed and artist’s rendition of mixed aggregate 43 with Me3N as a DMEA surrogate.

We probed the mechanism for the addition of NaHMDS to picolinate 3 in DMEA-pentane mixtures using in situ IR spectroscopy35 under pseudo-first-order conditions (excess NaHMDS) following the loss of monochelated dimer 40 (1737 cm−1) to form mixed dimer 43 (1584 cm−1). Clean first-order decays (Figure 4) afford pseudo-first-order rate constants, kobsd, that are independent of the initial concentration of 40. Plotting kobsd versus DMEA concentration with pentane cosolvent (Figure 5) and versus NaHMDS dimer 39 concentration (Figure 6) both show zeroth-order dependencies. The resulting rate law (eq 2) is consistent with a rate-limiting addition via a transition structure of stoichiometry A2S(substrate). (Recall that A an amide subunit and S is solvent.) The dimer-based reaction coordinate is a hallmark of alkali metal amides in poorly coordinating solvents.2,7c,36 Figure 7 depicts the computed transition structure TS-1 displaying a developing Na2(O)(N) mixed dimer core.

Figure 4.

Figure 4.

Addition of 0.10 M NaHMDS to methyl-2-picolinate (3, 0.005 M) at 25 °C measured with IR spectroscopy (1737 cm−1). The curve depicts an unweighted least-squares fit to y = aebx (a = 8.1 × 10−3 ± 0.1 × 10−3; b = 1.5 × 10−3 ± 0.1 × 10−3).

Figure 5.

Figure 5.

A plot of kobsd vs [DMEA] (M) in pentane for the addition of NaHMDS (0.10 M) to methyl-2-picolinate (3, 0.005 M) at 25 °C measured with IR spectroscopy (1737 cm−1). The curve depicts an unweighted least-squares fit to y = ax + b (a = 2.4 × 10−3 ± 0.2 × 10−3; b = 8.3 × 10−6 ± 0.3 × 10−6).

Figure 6.

Figure 6.

A plot of kobsd vs [NaHMDS] (M) for the addition of NaHMDS to methyl-2-picolinate (3, 0.050 M) in 3.8 M (neat) DMEA at 25 °C measured with IR spectroscopy (1737 cm−1). The curve depicts an unweighted least-squares fit to y = ax + b (a = 2.7 × 10−3 ± 0.2 × 10−3; b = −1.3 × 10−3 ± 0.1 × 10−3).

Figure 7.

Figure 7.

DFT-computed transition structure TS-1 for the NaHMDS-based addition to methyl-2-picolinate (3) with Me3N as a DMEA surrogate.

-d[40]/dt=k[40]1A2S20[S]0 (2)

Probing the reaction coordinate en route to observable product 43 (carboxamide 13 after workup) and Me3SiOMe using DFT computations served up a nuanced and quite complex story (Scheme 4). To reiterate, Me3N is used as a surrogate of DMEA to reduce unnecessary degrees of freedom while taking the liberty of reusing the original numbers for 40, 43, and 46. Transition structures TS-1, TS-2, TS-3, and TS-4 all have single negative frequencies. Intrinsic reaction coordinate (IRC) calculations37 were used to determine the minima flanking each transition structure. Minima connected without connecting transition structures represent structural minima corresponding to reorganizations of insufficient interest to probe further. There are likely more minima and barriers resulting from other minor adjustments. These same caveats also apply to the reaction coordinate depicted in Scheme 6 (below).

Scheme 4.

Scheme 4.

DFT-computed reaction coordinate for the aminolysis of methyl-2-picolinate (3) by NaHMDS using Me3N as a DMEA surrogate.

Scheme 6.

Scheme 6.

DFT-computed reaction coordinate for the aminolysis of methyl-2-picolinate (3) with NaHMDS in THF with the +Na(THF)6 omitted for all structures.

Following the formation of open dimer 45,36 the 1,2-addition proceeds via rate-limiting transition structure TS-1 (depicted in Figure 7) to form the tetrahedral adduct as mixed aggregate 46. Structures beyond TS-1 represent kinetically and spectroscopically invisible post-rate-limiting events. The collapse of the adduct 46 via transition structure TS-2 affords NaOMe–NaHMDS mixed dimer 47. Following a readjustment via TS-3 to position the methoxy as a MeO–Si Lewis acid-base complex 48, silyl transfer via transition structure TS-4 extrudes Me3SiOMe via complex 49 to form mixed dimer 50 (same structure compared to mixed aggregate 43 instead of utilizing Me3N as a solvent surrogate in DFT-computation) as the spectroscopically observable product (albeit with an exothermic addition of a second solvent.)

As drawn, the NaOMe silylation is an intra-aggegate reaction. We cannot exclude the possibility that NaOMe is released into solution before silyl transfer but have no reason to invoke it either. We also probed for a methoxy-silyl coupling directly from the tetrahedral adduct 46 without first expelling NaOMe (see 51, path a) but were unable to locate such a transfer. Similarly, a silyl transfer to extrude NaOSiMe3 from 46 (see 51, path b) was located but had an unreasonably high (23 kcal/mol) barrier consistent with our failure to observe Me3SiONa spectroscopically.

graphic file with name nihms-2015774-f0032.jpg

Part 2: Mechanism of aminolysis of methyl-2-picolinate (3) with NaHMDS/THF.

The aminolysis in THF follows a distinctly different pathway. The spectroscopically observable forms summarized in Scheme 5 bely mechanistic complexity lurking beneath the surface. At the outset, we note that the excess NaHMDS undergoes a THF-dependent deaggregation of disolvated dimer 52 to provide tetrasolvated monomer 53 described previously,7a,b which proves to be mechanistically important. Picolinate-complexed dimer 54 (analogous to 9) and THF-complexed monomer 55 can be observed at −110 °C at both low and high THF concentrations, respectively, using pentane cosolvent. Evidence of complexation is gleaned from distinct chemical shift differences in the 1H and 29Si NMR when compared with the uncomplexed substrate 3 and free NaHMDS (Supporting Information). The assigned THF solvation numbers derive from DFT computations.

Scheme 5.

Scheme 5.

Spectroscopically observed species for the reaction of NaHMDS to methyl-2-picolinate 3 in THF.

Warming mixtures of 3 and NaHMDS to room temperature affords imino ether 56, analogous to that isolated previously,13 within 1.0 min. The concomittant formation of Me3SiONa was confirmed by comparison with an authentic sample (17.70 ppm). In the absence of excess NaHMDS, 56 gives nitrile 6 over 24 h at RT or in 30 minutes at 50 °C in good yield (Table 1, entry 7) with concomitant formation of Me3SiOMe, confirmed with an authentic sample. The conversion of 56 to 6 is somewhat mysterious in that it does not appear to require NaHMDS and is not accelerated by excess NaHMDS, yet it does not occur by heating a crude isolated sample of 56.

Reaction of 3 with excess NaHMDS renders nitrile 6 unobservable because it is quickly scavenged to give amidinate 5738 with noteworthy spectroscopic properties. The 29Si resonance of [15N2]57 appears as a clean triplet owing to second-order effects referred to as “virtual coupling.” The two magnetically equivalent 29Si nuclei experience equivalent coupling by the 15N nuclei despite one being proximal and one being distal to each silicon nucleus (Figure 8). Similar virtual coupling is observed in bridging transition metal phosphides (M2P2 core)39 and has been observed in a Li2P2 lithium phosphide dimer.40 In theory, we could have observed this in NaHMDS (Na2N2) dimers as well but have not.7 If the symmetry is broken by generating [15N1]57 from unlabelled nitrile 6 and [15N]NaHMDS, the 29Si resonance appears as a doublet of doublets owing to a large 1JN-Si coupling and small 3JN-Si coupling with a slight isotopic perturbation visible.41 The requisite silyl transfer from the initially formed unsymmetric N,N-disilylamidine to N,N’-disilylamidine 56 is discussed in the context of DFT computations below. The depicted 657 nitrile-amidinate equilibrium cannot be observed directly, but adding [14N]NaHMDS to labelled amidinate 57 results in label exchange (eq 3).

Figure 8.

Figure 8.

29Si NMR spectra of [15N2]57 showing virtual coupling (1JN–Si = 8.8 Hz) and standard coupling in [15N1]57 (1JN–Si = 8.9 Hz and 3JN–Si = 2.8 Hz).

graphic file with name nihms-2015774-f0033.jpg (3)

An analogous addition of NaHMDS to PhCO2Me (Table 1, entry 13) to generate phenyl-substituted amidinate 58 was approximately 60-fold (see in supporting information) slower than for 3. It also revealed the elusive nitrile-amidinate exchange as a temperature-dependent equilibrium (eq 4). Heating excess NaHMDS (3.0 equiv) and PhCO2Me to 80 °C followed by rapid thermal quenching to −80 °C affords predominantly benzonitrile (17). Alternatively, cooling the reaction to 25 °C and letting it stand for 12 h generates amidinate 58. Heating-cooling cycles show the changes are temperature dependent and reversible. Adding [15N]NaHMDS to unlabelled benzonitrile, 17, shows label incorporation in both 17 and 58. Thus, benzonitrile can be isolated in 86% yield using 1.1 equiv of NaHMDS and elevated temperatures (Table 1, entry 13), while amidine 58 can be isolated in 98% yield using 3.0 equiv NaHMDS at 80 °C for 3 h, then 25 °C for 12 h. This odd temperature dependence in which heating accelerates the aminolysis but retards addition to the intermediate nitrile might be why nitrile-to-amidine conversions largely exploit LiHMDS15 rather than NaHMDS.16 Alternatively, it could just be a cultural preference for LiHMDS. There are no extraneous silazide signals in the 29Si spectrum that would implicate a mixed aggregate.

graphic file with name nihms-2015774-f0034.jpg (4)

Rate studies for the addition of NaHMDS to picolinate 3 to afford imino ether 56 provided yet another nuanced story. We anticipated that the first-order dependence—the first-order decay—on complexed monomer 55 measured in neat THF would be accompanied by a zeroth-order NaHMDS dependence. To the contrary, a plot of kobsd vs NaHMDS shows a half-order dependence on the free NaHMDS monomer (Figure 9). The partial rate law (deferring discussion of solvent-concentration dependencies momentarily) is described by eq 5. To be clear on a critical point: A dimer-based addition involving contributions of a standard NaHMDS monomer would manifest a first-order NaHMDS dependence.42 The half-order on a monomer demands a full ionization of the sodium cation.43

d[55]/dt=k[55]1AS41/2 (5)

such that 55=AS2(3)

Figure 9.

Figure 9.

The plot of kobsd vs [NaHMDS] (M) for the addition of NaHMDS to methyl-2-picolinate (3, 0.005 M) in 12.8 M (neat) THF at 0 °C measured with IR spectroscopy (1731 cm−1). The curve depicts an unweighted least-squares fit to y = axn (a = 2.4 × 10−3 ± 0.1 × 10−3; n = 0.44 ± 0.01).

We may not yet fully understand the role of solvation. The forthcoming discussion may even seem excessive, but we are still groping to understand how organosodium solvation and aggregation differ from those of organolithiums. A plot of kobsd vs THF in pentane (Figure 10, curve A) manifests a maximum in the rates at intermediate THF concentrations. By contrast, analogous data using 2,5-dimethyltetrahydrofuran (2,5-Me2THF) or the decidedly more expensive 2,2,5,5-tetramethyltetrahydrofuran (2,2,5,5-Me4THF) as poorly coordinating polar cosolvents4447 eliminate the maximum. The solvent dependencies suggest that medium effects are in play. We remind the reader that two deaggregations are occurring concurrently. At low THF concentrations, the dimeric reactant A2S(substrate) (54) exists concurrently with A2S2 (NaHMDS dimer 52), and the rise in rates indicates they are collectively under-solvated relative to the rate-limiting transition structure. At high THF concentrations in polar cosolvents, the observable reactants AS2(substrate) (54) and AS47c (53) appear optimally solvated—six THF ligands total—but under-aggregated relative to the dimeric rate-limiting transition structure.

Figure 10.

Figure 10.

A plot of kobsd vs [THF] (M) for the addition of NaHMDS to methyl-2-picolinate (3, 0.005 M) in 0.10 M NaHMDS at 0 °C measured with IR spectroscopy (1732 cm−1). Curve A derives from pentane cosolvent and depicts an unweighted least-squares fit to y = [axn/(1+bxn)](1/1+cx2) (a = 2.4 × 10−3 ± 0.1 × 10−3; b = 1.3 × 10−1 ± 0.1 × 10−1; c = 0.6 × 10−3 ± 0.1 × 10−4; n = 5). Curve B derives from 2,5-Me2THF cosolvent and depicts an unweighted least-squares fit to y = axn/(1+bxn) (a = 1.6 × 10−1 ± 0.1 × 10−1; b = 1.6 × 10−1 ± 0.1 × 10−1; n = 5). Curve C derives from 2,2,5,5-Me4THF cosolvent and depicts an unweighted least-squares fit to y = axn/(1+bxn) (a = 8.2 × 10−4 ± 0.1 × 10−1; b = 7.9 × 10−2 ± 0.1 × 10−2; n = 2.2 ± 0.2).

The problem we confronted in this study was that 2,5-Me2THF was binding cooperatively with THF to promote monomer formation as illustrated in eq 6. At low THF concentration, an unusually elevated concentration of NaHMDS monomer in 2,5-Me2THF compared with pentane implicates a

(Me3Si)2N(THF)x(Me2THF)y mixed solvate. However, there is almost no chance that 2,2,5,5-Me4THF cooperatively solvates monomers, and DFT computations support this assertion. We still, however, observe the promotion of the monomer. Thus, the flattening of the curve is at least partially attributable to such an unproductive side equilibrium to form a mixed solvate according to the Principle of Detailed Balance.48 We must confess that the function used to fit the THF/pentane data (caption in Figure 10) contains a correction for inhibitory medium effects, 1/(1+cx2), that was derived empirically with no formal molecular basis. The challenge posed by these medium effects rears its ugly head below.

graphic file with name nihms-2015774-f0035.jpg (6)

With all that said, the rate law in eq 5 provides an adequate picture of a mechanism requiring 6 solvents and two silazide subunits—a hexasolvated dimer-based transition structure TS-5 (Figure 11) with a +Na(THF)6 gegenion (Figure 12).36

Figure 11.

Figure 11.

DFT-computed transition structure TS-5 for the NaHMDS-based addition to methyl-2-picolinate (3) solvated by THF. The implicit +Na(THF)6 cation7b (Figure 12) is not included.

Figure 12.

Figure 12.

DFT computed +Na(THF)6.

We computationally examined the overall reaction coordinate using a triple-ion-based framework (Scheme 6). The cation (Figure 12) is omitted for all structures in Scheme 6. Triple ions are well precedented,36 including for LiHMDS. Inspection of Figure 11, however, reveals that it is a triple-ion-based reference state (59) because the lowest-energy rate-limiting transition structure TS-5 has lost all semblance of the N–Na–N triple ion connectivity as has the IRC-derived minimum 61 preceding it.

There is plenty of room for alternative interpretation. One could, for example, imagine 61 stemming directly from a chelated monomer and ionized NaHMDS fragment. It is also possible, however, that the conversion of 60 to transition structure TS-5 proceeds via 61 in which the (Me3Si)2N in 61 remains “bound.” (The dissociation of 61 is calculated to be +5.5 kcal/mol.) Transition structure TS-5 gives way to tetrahedral adduct 62. The sequence follows an aza-Brook-like silicon transfer akin to that presumed to occur in an aza-Peterson-like imine formation49 via transition structure TS-6, elimination of 63 via transition structure TS-7 to form arene π complex 64, and arene dissociation to afford the observed imino ether 56 and unobserved (fleeting) mixed triple ion 65.36,50

Part 3: Mechanism of aminolysis of 2-pyridinecarbonitrile (6) with NaHMDS/DMEA.

Motivated as much by an obsessive need to complete the story as by curiosity, we examined the addition of NaHMDS to nitrile 6 to form amidinate 57 (Scheme 5) and were rewarded for our persistence. IR and NMR spectroscopy showed no evidence that nitrile 6 binds to NaHMDS at low or high DMEA concentrations. The addition occurs to give mixed amidinate 66 within seconds at 25 °C (eq 7) manifesting a silazide fragment that is well resolved from NaHMDS homodimer 39 displaying dimer-like coupling (1JN–Si = 8.5 Hz; Figure 13). As described above, further heating at 50 °C causes scrambling of the 14N–15N isotopes.

Figure 13.

Figure 13.

29Si NMR spectra of [15N2]NaHMDS dimer (39, 1JN–Si = 8.5 Hz) and mixed aggregate [15N1]66 (1JN–Si = 7.3 Hz, 3JN–Si = 2.0 Hz).

graphic file with name nihms-2015774-f0036.jpg (7)

Rate studies show an inverse-second-order dependence on DMEA (Figure 14) and first-order dependence on NaHMDS dimer 39 (Figure 15) consistent with a disolvated-dimer-based 1,2-addition. A full computational workup is found in the Supporting Information. An abbreviated version is illustrated in Scheme 7. For example, the conversion of 70 to 66 requires several additional mundane readjustments. Also, we examined the role of solvation of the intermediates and found minima that were higher energy than the unsolvated forms except for 70, whose Me3N-solvated form is −2.5 kcal/mol more stable.

Figure 14.

Figure 14.

A plot of kobsd vs [DMEA] (M) in pentane for the addition of NaHMDS to pyridinecarbonitrile (6, 0.005 M) at −40 °C measured with IR spectroscopy (1584 cm−1). The curve depicts an unweighted least-squares fit to y = axn/(1+bxn) (n = −1.9 ± 0.1; a = 2.7 × 10−2 ± 0.1 × 10−2; b = 7.1 × 10−3 ± 0.1 × 10−2).

Figure 15.

Figure 15.

Plot of kobsd vs [NaHMDS] (M) for the addition of NaHMDS to 2-pyridinecarbonitrile (6, 0.005 M) in 3.85 M (neat) DMEA at −40 °C measured with IR spectroscopy (1584 cm−1). The curve depicts an unweighted least-squares fit to y = axn (a = 2.1 × 10−2 ± 0.1 × 10−2; n = 1.1 ± 0.1).

Scheme 7.

Scheme 7.

DFT-computed reaction coordinate for the aminolysis of 2-pyridinecarbonitrile (6) by NaHMDS using Me3N as a DMEA surrogate.

Part 4. Mechanism of aminolysis of 2-pyridinecarbonitrile (6) with NaHMDS/THF.

IR and NMR spectroscopy showed no evidence that nitrile 6 binds to NaHMDS (eq 8). The only species observable by 1H and 29Si NMR spectroscopy was free NaHMDS (52 and 53, Scheme 5) and sodium amidinate 57 to the exclusion of any mixed aggregate. DFT computations suggest 57 is a disolvate.

graphic file with name nihms-2015774-f0037.jpg (8)

Rate studies showed first-order decays and concentration-independent values of kobsd, consistent with a first-order dependence on nitrile 6. A plot of kobsd vs THF concentration shows a distinct sigmoidal dependence affording a pronounced inhibition at high THF (Figure 16, curve A). Any doubt that this correlates with the NaHMDS dimer-monomer deaggregation (52 and 53, Scheme 5) was put to rest by superimposing the rate data on a plot of the equilibrium population of dimer (curve B) measured in 2020 by a different experimentalist.7b Clearly, deaggregation inhibits the reaction. A plot of kobsd vs NaHMDS in neat THF where NaHMDS is >98% monomer shows a clean second-order dependence (Figure 17), consistent with a requisite monomer-to-dimer aggregation before rate-limiting addition. The crudely defined zeroth-order THF dependence implicates the partial rate law in eq 9, pointing to the overall dimer-based transition structure, [A2S2(ArCN)] .51,52 The DFT computed transition structure TS-10 is illustrated in Figure 18.

Figure 16.

Figure 16.

A plot of kobsd vs [THF] (M) for the addition of NaHMDS to 2-pyridinecarbonitrile in pentane cosolvent (curve A, red) and 29Si chemical shift (curve B, green) plotted versus [THF] in 2:1 pentane/toluene as cosolvent measured at −20 °C. The latter function fits to a model based on an A2S2−AS4 equilibrium (Supporting Information).7b,52 The blue data (Curve C) derives from 2,5-Me2THF rather than pentane as cosolvent (vide infra).

Figure 17.

Figure 17.

Plot of kobsd vs [NaHMDS] (M) for the addition of NaHMDS to picolinonitrile (6, 0.005 M) in 12.8 M (neat) THF at −20 °C measured with IR spectroscopy (1584 cm−1). The curve depicts an unweighted least-squares fit to y = axn (a = 4.3 × 10−2 ± 0.1 × 10−2; n = 2.0 ± 0.1).

Figure 18.

Figure 18.

DFT computed transition structure TS-10 for the NaHMDS-dimer-based addition to nitrile 6.

d[6]/dt=k[6]1AS42 (9)

The DFT-computed reaction coordinate for addition to nitrile 6 in THF shows many parallels with that in DMEA that are relegated to the supporting information. The notable differences are the exothermic solvation of all key species including transition structure TS-10 in Figure 18, mixed dimer-based transition structure TS-9 for silicon transfer, and unobserved mixed aggregate 71 whose analog (66, eq 7) was fully characterized in DMEA.53

graphic file with name nihms-2015774-f0038.jpg

We have accumulated evidence of pre-aggregation-based reactions originating from monomers, but none have ever been so poignant.36 Why is the dimer-based addition to nitriles dominant? A partial answer is that the trajectory of the silazide attack on the nitrile π system in TS-10 (Figure 18) appears optimal. The most stable transition structure for monomer-based addition (TS-11, Figure 19) is less favorable by 4.4 kcal/mol using a monomer reference state, which appears to be a consequence of an inferior trajectory.

Figure 19.

Figure 19.

DFT computed transition structure TS-11 for the unfavorable NaHMDS-monomer-based addition to nitrile 6.

We would be remiss to not comment on the influence of 2,5-Me2THF rather than pentane as the cosolvent for nitrile addition (curve C in Figure 16). It provided further evidence of cooperative solvation. While checking for polarity effects we detected and subsequently documented cooperative monomer solvation, (Me3Si)2NNa(THF)x(Me2THF)y, noted in eq 6. The details are obscured in Figure 16 by an overlaying THF concentration dependence. The complete suppression of the rate suggested that the mixed solvated monomers persisted as an unproductive side equilibrium at stunningly low THF concentrations. Indeed, this was caused by a steep temperature dependence on the deaggregation. The message for us is that 2,5-Me2THF is a far better non-coordinating THF surrogate for lithium than for sodium.

Summary.

The studies described above proved to be four discrete mechanistic studies in which the choice of solvent and conditions markedly influence the outcome of the reactions. We offer an overview in which the results are grouped according to choice of solvent—THF or DMEA. This summary is organosodium centric; there is a considerable body organolsilicon chemistry (largely addressed computationally) that is left to the four parts described above. The arrows in Schemes 8 and 9 encompass many discrete steps including the organosilicon chemistry.

Scheme 8.

Scheme 8.

Summary of NaHMDS mechanistic studies in THF solution.

Scheme 9.

Scheme 9.

Summary of NaHMDS mechanistic studies in DMEA solution.

The THF results are summarized in Scheme 8. While NaHMDS in neat THF solution exists exclusively as tetrasolvated monomer 53, it reacts with methyl-2-picolinate, 3, via a dimer-based pathway in which a triple ion motif TS-5 is invoked at the rate-limiting transition structure. The critical chemoselectivity stems from a post-rate-limiting extrusion of NaOSiMe3 to form spectroscopically observable imino ether 56, which reacts further to give nitrile 6 and sodium amidinate 57 en route to a high-yielding one-pot synthesis of amidine 7. This ester-to-amidine conversion seems potentially important synthetically. Addition of NaHMDS to nitrile 6 is also dimer-based as depicted in computationally viable transition structure TS-10. The dominance of dimer-based reactivity under conditions affording monomeric NaHMDS should pique the interest of those interested in mechanistic alkali metal chemistry.

The chemistry of NaHMDS in poorly coordinating DMEA7b (Scheme 9) shows some parallels with the results in THF, but the differences are consequential. Although we do not consider a dimeric NaHMDS to be the proximate cause of dimer-based reactivity—many if not most monomer-based organoalkali metal reactions emanate from aggregates—the dominance of a dimer-based transition structure, TS-1, leading to observable mixed aggregate 43 aligns with previous studies of both LiHMDS and NaHMDS in DMEA.7c,54 In contrast to THF, however, a critical post-rate-limiting extrusion of MeOSiMe3 rather than NaOSiMe3 dictates the chemoselective formation of carboxamide 4. Observable mixed aggregate 50 is robust but finds a path to 66 over a month at 25 °C. Although this is clearly an inferior route to amidine 7 we examined the addition to nitrile 6 and, once again, uncovered dimer-based chemistry via an unsolvated transition structure, TS-8.

Conclusion

While casting about for case studies to investigate solvent-structure-reactivity relationships in NaHMDS we happened across several potentially useful reactions that are poorly represented in the literature and provided mechanistic insights that were wholly unexpected. The potential utility stems from C-N bond formations directly from aryl methyl esters, bypassing more activated forms. For us, however, they provided clean examples illustrating the role of solvation and aggregation in the chemistry of sodium amides in general and NaHMDS in particular. The NaHMDS itself serves as a preface to ongoing studies of sodium alkyl(trimethylsilyl) amides with potentially greater utility as strong bases and for C-N bond-forming reactions.

The dimer- and mixed dimer-based reactivity has precedent,7c,36 but the total dominance of aggregate-based reactivity still came as a surprise. From a mechanistically tactical perspective, the work underscores both the power of 29Si–15N coupling to determine solution structures and the general merits of 29Si NMR spectroscopy. (We asserted previously that 29Si NMR spectroscopy is underutilized by those outside the organosilicon community.) We offer a final caveat to those who might be tempted to probe primary- versus secondary-shell solvation by using 2,5-Me2THF as a polar, non-coordinating cosolvent. Despite its success in organolithium chemistry, the larger sodium ion may call for the more expensive 2,2,5,5-Me4THF. On the truism that an effect is either sterics or electronics, ours’ and others’ experience with metal ion solvation is that sterics dominates.55

Experimental

Reagents and solvents.

NaHMDS and [15N]NaHMDS were prepared as white crystalline solids.7b Toluene, hexanes, THF, MTBE, cyclopentane, 2,5-Me2THF, and HMPA were distilled from blue or purple solutions containing sodium benzophenone ketyl. All substrates and products in Table 1 are commercially available.31,32

General Procedure A: picolinamide 4.

Solid sodium hexamethyldisilazide (NaHMDS, 165 mg, 0.9 mmol) was dissolved in 2 mL of DMEA at 25 °C. 2 mL of the NaHMDS solution was added to a dry 5 mL Kimble vial equipped with a magnetic stir bar. Methyl picolinate (3, 36 μl, 0.30 mmol) was then added to the reaction solution. The reaction was stirred at 25 °C for 0.3 h. DI water (1 mL) was added, and the resulting biphasic mixture was partitioned between water (1 mL) and ethyl ether (2 mL). The aqueous layer was separated and extracted further with three 2 mL portions of ethyl ether. The combined organic layers were dried over anhydrous magnesium sulfate and then concentrated. Purification of the residue by flash column chromatography (80 % ethyl acetate in hexanes) afforded picolinamide 4 as a white solid. 1H NMR (500 MHz, CDCl3) δ 8.56 (d, J = 4.6 Hz, 1H), 8.20 (d, J = 7.8 Hz, 1H), 7.89 (s, 1H), 7.84 (td, J = 7.7, 1.8 Hz, 1H), 7.43 (dd, J = 7.6, 4.8 Hz, 1H), 6.23 (s, 1H). 13C{1H} NMR (126 MHz, CDCl3) δ 166.99, 145.59, 148.32, 137.31, 126.47, 122.44. HRMS (DART) Calc. for C6H6N2O (M+H+): 123.04746; found: 123.05493.

General Procedure B: amidine 7.

Solid sodium hexamethyldisilazide (NaHMDS, 165 mg, 0.9 mmol) was dissolved in 1.5 mL of THF at 25 °C. 1.5 mL of the NaHMDS solution was added to a dry 5 mL Kimble vial equipped with a magnetic stir bar. Methyl picolinate (3, 36 μl, 0.30 mmol) was then added to the reaction solution. The reaction was stirred at 50 °C for 0.3 h. DI water (1 mL) was then added. 2M HCl aqueous solution was added to adjust pH = 1, which then was stirred at 25 °C for 1 h. Saturated NaOH solution was added until pH = 12, and the mixture was partitioned between water (1 mL) and chloroform (4 mL). The aqueous layer was separated and extracted further with three 4 mL portions of chloroform. The combined organic layers were dried over anhydrous magnesium sulfate and then concentrated to afford amidine 7 as a yellow oil. 1H NMR (500 MHz, CDCl3) δ 8.56 (m, 1H), 8.11 (d, J = 7.9 Hz, 1H), 7.78 (td, J = 7.7, 1.8 Hz, 1H), 7.35 (dd, J = 7.5, 4.9 Hz, 1H), 5.94 (s, 3H). 13C{1H} NMR (126 MHz, CDCl3) δ 161.06, 150.79, 148.35, 137.03, 125.21, 120.81. HRMS (DART) Calc. for C6H7N3 (M+H+): 122.06345; found: 122.07097.

NMR spectroscopic analyses.

An NMR tube fitted with a double-septum under vacuum was flame-dried on a Schlenk line and allowed to passively cool to room temperature, backfilled with argon, and placed in a dry ice/acetone cooling bath. Individual stock solutions of the substrate and NaHMDS or [15N]NaHMDS were prepared at room temperature. The appropriate amounts of the substrate, NaHMDS, solvent, and co-solvent were added sequentially to the tube cooled to −78 °C via a gastight syringe. The tube was flame-sealed under a partial vacuum while cold to minimize evaporation in some cases and left unsealed for incremental additions. The tubes were mixed with a vortex mixer for approximately 10 secs to minimize warming. Standard 1H, 13C, and 29Si direct detection spectra were recorded at 500, 125.79, and 99.36 MHz, respectively, and referenced to Me4Si (0.0 ppm). Integration of the NMR signals was determined using the line-fitting method included in MNova (Mestrelab research S.L.).

Rate Studies.

IR spectra were recorded with an in situ IR spectrometer fitted with a 30-bounce, silicon-tipped probe. The spectra were acquired at a gain of 1 and a resolution of 4 cm−1. All tracked reactions were conducted under a positive flow of argon from a Schlenk line. A representative reaction was carried out as follows: The IR probe was inserted through a Teflon adapter and O-ring seal into an oven-dried, cylindrical flask fitted with a magnetic stir bar and a T-joint. The T-joint was capped with a septum for injections and an argon line. After evacuation under full vacuum, heating, and flushing with argon, the flask was charged with the solvent mixtures of our choice (toluene, DMEA, THF, 2,5-Me2THF, and 2,2,5,5-Me4THF) and cooled to −78 °C in a dry ice−acetone bath. A set of 256 baseline scans were collected and IR spectra were recorded every 15 seconds from 30 scans. The reaction vessel was charged stock solutions of NaHMDS and additional cosolvent through the septum. One set of scans was collected before the addition of substrate through the septum, and the baseline is zeroed. The substrate was added neat or in highly concentrated solutions and was tracked to completion (1731 cm−1 for 7, for example). The spectrometer was configured to collect spectra every 5 seconds from 16 scans. The reaction was tracked over 3–5 half-lives monitoring the disappearance of the starting material and the appearance of the product. The former was used for the rate studies.

Density functional theory (DFT) computations.

All DFT calculations were carried out using Gaussian 16.29 Prompted by a recent benchmarking of modern density functionals, all calculations were conducted at the M06–2X level of theory.30ac A pruned (99, 590) integration grid (equivalent to Gaussian’s “UltraFine” option) was used for all calculations. The Ahlrichs basis set def2-svp was used for geometry optimizations and the expanded def2-tsvp basis set for single-point energy calculations.30d Ball-and-stick models were rendered using CYLview 1.0b.30e A large number of DFT-computed energies are archived in the Supporting Information.

Supplementary Material

SI

Acknowledgments.

We thank the National Institutes of Health (GM131713) for support.

Footnotes

The authors declare no competing financial interests.

Supporting Information: The Supporting Information is available free of charge at http://pubs.acs.org. Synthetic and experimental procedures, 1H and 29Si NMR spectroscopic, rate, and computational data.

References and Footnotes

  • 1.(a) Mulvey RE; Robertson SD Synthetically Important Alkali-Metal Utility Amides: Lithium, Sodium, and Potassium Hexamethyldisilazides, Diisopropylamides, and Tetramethylpiperidides. Angew. Chem., Int. Ed. 2013, 52, 11470. [DOI] [PubMed] [Google Scholar]; (b) Robertson SD; Uzelac M; Mulvey RE Alkali-Metal-Mediated Synergistic Effects in Polar Main Group Organometallic Chemistry. Chem. Rev. 2019, 119, 8332. [DOI] [PubMed] [Google Scholar]; (c) McLellan R; Uzelac M; Bole LJ; Gil-Negrete JM; Armstrong DR; Kennedy AR; Mulvey RE; Hevia E Alkali Metal Effects in Trans-​Metal-​Trapping (TMT)​: Comparing LiTMP with NaTMP in Cooperative MTMP​/Ga(CH2SiMe3)​3 Metalation Reactions Synthesis 2019, 51, 1207. [Google Scholar]; (d) Garden JA; Armstrong DR; Clegg W; Garcia-Alvarez J; Hevia E; Kennedy AR; Mulvey RE; Robertson SD; Russo L Donor-​Activated Lithiation and Sodiation of Trifluoromethylbenzene: Structural, Spectroscopic, and Theoretical Insights. Organometallics 2013, 32, 5481. [Google Scholar]; (e) Wong HC Is Sodium Finally Coming of Age? Nat. Catal. 2019. 2, 282. [Google Scholar]; (f) Lochmann L; Janata M 50 Years of Superbases Made from Organolithium Compounds and Heavier Alkali Metal Alkoxides. Eur. J. Chem. 2014, 12, 537. [Google Scholar]; (g) Seyferth D Alkyl and Aryl Derivatives of the Alkali Metals: Strong Bases and Reactive Nucleophiles. 2. Wilhelm Schlenk’s Organoalkali-Metal Chemistry. The Metal Displacement and the Transmetalation Reactions. Metalation of Weakly Acidic Hydrocarbons. Superbases. Organometallics 2009, 28, 2. [Google Scholar]; (h) The quest for Organo-Alkali Metal Monomers: Unscrambling the Structure–Reactivity Relationship. Davison N; Lu E Dalton Trans. 2023, 52, 8172. [DOI] [PubMed] [Google Scholar]
  • 2.Structure, Reactivity, and Synthetic Applications of Sodium Diisopropylamide. Woltornist RA; Ma Y Algera RF; Zhou Y; Zhang Z; Collum DB Synthesis 2020, 52, 1478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.(a) For a recent studies of illustrating the synthetic potential of organosodium chemistry, see: Gentner TX; Mulvey RE Alkali-Metal Mediation: Diversity of Applications in Main-Group Organometallic Chemistry. Angew. Chem., Int. Ed. 2021, 60, 9247. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Anderson DE; Tortajada A; Hevia E Highly Reactive Hydrocarbon Soluble Alkylsodium Reagents for Benzylic Aroylation of Toluenes Using Weinreb Amides. Angew. Chem., Int. Ed. 2023, 62, e202218498. [DOI] [PubMed] [Google Scholar]; (c) Harenberg JH; Reddy R; Reddy A; Karaghiosoff K; Knochel P Continuous Flow Preparation of Benzylic Sodium Organometallics. Angew. Chem., Int. Ed. 2022, 61, e202203807. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Davison N; McMullin CL; Zhang L; Hu S-X; Waddell PG; Wills C; Dixon C; Lu E Li vs Na: Divergent Reaction Patterns between Organolithium and Organosodium Complexes and Ligand-Catalyzed Ketone/Aldehyde Methylenation. J. Am. Chem. Soc. 2023, 145, 6562. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) De PB; Asako S; Ilies L Recent Advances in the Use of Sodium Dispersion for Organic Synthesis. Synthesis 2021, 53, 3180. [Google Scholar]
  • 4.Ma Y; Lui NM; Keresztes I; Woltornist RA; Collum DB Sodium Isopropyl(trimethylsilyl)amide (NaPTA): A Stable and Highly Soluble Lithium Diisopropylamide Mimic. J. Org. Chem. 2022, 87, 14223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.“Sodium hexamethyldisilazide.” Watson BT; Lebel H In e-EROS Encyclopedia of Reagents for Organic Synthesis; John Wiley & Sons, New York; 2005, p 1–10. [Google Scholar]
  • 6.(a) Driess M; Pritzkow H; Skipinski M; Winkler U Synthesis and Solid State Structures of Sterically Congested Sodium and Cesium Silyl(fluorosilyl)phosphanide Aggregates and Structural Characterization of the Trimeric Sodium Bis(trimethylsilyl)amide. Organometallics 1997, 16, 5108. [Google Scholar]; (b) Kennedy AR; Mulvey RE; O’Hara CT; Robertson SD; Robertson GM Catena-Poly[Sodium-μ2-(N,N,N′,N′- Tetramethylethane-1,2-Diamine)-κ2-N,N′-Sodium- Bis[μ2-Bis(Trimethylsilyl)Azanido-κ2 N:N]]. Acta Crystallogr. Sect. E Struct. Rep. Online 2012, 68, m1468. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Schü P; Görls H; Westerhausen M; Kriec S Bis(trimethylsilyl)amide Complexes of s-Block Metals with Bidentate Ether and Amine Ligands. Dalton Trans. 2019, 48, 8966. [DOI] [PubMed] [Google Scholar]; (d) Ojeda-Amador AI; Martínez-Martínez AJ; Kennedy AR; Armstrong DR; O’Hara CT Monodentate Coordination of the Normally Chelating Chiral Diamine (R,R)-TMCDA. Chem. Commun. 2017, 53, 324. [DOI] [PubMed] [Google Scholar]; (e) Sarazin Y; Coles SJ; Hughes DL; Hursthouse MB; Bochmann M Cationic Brønsted Acids for the Preparation of Sn(IV) Salts: Synthesis and Characterisation of [Ph3Sn(OEt2)][H2N{B(C6F5)3}2],[Sn(NMe2)3(HNMe2)2][B(C6F5)4] and [Me3Sn(HNMe2)2][B(C6F5)4]. Eur. J. Inorg. Chem. 2006, 2006, 3211. [Google Scholar]; (f) Karl M; Seybert G; Massa W; Harms K; Agarwal S; Maleika R; Stelter W; Greiner A; Neumüller WH; Dehnicke K Amidometallate von Seltenerdelementen. Synthese Und Kristallstrukturen von [Na(12-Krone-4)2][M{N(SiMe3)2}3(OSiMe3)] (M = Sm, Yb), [Na(THF)3Sm{N(SiMe3)2}3(C≡C-Ph)], [Na(THF)6][Lu2(μ-NH2)(μ-NSiMe3){N(SiMe3)2}4] Sowie von [NaN(SiMe3)2(THF)]2. Z. Anorg. Allg. Chem. 1999, 625, 1301. [Google Scholar]; (g) Neufeld R; Michel R; Herbst-Irmer R; Schöne R; Stalke D Introducing a Hydrogen-Bond Donor into a Weakly Nucleophilic Brønsted Base: Alkali Metal Hexamethyldisilazides (MHMDS, M=Li, Na, K, Rb, and Cs) with Ammonia. Chem. - Eur. J. 2016, 22, 12340. [DOI] [PubMed] [Google Scholar]; (h) Edelmann FT; Pauer F; Wedler M; Stalke D Preparation and Structural Characterization of Dioxane Coordinated Alkali Metal Bis(Trimethylsilyl)Amides. Inorg. Chem. 1992, 31, 4143. [Google Scholar]; (i) Also, see reference 10a. [Google Scholar]
  • 7.(a) Sodium Hexamethyldisilazide: Using 15N–29Si Scalar Coupling to Determine Aggregation and Solvation States. Woltornist RA; Collum DB J. Am. Chem. Soc. 2020, 142, 6852. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Aggregation and Solvation of Sodium Hexamethyldisilazide: Across the Solvent Spectrum. Woltornist RA; Collum DB J. Org. Chem. 2021, 86, 2406. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Ketone Enolization with Sodium Hexamethyldisilazide: Solvent- and Substrate-Dependent E–Z Selectivity and Affiliated Mechanisms. Woltornist RA and Collum DB J. Am. Chem. Soc. 2021, 143, 17452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.(a) Brown DG Boström J Analysis of Past and Present Synthetic Methodologies on Medicinal Chemistry: Where Have All the New Reactions Gone? J. Med. Chem. 2016, 59, 4443. [DOI] [PubMed] [Google Scholar]; (b) Guan Y; Lee T; Wang K; McWilliams CJ SNAr Regioselectivity Predictions: Machine Learning Triggering DFT Reaction Modeling through Statistical Threshold. J. Chem. Inf. Model. 2023, 63, ASAP. [DOI] [PubMed] [Google Scholar]
  • 9.(a) Harrington A; Tal-Gan Y The Importance of Amide Protons in Peptide Drug Development. Future Med. Chem. 2019, 11, 2759. [DOI] [PubMed] [Google Scholar]; (c) Kumari S; Carmona AV; Tiwari AK; Trippier PC Amide Bond Bioisosteres: Strategies, Synthesis, and Successes. J. Med. Chem. 2020, 63, 12290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.(a) For a NaHMDS-sodium enolate mixed aggregates, see: Williard PG; Hintze MJ Mixed Aggregates: Crystal Structures of a Lithium Ketone Enolate/Lithium Amide and of a Sodium Ester Enolate/Sodium Amide. J. Am. Chem. Soc. 1990, 112, 8602. [Google Scholar]; (b) Zhang Z; Collum DB Structures and Reactivities of Sodiated Evans Enolates: Role of Solvation and Mixed Aggregation on the Stereochemistry and Mechanism of Alkylations. J. Am. Chem. Soc. 2019, 141, 388. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Ojeda-Amador AI; Martinez-Martinez AJ; Kennedy AR; O’Hara CT Synthetic and Structural Studies of Mixed Sodium bis(Trimethylsilyl) Amide/Sodium Halide Aggregates in the Presence of η2-N,N-, η3-N,N,N/N,O,N-, and η4-N,N,N,N-Donor Ligands. Inorg. Chem. 2015, 54, 9833. [DOI] [PubMed] [Google Scholar]; (d) Knapp C; Lork E; Borrmann T; Stohrer W-D; Mews R Versuche Zur Darstellung Vont-BuCN5S3 Und Die Unerwartete Isolierung Einer Kovalenten Modifikation von Tetraschwefelpentastickstoff-Chlorid S4N5Cl. Z. Anorg. Allg. Chem. 2005, 631, 1885. [Google Scholar]; (e) Clark NM; García-Álvarez P; Kennedy AR; O’Hara CT; Robertson GM Reactions of (−)-Sparteine with Alkali Metal HMDS Complexes: Conventional Meets the Unconventional. Chem. Commun. 2009, 39, 5835. [DOI] [PubMed] [Google Scholar]; (f) Williard PG; Nichols MA Structural Characterization of Mixed Alkali Metal Bis(trimethylsilyl) Amide Bases. J. Am. Chem. Soc. 1991, 113, 9671. [Google Scholar]
  • 11.. We suspect that our failures and other failures to undergo 1,4-addition derives from a thermodynamic problem in which the reaction is endothermic. Products of cinnamate self-condensation of the tert-butylcinnamate dominated reaction contents. Ongoing studies of enolization with NaHMDS/THF-based enolizations suggest a delicately balanced equilibrium for some substrates (unpublished). 1,4-Addition to the malonate analog [PhCH=C(CO2Me)2, by contrast, are successful (unpublished). A 1,4-addition of LiHMDS to an unsaturated ester has been reported.12 [Google Scholar]
  • 12.Rico JG Synthesis of Novel β-Amino Acid Precursors: β-Amino-Hydrocoumarins as Unusual Aspartic Acid Mimetics Used in Fibrinogen Receptor Antagonists. Tetrahedron Lett. 1994, 35, 6599. [Google Scholar]
  • 13.(a) Krüger C; Rochow EG; Wannagat U Uber die Reaktion von Natrium-bis-trimethylsilyl-amid mit Derivaten der Benzoesaure. Chem. Ber. 1963, 96, 2138. [Google Scholar]; (b) Afifi MS; El-Sayed BA Spectroscopic Investigation of the Reaction Products Between the Sodium Salt of Hexamethyldisilazane and Ethyl Salicylate. Mass Spectra, Proton Magnetic Resonance and Infrared Analysis. Egypt. J. Chem. 1983, 26, 551. [Google Scholar]
  • 14.Hwu JR; Hsu CH; Wong FF; Chung C-S; Hakimelahi GH Sodium Bis(trimethylsilyl)amide in the “One-Flask” Transformation of Aromatic Esters to Nitriles. Synthesis, 1998, 329. [Google Scholar]
  • 15.(a) Abou-Elkhair RAI; Hassan AEA; Boykin DW; Wilson WD Lithium Hexamethyldisilazane Transformation of Transiently Protected 4-Aza/benzimidazole Nitriles to Amidines and Their Dimethyl Sulfoxide-Mediated Imidazole Ring Formation. Org. Lett. 2016, 18, 4714. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Alharbi N; Díaz-Moscoso A; Tizzard GJ; Coles SJ; Cook MJ; Cammidge AN Improved Syntheses of meso-Aryl Tetrabenzotriazaporphyrins (TBTAPs). Tetrahedron, 2014, 70, 7370. [Google Scholar]
  • 16.(a) Baumann M; Baxendale IR Bioorg. Med. Chem. 2017, 25, 6218. [DOI] [PubMed] [Google Scholar]; (b) Dorsey BD; Dugan BJ; Fowler KM; Hudkins RL; Mesaros EF; Monck NJ; Morris EL; Olowoye I; Ott GR; Pave GA; Roffey JRA; Soudy CN; Zificsak CA; Zulli AL Preparation of Azaquinazoline Inhibitors of Atypical Protein Kinase C. WO2015148597 October 1, 2015. [Google Scholar]
  • 17.Chen J; Xu Z; Wang T; Lyssikatos JP; Ndubaku CO A Versatile Annulation Route to Primary-Amino-Substituted Naphthyridine Esters. SynLett 2014, 25, 89. [Google Scholar]
  • 18.Guedira NE; Beugelmans R Ambident Behavior of Ketone Enolate Anions in SNAr Substitutions on Fluorobenzonitrile Substrates. J. Org. Chem. 1992, 57, 5577. [Google Scholar]
  • 19.Sun K; Sagisaka K; Peng L; Watanabe H,; Xu, F.; Pawlak, R.; Meyer, E.; Okuda, Y.; Orita, A.; Kawai, S. Head-to-Tail Oligomerization by Silylene-Tethered Sonogashira Coupling on Ag(111). Angew. Chem., Int. Ed. 2021, 60, 19598. [DOI] [PubMed] [Google Scholar]
  • 20.Li G; Ji C; Hong X; Szostak M Highly Chemoselective, Transition-Metal-Free Transamidation of Unactivated Amides and Direct Amidation of Alkyl Esters by N–C/O–C Cleavage J. Am. Chem. Soc. 2019, 141, 11161 and references cited therein. [DOI] [PubMed] [Google Scholar]
  • 21.Hwu JR; Chuang K-S; Chuang SH; Tsay S-C 1,2-Eliminations in a Novel Reductive Coupling of Nitroarenes to Give Azoxy Arenes by Sodium Bis(trimethylsilyl)amide. Org. Lett. 2005, 7, 3211. [DOI] [PubMed] [Google Scholar]
  • 22.For a discussion and leading references to the challenges of metalating nitroarenes, see: Nagaki A; Kim H; Yoshida J.-i. Nitro-Substituted Aryl Lithium Compounds in Microreactor Synthesis: Switch Between Kinetic and Thermodynamic Control. Angew. Chem., Int. Ed. 2009, 48, 8063. [DOI] [PubMed] [Google Scholar]
  • 23.For a similar trimethylsilyl transfer that occurs when benzyne is trapped by metal silazides, see: Ikawa T; Masuda S; Akai S One-Pot Generation of Benzynes from Phenols: Formation of Primary Anilines by the Deoxyamination of Phenols. Chem. Eur. J. 2020, 26, 4320. [DOI] [PubMed] [Google Scholar]
  • 24.Schnürch M; Spina M; Khan AF; Mihovilovic MD; Stanetty P Halogen Dance Reactions—A Review. Chem. Soc. Rev. 2007, 36, 1046. [DOI] [PubMed] [Google Scholar]
  • 25.Estel L Marsais F; Queguiner G Metalation/SRN1 Coupling in Heterocyclic Synthesis. A Convenient Methodology for Ring Functionalization. J. Org. Chem. 1988, 53, 2740. [Google Scholar]
  • 26.Kadambar VK; Bachu S; Reddy MR; Torlikonda V; Manjunatha SG; Ramasubramanian S; Nambiar S; Howell GB Withnall, J.; Murugan, A. Regioselective Synthesis of 1,2-Aminoalcohols from Epoxides and Chlorohydrins. Tetrahedron Lett. 2012, 53, 5739. [Google Scholar]
  • 27.(a) Wang J; Rosingana M; Discordia RP; Soundararajan N; Polniaszek R Aminolysis of Esters or Lactones Promoted by NaHMDS - A General and Efficient Method for the Preparation of N-Aryl Amides. Synlett 2001, 12, 1485. [Google Scholar]; (b) Jones CD; Bunyard P; Pitt G; Byrne L; Pesnot T; Guisot N Heterocyclylamino-substituted Triazoles as Modulators of rho-Associated Protein Kinase. WO2019145729, August 1, 2019. [Google Scholar]
  • 28.(a) Several general-purpose reviews on determining reaction mechanism: Meek SJ; Pitman CL; Miller AJM Deducing Reaction Mechanism: A Guide for Students, Researchers, and Instructors. J. Chem. Educ. 2016, 93, 275. [Google Scholar]; (b) Simmons EM; Hartwig JF On the Interpretation of Deuterium Kinetic Isotope Effects in C–H Bond Functionalizations by Transition Metal Complexes. Angew. Chem., Int. Ed. 2012, 51, 3066. [DOI] [PubMed] [Google Scholar]; (c) Collum DB; McNeil AJ; Ramírez A Lithium Diisopropylamide: Solution Kinetics and Implications for Organic Synthesis. Angew. Chem., Int. Ed. 2007, 46, 3002. [DOI] [PubMed] [Google Scholar]; (d) Algera RF; Gupta L; Hoepker AC; Liang J; Ma Y; Singh KJ; Collum DB Lithium Diisopropylamide: Non-Equilibrium Kinetics and Lessons Learned about Rate Limitation. J. Org. Chem. 2017, 82, 4513. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.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]
  • 30.(a) Mardirossian N; Head-Gordon M Thirty Years of Density Functional Theory in Computational Chemistry: An Overview and Extensive Assessment of 200 Density Functionals. Mol. Phys. 2017, 115, 2315. [Google Scholar]; (b) Wang Y; Verma P; Jin X; Truhlar DG; He X Revised M06 Density Functional for Main-Group and Transition-Metal Chemistry. Proc. Nat. Acad. Sci. 2018, 115, 10257. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Zhao Y; Truhlar DG The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Function. Theor. Chem. Acc. 2008, 120, 215. [Google Scholar]; (d) Weigend F; Ahlrichs R Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297. [DOI] [PubMed] [Google Scholar]; (e) CYLview, 1.0b; C. Y. Legault, Université de Sherbrooke; 2009 (http://www.cylview.org; accessed 2023-7-6). [Google Scholar]
  • 31.Chen J; Xia Y; Lee S Transamidation for the Synthesis of Primary Amides at Room Temperature. Org. Lett. 2020, 22, 3504. [DOI] [PubMed] [Google Scholar]
  • 32.Jones AC; Williams MTJ; Morrill LC; Browne DL Mechanical Activation of Zero-Valent Metal Reductants for Nickel-Catalyzed Cross-Electrophile Coupling. ACS Catal. 2022, 12, 13681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Amidine 31 (Scheme 5) can be isolated after a month at 25 °C.
  • 34.The computations use the Gaussian standard state of 1.0 atm. If the THF concentration is corrected to neat THF (approximately 13 M), each solvation step benefits from approximately 2.0 kcal/mol of additional stabilization at −78 °C (195 K). Pratt LM; Merry S; Nguyen SC; Quan P; Thanh BT A Computational Study of Halomethyllithium Carbenoid Mixed Aggregates with Lithium Halides and Lithium Methoxide. Tetrahedron 2006, 62, 10821. [Google Scholar]
  • 35.Rein AJ; Donahue SM; Pavlosky MA In Situ FTIR Reaction Analysis of Pharmaceutical-Related Chemistry and Processes. Curr. Opin. Drug Discov. Dev. 2000, 3, 734. [PubMed] [Google Scholar]
  • 36.For leading references to open-dimer-based mechanisms, mechanisms that appear to involve monomer-to-dimer pre-aggregation, triple ion-based mechanisms, and solvation of sodium cation, see ref 7c and references cited therein.
  • 37.Intrinsic reaction coordinate (IRC) calculations are defined as “the minimum energy reaction pathway (MERP) in mass-weighted cartesian coordinates between the transition state of a reaction and its reactants and products.” They show the minima preceding and following transition state. [Google Scholar]
  • 38.For crystal structure of a lithium amidinates akin to 30 characterized crystallographically and excellent references to the applications of such amidinates, see: N,N′-Bis-Silylated Lithium Aryl Amidinates: Synthesis, Characterization, and the Gradual Transition of Coordination Mode from σ Toward π Originated by Crystal Packing Interactions. Aharonovich S; Kapon M; Botoshanski M; Eisen MS Organometallics 2008, 27, 1869 [Google Scholar]
  • 39.Verstuyft AW; Nelson JH; Cary LW Utility of Virtual Coupling in the Carbon-13 {proton} Nuclear Magnetic Resonance Spectra of Bis-phosphite Complexes of Palladium and Platinum. Algebraic Cancellation of Spin-Spin Coupling Inorg.Chem. 1976, 15, 732. [Google Scholar]
  • 40.Reich HJ; Dykstra RR Solution Structure of Lithium Benzeneselenolate and Lithium Diphenylphosphide: NMR Identification of Cyclic Dimers and Mixed Dimers. Organometallics 1994, 13, 4578. [Google Scholar]
  • 41.Perrin CL; Shrinidhi A; Burke KD Isotopic-Perturbation NMR Study of Hydrogen-Bond Symmetry in Solution: Temperature Dependence and Comparison of OHO and ODO Hydrogen Bonds. J. Am. Chem. Soc. 2019, 141, 17278. [DOI] [PubMed] [Google Scholar]
  • 42.Edwards JO; Greene EF; Ross J From Stoichiometry and Rate Law to Mechanism. J. Chem. Educ. 1968, 45, 381. [Google Scholar]
  • 43.Ashby EC; Dobbs FR; Hopkins HP Jr. Composition of Complex Aluminum Hydrides and Borohydrides, as Inferred from Conductance, Molecular Association, and Spectroscopic Studies. J. Am. Chem. Soc. 1973, 95, 2823. [Google Scholar]
  • 44.Me4THF binds to NaHMDS, CsHMDS, and RbHMDS at least reluctantly as evidenced by x-ray crystal structures of the disolvated dimers showing significant distortions when compared with the THF solvates. LiHMDS crystallizes from Me4THF as a solvent-free tetramer, although there is evidence it binds weakly in solution.45 Krieck, S.; Schüler, P.; Görls, H.; Westerhausen, M. Straightforward Synthesis of Rubidium Bis(trimethylsilyl)amide and Complexes of the Alkali Metal Bis(trimethylsilyl)amides with Weakly Coordinating 2,2,5,5- Tetramethyltetrahydrofuran. Dalton Trans. 2018, 47, 12562. [DOI] [PubMed] [Google Scholar]
  • 45.Lucht BL; Collum DB Ethereal Solvation of Lithium Hexamethyldisilazide (LiHMDS): Unexpected Relationships of Solvation Number, Solvation Energy, and Aggregation State. J. Am. Chem. Soc. 1995, 117, 9863. [Google Scholar]
  • 46.(a) Bates RB; Kroposki LM; Potter DE Cycloreversions of Anions from Tetrahydrofurans. A Convenient Synthesis of Lithium Enolates of Aldehydes. J. Org. Chem. 1972, 37, 560. [Google Scholar]; (b) Byrne F; Forier B; Bossaert G; Hoebers C; Farmer TJ; Clark JH; Hunt AJ 2,2,5,5-Tetramethyltetrahydrofuran (TMTHF): a Non-polar, Non-peroxide Forming Ether Replacement for Hazardous Hydrocarbon Solvents. Green Chem. 2017, 19, 3671. [Google Scholar]
  • 47.(a) The dielectric constants of substituted tetrahydrofurans are slightly lower than THF. Harada Y; Salomon M; Petrucci S Molecular Dynamics and Ionic Associations of Lithium Hexafluoroarsenate (LiAsF6) in 4-Butyrolactone Mixtures with 2-Methyltetrahydrofuran. J. Phys. Chem. 1985, 89, 2006. [Google Scholar]; (b) Carvajal C; Tolle KJ; Smid J; Szwarc M Studies of Solvation Phenomena of Ions and Ion Pairs in Dimethoxyethane and Tetrahydrofuran. J. Am. Chem. Soc. 1965, 87, 5548. [Google Scholar]
  • 48.The Principle of Detailed Balance asserts that individual equilibria within an ensemble of equilibria are maintained. It is particularly useful in understanding the complex equilibria observed in organolithium chemistry. Alberty RA Principle of Detailed Balance in Kinetics. J. Chem. Educ. 2004, 81, 1206. [Google Scholar]
  • 49.Linder T; Sutherland TC; Baumgartner T; Extended 2,5-Diazaphosphole Oxides: Promising Electron-Acceptor Building Blocks for π-Conjugated Organic Materials. Chem.–Eur. J. 2010, 16, 7101. [DOI] [PubMed] [Google Scholar]
  • 50.Substitution of the π-bound substrate in 43 with benzene is endothermic by 6.1 kcal/mol.
  • 51.The mathematics is in the Supporting Information of ref 7b.
  • 52.The order in THF cannot be determined from a fit with values of −4 through −6 all providing adequate fits to the data. The presumed value of −6 comes from the assigned solvation states of NaHMDS monomer as tetrasolvated,7b and transition structure 50 as a disolvate.
  • 53.A monomer-based N-to-N’ silicon transfer analogous to that depicted in Figure 18 is substantially less favorable.
  • 54.(a) Zhao P; Collum DB Ketone Enolization by Lithium Hexamethyldisilazide: Structural and Rate Studies of the Accelerating Effects of Trialkylamines. J. Am. Chem. Soc. 2003, 125, 14411. [DOI] [PubMed] [Google Scholar]; (b) Mack KA; McClory A; Zhang H; Gosselin F; Collum DB Lithium Hexamethyldisilazide-Mediated Enolization of Highly Substituted Aryl Ketones: Structural and Mechanistic Basis of the E/Z Selectivities. J. Am. Chem. Soc. 2017, 139, 12182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Structures of Lithium N-Monosubstituted Anilides: Trisolvated Monomer to Tetrasolvated Dimer. Su C; Guang J; Williard PG J. Org. Chem. 2014, 79, 1032. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

SI

RESOURCES