Abstract
Bicyclic triazolium scaffolds are widely employed in N-heterocyclic carbene (NHC) organocatalysis. While the incorporation of a fused ring was initially for synthetic utility in accessing chiral, modular triazolyl scaffolds, recent results highlight the potential for impact upon reaction outcome with the underpinning origins unclear. The common first step to all triazolium-catalyzed transformations is C(3)-H deprotonation to form the triazolylidene NHC. Herein, we report an analysis of the impact of size of the fused (5-, 6-, and 7-membered, n = 1, 2, and 3, respectively) ring on the C(3) proton transfer reactions of a series of bicyclic triazolium salts. Rate constants for the deuteroxide-catalyzed C(3)-H/D-exchange of triazolium salts, kDO, were significantly influenced by the size of the adjacent fused ring, with the kinetic acidity trend, or protofugalities, following the order kDO (n = 1) > kDO (n = 2) ≈ kDO (n = 3). Detailed analyses of X-ray diffraction (XRD) data for 20 triazolium salts (including 16 new structures) and of computational data for the corresponding triazolylidene NHCs provide insight on structural effects of alteration of fused ring size. In particular, changes in internal triazolyl NCN angle and positioning of the most proximal CH2 with variation in fused ring size are proposed to influence the experimental protofugality order.
Introduction
Since Ukai first utilized naturally occurring thiazolium salt 1, or vitamin B1, to catalyze the benzoin reaction of aldehydes,1 and later seminal studies by Breslow in establishing the catalytic mechanism,2 a broad range of related transformations catalyzed by different heterocyclic azolium salts continue to be investigated.3,4 As organocatalysts, azolium salts have been used to catalyze a diverse spread of reactions, including the benzoin and related acyloin reactions,2,5 Stetter reaction,6 cycloadditions,7 dearomatizations,8 among many others. Although a range of heterocyclic azolium classes have been employed in organocatalysis, triazolium precatalysts 2 are most often utilized9 (Figure 1). Typically, the reactions are initiated by a deprotonation step, which involves in situ conversion of the azolium salt 2 to the reactive N-heterocyclic carbene (NHC) 3/ylide 3′. In particular, chiral bicyclic triazolylidene scaffolds, initially developed by Knight and Leeper,10 have been widely investigated, with the development of modular syntheses allowing improved stereoselectivities.3a A range of fused ring triazolium salt precatalysts have been employed, with those derived from pyrrolidinone,11 morpholinone,12 or oxazolidinone13 scaffolds (4–6, respectively) common.
Figure 1.
(a) Structure of thiamin 1 (vitamin B1); (b) C-(3) deprotonation of triazolium precatalyst 2; (c) pyrrolidinone-, morpholinone-, and oxazolidinone-derived triazolium salts 4–6.
Recent investigations have recognized that the role of the fused ring is not solely limited to providing a structural scaffold for chiral modification toward asymmetric catalysis. Gravel and co-workers reported that the chemoselectivity of the benzoin condensation could be influenced by a simple change in both the size and composition of the fused ring for bicyclic triazolium salts 7k, 8k, and 9k (Figure 2a).14 Piperidinone-derived triazolium salt 8k (n = 2) with a N-pentafluorophenyl substituent was shown to catalyze highly chemoselective cross-benzoin reactions between aliphatic and aromatic aldehydes. Although still yielding the same major product, analogous pyrrolidinone- and ε-caprolactam-derived triazolium salts, 7k (n = 1) and 9k (n = 3), showed significantly reduced chemoselectivities under these reaction conditions. This remains one of the few examples of a chemoselective cross-benzoin reaction that is catalyst- rather than substrate-controlled. Notably, selectivity and yield were shown to be invariant to either base or solvent choice. In a follow-on computational study supported by 1H NMR and crossover experiments, Gravel, Legault, and co-workers explored the origin of the chemoselectivity differences for 7k, 8k, and 9k, although this study was not extended to triazolium salts with additional N-aryl substituents.14b The superior selectivity observed for the piperidinone-derived catalyst 8k was attributed to the kinetic control of reaction conditions. Furthermore, it was suggested that the steric interactions of proximal methylenes on the catalyst backbone with the aldehyde could play a significant role in obtaining chemoselectivity.
Figure 2.
(a) Previous work: demonstration by Gravel et al. of catalyst control of product selectivity in the cross-benzoin reaction by variation of triazolium fused ring size. (b) This work.
Despite the enormous current interest in NHC-organocatalysis, an understanding of chemoselectivity (as represented by the cross-benzoin reaction) remains one of the key unsolved challenges, and so the influence of the fused ring is potentially of importance across many transformations. As an initial deprotonation step is common to the majority of these transformations, we herein report an analysis of the impact of fused ring size on the C(3) proton transfer reactions of a series of triazolium salts (Figure 2b). Following on from our earlier studies,15 we undertook a hydrogen–deuterium exchange study to evaluate the effect of ring size on the kinetic acidities, or protofugalities,16 of the C(3)-hydrogens in a series of bicyclic triazolium ions. Experimental second-order rate constants for deuteroxide-catalyzed H/D-exchange, kDO (M–1 s–1), could be used in Hammett structure–activity analysis in addition to providing estimates of C(3)-H pKa values. We have performed X-ray diffraction and computational studies to afford additional insight. Despite the substantive literature focused on triazolium catalysis, there is no report to our knowledge comparing structural data for a large series of triazolium salts. We have analyzed experimental X-ray diffraction data for 20 triazolium salts (Figure 2b), which includes 16 previously unreported X-ray crystal structures. In addition, a DFT computational evaluation of the structures of these 20 triazolium ions and corresponding triazolylidenes has been performed. These data provide new insights into the role of fused ring size on the initial proton transfer step common to all triazolium-catalyzed transformations.
Results and Discussion
Hydrogen–Deuterium Exchange Kinetic Studies
Hydrogen–deuterium or hydrogen–tritium exchange has long been used as a method to estimate the kinetic lability of protons attached to carbon17 and to provide insight into factors influencing carbon acidity. For catalytic applications, the knowledge of rates of proton transfer between active species, and the corresponding activation barriers, is equally as important as pKa. In early studies, several groups reported rate constants for the base-catalyzed exchange of the C(2)-H of thiamin 1 providing initial evidence for the role of the conjugate base NHC/ylide as the active form of catalyst.17c−17f,18 The later isolation and structural characterization of a range of NHCs by Bertrand and Arduengo,19 including thiazolylidenes20 and triazolylidenes,21 provided further support for their relatively long lifetimes versus many other carbene classes. We have previously reported H/D-exchange kinetic studies for a range of conjugate acids of NHCs/ylides.15a−15d This included rate-pD profiles for the C(3)-H/D exchange of pyrrolidinone-derived triazolium salts 7a–c,i,k with 5-membered fused rings.15a−15c
In order to analyze the effect of ring size on proton transfer, analogous H/D-exchange kinetic studies have been performed for piperidinone-derived triazolium salts 8a–c,k (with 6-membered fused rings) and for ε-caprolactam-derived triazolium salts 9a–c,k (with 7-membered fused rings). To ensure consistency and accuracy, hydrogen–deuterium exchange experiments for 7a–c,k are repeated herein and showed excellent agreement with our earlier data.22 The kinetic methods utilized (section S1.4, Supporting Information) were identical to those employed in our previous work. First-order rate constants for H/D-exchange were determined in aqueous solution at a range of pD values with ionic strength I = 1.0 (KCl) at 25 °C and with tetramethylammonium deuteriosulfate as internal standard. Using 1H NMR spectroscopy, the decrease in the area of the C(3)-proton of each substrate could be quantified over time. Figures S1–S12 include representative spectral overlays of reaction progress for 7a–c,k, 8a–c,k, and 9a–c,k. The reactions were analyzed in the pD range of 0.5–3.5, because outside of this range the exchange reactions were too fast (pD > 3.5) or slow (pD < 0.5) for NMR kinetic analysis. During these experiments there was no decrease of other NMR resonances or appearance of additional signals, confirming that no side reactions, such as hydrolysis or decomposition of triazolium salt substrates, were occurring under these conditions. The observed pseudo first-order rate constants for deuterium exchange of the C(3)-proton, kex (s–1), were obtained from the slopes of semilogarithmic plots (Figures S13–S24) of reaction progress against time according to eq 1, where f(s) is the value of the fraction of unexchanged substrate. Tables S1–S12 summarize all first-order rate constants, kex (s–1), as a function of pD, and Figures S25–S28 include the corresponding log kex–pD profiles. No corrections were applied for data acquired in buffer solution, because in all previous studies no evidence of buffer catalysis was observed.
| 1 |
| 2 |
![]() |
3 |
Figure 3 compares the log kex–pD profiles for the deuterium exchange reactions of 7c, 8c, and 9c with those for 7k, 8k, and 9k. For N-4-fluorophenyl triazolium salts 7c, 8c, and 9c, good linear fits of log kex–pD data to eq 2 were observed; in eq 2, kDO (M–1 s–1) is the second-order rate constant for deuteroxide-catalyzed exchange, Kw the ion product of D2O at 25 °C,23 and γDO the activity coefficient for deuteroxide ion under our experimental conditions. Data for all other triazolium salts studied herein also showed good linear fits to eq 2 (Figures S25–S27) with the exception of 7k, 8k, and 9k. This increase in log kex with pD, and first-order dependence on deuteroxide ion, is consistent with a single mechanism for deuteroxide-catalyzed deuterium exchange as shown in Scheme 1. In this mechanism, deprotonation of the triazolium salts 7–9 by deuteroxide results in the formation of a complex between NHCs 7′–9′ and a molecule of HOD (in Scheme 1, the ylidic resonance structures of NHCs 7′–9′ have been excluded for clarity). Subsequent reorganization of 7′–9′·HOD to 7′–9′·DOL (L = H or D) to allow for delivery of deuterium, followed by deuteration, leads to exchange product 10.15a,15d,24 Owing to the large excess of bulk solvent D2O over substrate, the deuteration step is effectively irreversible; thus, kex reflects rate-limiting formation of solvent-equilibrated NHC from the triazolium salt and deuteroxide ion.
Figure 3.
Comparison of pD rate profiles of C(3)-H/D exchange for triazolium salts 7–9c and 7–9k in D2O solution at 25 °C.
Scheme 1. Mechanism for C(3)-H/D Exchange for Triazolium Salts 7–9.
As a result of a change from a simple first-order dependence on deuteroxide ion at lower pD values, and hence from a slope of unity, log kex–pD data for N-pentafluorophenyl salts 7k, 8k, and 9k instead show excellent fits to eq 3. The altered dependence of log kex on pD is consistent with the onset of alternative pathways for deuterium exchange, which we have discussed in detail previously.15a,15b,15f The most likely mechanistic explanation is a pathway via N(1)-deuteration at lower pD values allowing for hydrogen–deuterium exchange of the N(1)-deuterated dicationic triazolium salt (Scheme S2, section S1.4.4.4, Supporting Information). The additional terms in eq 3 (versus eq 2) include k′DO (M–1 s–1), the second-order rate constant for deuteroxide-catalyzed C(3)-H/D exchange of the dicationic salt, and KaN1, the equilibrium constant of N(1)-protonation. The extent of dominance of this alternative pathway at lower pDs is dependent on pKaN1. The similar overall behavior observed for 7k, 8k, and 9k suggests that the biggest influence on the onset of alternative H/D-exchange mechanisms is the N-aryl substituent rather than the size of the fused ring within the triazolium salt.
Table 1 summarizes values of the second-order rate constant for deuteroxide-catalyzed exchange, kDO (M –1s–1), for all triazolium salts studied herein. By definition, the second-order rate constant, kDO (M–1 s–1), is the observed kex value in 1 M DO– solution (pD ≈ 14). Experimentally, hydrogen–deuterium exchange for all triazolium ions is many orders of magnitude too fast to directly monitor in 1 M DO– (half-lives ≈ nanoseconds); thus, kDO values are obtained by assessment of a range of kex values at lower pD’s as described above. The comparison of reactivities toward deprotonation by a common base, kDO, allows for comparison of the kinetic acidities, or protofugalities,16 of the C(3)-hydrogens in the series of bicyclic triazolium ions. As DO– is the relevant conjugate base of water and of similar basicity to widely used alkoxide ions, and with an increasing general focus on organocatalysis in more sustainable solvent media such as water,25 it is an appropriate choice as a reference base.
Table 1. Summary of Second-Order Rate Constants for C(3)-H/D Exchange (kDO) for 7a–c,k (n = 1); 8a–c,k (n = 2); 9a–c, 9k (n = 3) and Calculated C(3)-H pKa Values.

| triazolium salt | n | R = | kDO (M–1 s–1)a | kDOreld | pKae |
|---|---|---|---|---|---|
| 7a | 1 | 4-OMe | 4.55 (±0.36) × 107 | 2.26 | 17.7 |
| 7a | 4.20 (±0.23) × 107b | 2.09 | 17.8 | ||
| 8a | 2 | 2.11 (±0.18) × 107 | 1.05 | 17.9 | |
| 9a | 3 | 2.01 (±0.20) × 107 | 1 | 18.0 | |
| 7b | 1 | H | 6.99 (±0.18) × 107 | 2.12 | 17.5 |
| 7b | 6.82 (±0.13) × 107b | 2.07 | 17.5 | ||
| 8b | 2 | 3.43 (±0.15) × 107 | 1.04 | 17.8 | |
| 9b | 3 | 3.29 (±0.15) × 107 | 1 | 17.8 | |
| 7c | 1 | 4-F | 8.97 (±0.27) × 107 | 2.04 | 17.4 |
| 7c | 8.66 (±0.11) × 107b | 1.97 | 17.4 | ||
| 8c | 2 | 5.00 (±0.31) × 107 | 1.14 | 17.7 | |
| 9c | 3 | 4.39 (±0.16) × 107 | 1 | 17.7 | |
| 7k | 1 | –F5 | 3.52 (±0.37) × 108c | 0.87 | 16.8 |
| 7k | 6.82 (±0.25) × 108b,c | 1.69 | 16.5 | ||
| 8k | 2 | 4.16 (±0.19) × 108 | 1.03 | 16.7 | |
| 9k | 3 | 4.04 (±0.19) × 108 | 1 | 16.7 |
This work.
Values of kDO (M–1 s–1) obtained previously.
The error quoted for 7k is that obtained based on the overall fit to eq 3 for all of the data. The larger variance from our previous value for 7k is due to the unavoidably small number of data points in the region of unit slope. The kDO value obtained previously is deemed more reliable owing to the lower overall error in fitting to eq 3 as a result of having a greater number of data points.
Calculated as kDOrel = kDOn/kDOn=3.
pKa values calculated using experimental kDO values (vide infra).
Obeying the common trend observed previously,15a,15c,15d electron-withdrawing N-aryl substituents increase kDO, regardless of the size of the fused ring, with a span of ∼26-fold across the series in Table 1. Thus, triazolium salts 7k, 8k, and 9k bearing the strongly electron-withdrawing N-pentafluorophenyl substituent have the highest kDO values, whereas analogues 7a, 8a, and 9a with the electron-donating N-4-methoxyphenyl substituent have the lowest protofugalities. Electron-withdrawing N-aryl substituents destabilize the cationic triazolium carbon acid 7–9 relative to the formally neutral NHC conjugate base 7′–9′ (Scheme 1), thus favoring the deprotonation process.
Table 1 also includes krel (= kDOn/kDOn=3) values, which allow for a comparison of the effect of the fused ring for a given N-aryl substituent. Notably, despite the distance from C(3)-H, the size of the fused ring clearly alters the protofugalities of the triazolium salts. In all cases, the 5-membered ring fused triazolium salts 7 have the highest rate constants for deuteroxide-catalyzed exchange, while the 6- and 7-membered ring fused salts 8 and 9 have lower values (n = 1 > n = 2 ≈ n = 3). The differences observed are relatively small (krel ≤ 2-fold); however, they are substantially outside experimental error, with added confidence provided from the observation of the same trend for different N-aryl substituents. In addition, the first-order rate constants for H/D-exchange (kex), which are used to calculate kDO, are consistently higher for triazolium salts 7 compared to 8 and 9 (see log kex–pD profiles Figures S25–S28).
Figure S29 includes a Hammett analysis of protofugalities (kDO) as a function of fused ring size. In the case of the N-pentafluorophenyl substituent, a σ substituent constant calculated by Taft26 is utilized. Hammett ρ values, obtained as slopes of the correlations in Figure S29, are positive for the three series as expected for a process favored by electron-withdrawing substituents. The Hammett substituent dependencies demonstrated by 5-, 6-, and 7-membered fused rings are closely similar in magnitude (ρ = 0.65n=1, ρ = 0.65n=2, 0.67n=3). Notably, these experimental ρ values are all less than unity, indicating a similar but smaller N-aryl substituent dependence than for the reference acid dissociation of benzoic acids despite having the same three bond distance from the acidic hydrogen to the ipso-aryl carbon. For 7–9, the inherent stabilization provided by the three nitrogen atoms within the central NHC heterocycle could decrease the dependence on the nature of the N-aryl substituent. Furthermore, the closely similar experimental ρ values suggest that the N-aryl substituent effect is independent of the fused ring.
X-ray Crystallographic Studies
During the syntheses of bicyclic triazolium salts for our kinetic studies, we determined 20 single-crystal X-ray structures (Figure 4), which includes 16 previously unreported structures (section S1.6, Supporting Information, and CCDC 2124937–2124950; 2124952–2124958). The crystals of triazolium salts 7a–g, 7i–k, 8a–c, 8i, and 9a–b, 9k were obtained by slow evaporation of methanol. The growth of single crystals of 7h, 8k, and 9c suitable for single-crystal X-ray diffraction analysis was performed via a modified high-throughput encapsulated nanodroplet crystallization (ENaCt) approach (section S1.6.1).27
Figure 4.
Summary of the single-crystal X-ray structures of 20 triazolium salts determined herein. (*Electron deficiency ordering based on Hammett substituent parameters; Table S13.)
The X-ray crystallographic data was analyzed for structural insight on the role of the size of the fused ring and potential structure–kinetic activity trends. The bond lengths and bond angles of individual triazolium salt core structures are summarized in Tables S17 and S18. In general, triazolium salts with the same fused ring size were found to result in similar core structural parameters irrespective of the N-aryl substituent. Within each series (n = 1 (11 structures), n = 2 (5 structures), and n = 3 (4 structures)) the average bond angles and bond lengths of triazolium salts were calculated and are summarized in Tables 2 and S19. The differences in average angles and distances between series (n = 1 vs n = 2 and n = 2 vs n = 3) are also included in Table 2.
Table 2. Summary of Average Bond Angles and Distances of Triazolium Salts 7a–k (n = 1); 8a–c, 8i, 8k (n = 2); 9a–c, 9k (n = 3) Obtained from Single-Crystal X-ray Structural Analysis.
| triazolium
salts: average angles and distancesb |
||||||
|---|---|---|---|---|---|---|
| n = 1 | n = 2 | n = 3 | n = 1 vs n = 2c | n = 2 vs n = 3d | ||
| bond angles (deg)a | CβC5N4 | 110.6 [0.5] | 122.2 [0.5] | 124.2 [0.5] | 11.6 | 2.0 |
| C5N4Cα | 113.7 [0.4] | 125.6 [0.7] | 127.6 [0.3] | 11.8 | 2.0 | |
| N4C5N1 | 111.9 [0.2] | 111.3 [0.6] | 110.7 [0.2] | –0.6 | –0.6 | |
| C3N4C5 | 107.4 [0.2] | 106.9 [0.5] | 106.9 [0.3] | –0.5 | 0.0 | |
| N1N2C3 | 111.9 [0.4] | 111.7 [0.4] | 110.8 [0.7] | –0.2 | –0.9 | |
| N2C3N4 | 106.1 [0.3] | 106.8 [0.3] | 107.4 [0.3] | 0.8 | 0.6 | |
| C5N1N2 | 102.8 [0.3] | 103.3 [0.1] | 104.1 [0.4] | 0.6 | 0.7 | |
| distance (Å) | H1H2e | 3.06 [0.04] | 2.76 [0.02] | 2.49 [0.03] | –0.31 | –0.26 |
| H1H3e | 3.29 [0.05] | 3.09 [0.03] | 3.37 [0.05] | –0.19 | 0.28 | |
| torsion angle (deg) | H1C3CαH2e | 43.6 [1.9] | 38.2 [2.3] | 1.7 [0.9] | –5.4 | –36.5 |
| H1C3CαH3e | 69.6 [1.6] | 70.9 [2.2] | 105.3 [1.6] | 1.3 | 34.4 | |
Average of bond angle values obtained by X-ray diffraction measurement for triazolium salts 7a–k (n = 1); 8a–c, 8i, 8k (n = 2); 9–c, 9k (n = 3).
Standard deviation of average values is shown in square brackets.
Difference = Average(n = 2) – Average(n = 1).
Difference = Average(n = 3) – Average(n = 2).
The Cα hydrogens with the shorter and longer through-space distances from H1 are labeled as H2 and H3, respectively.
Focusing first on average bond angles in the triazolyl and fused rings, it can be observed that significant changes in bond angles occur between the 5-, 6-, and 7-membered ring series (Table 2 and Figure 5). With expansion of the size of the fused ring (from 5- through to 6- and 7-membered ring series), bond angles at the fusion point within the fused ring increase, and this impacts the bond angles in the adjacent triazolyl ring. Changing from 5- to 6-membered rings (n = 1 to 2) results in larger increases for the CβC5N4 and C5N4Cα average angles by 11.6° and 11.8°, with smaller further increases of 2.0° for both angles upon changing from 6- to 7-membered rings (n = 2 to n = 3, Figure 5a). As a result, three of the five angles of the central triazolyl ring decrease (C3N4C5, N4C5N1, and N1N2C3; Figure 5b), while the remaining two angles increase (N2C3N4 and C5N1N2; Figure 5c) owing to the increasing fused ring size.
Figure 5.
Bond angles (a) CβC5N4 and C5N4Cα increase; (b) N1N2C3, C3N4C5, and N4C5N1 decrease; and (c) C5N1N2 and N2C3N4 increase as n increases.
The observed angular changes at the fusion point within the fused ring upon moving from the 5- to 6-membered ring series are substantially larger (∼11–12° from n = 1 to n = 2) than between the 6- and 7-membered ring series (∼2° from n = 2 to n = 3), which parallels the observed effects of ring size on protofugality. Irrespective of N-aryl substituent, 5-membered ring fused triazolium salts 7 have the highest kDO values, while 6- and 7-fused salts 8 and 9 have smaller but similar values (n = 1 > n = 2 ≈ n = 3). The angular changes within the triazolyl ring appear more similar with increasing fused ring size; however, its significance cannot be unequivocally ascertained owing to the smaller changes involved. Of most relevance to the H/D exchange studies herein, there is a small but significant increase in the N2C3N4 angle at the acidic C(3)-H position from n = 1 to n = 3 in all cases. This ranged from an increase of 1.8° for 7–9a, 1.8° for 7–9b, 1.3° for 7–9c, and 1.0° for 7–9k.
We also analyzed the through-space distances from the C(3)-hydrogen (H1) to the most proximal methylene hydrogens (Cα-H2 and H3) with changing fused ring size, but note the caveat of a greater uncertainty in hydrogen atom positioning from standard X-ray diffraction analysis. The through-space distances are reproduced in our DFT analysis of all the triazolium ions (vide infra), giving confidence in these values. The Cα hydrogens with the shorter and longer distances from H1 are labeled as H2 and H3, respectively (Table 2). These H1H2 and H1H3 distances are used as one measure of differences in the magnitude of the steric effect of the proximal methylene group as a function of changing fused ring size. The average H1H2 distance decreases from 3.06 to 2.49 Å with expansion of the size of the fused ring indicative of a corresponding increased steric effect of the methylene from 5- through to 6- and 7-membered ring series. This parallels the overall observed decrease in protofugality, and rate constants for deuteroxide-catalyzed C(3)-H/D exchange, with increasing fused ring size presumably owing to greater steric hindrance of the approach of DO– to H1 when closer to H2.
The through-space distances of H1 from H2 and H3 are linked with changes in the torsion angles H1C3CαH2 and H1C3CαH3, respectively. Notably, CαH2 is almost coplanar with C3H1 when n = 3 as the average H1C3CαH2 dihedral angle is only 1.7°. By contrast, CαH2 is not coplanar with C3H1 when n = 1 and n = 2 with similar average H1C3CαH2 torsion angles of 43.6° and 38.2° for the 5- and 6-membered series. We postulate that the almost coplanar relationship of C3H1 and CαH2 (n = 3), in addition to the shorter through-space distance of H1 and H2 in this case, may contribute to an increased steric hindrance on the approach of the base and a resultant decrease in protofugalities. While these through-space distance and torsion angle trends align with the overall decrease in protofugalities from n = 1 to n = 3, clearly additional factors are needed to explain the observation of higher kDO values for the 5-fused triazolium series with similar, smaller values for 6- and 7-fused salts 8 and 9 (n = 1 > n = 2 ≈ n = 3).
Finally, alteration of the triazolium N-aryl group changes the N1N2CγCδ dihedral angle in the solid-state structure (Tables S17 and S18). Typically, N-aryl groups without ortho-substituents are close to coplanar with the central triazolium ring and have N1N2CγCδ dihedral angles of ∼0°. Unsurprisingly, the introduction of bulky ortho-substituents on the N-aryl ring increases N1N2CγCδ with the largest dihedral angle of 91.2° observed for di-ortho-ipropylphenyl-substituted 7h. Alteration of the N-aryl group is external to the triazolium core structure and removed from the fused ring. Thus, changes in the N1N2CγCδ dihedral angle are mainly dictated by the N-aryl group rather than fused ring size (n = 1, 2, 3). This is consistent with the observation of closely similar ρ-values in our Hammett analysis of kDO for 7, 8, and 9, respectively.
DFT Calculations
To further understand the key structural factors underpinning observed differences in protofugalities with fused ring size, the 20 triazolium salts 7–9 in Figure 2 and their corresponding carbenes 7′–9′, were studied computationally (Supporting Information, section S2). The B3LYP and M062X functional with the polarized diffuse split-valence 6-311++g(d,p) basis set and the polarizable continuum model (PCM) with the solvent option of water were used in all cases.28 Several conformations of each molecule as starting geometries were optimized to find the most stable (lowest energy) geometry. The vibrational frequencies of these most stable geometries were then calculated at the same level of theory and revealed no imaginary frequencies, confirming them as true minima. The calculated bond angles and bond lengths of the core bicyclic structures of the 20 individual triazolium salts 7–9, and corresponding carbenes 7′–9′, are listed in Tables S21–S24 and S28–S31. The lowest-energy conformations calculated computationally for the triazolium ions 7–9 matched the solid-state experimental X-ray structures. Pleasingly, the trends in computed average angles and distances for the triazolium series 7–9 (Table S25 and S26) also match closely with the experimental values in Table 2. To our knowledge, while there is a published X-ray crystal structure for a 2,4,5-triphenyltriazolylidene,29 there are no reported experimental structural data for triazolylidenes 7′–9′. The excellent computational reproduction of experimental trends for the fused ring effect in the triazolium series 7–9 gives confidence in a similar analysis of only computational data for the corresponding triazolyl NHCs 7′–9′.
The average bond angles of triazolylidenes 7′–9′ obtained by calculation are summarized in Tables 3, S32, and S33. For a given size of the fused ring, bond angles within the heterocyclic ring are significantly influenced by conversion of the triazolium conjugate acid 7–9 to the corresponding triazolyl carbene 7′–9′. The largest changes in average bond angle is for N2C3N4, which decreases by ∼6° upon formation of the carbene (Table 2 versus Table 3).
Table 3. Summary of Average Bond Angles and Distances of Triazolylidenes 7′a–k (n = 1); 8′a–c, 8′i, 8′k (n = 2); 9′a–c, 9′k (n = 3) Obtained from DFT Calculation (M062X).
| triazolyl
carbene: calculated average angles and
distancesb |
||||||
|---|---|---|---|---|---|---|
| n = 1 | n = 2 | n = 3 | n = 1 vs n = 2c | n = 2 vs n = 3d | ||
| bond angle (deg)a | CβC5N4 | 111.1 [0.0] | 122.9 [0.0] | 124.5 [0.1] | 11.9 | 1.6 |
| C5N4Cα | 112.7 [0.1] | 124.1 [0.1] | 124.7 [0.1] | 11.5 | 0.5 | |
| N4C5N1 | 110.9 [0.1] | 110.1 [0.1] | 110.0 [0.1] | –0.8 | –0.1 | |
| C3N4C5 | 111.6 [0.1] | 110.8 [0.2] | 110.8 [0.1] | –0.8 | 0.0 | |
| N1N2C3 | 115.9 [0.3] | 115.4 [0.3] | 115.3 [0.3] | –0.5 | –0.1 | |
| N2C3N4 | 99.6 [0.2] | 100.5 [0.2] | 100.6 [0.3] | 0.9 | 0.1 | |
| C5N1N2 | 102.1 [0.2] | 103.2 [0.3] | 103.3 [0.3] | 1.1 | 0.2 | |
| distance (Å) | C3H2e | 2.92 [0.00] | 2.69 [0.00] | 2.53 [0.00] | –0.22 | –0.16 |
| C3H3e | 3.05 [0.00] | 2.91 [0.00] | 3.18 [0.00] | –0.14 | 0.28 | |
| torsion angle (deg) | N1C3CαH2e | 50.6 [0.1] | 42.1 [0.3] | 2.9 [0.3] | –8.5 | –39.2 |
| N1C3CαH3e | 85.4 [0.1] | 82.6 [0.2] | 121.5 [0.3] | –2.7 | 38.8 | |
Average of bond angle values obtained by DFT calculation (M062X) for triazolylidenes 7′a–k (n = 1); 8′a–c, 8′i, 8′k (n = 2); 9′a–c, 9′k (n = 3).
Standard deviation of average values is shown in square brackets.
Difference = Average(n = 2) – Average(n = 1).
Difference = Average(n = 3) – Average(n = 2).
The Cα hydrogens with the shorter and longer distances from C3 are labeled as H2 and H3, respectively.
Closely similar effects are observed upon increase in fused ring size for the bicyclic triazolyl carbenes 7′–9′ as observed for the triazolium ions 7–9. Changing from 5- to 6-membered fused ring series (n = 1 to 2) results in large increases in the average CβC5N4 and C5N4Cα angles of the triazolylidenes 7′–9′ (11.8° and 11.4°, respectively), with smaller further increases of 1.6° (CβC5N4) and 0.6° (C5N4Cα) upon changing to 6- and 7-membered series (n = 2 to n = 3). Three of the five angles of the central triazolyl ring decrease (C3N4C5, N4C5N1, and N1N2C3), while the remaining two angles increase (N2C3N4 and C5N1N2) owing to the increasing fused ring size. From XRD data, the average N2C3N4 angle of triazolium salts 7 with n = 1 equals 106.1° (±0.3°), and this number increases to 106.8° (±0.3°) and 107.4° (±0.3°) with n = 2 and 3 for 8 and 9. From computational data for the corresponding NHCs 7′–9′ there is a similar increase in the average N2C3N4 angle from 99.6° (±0.2°) for n = 1 to 100.5° (±0.2°) and 100.6° (±0.3°) for n = 2 and n = 3.
X-ray structural analysis allowed for an evaluation of the distance of the most proximal methylene hydrogens of the fused ring (Cα-H2 and H3) from the C(3)-H1, and changes in the dihedral angles H1C3CαH2 and H1C3CαH3, with changing fused ring size in the triazolium series 7–9. Similar conclusions can be drawn from an analysis of the computational data for the triazolylidenes 7′–9′ in this case represented by the through-space distances of H2 and H3 to the carbenic C(3)-center and torsional angles N1C3CαH2 and N1C3CαH3 (Table 3). Again, the methylene hydrogens H2 and H3 are closer to the carbenic position (C3) for the largest fused ring size (n = 3). Using these through-space distances and torsion angles as one measure of the steric effect of the most proximal methylene on the fused ring suggests that the steric impediment to approach of another reagent (e.g., DO–) to either the triazolium or carbene is largest when n = 3 and falls off for the smaller fused rings sizes (n = 2 ∼ n = 1).
As the structural and computational data discussed thus far relate only to the lowest-energy conformations of the triazolium ions and triazolylidenes, we have additionally evaluated the role of conformational flexibility in the fused ring through DFT modeling (section S2.5). By fixing the torsion angle between C3H1 and CαH2 for 7–9b and between N1C3 and CαH2 for 7′–9′b, the energy changes caused by conformation in the fused ring in the vicinity of the carbenic position could be evaluated (Tables S35–S40). The Boltzmann distribution of conformer mole fraction can then be obtained from the energy difference compared with the lowest-energy conformation (Figures S51 and S52). These plots clearly highlight the favored coplanar conformation of Cα-H2 and the carbenic position for 9b and 9′b (n = 3), whereas for 7/7′b (n = 1) and 8/8′b (n = 2) the minimum energy preference is for a larger torsion angle of 35–42°. Despite having different minimum energy starting points, the energy-torsion angle profiles for all structures probed are essentially superimposable after applying starting point corrections (Figure S53).30
NCN Angle Effects on Proton Transfer
In our previous H/D-exchange studies we observed substantial decreases in kDO with increase in the internal NCN angle of the NHC. For example, protofugalities for imidazolium ions 10 and 11, which are the conjugate acids of imidazolyl NHCs, are 3–8 orders of magnitude higher than for tetrahydropyrimidinium ions 12. In the imidazolyl case the internal NCN angles within the five-membered heterocyclic ring are substantially smaller than within the six-membered ring of 12 (Figure 6). Specifically, comparing two examples with N-alkyl substituents, 11 and 12, for an increase in NCN angle of ∼13° there is a corresponding drop in kDO of 720-fold. Similarly, comparing 12 and 13 with N-iPr substituents, an increase in NCN angle of ∼7° is accompanied by a drop in kDO by ∼50-fold. Although other factors, including the number and type of NHC heteroatoms, aromaticity, and N-substitution, influence protofugalities, the NCN angle changes are clearly also important. We previously argued that an enforced increase in internal NCN angle through alteration of heterocyclic ring size is better accommodated by the conjugate acid (azolium ion) than the conjugate base (NHC), thus contributing to a decrease in acidity.15d This argument is supported by closely similar trends for carbene proton affinity (PA) measurements with higher PAs observed for NHCs with larger ring sizes (Figure 7).15e
Figure 6.
Decrease in protofugality (kDO) with increasing NCN angle.15d,31
Figure 7.
Increase in proton affinity (PA) with increasing NCN angle.15e,32
Although substantially smaller in the present case, the increase in N2C3N4 bond angle (by ∼2°) at the acidic position with changing fused ring size from n = 1 to n = 3 also correlates with decreases in kDO (kDOrel ≤ 2). The increase in N2C3N4 angle with changing fused ring size is observed consistently for different N-aryl substituents, supporting the generality of this trend.
In general an increase in angle toward 180° at a carbene center is associated with an increase in p-character of the nonbonding orbitals bearing the negative charge. Similarly, in the present case the small increase in NCN bond angle moving from n = 1 to n = 3 could be qualitatively associated with a concomitant increase in p-character, or decreased s-character, of the hybrid orbital at the carbenic center.33 Parallels can be drawn with the common textbook explanation for the increase in carbon acidity along the alkane (sp3C-H, pKa ∼ 45–70), alkene (sp2C-H, pKa ∼ 40–45), and alkyne (sp C-H, pKa ∼ 20–25) hydrocarbon series, where the negative charge formed upon deprotonation is more stable in an orbital of greater s-character.34 Using this argument, we can speculate that the higher protofugalities of the 5-fused triazolyl series might result from increased s-character at the C(3)-position stemming from smaller NCN angles relative to the 6- and 7-fused series.
Estimation of Carbon Acid pKa Values
There is an increasingly large body of work on the conjugate acid pKa’s of NHCs in a range of solvents, including water, acetonitrile, and DMSO, in addition to related gas-phase acidities and proton affinities.15a−15d,24,32b,35 Several approaches may be employed to access these carbon acid pKa values, which we have reviewed previously.15e In aqueous solution, the main difficulty associated with the direct determination of conjugate acid pKa values of NHCs is the well-established leveling effect. Owing to the substantially higher basicities of most NHCs relative to the conjugate base of solvent (HO–), quantitative deprotonation of solvent occurs.
In the determination of aqueous carbon acidities of weakly acidic species, one solution long-employed to circumvent this problem is the calculation of pKa from the rate constants for the forward and reverse directions of the proton transfer equilibrium.36 We and others have previously employed this kinetic method for the determination of carbon acid pKa values for deprotonation at C(3) for a series of triazolium salts including 7 using eq 4 derived for Scheme 2.15a−15d,24 In this equation, kHO (M–1 s–1) is the second-order rate constant for deprotonation at C(3) by hydroxide ion, which may be calculated from the corresponding kDO value using a value of kDO/kHO = 2.4 for the secondary solvent isotope effect on the basicity of HO– in H2O versus DO– in D2O. As discussed previously,15a,15d the absence of significant general base catalysis of deuterium exchange provides good evidence that the reverse protonation of the triazol-3-ylidene NHC by water is equal or close to the limiting rate constant for the physical process of dielectric relaxation of the solvent (kHOH ≤ kreorg = 1011 s–1). Rate constants of this magnitude are challenging to access experimentally, generally requiring femtosecond transient absorption spectroscopy; however, values for the dielectric relaxation of water have been determined reliably.37 Because the main error in pKa determination using this method is associated with the value assumed for kHOH, these pKa values provide upper limit estimations.
| 4 |
Scheme 2. Equilibrium for Acid Dissociation of Triazolium Ions.
Using the kDO values determined herein, and by application of eq 4, we have calculated pKa values for triazolium salts 8a–c,k with 6-fused rings and 9a–c,k with 7-fused rings (Table 1). Owing to the ∼2-fold effect of fused ring size on kDO, and the logarithmic relationship to pKa in eq 4, the resulting pKa values for a given N-aryl substituent are closely similar. These triazolium pKa’s in the 16.7–18 range are ∼4 and 6 units below typical aqueous values for diaryl- and dialkylimidazolium ions, respectively, whereas only ∼1–2 units below typical N-alkylthiazolium pKa’s.15a−15d,24 The overall effect of the central ring heteroatoms on pKa is larger than either N-aryl or fused ring substituent effects. The influence of fused ring size on kinetic acidities or protofugalities, however, is significantly larger than on pKa. Thus, rates and half-lives of NHC generation by C(3)-deprotonation of the conjugate acid, which is of direct relevance to catalysis, will be significantly influenced by the nature of the fused ring size in the bicyclic triazolium precursor.
Conclusion
In conclusion, this work has probed the effect of the size of the fused rings (5-, 6-, and 7-membered; n = 1, 2, and 3 respectively) within a range of bicyclic triazolium salts on their kinetic acidities, or protofugalities. Rate constants for the deuteroxide-catalyzed H/D-exchange of triazolium salts, kDO, were observed to be significantly influenced by the size of the adjacent fused ring, with the trend following the order kDO(5-membered ring, n = 1) > kDO(6-membered ring, n = 2) ≈ kDO(7-membered ring, n = 3). Hammett ρ values, obtained as slopes of Hammett correlations of kDO values with varying NHC N-aryl substituent, were positive for the three series (5-, 6-, and 7-membered rings, n = 1–3) consistent with a deprotonation process favored by electron-withdrawing substituents. Hammett N-aryl substituent dependencies were found to be closely similar (ρ = 0.65n=1, ρ = 0.64n=2, 0.67n=3). Detailed analyses of X-ray diffraction structural data and computational analysis for 20 triazolium salts/triazolylidenes provide insight on the structural effects of altering fused ring size, which could be aligned with observed effects on kDO. The most proximal methylene of the fused ring is closest and almost coplanar with the carbenic center when n = 3, indicative of a greater steric effect; this parallels our observation of a decrease in rates of deprotonation by DO– when n = 3. Changing from 5- to 6-membered fused ring (n = 1 to 2) results in significant increases (by ∼11–12°) in the CβC5N4 and C5N4Cα angles within the fused ring, with smaller further increases of ∼2° upon changing from 6- to 7-membered ring (n = 2 to n = 3). A resultant change in internal angles in the adjacent NHC ring is also observed. In particular, the resulting ∼2° increase in internal NCN angle is consistent with a drop in acidity (lower kDO values) with increased fused ring size. Finally, carbon acid pKa values of the bicyclic triazolium salts were estimated using a kinetic method.
Given the prevalence of bicyclic structures in NHC organocatalytic scaffolds and their importance for many C–C bond-forming reactions, we hope this work provides quantitative and structural insight into the role of a simple incremental change in the number of methylenes in the fused ring. Clearly, for the proton transfer step between a triazolium salt and triazolyl carbene in aqueous solvent, the size of the fused ring size is important, with five membered fused rings (n = 1) giving significantly higher rates of deprotonation. Ongoing studies in our laboratories are focused on kinetic analysis of the role of the fused ring size in later reaction steps in triazolium-catalyzed benzoin and related acyloin reactions.
Acknowledgments
We thank the EPSRC (J.Z. and C.M.Y., EP/S020713/1; P.Q., EP/M506321/1), CSC St Andrews Ph.D. studentship (Z.D.), Universidad de Castilla-La Mancha and the European Regional Development Fund (I.M.), and the BBSRC (M.J.H., M.R.P., and A.R.T., BB/L013754/1) for funding. We are grateful to Durham Chemistry NMR service for their ongoing help and support.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.joc.1c03073.
Synthetic procedures, 1H NMR and 13C spectral data, and HRMS data; kinetic procedures, fitting, and data; single-crystal X-ray diffraction data; computational details (PDF)
Accession Codes
CCDC 2124937–2124950 and 2124952–2124958 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.
The authors declare no competing financial interest.
Supplementary Material
References
- Ukai T.; Tanaka R.; Dokawa T. A new catalyst for acyloin condensation. J. Pharm. Soc. Jpn. 1943, 63, 296–300. 10.1248/yakushi1881.63.6_296. [DOI] [Google Scholar]
- Breslow R. On the mechanism of thiamine action. IV. 1 Evidence from studies on model systems. J. Am. Chem. Soc. 1958, 80, 3719–3726. 10.1021/ja01547a064. [DOI] [Google Scholar]
- a Flanigan D. M.; Romanov-Michailidis F.; White N. A.; Rovis T. Organocatalytic reactions enabled by N-heterocyclic carbenes. Chem. Rev. 2015, 115, 9307–9387. 10.1021/acs.chemrev.5b00060. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Nair V.; Vellalath S.; Babu B. P. Recent advances in carbon–carbon bond-forming reactions involving homoenolates generated by NHC catalysis. Chem. Soc. Rev. 2008, 37, 2691–2698. 10.1039/b719083m. [DOI] [PubMed] [Google Scholar]; c Zhao C.; Blaszczyk S. A.; Wang J. Asymmetric reactions of N-heterocyclic carbene (NHC)-based chiral acyl azoliums and azolium enolates. Green Synth. Catal. 2021, 2, 198–215. 10.1016/j.gresc.2021.03.003. [DOI] [Google Scholar]; d Song R.; Jin Z.; Chi Y. R. NHC-catalyzed covalent activation of heteroatoms for enantioselective reactions. Chem. Sci. 2021, 12, 5037–5043. 10.1039/D1SC00469G. [DOI] [PMC free article] [PubMed] [Google Scholar]; e Dai L.; Ye S. Recent advances in N-heterocyclic carbene-catalyzed radical reactions. Chin. Chem. Lett. 2021, 32, 660–667. 10.1016/j.cclet.2020.08.027. [DOI] [Google Scholar]; f Enders D.; Niemeier O.; Henseler A. Organocatalysis by N-heterocyclic carbenes. Chem. Rev. 2007, 107, 5606–5655. 10.1021/cr068372z. [DOI] [PubMed] [Google Scholar]; g Grossmann A.; Enders D. N-heterocyclic carbene catalyzed domino reactions. Angew. Chem., Int. Ed. 2012, 51, 314–325. 10.1002/anie.201105415. [DOI] [PubMed] [Google Scholar]; h Chen X. Y.; Liu Q.; Chauhan P.; Enders D. N-Heterocyclic carbene catalysis via azolium dienolates: an efficient strategy for remote enantioselective functionalizations. Angew. Chem., Int. Ed. 2018, 57, 3862–3873. 10.1002/anie.201709684. [DOI] [PubMed] [Google Scholar]
- a Yatham V. R.; Harnying W.; Kootz D.; Neudörfl J. r.-M.; Schlörer N. E.; Berkessel A. 1,4-Bis-dipp/mes-1,2,4-triazolylidenes: carbene catalysts that efficiently overcome steric hindrance in the redox esterification of α-and β-substituted α,β-enals. J. Am. Chem. Soc. 2016, 138, 2670–2677. 10.1021/jacs.5b11796. [DOI] [PubMed] [Google Scholar]; b Wu J.; Zhao C.; Wang J. Enantioselective intermolecular enamide-aldehyde cross-coupling catalyzed by chiral N-heterocyclic carbenes. J. Am. Chem. Soc. 2016, 138, 4706–4709. 10.1021/jacs.5b13501. [DOI] [PubMed] [Google Scholar]; c Davidson R. W.; Fuchter M. J. Direct NHC-catalysed redox amidation using CO2 for traceless masking of amine nucleophiles. Chem. Commun. 2016, 52, 11638–11641. 10.1039/C6CC04639H. [DOI] [PubMed] [Google Scholar]; d Yasui K.; Kamitani M.; Fujimoto H.; Tobisu M. N-Heterocyclic carbene-catalyzed truce–smiles rearrangement of N-arylacrylamides via the cleavage of unactivated C(aryl)–N bonds. Org. Lett. 2021, 23, 1572–1576. 10.1021/acs.orglett.0c04281. [DOI] [PubMed] [Google Scholar]; e Satyam K.; Ramarao J.; Suresh S. N-Heterocyclic carbene (NHC)-catalyzed intramolecular benzoin condensation–oxidation. Org. Biomol. Chem. 2021, 19, 1488–1492. 10.1039/D0OB02606A. [DOI] [PubMed] [Google Scholar]
- a Hachisu Y.; Bode J. W.; Suzuki K. Catalytic intramolecular crossed aldehyde--ketone benzoin reactions: a novel synthesis of functionalized preanthraquinones. J. Am. Chem. Soc. 2003, 125, 8432–8433. 10.1021/ja035308f. [DOI] [PubMed] [Google Scholar]; b Li G.-Q.; Dai L.-X.; You S.-L. Thiazolium-derived N-heterocyclic carbene-catalyzed cross-coupling of aldehydes with unactivated imines. Chem. Commun. 2007, 852–854. 10.1039/B611646A. [DOI] [PubMed] [Google Scholar]; c Delany E. G.; Connon S. J. Enantioselective N-heterocyclic carbene-catalysed intermolecular crossed benzoin condensations: improved catalyst design and the role of in situ racemisation. Org. Biomol. Chem. 2021, 19, 248–258. 10.1039/D0OB02017F. [DOI] [PubMed] [Google Scholar]; d Rose C. A.; Gundala S.; Fagan C.-L.; Franz J. F.; Connon S. J.; Zeitler K. NHC-catalysed, chemoselective crossed-acyloin reactions. Chem. Sci. 2012, 3, 735–740. 10.1039/C2SC00622G. [DOI] [Google Scholar]; e O’Toole S. E.; Rose C. A.; Gundala S.; Zeitler K.; Connon S. J. Highly chemoselective direct crossed aliphatic–aromatic acyloin condensations with triazolium-derived carbene catalysts. J. Org. Chem. 2011, 76, 347–357. 10.1021/jo101791w. [DOI] [PubMed] [Google Scholar]; f Ramanjaneyulu B. T.; Mahesh S.; Vijaya Anand R. N-Heterocyclic carbene catalyzed highly chemoselective intermolecular crossed acyloin condensation of aromatic aldehydes with trifluoroacetaldehyde ethyl hemiacetal. Org. Lett. 2015, 17, 6–9. 10.1021/ol502581b. [DOI] [PubMed] [Google Scholar]
- a Stetter H. Catalyzed addition of aldehydes to activated double bonds—a new synthetic approach. Angew. Chem., Int. Ed. 1976, 15, 639–647. 10.1002/anie.197606391. [DOI] [Google Scholar]; b Enders D.; Breuer K.; Runsink J.; Teles J. H. The first asymmetric intramolecular Stetter reaction. Preliminary communication. Helv. Chim. Acta 1996, 79, 1899–1902. 10.1002/hlca.19960790712. [DOI] [Google Scholar]; c Debiais M.; Hamoud A.; Drain R.; Barthélémy P.; Desvergnes V. Bio-inspired NHC-organocatalyzed Stetter reaction in aqueous conditions. RSC Adv. 2020, 10, 40709–40718. 10.1039/D0RA08326G. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Heravi M. M.; Zadsirjan V.; Kafshdarzadeh K.; Amiri Z. Recent advances in Stetter reaction and related chemistry: an update. Asian J. Org. Chem. 2020, 9, 1999–2034. 10.1002/ajoc.202000378. [DOI] [Google Scholar]; e Geng H.; Chen X.; Gui J.; Zhang Y.; Shen Z.; Qian P.; Chen J.; Zhang S.; Wang W. Practical synthesis of C1 deuterated aldehydes enabled by NHC catalysis. Nat. Catal. 2019, 2, 1071–1077. 10.1038/s41929-019-0370-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a He L.; Lv H.; Zhang Y. R.; Ye S. Formal cycloaddition of disubstituted ketenes with 2-oxoaldehydes catalyzed by chiral N-heterocyclic carbenes. J. Org. Chem. 2008, 73, 8101–8103. 10.1021/jo801494f. [DOI] [PubMed] [Google Scholar]; b Zhang Y. R.; Lv H.; Zhou D.; Ye S. Chiral N-heterocyclic carbene-catalyzed formal [4 + 2] cycloaddition of ketenes with enones: highly enantioselective synthesis of trans-and cis-δ-lactones. Chem.—Eur. J. 2008, 14, 8473–8476. 10.1002/chem.200801165. [DOI] [PubMed] [Google Scholar]; c Wang Y.; Qu L.-B.; Wei D.; Lan Y. Unveiling the chemo-and stereoselectivities of NHC-catalyzed reactions of an aliphatic ester with aminochalcone. J. Org. Chem. 2020, 85, 8437–8446. 10.1021/acs.joc.0c00769. [DOI] [PubMed] [Google Scholar]
- a Wang Y.; Qu L. B.; Lan Y.; Wei D. Origin of regio-and stereoselectivity in the NHC-catalyzed reaction of alkyl pyridinium with aliphatic enal. ChemCatChem. 2020, 12, 1068–1074. 10.1002/cctc.201901965. [DOI] [Google Scholar]; b Wang Y.; Wu Q.-Y.; Lai T.-H.; Zheng K.-J.; Qu L.-B.; Wei D. Prediction on the origin of selectivities of NHC-catalyzed asymmetric dearomatization (CADA) reactions. Catal. Sci. & Technol. 2019, 9, 465–476. 10.1039/C8CY02238K. [DOI] [Google Scholar]; c Guo C.; Fleige M.; Janssen-Müller D.; Daniliuc C. G.; Glorius F. Switchable selectivity in an NHC-catalysed dearomatizing annulation reaction. Nat. Chem. 2015, 7, 842. 10.1038/nchem.2337. [DOI] [PubMed] [Google Scholar]; d Janssen-Müller D.; Fleige M.; Schlüns D.; Wollenburg M.; Daniliuc C. G.; Neugebauer J.; Glorius F. NHC-catalyzed enantioselective dearomatizing hydroacylation of benzofurans and benzothiophenes for the synthesis of spirocycles. ACS Catal. 2016, 6, 5735–5739. 10.1021/acscatal.6b01852. [DOI] [Google Scholar]; e Flanigan D. M.; Rovis T. Enantioselective N-heterocyclic carbene-catalyzed nucleophilic dearomatization of alkyl pyridiniums. Chem. Sci. 2017, 8, 6566–6569. 10.1039/C7SC02648J. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Nguyen X. B.; Nakano Y.; Duggan N. M.; Scott L.; Breugst M.; Lupton D. W. N-Heterocyclic carbene catalyzed (5 + 1) annulations exploiting a vinyl dianion synthon strategy. Angew. Chem., Int. Ed. 2019, 131, 11607–11614. 10.1002/ange.201905475. [DOI] [PubMed] [Google Scholar]; b Draskovits M.; Kalaus H.; Stanetty C.; Mihovilovic M. D. Intercepted dehomologation of aldoses by N-heterocyclic carbene catalysis–a novel transformation in carbohydrate chemistry. Chem. Commun. 2019, 55, 12144–12147. 10.1039/C9CC05906G. [DOI] [PubMed] [Google Scholar]; c Zhao L.-L.; Li X.-S.; Cao L.-L.; Zhang R.; Shi X.-Q.; Qi J. Access to dihydropyridinones and spirooxindoles: application of N-heterocyclic carbene-catalyzed [3 + 3] annulation of enals and oxindole-derived enals with 2-aminoacrylates. Chem. Commun. 2017, 53, 5985–5988. 10.1039/C7CC02753B. [DOI] [PubMed] [Google Scholar]; d Schedler M.; Wurz N. E.; Daniliuc C. G.; Glorius F. N-Heterocyclic carbene catalyzed umpolung of styrenes: mechanistic elucidation and selective tail-to-tail dimerization. Org. Lett. 2014, 16, 3134–3137. 10.1021/ol501256d. [DOI] [PubMed] [Google Scholar]; e Enders D.; Breuer K.; Raabe G.; Runsink J.; Teles J. H.; Melder J. P.; Ebel K.; Brode S. Preparation, structure, and reactivity of 1,3,4-triphenyl-4, 5-dihydro-1H-1,2,4-triazol-5-ylidene, a new stable carbene. Angew. Chem., Int. Ed. 1995, 34, 1021–1023. 10.1002/anie.199510211. [DOI] [Google Scholar]; f Enders D.; Breuer K.; Teles J.; Ebel K. 1,3,4-Triphenyl-4,5-dihydro-1H-1,2,4-triazol-5-ylidene–applications of a stable carbene in synthesis and catalysis. J. Prakt. Chem. 1997, 339, 397–399. 10.1002/prac.19973390170. [DOI] [Google Scholar]; g Enders D.; Breuer K.; Raabe G.; Simonet J.; Ghanimi A.; Stegmann H. B.; Teles J. H. A stable carbene as π-acceptor electrochemical reduction to the radical anion. Tetrahedron Lett. 1997, 38, 2833–2836. 10.1016/S0040-4039(97)00493-0. [DOI] [Google Scholar]
- Knight R. L.; Leeper F. J. Comparison of chiral thiazolium and triazolium salts as asymmetric catalysts for the benzoin condensation. J. Chem. Soc., Perkin Trans. 1 1998, 1891–1894. 10.1039/a803635g. [DOI] [Google Scholar]
- a Baragwanath L.; Rose C. A.; Zeitler K.; Connon S. J. Highly enantioselective benzoin condensation reactions involving a bifunctional protic pentafluorophenyl-substituted triazolium precatalyst. J. Org. Chem. 2009, 74, 9214–9217. 10.1021/jo902018j. [DOI] [PubMed] [Google Scholar]; b DiRocco D. A.; Oberg K. M.; Dalton D. M.; Rovis T. Catalytic asymmetric intermolecular Stetter reaction of heterocyclic aldehydes with nitroalkenes: backbone fluorination improves selectivity. J. Am. Chem. Soc. 2009, 131, 10872–10874. 10.1021/ja904375q. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Campbell C. D.; Collett C. J.; Thomson J. E.; Slawin A. M.; Smith A. D. Organic base effects in NHC promoted O-to C-carboxyl transfer; chemoselectivity profiles, mechanistic studies and domino catalysis. Org. Biomol. Chem. 2011, 9, 4205–4218. 10.1039/c1ob05160a. [DOI] [PubMed] [Google Scholar]
- a Enders D.; Kallfass U. An efficient nucleophilic carbene catalyst for the asymmetric benzoin condensation. Angew. Chem., Int. Ed. 2002, 41, 1743–1745. . [DOI] [PubMed] [Google Scholar]; b Concellon C.; Duguet N.; Smith A. D. N-Heterocyclic carbene-mediated enantioselective addition of phenols to unsymmetrical alkylarylketenes. Adv. Synth. Catal. 2009, 351, 3001–3009. 10.1002/adsc.200900538. [DOI] [Google Scholar]; c Liu L.; Guo D.; Wang J. NHC-catalyzed asymmetric α-regioselective [4 + 2] annulation to construct α-alkylidene-δ-lactones. Org. Lett. 2020, 22, 7025–7029. 10.1021/acs.orglett.0c02573. [DOI] [PubMed] [Google Scholar]
- a Langdon S. M.; Wilde M. M.; Thai K.; Gravel M. Chemoselective N-heterocyclic carbene-catalyzed cross-benzoin reactions: importance of the fused ring in triazolium salts. J. Am. Chem. Soc. 2014, 136, 7539–7542. 10.1021/ja501772m. [DOI] [PubMed] [Google Scholar]; b Langdon S. M.; Legault C. Y.; Gravel M. Origin of chemoselectivity in N-heterocyclic carbene catalyzed cross-benzoin reactions: DFT and experimental insights. J. Org. Chem. 2015, 80, 3597–3610. 10.1021/acs.joc.5b00301. [DOI] [PubMed] [Google Scholar]
- a Massey R. S.; Collett C. J.; Lindsay A. G.; Smith A. D.; O’Donoghue A. C. Proton transfer reactions of triazol-3-ylidenes: kinetic acidities and carbon acid pKa values for twenty triazolium salts in aqueous solution. J. Am. Chem. Soc. 2012, 134, 20421–20432. 10.1021/ja308420c. [DOI] [PubMed] [Google Scholar]; b Massey R. S.; Quinn P.; Zhou S.; Murphy J. A.; O’Donoghue A. C. Proton transfer reactions of a bridged bis-propyl bis-imidazolium salt. J. Phys. Org. Chem. 2016, 29, 735–740. 10.1002/poc.3567. [DOI] [Google Scholar]; c Tucker D. E.; Quinn P.; Massey R. S.; Collett C. J.; Jasiewicz D. J.; Bramley C. R.; Smith A. D.; O’Donoghue A. C. Proton transfer reactions of N-aryl triazolium salts: unusual ortho-substituent effects. J. Phys. Org. Chem. 2015, 28, 108–115. 10.1002/poc.3399. [DOI] [Google Scholar]; d Higgins E. M.; Sherwood J. A.; Lindsay A. G.; Armstrong J.; Massey R. S.; Alder R. W.; O’Donoghue A. C. pKas of the conjugate acids of N-heterocyclic carbenes in water. Chem. Commun. 2011, 47, 1559–1561. 10.1039/C0CC03367G. [DOI] [PubMed] [Google Scholar]; e O’Donoghue A. C.; Massey R. S. In Contemporary Carbene Chemistry; Moss R. A., Doyle M. P., Eds.; John Wiley & Sons, 2013. [Google Scholar]; f Quinn P.; Smith M. S.; Zhu J.; Hodgson D. R.; O’Donoghue A. C. Triazolium salt organocatalysis: mechanistic evaluation of unusual ortho-substituent effects on deprotonation. Catalysts 2021, 11, 1055. 10.3390/catal11091055. [DOI] [Google Scholar]
- Mayr H.; Ofial A. R. Philicities, fugalities, and equilibrium constants. Acc. Chem. Res. 2016, 49, 952–965. 10.1021/acs.accounts.6b00071. [DOI] [PubMed] [Google Scholar]
- a Gronert S. In Reactive intermediate chemistry; Moss R. A., Platz M., Jones M.; Wiley Online Library, 2004; p 69. [Google Scholar]; b Stivers J. T.; Washabaugh M. W. C(α)-proton transfer from thiazolium ions: Structure-reactivity correlations and the C(α)-H pKa of 2-(1-hydroxyethyl) thiamin. Bioorg. Chem. 1992, 20, 155–172. 10.1016/0045-2068(92)90036-3. [DOI] [Google Scholar]; c Kluger R. Thiamin diphosphate: a mechanistic update on enzymic and nonenzymic catalysis of decarboxylation. Chem. Rev. 1987, 87, 863–876. 10.1021/cr00081a001. [DOI] [Google Scholar]; d Washabaugh M. W.; Jencks W. P. Thiazolium C (2)-Proton exchange: Isotope effects, internal return, and a small intrinsic barrier. J. Am. Chem. Soc. 1989, 111, 683–692. 10.1021/ja00184a043. [DOI] [Google Scholar]; e Washabaugh M. W.; Jencks W. P. Thiazolium C(2)-proton exchange: structure-reactivity correlations and the pKa of thiamin C (2)-H revisited. Biochemistry 1988, 27, 5044–5053. 10.1021/bi00414a015. [DOI] [PubMed] [Google Scholar]; f Kemp D. S.; O’Brien J. Base catalysis of thiazolium salt hydrogen exchange and its implications for enzymic thiamine cofactor catalysis. J. Am. Chem. Soc. 1970, 92, 2554–2555. 10.1021/ja00711a062. [DOI] [PubMed] [Google Scholar]
- a Gish G.; Smyth T.; Kluger R. Thiamin diphosphate catalysis. Mechanistic divergence as a probe of substrate activation of pyruvate decarboxylase. J. Am. Chem. Soc. 1988, 110, 6230–6234. 10.1021/ja00226a044. [DOI] [PubMed] [Google Scholar]; b Crosby J.; Lienhard G. E. Mechanisms of thiamine-catalyzed reactions. Kinetic analysis of the decarboxylation of pyruvate by 3, 4-dimethylthiazolium ion in water and ethanol. J. Am. Chem. Soc. 1970, 92, 5707–5716. 10.1021/ja00722a027. [DOI] [PubMed] [Google Scholar]; c Breslow R. The mechanism of thiamine action: predictions from model experiments. Ann. N.Y. Acad. Sci. 1962, 98, 445–452. 10.1111/j.1749-6632.1962.tb30565.x. [DOI] [PubMed] [Google Scholar]
- a Igau A.; Grutzmacher H.; Baceiredo A.; Bertrand G. Analogous. alpha.,. alpha.’-bis-carbenoid, triply bonded species: synthesis of a stable. lambda. 3-phosphino carbene-. lambda. 5-phosphaacetylene. J. Am. Chem. Soc. 1988, 110, 6463–6466. 10.1021/ja00227a028. [DOI] [Google Scholar]; b Arduengo A. J. III; Harlow R. L.; Kline M. A stable crystalline carbene. J. Am. Chem. Soc. 1991, 113, 361–363. 10.1021/ja00001a054. [DOI] [Google Scholar]
- Arduengo A. J. III; Goerlich J. R.; Marshall W. J. A Stable thiazol-2-ylidene and its dimer. Liebigs Ann. 1997, 1997, 365–374. 10.1002/jlac.199719970213. [DOI] [Google Scholar]
- a Korotkikh N. I.; Glinyanaya N. V.; Cowley A. H.; Moore J. A.; Knishevitsky A. V.; Pekhtereva T. M.; Shvaika O. P. Tandem transformations of 1,2,4-triazol-5-ylidenes into 5-amidino-1, 2,4-triazoles. ARKIVOC 2008, 2007, 156. 10.3998/ark.5550190.0008.g17. [DOI] [Google Scholar]; b Korotkikh N. I.; Rayenko G. F.; Shvaika O. P.; Pekhtereva T. M.; Cowley A. H.; Jones J. N.; Macdonald C. L. Synthesis of 1, 2,4-triazol-5-ylidenes and their interaction with acetonitrile and chalcogens. J. Org. Chem. 2003, 68, 5762–5765. 10.1021/jo034234n. [DOI] [PubMed] [Google Scholar]; c Glinyanaya N. V.; Saberov V. S.; Korotkikh N. I.; Cowley A. H.; Butorac R. R.; Evans D. A.; Pekhtereva T. M.; Popov A. F.; Shvaika O. P. Syntheses of sterically shielded stable carbenes of the 1,2,4-triazole series and their corresponding palladium complexes: efficient catalysts for chloroarene hydrodechlorination. Dalton Trans. 2014, 43, 16227–16237. 10.1039/C4DT01353K. [DOI] [PubMed] [Google Scholar]; d Marchenko A. P.; Koidan H. N.; Zarudnitskii E. V.; Hurieva A. N.; Kirilchuk A. A.; Yurchenko A. A.; Biffis A.; Kostyuk A. N. Stable N-phosphorylated 1,2,4-triazol-5-ylidenes: novel ligands for metal complexes. Organometallics 2012, 31, 8257–8264. 10.1021/om300872g. [DOI] [Google Scholar]
- Greater variation is observed in the kDO value for 7k owing to the limited datapoints in the small region of unit slope. This is unavoidable as a result of competition from alternative pathways for H/D-exchange. Clearly all the observed first-order rate constants, kex, for 7k are consistently higher than for 8k and 9k.
- Covington A.; Robinson R.; Bates R. G. The ionization constant of deuterium oxide from 5 to 50. J. Phys. Chem. 1966, 70, 3820–3824. 10.1021/j100884a011. [DOI] [Google Scholar]
- Amyes T. L.; Diver S. T.; Richard J. P.; Rivas F. M.; Toth K. Formation and stability of N-heterocyclic carbenes in water: the carbon acid pKa of imidazolium cations in aqueous solution. J. Am. Chem. Soc. 2004, 126, 4366–4374. 10.1021/ja039890j. [DOI] [PubMed] [Google Scholar]
- van Der Helm M. P.; Klemm B.; Eelkema R. Organocatalysis in aqueous media. Nat. Rev. Chem. 2019, 3, 491–508. 10.1038/s41570-019-0116-0. [DOI] [Google Scholar]
- a Hansch C.; Leo A.; Taft R. A survey of Hammett substituent constants and resonance and field parameters. Chem. Rev. 1991, 91, 165–195. 10.1021/cr00002a004. [DOI] [Google Scholar]; b Korenaga T.; Kadowaki K.; Ema T.; Sakai T. Reestimation of the Taft’s substituent constant of the pentafluorophenyl group. J. Org. Chem. 2004, 69, 7340–7343. 10.1021/jo048939g. [DOI] [PubMed] [Google Scholar]
- Tyler A. R.; Ragbirsingh R.; McMonagle C. J.; Waddell P. G.; Heaps S. E.; Steed J. W.; Thaw P.; Hall M. J.; Probert M. R. Encapsulated nanodroplet crystallization of organic-soluble small molecules. Chem. 2020, 6, 1755–1765. 10.1016/j.chempr.2020.04.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Petersson G. A.; Nakatsuji H.; Li X.; Caricato M.; Marenich A. V.; Bloino J.; Janesko B. G.; Gomperts R.; Mennucci B.; Hratchian H. P.; Ortiz J. V.; Izmaylov A. F.; Sonnenberg J. L.; Williams-Young D.; Ding F.; Lipparini F.; Egidi F.; Goings J.; Peng B.; Petrone A.; Henderson T.; Ranasinghe D.; Zakrzewski V. G.; 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 J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M. J.; Heyd J. J.; Brothers E. N.; Kudin K. N.; Staroverov V. N.; Keith T. A.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A. P.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Millam J. M.; Klene M.; Adamo C.; Cammi R.; Ochterski J. W.; Martin R. L.; Morokuma K.; Farkas O.; Foresman J. B.; Fox D. J.. Gaussian Inc.: Wallingford, CT, 2016.; b Lee C.; Yang W.; Parr R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785. 10.1103/PhysRevB.37.785. [DOI] [PubMed] [Google Scholar]; c Zhao Y.; Truhlar D. G. 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 functionals. Theor. Chem. Acc. 2008, 120, 215–241. 10.1007/s00214-007-0310-x. [DOI] [Google Scholar]; d Tomasi J.; Mennucci B.; Cammi R. Quantum mechanical continuum solvation models. Chem. Rev. 2005, 105, 2999–3094. 10.1021/cr9904009. [DOI] [PubMed] [Google Scholar]; e Krishnan R.; Binkley J. S.; Seeger R.; Pople J. A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Phys. Chem. 1980, 72, 650–654. 10.1063/1.438955. [DOI] [Google Scholar]; f Clark T.; Chandrasekhar J.; Spitznagel G. W.; Schleyer P. V. R. Efficient diffuse function-augmented basis sets for anion calculations. III. The 3-21+ G basis set for first-row elements, Li–F. J. Comput. Chem. 1983, 4, 294–301. 10.1002/jcc.540040303. [DOI] [Google Scholar]; g Li Y.; Zhang Z. A DFT study on NHC-catalyzed [4 + 2] annulation of 2H-azirines with ketones: Mechanism and selectivity. Int. J. Quantum Chem. 2021, 121, e26557. 10.1002/qua.26557. [DOI] [Google Scholar]; h Pause L.; Robert M.; Heinicke J.; Kühl O. Radical anions of carbenes and carbene homologues. DFT study and preliminary experimental results. J. Chem. Soc., Perkin Trans. 2. 2001, 8, 1383–1388. 10.1039/b101228m. [DOI] [Google Scholar]; i Wang Y.; Wei D.; Zhang W. Recent advances on computational investigations of N-heterocyclic carbene catalyzed cycloaddition/annulation reactions: mechanism and origin of selectivities. ChemCatChem. 2018, 10, 338. 10.1002/cctc.201701119. [DOI] [Google Scholar]
- Enders D.; Breuer K.; Raabe G.; Runsink J.; Teles J. H.; Melder J. P.; Ebel K.; Brode S. Darstellung, struktur und reaktivität von 1,3,4-triphenyl-4,5-dihydro-1H-1,2,4-triazol-5-yliden, einem neuen stabilen carben. Angew. Chem., Int. Ed. 1995, 107, 1119–1122. 10.1002/ange.19951070926. [DOI] [Google Scholar]
- For calculation of ring strain energies of cyclopentene, cyclohexene, and cycloheptene using a homodesmotic computational approach please see:; Khoury P. R.; Goddard J. D.; Tam W. Ring strain energies: substituted rings, norbornanes, norbornenes and norbornadienes. Tetrahedron 2004, 60, 8103–8112. 10.1016/j.tet.2004.06.100. [DOI] [Google Scholar]; The ring strain energies (RSEs) of cyclopentene, cyclohexene, and cycloheptene follow the order 6 < 5∼7, and this order is invariant with the computational method employed. This order does not correlate with our observed protofugality (kinetic acidity) order of 5 > 6∼7; thus, ring strain cannot directly provide an explanation for the origin of the trends observed.
- a Wan Y.; Xin H.; Chen X.; Xu H.; Wu H. 1, 3-Bis (4-methoxyphenyl) imidazolidium chloride monohydrate. Acta Crystallogr. Sect. D-Biol. Crystallogr. 2008, 64, o2159–o2159. 10.1107/S1600536808033965. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Garden S. J.; Gama P. E.; Tiekink E. R.; Wardell J. L.; Wardell S. M.; Howie R. A. 1, 3-Bis (4-bromophenyl) imidazolium chloride dihydrate. Acta Crystallogr. Sect. D-Biol. Crystallogr. 2010, 66, o1438–o1439. 10.1107/S1600536810018581. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Arduengo A. J. III; Dias H. R.; Harlow R. L.; Kline M. Electronic stabilization of nucleophilic carbenes. J. Am. Chem. Soc. 1992, 114, 5530–5534. 10.1021/ja00040a007. [DOI] [Google Scholar]; d Sioriki E.; Lordan R.; Nahra F.; Van Hecke K.; Zabetakis I.; Nolan S. P. In vitro anti-atherogenic properties of N-heterocyclic carbene aurate (I) compounds. ChemMedChem. 2018, 13, 2484. 10.1002/cmdc.201800601. [DOI] [PubMed] [Google Scholar]; e Arduengo A. J. III; Calabrese J.; Davidson F.; Rasika Dias H.; Goerlich J. R.; Krafczyk R.; Marshall W. J.; Tamm M.; Schmutzler R. C–H insertion reactions of nucleophilic carbenes. Helv. Chim. Acta 1999, 82, 2348–2364. . [DOI] [Google Scholar]; f Alder R. W.; Blake M. E.; Bufali S.; Butts C. P.; Orpen A. G.; Schütz J.; Williams S. J. Preparation of tetraalkylformamidinium salts and related species as precursors to stable carbenes. J. Chem. Soc., Perkin Trans. 1 2001, 1586–1593. 10.1039/b104110j. [DOI] [Google Scholar]; g Alder R. W.; Blake M. E.; Bortolotti C.; Bufali S.; Butts C. P.; Linehan E.; Oliva J. M.; Guy Orpen A.; Quayle M. J. Complexation of stable carbenes with alkali metals. Chem. Commun. 1999, 241–242. 10.1039/a808951e. [DOI] [Google Scholar]; h Wallbaum L.; Weismann D.; Löber D.; Bruhn C.; Prochnow P.; Bandow J. E.; Siemeling U. Stable and persistent acyclic diaminocarbenes with cycloalkyl substituents and their transformation to β-lactams by uncatalysed carbonylation with CO. Chem.—Eur. J. 2019, 25, 1488–1497. 10.1002/chem.201805307. [DOI] [PubMed] [Google Scholar]
- a Kirmse W. In Advances in carbene chemistry; Brinker U. H., Ed.; Elsevier, 2000; Vol. 3, pp 1–52. [Google Scholar]; b Magill A. M.; Cavell K. J.; Yates B. F. Basicity of nucleophilic carbenes in aqueous and nonaqueous solvents theoretical predictions. J. Am. Chem. Soc. 2004, 126, 8717–8724. 10.1021/ja038973x. [DOI] [PubMed] [Google Scholar]; c Stein T. H.; Vasiliu M.; Arduengo A. J. III; Dixon D. A. Lewis acidity and basicity: another measure of carbene reactivity. J. Phys. Chem. A 2020, 124, 6096–6103. 10.1021/acs.jpca.0c03877. [DOI] [PubMed] [Google Scholar]; d Liu M.; Yang I.; Buckley B.; Lee J. K. Proton affinities of phosphines versus N-heterocyclic carbenes. Org. Lett. 2010, 12, 4764–4767. 10.1021/ol101864p. [DOI] [PubMed] [Google Scholar]; e Liu M.; Chen M.; Zhang S.; Yang I.; Buckley B.; Lee J. K. Reactivity of carbene• phosphine dimers: proton affinity revisited. J. Phys. Org. Chem. 2011, 24, 929–936. 10.1002/poc.1890. [DOI] [Google Scholar]; f Gronert S.; Keeffe J. R.; More O’Ferrall R. A. Stabilities of carbenes: independent measures for singlets and triplets. J. Am. Chem. Soc. 2011, 133, 3381–3389. 10.1021/ja1071493. [DOI] [PubMed] [Google Scholar]
- NBO computational analysis (section S2.6) revealed a small decrease in s-character from n = 1 to n = 3 for triazolium salts 7-9c, however, the differences are not significantly outside the error of the calculations.
- Anslyn E. V.; Dougherty D. A. In Modern physical organic chemistry; 2006; p 91. [Google Scholar]
- a Konstandaras N.; Dunn M. H.; Guerry M. S.; Barnett C. D.; Cole M. L.; Harper J. B. The impact of cation structure upon the acidity of triazolium salts in dimethyl sulfoxide. Org. Biomol. Chem. 2020, 18, 66–75. 10.1039/C9OB02258A. [DOI] [PubMed] [Google Scholar]; b Wang N.; Xu J.; Lee J. K. The importance of N-heterocyclic carbene basicity in organocatalysis. Org. Biomol. Chem. 2018, 16, 8230–8244. 10.1039/C8OB01667D. [DOI] [PubMed] [Google Scholar]; c Paul M.; Detmar E.; Schlangen M.; Breugst M.; Neudörfl J. M.; Schwarz H.; Berkessel A.; Schäfer M. Intermediates of N-heterocyclic carbene (NHC) dimerization probed in the gas phase by ion mobility mass spectrometry: C–H···:C hydrogen bonding versus covalent dimer formation. Chem.—Eur. J. 2019, 25, 2511–2518. 10.1002/chem.201803641. [DOI] [PubMed] [Google Scholar]; d Kim Y.-J.; Streitwieser A. Basicity of a stable carbene, 1,3-di-tert-butylimidazol-2-ylidene, in THF1. J. Am. Chem. Soc. 2002, 124, 5757–5761. 10.1021/ja025628j. [DOI] [PubMed] [Google Scholar]; e Nelson D. J.; Nolan S. P. Quantifying and understanding the electronic properties of N-heterocyclic carbenes. Chem. Soc. Rev. 2013, 42, 6723–6753. 10.1039/c3cs60146c. [DOI] [PubMed] [Google Scholar]; f Alder R. W.; Allen P. R.; Williams S. J. Stable carbenes as strong bases. J. Chem. Soc., Chem. Commun. 1995, 1267–1268. 10.1039/c39950001267. [DOI] [Google Scholar]; g Chu Y.; Deng H.; Cheng J.-P. An acidity scale of 1,3-dialkylimidazolium salts in dimethyl sulfoxide solution. J. Org. Chem. 2007, 72, 7790–7793. 10.1021/jo070973i. [DOI] [PubMed] [Google Scholar]; h Li Z.; Li X.; Cheng J.-P. An acidity scale of triazolium-based NHC precursors in DMSO. J. Org. Chem. 2017, 82, 9675–9681. 10.1021/acs.joc.7b01755. [DOI] [PubMed] [Google Scholar]; i Dunn M. H.; Konstandaras N.; Cole M. L.; Harper J. B. Targeted and systematic approach to the study of pKa values of imidazolium salts in dimethyl sulfoxide. J. Org. Chem. 2017, 82, 7324–7331. 10.1021/acs.joc.7b00716. [DOI] [PubMed] [Google Scholar]
- a Eigen M. Proton transfer, acid-base catalysis, and enzymatic hydrolysis. Part I: elementary processes. Angew. Chem., Int. Ed. 1964, 3, 1–19. 10.1002/anie.196400011. [DOI] [Google Scholar]; b Keeffe J. R.; Kresge A. J. I. In Chemistry of Enols; Rappoport Z., Ed.; J. Wiley, 1990; pp 399–480. [Google Scholar]; c Amyes T. L.; Richard J. P. Generation and stability of a simple thiol ester enolate in aqueous solution. J. Am. Chem. Soc. 1992, 114, 10297–10302. 10.1021/ja00052a028. [DOI] [Google Scholar]; d Amyes T. L.; Richard J. P. Determination of the pKa of ethyl acetate: Brønsted correlation for deprotonation of a simple oxygen ester in aqueous solution. J. Am. Chem. Soc. 1996, 118, 3129–3141. 10.1021/ja953664v. [DOI] [Google Scholar]; e Richard J. P.; Williams G.; Gao J. Experimental and computational determination of the effect of the cyano group on carbon acidity in water. J. Am. Chem. Soc. 1999, 121, 715–726. 10.1021/ja982692l. [DOI] [Google Scholar]; f Richard J. P.; Amyes T. L.; Toteva M. M. Formation and stability of carbocations and carbanions in water and intrinsic barriers to their reactions. Acc. Chem. Res. 2001, 34, 981–988. 10.1021/ar0000556. [DOI] [PubMed] [Google Scholar]; g Richard J. P.; Williams G.; O’Donoghue A. C.; Amyes T. L. Formation and stability of enolates of acetamide and acetate anion: An Eigen plot for proton transfer at α-carbonyl carbon. J. Am. Chem. Soc. 2002, 124, 2957–2968. 10.1021/ja0125321. [DOI] [PubMed] [Google Scholar]
- Peon J.; Polshakov D.; Kohler B. Solvent reorganization controls the rate of proton transfer from neat alcohol solvents to singlet diphenylcarbene. J. Am. Chem. Soc. 2002, 124, 6428–6438. 10.1021/ja017485r. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.













