Abstract
Nitrogen heterocycles and aromatic amines are well-known and widely used compounds that are usually not seen as acids, although in organic solvents like acetonitrile (MeCN) or dimethyl sulfoxide (DMSO) their acidic properties can be observed. In this work, the acidities (pKa values) of 37 such weak NH acids were determined in acetonitrile and presented together with computational gas-phase acidities and literature pKa values in DMSO. In the course of the work the weakest acids range (from pKa 29 upward) of the MeCN acidity scale has been revised and expanded to around 33.5 by adding 31 compounds in that specific region and the span of experimental pKa values in MeCN is now more than 30 orders of magnitude. The relations between the structure and acidity of a selection of the studied compounds have been investigated in MeCN and DMSO. The measured pKa values in MeCN and the gathered pKa values in DMSO were used for a correlation analysis between the two solvents and for assessing the quality of pKa conversion equations. A number of pKa values have been predicted in MeCN from pKa values in DMSO via the correlation analysis and pKa conversion equations.
Keywords: acidity, pKa, nitrogen heterocycles, aromatic amines, acetonitrile, gas-phase acidity, UV−vis spectroscopy
Introduction
In most chemical processes, the acidities or basicities of the involved compounds are essential, particularly when charged species are involved. The most widely used representation of the acidity of a molecule in the liquid phase is based on the Brønsted theory1 and is quantitatively expressed by the equilibrium constant Ka of the acid dissociation reaction in solvent S (eq 1) or, more often, its negative logarithm pKa (eq 2).
| 1 |
| 2 |
When dealing with chemical reactions in the presence of strong bases, in order to control the reaction path and outcomes it is important to understand which of the solution components can get deprotonated by the chosen base. For that, knowing the pKa values of even very weak acids is crucial. As the majority of such reactions are carried out in nonaqueous media, it is the pKa values in organic solvents such as acetonitrile (MeCN) and dimethyl sulfoxide (DMSO) that are relevant.
One significant type of acids are NH acids, in which the acidic proton is directly attached to the nitrogen atom. NH acids cover a wide range of pKa values, from superacidic catalysts2 to the weakly acidic indole and its derivatives.3−5 In MeCN, the pKa values of numerous superacidic NH have been reported,6 and the Bordwell group has published the pKa values of numerous weak NH acids in DMSO.3,7 However, to the best of our knowledge, only very limited information is available regarding the acidities of very weak acids (pKa > 30) in MeCN.
Prominent classes of weak acids lacking experimental pKa data are conjugated nitrogen heterocycles and aromatic amines - compounds that are usually better known as bases. Compounds such as benzimidazole, benzotriazole, indazole, carbazole, azaindoles and especially their derivatives have numerous applications: these moieties are found in pharmaceuticals,8 pesticides, bioactive compounds,9 etc. Nitrogen heterocycles are one of the most significant structural fragments in pharmaceuticals10 as their derivatives exhibit significant anti-inflammatory, antiviral, antihistaminic, antiparasitic, antifungal and anticancer activity.11−13 Despite the available acidities in DMSO, their pKa values in a less basic solvent like MeCN are also of interest. Yet, no such data have been published. Some acids of interest also have no pKa values in DMSO. For instance, to the best of our knowledge, the acidities of azaindoles and benzocarbazoles have not been measured in any nonaqueous solvents before.
Besides the above-mentioned fields of application, weak NH acids are well-suited additions to the previously published comprehensive self-consistent acidity scale in acetonitrile, which contains pKa values of 231 acids and covers close to 30 orders of magnitude of acidity.6 Despite the substantial number of acids in it, the pKa scale in MeCN is scarce in compounds in the region of the weakest acids. It contains only five acids with pKa values between 29 and 32.57 (the highest pKa value on the scale). For comparison, between pKa values of 25.4 and 29, there are 34 acids on the scale. The available acidity data in MeCN show that the pKa values of aromatic amines, such as substituted diphenylamines and anilines, can vary over a wide range depending on the number and properties of the substituents.6 Thus, aromatic amines with a suitable structure and substituent groups should have pKa values in the desired region.
MeCN as a solvent is generally not considered as suitable as DMSO for studying very weak acids because of its lower basicity.14 Nevertheless, numerous acids are expected to have pKa values above 29 in MeCN, enabling supplementing and expanding the current acidity scale in MeCN. As the highest pKaH (pKa value of the conjugated acid of a base)14 value of the MeCN basicity scale is 33.14,15 extending the acidity scale to at least a similarly high pKa value should be possible.
The aim of this study was to present a number of new previously unavailable pKa values of weak NH acids in MeCN and to expand the pKa scale in MeCN. In addition, this study aims to evaluate the previously published simple conversion equations for pKa estimations between DMSO and MeCN,6 specifically for weak NH acids, and to predict a number of additional pKa values in MeCN. Also, computational gas-phase acidity (GA) values were calculated for all the compounds studied in this work.
Results and Discussion
Expanded pKa Scale in Acetonitrile
The pKa values of 37 weak NH acids, mainly conjugated nitrogen heterocycles and aromatic amines, were determined in MeCN by 141 relative acidity measurements. As a result, the least acidic section of the MeCN pKa scale,6 with pKa values over 29, was rebuilt and expanded by adding 31 new compounds. Consequently, the highest pKa value on the scale is now 33.51 belonging to 2,3,5,6-Cl4-aniline. The previously published pKa values for indole (32.57 → 32.78), 3-MeCOO-indole (29.97 → 30.2) and 7-NO2-indole (29.99 → 30.12) were revised. The results are presented in Table 1. The pKa values of the aromatic amines presented in this work include some of the highest ever experimentally determined pKa values of neutral NH acids in MeCN. In addition to the results in Table 1, the pKa value of benzotriazole in MeCN was determined to be 22.98 (see Table S1 in the SI).
Table 1. Acidity Measurements and Assigned pKa Values of Weak NH Acids in MeCNa.

Compounds with a gray background serve as anchor compounds with fixed pKa values from ref (6) Previously published pKa values (from ref (6)) are given for indole, 3-MeCOO-indole and 7-NO2-indole in parentheses for comparison. Black, brown and red arrows denote high-, medium- and low-reliabiliy ΔpKa measurements, respectively, depending on the type of acids. See text for more details.
The pKa scale in Table 1 was built by performing relative acidity (relative pKa or ΔpKa) measurements. Each double-headed arrow corresponds to a ΔpKa determination between a pair of acids (described in more detail in the Experimental Section and SI). Absolute pKa values of the analyzed acids were assigned using the “ladder” approach,6,16 i.e. the following least-squares minimization procedure with the experimentally determined relative acidities.
| 3 |
According to eq 3, the sum of squares of the differences between all the experimentally determined relative acidities ΔpKa and the differences between the assigned absolute pKa values of acids HA2 and HA1 over all the relative acidity determination series nm was minimized. If absolute pKa values are desired as a result of the minimization then the ladder has to contain at least one reference acid (anchor acid) with a previously known and reliable pKa value. This pKa value is kept constant while the pKa values of the remaining compounds are varied until the minimal SSD is reached. The acids tBu4Box2CH2,17 4–CN–2,3,5,6–F4–aniline, 9–C6F5–fluorene, C6H5–CH2–SO2CF3 and (4–Me–C6F4)(C6H5)CHCN have known reliable pKa values6 and were used as reference acids in the least-squares minimization process.
Before the minimization, the analyzed acids were divided into two groups – “backbone” and “secondary” acids–according to the reliability of the relative acidity measurements conducted with them. The acids with the most convenient properties that enable high-quality ΔpKa measurements were used as the backbone acids. As described in a previous publication, the reliability of the measurements was evaluated by considering the spectral properties of the acids, consistency of values obtained within a measurement series, self-consistency of the ΔpKa values obtained for a given compound against different reference compounds and the possibility of unwanted side processes.6 Aromatic amines were chosen as the “backbone” acids because of their very suitable spectral properties, good within-series agreement, and good consistency of measurements. As a rule, nitrogen heterocycles displayed somewhat worse performance in pKa measurements and were categorized as “secondary” acids. The third category consists of indolo[2,3-a]carbazole and 3-MeCOO-indole. Even after additional purification, the relative acidity measurements of these compounds were found to be less reliable than those of the other studied compounds. This could be caused by possible side reactions or the presence of impurities, which influenced the UV–vis spectra of these compounds during the spectrophotometric titration in such a way that the self-consistency and accuracy of the experimental results were lowered.
In total, three minimization steps were performed to assign pKa values for all analyzed compounds. All ΔpKa measurements involving “backbone” acids are presented as black-colored double-sided arrows in Table 1. The arrows of “secondary” acids are brown, and those involving indolo[2,3-a]carbazole or 3-MeCOO-indole are red.
For the first minimization step, only the relative pKa values between aromatic amines and relative pKa values between aromatic amines and the above-mentioned reference acids were used. As a result of this minimization, pKa values for the “backbone” acids were assigned. In the second minimization step the acidities of the “backbone” acids were kept constant and used as reference to obtain the pKa values for the “secondary” acids. The third minimization was performed to obtain the pKa values of 3-MeCOO-indole and indolo[2,3-a]carbazole using all other compounds, which had pKa values assigned in the previous steps, as reference acids.
The quality of a pKa scale formed by the relative pKa determination approach can be evaluated using the consistency standard deviation s of the scale, according to the following equation:
| 4 |
In eq 4, SSD is the sum of squares found in eq 3, nm is the number of measurement series, and nc is the number of compounds for which an absolute pKa value was assigned. Strictly speaking, the parameter s is not interpretable as the uncertainty of an individual pKa value in the scale. However, it serves as a general estimate of the quality of the pKa values forming the acidity scale, can be loosely interpreted as the average standard uncertainty of the pKa values with respect to the scale, and can be used for comparing the strengths of the acids belonging to the scale. The consistency standard deviation of the ΔpKa measurements involving aromatic amines (“backbone” acids) is 0.03. It is the same consistency as that of the “backbone” acids forming the overall pKa scale in MeCN spanning almost 30 orders of magnitude and can be rated as good.6 The s value of the ΔpKa measurements involving the “secondary” acids is 0.06 and the s value of the ΔpKa measurements involving indolo[2,3-a]carbazole or 3-MeCOO-indole is 0.09, which can both be considered satisfactory.
The acidity scale published in 2021 by Kütt et al.6 was scarce in compounds with pKa values over 30. The compounds indole, 3-MeCOO-indole and 7-NO2-indole were already present in that scale. However, because of the lack of other compounds with similar pKa values, measurements of considerable pKa differences (ΔpKa values around or above 2) were used to assign pKa values for these compounds. Usually, experimental ΔpKa values higher than 1.5 are less accurate than lower ΔpKa values because of the narrow (or nonexistent) range within which both measured acids have accurately measurable degrees of dissociation (i.e., degrees of dissociation between 0.1 and 0.9). Thus, such results need confirmatory measurements. The present work adds a large number of acids to the pKa range of 30 to 33, completely eliminating the need for measuring large ΔpKa values. The results from the present work indicate that such higher than recommended ΔpKa measurements caused a contraction of the upper part of the acidity scale. The following four earlier experimental relative acidity measurements6 seem to be the main cause of this effect (directly measured ΔpKa values in parentheses): 3-MeCOO-indole vs 5-NO2-indole (ΔpKa = 1.82), 7-NO2-indole vs 9-C6F5-fluorene (ΔpKa = 1.88), 7-NO2-indole vs 6-NO2-indole (ΔpKa = 2.21) and 7-NO2-indole vs 5-NO2-indole (ΔpKa = 1.74). The corresponding ΔpKa values calculated from the results of this work (Table 1) are systematically higher by 0.1–0.2 pKa units. Thus, these four ΔpKa measurements were considered unreliable and were excluded from the data analysis. The previous6 directly measured ΔpKa values for the acid pairs indole vs 3-MeCOO-indole (ΔpKa = 2.58) and indole vs 7-NO2-indole (ΔpKa = 2.61) are in agreement with the absolute pKa values in the scale presented in Table 1. Nevertheless, these two previous ΔpKa measurements were also considered doubtful (in part because of the very high ΔpKa value–far above the reliable directly measured ΔpKa range of the used method of up to approximately 1.5 pKa units) and excluded from the data analysis.
Comparison of Acidities between Different Solvents and the Gas Phase
The experimental pKa values in DMSO and H2O, as well as the gas-phase acidity (GA) values (found in literature and/or calculated) are presented in Table 2 together with the pKa values determined in this work. The pKa values in DMSO and H2O have been published for less than half of the compounds of the present work. In the case of water, the obvious reason for the scarcity is that the aqueous pKa values of the majority of the investigated compounds are outside the common aqueous pKa range. In the case of DMSO, all the values are well within the experimentally achievable pKa range in this solvent. Data from the Bordwell3,7 group were preferred if pKa values from multiple sources were available. The available experimental gas-phase acidity (GA) data was even more scarce than in either of the two solvents. The GA values of only seven studied compounds were found in the NIST Chemistry WebBook,18 so computational GA values are also provided in Table 2. The root-mean-square difference between the available experimental GA values and their computational counterparts is 10 kJ mol–1 and the maximum difference is 15 kJ mol–1 (for comparison, a typical uncertainty of GA values reported in the NIST database is ±8.4 kJ mol–1). Most of the computational values are higher than the experimental values, displaying an average bias of 8 kJ mol–1. The possible reason for these comparatively high discrepancies could be a compound class-specific bias that some methods exhibit, method-specific systematic bias (which, in turn, could originate from overestimation of vibrational frequencies),19 or the interplay between the two.
Table 2. Acidities in Different Solvents and the Gas Phase.
| Compound | CAS RN | pKa (MeCN)a | pKa (DMSO) | pKa (H2O) | Exp. GA,b kJ mol-1 | Calc. GA,c kJ mol-1 |
|---|---|---|---|---|---|---|
| 2,3,5,6-Cl4-Aniline | 3481–20–7 | 33.51 | 21.020 | 19.2221 | 1426 | |
| 4-CF3SO2-Aniline | 473–27–8 | 33.11 | 21.87 | 1400 | ||
| 2,3,4,5,6-Cl5-Aniline | 527–20–8 | 32.79 | 1415 | |||
| Indole | 120–72–9 | 32.78 | 20.953 | 16.9722 | 1431;23 144024 | 1441 |
| 2-NO2-Aniline | 88–74–4 | 32.63 | 17.725 | 1436 | ||
| N-Me-4-NO2-Aniline | 100–15–2 | 32.62 | 18.225 | 1421 | ||
| 4-NO2-Aniline | 100–01–6 | 32.58 | 20.97 | 18.225 | 140726 | 1421 |
| (2-NO2–C6H4)(Ph)NHg | 119–75–5 | 31.79 | 19.2d (17.727) | 18.025 | 1421 | |
| 2-MeO-Carbazole | 6933–49–9 | 31.78 | 1417 | |||
| Carbazole | 86–74–8 | 31.74 | 19.93 | 16.728 | 141223 | 1420 |
| 2-NH2-4-NO2-Aniline | 99–56–9 | 31.75 | 1416 | |||
| 3-Cl-4-NO2-Aniline | 825–41–2 | 31.53 | 1403 | |||
| 4-Cl-2-NO2-Aniline | 89–63–4 | 31.06 | 18.93 | 17.121 | 1412 | |
| 5-Cl-2-NO2-Aniline | 1635–61–6 | 31.03 | 1411 | |||
| 1,3-Ph2-Urea | 102–07–8 | 30.96 | 19.63 | 1396 | ||
| (2-NO2–C6H4)(4–Cl-C6H4)NH | 23008–56–2 | 30.92 | 1402 | |||
| 2-Cl-6-NO2-Aniline | 769–11–9 | 30.84 | 1423 | |||
| 7-Azaindole | 271–63–6 | 30.79 | 12.129,f | 1437 | ||
| 4-Azaindolee | 272–49–1 | 30.49 | 15.530,e; 16.1,30,e | 1422 | ||
| 2-Cl-4-NO2-Aniline | 121–87–9 | 30.38 | 18.0631 | 1398 | ||
| Benzo[c]carbazole | 205–25–4 | 30.33 | 1401 | |||
| 3-MeCOO-Indole | 608–08–2 | 30.2 | 1427 | |||
| Benzo[a]carbazole | 239–01–0 | 30.20 | 1400 | |||
| 7-NO2–Indole | 6960–42–5 | 30.12 | 1415 | |||
| 4-NH2–C5F4N | 1682–20–8 | 30.12 | 18.7d (19.232) | 139233 | 1407 | |
| 5-Azaindole | 271–34–1 | 30.06 | 1414 | |||
| Indolo[2,3-a]carbazole | 60511–85–5 | 29.9 | 1387 | |||
| (5-Cl-2-NO2–C6H3)(Ph)NH | 25781–92–4 | 29.88 | 1397 | |||
| Indazole | 271–44–3 | 29.79 | 18.228 | 13.8628 | 142534 | 1433 |
| 6-Azaindole | 271–29–4 | 29.75 | 1413 | |||
| 2,3-Cl2-6-NO2-Aniline | 65078–77–5 | 29.43 | 1403 | |||
| Norharman | 244–63–3 | 29.33 | 14.5335 | 1398 | ||
| 2,4-Cl2-6-NO2-Aniline | 2683–43–4 | 29.26 | 16.3931 | 1401 | ||
| 2,5-Cl2-4-NO2-Aniline | 6627–34–5 | 29.20 | 17.47 | 16.0536 | 1380 | |
| (4-NO2–C6H4)(Ph)NHg | 836–30–6 | 28.87 | 16.857 | 15.625 | 137433,g | 1384 |
| 4-NO2-1-Naphthalenamine | 776–34–1 | 28.85 | 17.4d (18.037) | 1385 | ||
| 2,6-Cl2-4-NO2-aniline | 99–30–9 | 28.30 | 15.5536 | 1387 | ||
| Benzimidazole | 51–17–2 | 27.92 | 16.43 | 12.8638 | 1403 | |
| 2,4-(NO2)2-Aniline | 97–02–9 | 27.64 | 15.97 | 15.036 | 1367 | |
| 1H-Benzotriazole | 95–14–7 | 22.98 | 11.923 | 8.5739 | 138434 | 1383 |
Experimental GA values.
Computational GA values (this work).
Original values from the respective publications (shown in parentheses) were corrected in ref (20)
In the abstract and the main text of ref (30) two different pKa(H2O) values were presented for 4-azaindole and it was not possible to tell which one is the correct one.
Doubtful value.
The GA value of 1374 kJ mol–1, actually belonging to (4-NO2–C6H4)(Ph)NH, is erroneously assigned to (2-NO2–C6H4)(Ph)NH in the NIST Webbook.
Fifteen compounds studied in this work had pKa(DMSO) values available in the literature (ten aromatic amines, four nitrogen heterocycles and 1,3-diphenylurea). Using these values and the respective pKa(MeCN) values determined in the present work, a correlation between acidities in the two solvents was composed (Figure 1). Benzotriazole was excluded from the correlation because of its almost 5 orders of magnitude higher acidity compared to the other compounds. The following correlation equation was obtained:
| 5 |
Figure 1.
Correlation between pKa values in MeCN and DMSO.
Figure 1 shows that the acidities of the selection of NH acids correlate reasonably well between DMSO and MeCN without any strongly deviating compounds. Solvent–solvent correlations are expected to be good if the solvents have sufficiently similar properties and the involved acids are of a similar nature.14 Both requirements are fulfilled in this case. Better correlations could be obtained for heterocycles and aromatic amines separately. The adequacy of the regression equation is evidenced by the fact that if the pKa of benzotriazole in MeCN is predicted from it, the value will be 23.6, which is approximately 0.5 pKa units different from the experiment. Considering the significant extrapolation involved, the agreement can be considered good.
Equations for pKa Conversions between Solvents
Kütt et al.6 developed several equations to predict the pKa values of acids in different nonaqueous solvents based on pKa values in MeCN and simple structure-based descriptors. Equations 6 (eq 2.3 in Table 2 in ref (6)) and 7 (eq 2.1 in Table 2 in ref (6)) are the universal and NH acid-specific equations, respectively, to convert pKa(MeCN) values to pKa(DMSO) values. These equations have a root-mean-square error (RMSE) of prediction of 0.5 and 1.2, respectively.
| 6 |
| 7 |
The pKa values measured in this work offer a convenient possibility to assess the quality of these equations, as most compounds in this work (a) were not used in the training data set for developing the equations and (b) have higher pKa values than the majority of the training set compounds. Thus, they serve as a good and demanding test set. The accuracy of pKa(DMSO) values from the literature was critically evaluated, and the corrected pKa(DMSO) values of (2-NO2–Ph)(Ph)NH, 4-NH2–C5F4N, 4-NO2-1-Naphthalenamine were used (see Table 2).
The descriptors used in eqs 6 and 7 are the number of hydrogen bond donors (nHBD), the number of hydrogen atoms or sulfonyl groups attached directly to the acidity center (X-H, X-SO2), and the number of nitrogen atoms in the molecule (nN). The equations were rearranged in such a way as to predict the pKa values in MeCN from values in DMSO and the predicted values were compared to the experimental pKa values determined in this paper. The results are presented in Table 3.
Table 3. pKa(MeCN) Values Predicted Using Conversion Equations (Eqs 6 and 7)a.
| Prediction
from pKa(DMSO) |
|||||
|---|---|---|---|---|---|
| Compound | pKa(MeCN)d | NHb | Difference | Allc | Difference |
| 2,3,5,6-Cl4-Aniline | 33.51 | 33.30 | –0.21 | 33.09 | –0.42 |
| 4-CF3SO2-Aniline | 33.11 | 34.10 | 0.99 | 33.94 | 0.83 |
| 4-NO2-Aniline | 32.58 | 32.90 | 0.32 | 32.98 | 0.40 |
| (2-NO2–C6H4)(Ph)NH | 31.79 | 31.20 | –0.59 | 31.49 | –0.30 |
| Carbazole | 31.74 | 32.20 | 0.46 | 32.23 | 0.49 |
| 4-Cl-2-NO2-Aniline | 31.06 | 30.90 | –0.16 | 30.85 | –0.21 |
| 1,3-Ph2-Urea | 30.96 | 31.60 | 0.64 | 30.74 | –0.22 |
| 4-NH2–C5F4N | 30.12 | 30.70 | 0.58 | 30.64 | 0.52 |
| Indazole | 29.79 | 30.20 | 0.41 | 30.43 | 0.64 |
| 2,5-Cl2-4-NO2-Aniline | 29.20 | 29.40 | 0.20 | 29.26 | 0.06 |
| (4-NO2–C6H4)(Ph)NH | 28.87 | 28.85 | –0.02 | 28.99 | 0.12 |
| 4-NO2-1-Naphthalenamine | 28.85 | 29.40 | 0.55 | 29.26 | 0.41 |
| Benzimidazole | 27.92 | 28.40 | 0.48 | 28.51 | 0.59 |
| 2,4-(NO2)2-Aniline | 27.64 | 27.60 | –0.04 | 27.66 | 0.02 |
| Benzotriazole | 22.98 | 23.62 | 0.64 | 23.74 | 0.76 |
| RMSEe | 0.49 | 0.46 | |||
The RMSE of the pKa(MeCN) prediction of the selection of weak acids is 0.49 when using the NH acid-specific and 0.46 when using the universal equation suitable for all types of acids. These RMSE values are equal to or lower than the RMSE values of pKa(MeCN) to pKa(DMSO) conversion equations in the original publication (0.5 and 1.2, respectively). This result is made even more noteworthy by considering that these equations were created using different compound classes (mainly sulfonamides and diarylamines), which are typically stronger acids than the aromatic amines and nitrogen heterocycles studied in this work.
Although similar pKa conversion equations were published in ref (6) for the water-MeCN pair, these almost always yield less accurate results than conversions between DMSO and MeCN. Aqueous pKa values from Table 2 were used to predict the pKa values of the respective acids in MeCN, but the obtained RMSE was significantly over 2, which shows that the conversion equations between water and MeCN from ref (6) are unusable for such weak acids. One reason for this large discrepancy is the much different nature of both solvents. MeCN is an aprotic solvent with very weak hydrogen bond donor (HBD) and hydrogen bond acceptor (HBA) properties, whereas water is a protic solvent with strong HBD and HBA properties, thus having the ability to specifically solvate anions. Usually, better correlations of acidities are obtained between solvents with similar properties.14 Another reason is that the majority of the aqueous pKa values of the compounds studied in this paper are outside the conventional aqueous pH scale (0–14). This means that certain approximations (regarding transferability of pKa values from highly concentrated aqueous alkali solutions or aqueous organic mixtures into dilute aqueous solutions)25 have been involved in obtaining these values, which invariably make the values less reliable.
Due to the low RMSE of prediction on the basis of pKa(DMSO) values, the acidities of a selection of well-known heterocycles were estimated in MeCN using their pKa values in DMSO. The results are presented in Table 4. To our knowledge, no experimental pKa values have yet been determined in MeCN for pyrrole, pyrazole, imidazole, 1,2,3-triazole, and 1,2,4-triazole.
Table 4. Estimated pKa Values of Some Heterocyclic Compounds.
| pKa(MeCN) estimated
from pKa(DMSO) |
||||||||
|---|---|---|---|---|---|---|---|---|
| Compound | CAS RN | pKa (DMSO)a | eq 5 | eq 6 | eq 7 | Recommended pKa (MeCN)b | Exp. GAc [kJ mol-1] | Calc. GAd [kJ mol-1] |
| Pyrrole | 109–97–7 | 23.0 | 34.9 | 35.3 | 35.5 | 35.2 ± 0.5 | 1468;40 147241 | 1482 |
| Pyrazole | 288–13–1 | 19.8 | 31.6 | 31.8 | 32.1 | 31.8 ± 0.5 | 144942 | 1463 |
| Imidazole | 288–32–4 | 18.6 | 30.4 | 30.6 | 30.9 | 30.6 ± 0.5 | 143442 | 1443 |
| 1,2,4-Triazole | 288–88–0 | 14.75 | 26.5 | 26.5 | 26.8 | 26.6 ± 0.5 | 141018 | 1419 |
| 1H-1,2,3-Triazole | 288–36–8 | 13.9 | 25.6 | 25.6 | 25.9 | 25.7 ± 0.5 | 141934 | 1412 |
All experimental values from the Bordwell group.3
Recommended pKa values based on the arithmetic mean of estimations from DMSO, uncertainty estimates refer to standard uncertainty.
Experimental GA values.
Computational GA values from this work.
Structural Effects on Acidity
Scheme 1 shows the effects of structural differences on the pKa(MeCN) values of nitrogen (N) heterocycles related to indole. The compounds indazole, benzimidazole and benzotriazole can be viewed as indole derivatives with two or three ring nitrogen atoms. Adding N atoms into the 5-membered indole ring has an acidity-enhancing effect, the extent of the effect depending on their position. Indazole (N in position 2) has by 2.99 units lower pKa value than indole, whereas the pKa of benzimidazole (N in position 3) is by 4.86 units lower. Similar acidity differences were observed in DMSO, 2.75 and 4.55 pKa units, respectively.3 The increase in acidity is caused by a combination of the inductive and resonance effects, which stabilize the anion. Although in benzimidazole, the second N is further away from the acidity center, thus having a lower inductive effect, it has a more substantial resonance effect, resulting in a lower pKa value than in the case of indazole. Benzotriazole has by 9.80 units lower pKa than indole. The third N in the five-member ring further stabilizes the anion. Interestingly, the ΔpKa of indole and benzimidazole with 4.86 is almost the same as the ΔpKa of benzimidazole and benzotriazole with 4.94. In DMSO, the acidity enhancing effect of the two additional N atoms is even larger, amounting to 11.92 units.3
Scheme 1. Relations between Structural Features and Acidities of Indole Derivatives.
Absolute pKa values (determined in this work) are given below the compound names. pKa differences between compounds are shown on the arrows.
The acidifying effect of an N atom in the six-membered ring of azaindoles is weaker due to the longer distance from the acidity center. The variation in pKa values of the azaindoles with an N atom in different positions is small, spanning around 1 pKa unit. 7-Azaindole is by 1.99, 4-azaindole by 2.29, 5-azaindole by 2.72 and 6-azaindole by 3.03 pKa units stronger acid than indole. In the case of norharman, which has by 2.41 pKa units lower pKa value than carbazole, the extent of this acidifying effect is comparable to that of azaindoles.
Comparing the pKa values of indole and carbazole reveals that the acidity-enhancing effect of a fused aromatic ring is smaller than the effect of additional N atoms, and carbazole is only by 1.04 pKa units stronger acid than indole. The effect is almost identical in DMSO with 1.05 units.3 The effect of a fused aromatic ring in 6-azaindole is still lower, and accordingly, the pKa value of norharman is only 0.42 units lower than the pKa of 6-azaindole. Somewhat surprisingly, the effect of adding a second fused benzene ring into the carbazole molecule has a slightly higher impact on the pKa value. Benzo[c]carbazole and benzo[a]carbazole are more acidic than carbazole by 1.41 and 1.54 units, respectively. Although the benzene ring fused to the c side of carbazole is further away from the acidity center than the one fused to the a side, the difference between the pKa values of these two compounds is marginal. Indolo[2,3-a]carbazole can be viewed as a carbazole with an indole fused to its a side. Indolo[2,3-a]carbazole has by 0.3 units lower pKa value than benzo[a]carbazole. The NH fragment close to the deprotonated acidity center might have some anion-stabilizing effect.
2-NO2-Aniline and 4-NO2-aniline have similar pKa values in MeCN - 32.63 and 32.58, respectively. Replacing one of the hydrogens in their amino groups with a phenyl group has a drastically different impact on their pKa values. (2-NO2–Ph)(Ph)NH is by 0.84 pKa units stronger acid than 2-NO2-aniline, but (4-NO2–Ph)(Ph)NH is by 3.71 units (4.05 units in DMSO) stronger acid than 4-NO2-aniline. This difference is likely caused by the stabilizing intramolecular hydrogen bond (IMHB) in the neutral (2-NO2–Ph)(Ph)NH molecule, thus making it less acidic than (4-NO2–Ph)(Ph)NH where no IMHB is possible. The pKa of 4-NO2-1-naphthalenamine is almost identical to the pKa of (4-NO2–Ph)(Ph)NH.
In MeCN, the pKa of 4-NO2-aniline is by 0.53 pKa units lower than the pKa of 4-CF3SO2-aniline. The same trend can be observed in DMSO, where 4-NO2-aniline is a stronger acid than 4-CF3SO2-aniline by 0.9 pKa units. To compare the electron-withdrawing nature of both substituents, the resonance and field inductive effect substituent constants can be used. These demonstrate that the SO2CF3 substituent (σF = 0.83, σR = 0.26) is a stronger resonance and induction acceptor than the NO2 substituent (σF = 0.64, σR = 0.16),43 at odds with the above-described acidity order. The inversion of acidity order of NO2 and SO2CF3 derivatives, relative to the order of substituent constants, has been described previously for other compounds and seems to depend on the solvent. In the case of the compound pairs Ph–CH2–NO2 and Ph–CH2–SO2CF3, the nitro-substituted compound is a stronger acid in DMSO.44 Goumont et al.,45 on the other hand, have demonstrated that the relative strength of the acidifying effect of the NO2 and SO2CF3 groups can be different in water and DMSO. They have shown on the example of NO2–CH2–COOEt and CF3SO2–CH2–COOEt that in water, the former is a stronger acid and in DMSO the latter is a stronger acid.45
The Upper Limit of the Acidity Scale in MeCN
In the present work, the MeCN acidity scale was expanded by 0.94 units to a pKa value of 33.51 (2,3,5,6-Cl4-aniline). Initial experiments have shown that theoretically it would be possible to expand the MeCN acidity scale even further toward higher pKa values but there are some experimental limits. First, the anions of the studied acids need to be stable. All the NH acids studied in this work met that requirement. However, this was not valid for two CH acids whose pKa is expected to be in the studied region. Specifically, it was observed from their UV–vis spectra that the anions of phenylacetonitrile and 4-NO2-toluene were not stable under the conditions used for pKa determinations.
Second, a base is needed that is strong enough to deprotonate the acids of interest but not too strong to decompose the solvent. Previously the phosphazene bases t–Bu-N=P1(pyrr)3 [pKaH(MeCN) = 28.42]15 and Et-N=P2(dma)5 [pKaH(MeCN) = 32.94]46 have been used. Neither of them is sufficiently basic to fully deprotonate acids with pKa values higher than approximately 32.5. Also Et-N=P2(pyrr)5 has been used as a deprotonating compound46 in MeCN, but due to its commercial unavailability and its only marginally higher pKa value than Et-N=P2(dma)5, it was not considered. t-Bu-N=P4(dma)9 would be a suitable deprotonating base in terms of basicity and spectral properties but P3 and higher phosphazenes cause the self-condensation of acetonitrile and the formation of 4-NH2-2,6-Me2-pyrimidine.47 Thus, the phosphazene base HN=P1(tmg)3 was used in the present work. The pKaH (pKa of the conjugate acid) value of HN=P1(tmg)3 in MeCN has been estimated to be 37.2.46 This work marks the first time HN=P1(tmg)3 has been used as a basic titrant to deprotonate acids for the relative pKa determinations in MeCN. Although HN=P1(tmg)3 is strong enough to deprotonate very weak acids and is stable as a free base in MeCN, it has a downside. Differently from the other mentioned strong bases, the protonated form of HN=P1(tmg)3 absorbs at wavelengths up to 290 nm (Figure S2 in the SI), meaning that it is a suitable titrant for acids that absorb at longer wavelengths, thus essentially limiting its usage to acids which have in their structure chromophores conjugated to the acidity center, e.g. nitro-substituted aromatic amines. This limitation would be absent in pKa determinations with NMR, meaning that HN=P1(tmg)3 could be used in such pKa determinations in the future. The phosphazene base t–Bu-N=P1(pyrr)3 has already been successfully used in pKa determinations with NMR.48
Conclusion
As a result of this work, the weak acid region of the MeCN acidity scale (from pKa 29 upward), which previously contained only five acids, has been populated by 31 new compounds and extended to the pKa value of 33.5. Altogether, 37 new pKa values have been determined, which were among the highest experimental pKa values of neutral NH acids in MeCN that have been reported until now. Several of the previously reported pKa values have been revised. The experimental challenges encountered when measuring the pKa values of very weak acids in MeCN were described and analyzed. The relations between the structure and acidity of a number of weak NH acids–nitrogen heterocycles and aromatic amines–have been discussed.
The newly determined pKa values in MeCN were used in correlation with acidity data in DMSO to assess the quality of simple pKa conversion equations from the literature between pKa values in MeCN and DMSO. It was found that these equations yield pKa values with an RMSE of prediction of 0.5, which is accurate enough for many applications. On the basis of the available data and correlation and conversion equations, some pKa values have been predicted in MeCN. Together with computational gas-phase acidities of more than 40 compounds, this work presents over 90 new experimental or predicted acidity values—pKa values in MeCN and computational acidity values in the gas phase.
Experimental Section
pKa Determination Method
The pKa values in MeCN were determined using the previously developed spectrophotometric titration methodology. This methodology is based on the measurement of relative acidities (ΔpKa) of two acids (HA and HB). The studied equilibrium is the following:
| 8 |
The logarithm of the equilibrium constant of the reaction shown in eq 8 is the difference in pKa values of compounds HB and HA, i.e. ΔpKa, as defined in eq 9.
| 9 |
Methanesulfonic acid (CAS No. 75–75–2) was used as the acidic titrant during all relative acidity measurements. Phosphazene base P2-Et (CAS No. 165535–45–5) was used as the basic titrant for deprotonating weak acids with pKa values under 31. Acids with pKa values over 31 were deprotonated using the phosphazene base HN=P1(tmg)3 (CAS No. 874220–27–6). For some measurements Et-N=P2(pyrr)5 (CAS No. 874220–47–0) was used as the basic titrant. The concentrations of the acidic and basic titrant solutions were (2 × 10–3 – 6 × 10–3) mol L–1.
One ΔpKa measurement series consisted of three experimental steps. As the first step, a solution of acid HA in MeCN with a concentration of (1 × 10–5 – 2 × 10–4) mol L–1 was prepared into a 1 cm optical path length quartz cuvette and its UV–vis spectrum was registered. Then, a drop of an acidic titrant was added to the cuvette with a 100 μL Hamilton gastight syringe and a new spectrum was registered. This was done to make sure acid HA was in its neutral form in the solution. After that, a basic titrant was stepwise added until all of HA in the solution was converted to its anion A–. This was evidenced by the absence of changes in the absorbance spectrum of the solution after further addition of the titrant. This step yielded the UV–vis spectra of the fully protonated (HA) and deprotonated (A–) forms of the acid. In total, 5–10 spectra were registered. After the UV–vis spectrum of the fully deprotonated form A– was recorded, an acidic titrant was stepwise added to the solution to verify the reversibility of deprotonation of HA. Although only the spectra of neutral and anionic forms were used in the calculation of the result, the spectra of the partially deprotonated HA were important for diagnostic reasons. These spectra gave information on whether the compound was pure or the presence of possible side processes (e.g., if it was decomposing during the measurement).
As the second step of the ΔpKa measurement, the same operations were done with acid HB.
As the third step, a mixture containing both HA and HB was prepared, and the same titration procedure was applied. If it was found during the individual titrations that HA and HB were already initially in their neutral forms, no acidic titrant was added, and the mixture of HA and HB was only titrated with the basic titrant. During the titration of the mixture, 15–30 spectra were recorded. It was important to have more spectra of the mixture where HA and HB are partially deprotonated than during the titration of HA and HB individually. The reason is that a ΔpKa value was calculated for every solution formed, which in turn was used to assign the final ΔpKa value for the pair of acids.
After mathematically treating the spectral data obtained from the titration of the mixture, as well as the spectra of the solutions of the individual species HA, A–, HB and B– at multiple wavelengths using multilinear regression analysis, the dissociation levels αHA = [A–]/([A–] + [HA]) and αHB = [B–]/([B–] + [HB]) of both acids in all the mixtures formed during titration were obtained and were then in turn used to calculate the ΔpKa values of HA and HB according to eq 10:
| 10 |
A more detailed description of the derivation of equations used to calculate the results of experimental ΔpKa determinations is provided in the Supporting Information (SI). The UV–vis spectra of all the acids studied in this paper can be found in the SI (Figures S4–S51).
Every weak acid studied in this paper has relative acidity measurements done against at least three other acids. The agreement between these three measurements can be visualized from Table 1 by comparing the directly measured ΔpKa values with the differences of the assigned pKa values of the respective acids.
Computational Method
The geometry optimization and vibrational frequency calculations were carried out at DFT BP86/def2-TZVPP level of theory with DFT-D3(BJ) dispersion correction (software: Turbomole V7.849 and V7.750). The absence of imaginary frequencies in the spectra was taken as proof that local energy minimum was reached. For structures with many possible conformers the most stable one was identified and its geometry was used in the further calculations.
The complete basis set (CBS) energy values were determined based on the procedure developed by Helgaker et al.51 For each structure, DLPNO–CCSD(T)52 single-point calculations were carried out using basis sets cc-pVDZ, cc-pVTZ, and cc-pVQZ (software: ORCA53 versions 5.0.4 and 6.0.0). For most compounds, the CBS energy value was found as an intercept of energy vs X–3 plot, where X = 2 for cc-pVDZ, X = 3 for cc-pVTZ and X = 4 for cc-pVQZ. For compound tBu4Box2CH2, due to its size, the CBS energy was found from cc-pVDZ and cc-pVTZ calculations via two parameter expression with integer exponent 4.54,55
Gas-phase acidity (GA) values were computed as follows:
| 11 |
where ECBS is CBS energy computed as described above, CPTZVPP is chemical potential obtained from the TZVPP calculation using the freeh module in Turbomole, and G(H+) is Gibbs energy of the proton (−6.275 kcal mol–1). Chemical potential is equal to H – TS, where H is enthalpy, T is temperature and S is entropy.
For comparison, GA values of several compounds were calculated with G4MP256 method (software: Gaussian 16 Rev A.0357). The G4MP2 results were very well correlated with CBS results:
| 12 |
G4MP2 is one of the most accurate of the relatively computationally affordable methods for GA calculation.19 Our results suggest that for the studied compounds it does not fall short of the used combination of DLPNO–CCSD(T) and BP86/TZVPP.
The optimized geometries of the lowest-energy conformers and tabulated results of individual single-point calculations are available in the SI.
Instruments
An Agilent Cary 60 spectrophotometer (scanning speed 600 nm/min) connected with optical fiber cables to an external cell compartment inside a commercial MBraun Unilab glovebox filled with 99.999% pure argon (5.0, Linde Gas) was used for the pKa determinations in MeCN. This setup ensured that the moisture and oxygen contents inside the glovebox were usually under 1 ppm and always under 10 ppm during all titrations. The external cell compartment did not have a built-in thermostat but its temperature was monitored and was in the range of (24.5 ± 1.5) °C.
Chemicals
Solvent
Acetonitrile (Romil 190 SpS far UV/gradient quality) was used as solvent after drying with molecular sieves (3 Å) for at least 12 h, which lowered the water content to under 6 ppm. The water content of the solvent was monitored using coulometric Karl Fischer titration. For all measurements involving aromatic amines, additionally purified MeCN was used. After drying with 3 Å molecular sieves, distillation over CaH2 under argon was applied. The water content after the distillation process was found to be under 10 ppm and the solvent was stored in the argon glovebox. No molecular sieves were used for further drying the solvent in order to keep it as pure as possible.
Studied Compounds
The following commercially obtained chemicals were used without further purification: 2,3,5,6–Cl4–aniline (Dr. Ehrenstorfer GmbH, 99.0%), 4–CF3SO2–aniline (Apollo Scientific, 97%), 2,3,4,5,6–Cl5–aniline (Fluka, 99.1%), indole (Aldrich, ≥ 99%), N–Me–4–NO2–aniline (Sigma-Aldrich, 97%), (2–NO2–C6H4)(Ph)NH (Aldrich, 98%), 2–MeO–carbazole (BLD Pharmatech, 99.94%), 2–NH2–4–NO2–aniline (Aldrich, 98%), 5–Cl–2–NO2–aniline (Fluka, 99%), 1,3–diphenylurea (Aldrich, 98%), (2–NO2–C6H4)(4–Cl–C6H4)NH (BLD Pharmatech, 97%), 2–Cl–6–NO2–aniline (BLD Pharmatech, 97%), 7–azaindole (Thermo Scientific, 98%), 4–azaindole (Thermo Scientific, 97%), 2–Cl–4–NO2–aniline (Sigma-Aldrich, 99%), benzo[c]carbazole (BLD Pharmatech, 97%), benzo[a]carbazole (BLD Pharmatech, 98%), 5–azaindole (Thermo Scientific, 98%), (5–Cl–2–NO2–C6H3)(Ph)NH (Alfa Aesar, 98%), indazole (Aldrich, 98%), 6–azaindole (Thermo Scientific, ≥ 97%), 2,3–Cl2–6–NO2–aniline (BLD Pharmatech, 97%), norharman (Acros Organics, 98%), 2,4–Cl2–6–NO2–aniline (BLD Pharmatech, 98%), 2,5–Cl2–4–NO2–aniline (BLD Pharmatech, 98%), (4–NO2–C6H4)(Ph)NH (Sigma-Aldrich, 99%), 4–NO2–1–Naphthalenamine (BLD Pharmatech, 98.53%), 2,6–Cl2–4–NO2–aniline (BLD Pharmatech, 98.32%), benzimidazole (Aldrich, 98%), benzotriazole (Reakhim, “pure for analysis”). The following chemicals were kind gifts and they were used as received: 3–Cl–4–NO2–aniline (a kind gift from Peeter Talts, University of Tartu) and 4–Cl–2–NO2–aniline (a kind gift from the late prof. Ilmar Koppel, University of Tartu). The following chemicals were of the same origin as in previous publications and they were used without further purification: 7–NO2-indole,4tBu4Box2CH2,58 4–CN–2,3,5,6–F4–aniline,59 4–NH2–C5F4N,59 9–C6F5–fluorene,60 C6H5–CH2–SO2CF3,6 (4–Me–C6F4)(C6H5)CHCN,60 2,4–(NO2)2–aniline,61 4–NO2–aniline,62 2–NO2–aniline.62 The purity of the chemicals was assessed while observing their UV–vis spectra during titrations and the compounds were declared pure enough if no unknown absorbance changes were seen and if the isosbestic points (if present) were sharp. Carbazole (Aldrich) was recrystallized from ethanol, and 3–MeCOO-indole (Chemapol) was recrystallized twice from 40% methanol before usage. Indolo[2,3-a]carbazole was the same as used in a previous publication65 and it was additionally recrystallized from ethanol.
Titrants
Methanesulfonic acid (Aldrich, ≥ 99%) was used to prepare the acidic titrant solution. Phosphazene base P2-Et (Sigma-Aldrich, ≥ 98%), Et-N=P2(pyrr)5 (same as used in a previous publication63) or HN=P1(tmg)3 (prepared for this work) were used to prepare the basic titrant solution.
HN=P1(tmg)3
HN=P1(tmg)3·HBF3 (2.02 mmol; 963 mg), KHMDS (2.06 mmol; 411 mg) and hexane (12 mL) were added to an oven-dried (140 °C) ACE pressure tube inside a glovebox. The reaction mixture was stirred overnight at 70 °C outside the glovebox under argon. The mixture was then filtered in a glovebox and hexane was removed in vacuo. The raw product was recrystallized from hexane at −30 °C to give HN=P1(tmg)3 as colorless crystals (514 mg, 65%). Some of the substrate did not react. It is likely that a finer ground powder compared to what was used is preferred since the starting compound is not soluble in hexane. Or longer reaction time should be used. The obtained spectral data corresponded to that previously reported.63
Acknowledgments
This research was funded by the Estonian Ministry of Education and Research (TK210) and by the Estonian Research Council grants PRG690, PRG1736 and PRG1031. This work was carried out using the instrumentation at the Estonian Center of Analytical Chemistry (TT4, https://www.akki.ee). The calculations were carried out in the High Performance Computing Center of the University of Tartu.64 The authors are grateful to Professor Ingo Krossing for advice on computations and to Professor Reinhard Schwesinger for inspiring discussions.
Data Availability Statement
The data underlying this study are available in the published article and its Supporting Information.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsorginorgau.4c00095.
Optimized gas-phase geometries as a set of XYZ files (ZIP)
Computational energies in the gas phase (XLSX)
Theoretical background of the relative acidity determination method, pKa measurement results of benzotriazole, computational GA values of reference acids, UV–vis titration spectra of all studied compounds, UV–vis spectrum of HN=P1(tmg)3, acetonitrile for pKa determinations (PDF)
Author Contributions
CRediT: Märt Lõkov investigation, writing–original draft, formal analysis, supervision, data curation; Carmen Kesküla investigation; Sofja Tshepelevitsh investigation, formal analysis, data curation, visualization, writing - review and editing; Marta-Lisette Pikma investigation; Jaan Saame investigation; Dmitri Trubitsõn investigation; Tõnis Kanger funding acquisition, supervision; Ivo Leito: conceptualization, methodology, formal analysis, supervision, resources, writing–review and editing, project administration, funding acquisition.
The authors declare no competing financial interest.
Supplementary Material
References
- Cox B. G.Acids and Bases: Solvent Effects on Acid-Base Strength; Oxford University Press: Oxford, 2013. 10.1093/acprof:oso/9780199670512.001.0001. [DOI] [Google Scholar]
- Scharf M. J.; Tsuji N.; Lindner M. M.; Leutzsch M.; Lõkov M.; Parman E.; Leito I.; List B. Highly Acidic Electron-Rich Brønsted Acids Accelerate Asymmetric Pictet–Spengler Reactions by Virtue of Stabilizing Cation−π Interactions. J. Am. Chem. Soc. 2024, 28339 10.1021/jacs.4c09421. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bordwell F. G. Equilibrium Acidities in Dimethyl Sulfoxide Solution. Acc. Chem. Res. 1988, 21 (12), 456–463. 10.1021/ar00156a004. [DOI] [Google Scholar]
- Trubitsõn D.; Martõnova J.; Erkman K.; Metsala A.; Saame J.; Kõster K.; Järving I.; Leito I.; Kanger T. Enantioselective N-Alkylation of Nitroindoles under Phase-Transfer Catalysis. Synthesis 2020, 52 (07), 1047–1059. 10.1055/s-0039-1690751. [DOI] [Google Scholar]
- Debrauwer V.; Leito I.; Lõkov M.; Tshepelevitsh S.; Parmentier M.; Blanchard N.; Bizet V. Synthesis and Physicochemical Properties of 2-SF5-(Aza)Indoles, a New Family of SF5 Heterocycles. ACS Org. Inorg. Au 2021, 1 (2), 43–50. 10.1021/acsorginorgau.1c00010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kütt A.; Tshepelevitsh S.; Saame J.; Lõkov M.; Kaljurand I.; Selberg S.; Leito I. Strengths of Acids in Acetonitrile. Eur. J. Org. Chem. 2021, 2021 (9), 1407–1419. 10.1002/ejoc.202001649. [DOI] [Google Scholar]
- Bordwell F. G.; Algrim D. J. Acidities of Anilines in Dimethyl Sulfoxide Solution. J. Am. Chem. Soc. 1988, 110 (9), 2964–2968. 10.1021/ja00217a045. [DOI] [Google Scholar]
- Akhtar J.; Khan A. A.; Ali Z.; Haider R.; Shahar Yar M. Structure-Activity Relationship (SAR) Study and Design Strategies of Nitrogen-Containing Heterocyclic Moieties for Their Anticancer Activities. Eur. J. Med. Chem. 2017, 125, 143–189. 10.1016/j.ejmech.2016.09.023. [DOI] [PubMed] [Google Scholar]
- Joule J. A.; Mills K.. Heterocyclic Chemistry, 5th ed.; Wiley: Hoboken, N.J, 2009. [Google Scholar]
- Vitaku E.; Smith D. T.; Njardarson J. T. Analysis of the Structural Diversity, Substitution Patterns, and Frequency of Nitrogen Heterocycles among U.S. FDA Approved Pharmaceuticals: Miniperspective. J. Med. Chem. 2014, 57 (24), 10257–10274. 10.1021/jm501100b. [DOI] [PubMed] [Google Scholar]
- Keri R. S.; Hiremathad A.; Budagumpi S.; Nagaraja B. M. Comprehensive Review in Current Developments of Benzimidazole-Based Medicinal Chemistry. Chemical Biology & Drug Design 2015, 86 (1), 19–65. 10.1111/cbdd.12462. [DOI] [PubMed] [Google Scholar]
- Bansal Y.; Silakari O. The Therapeutic Journey of Benzimidazoles: A Review. Bioorg. Med. Chem. 2012, 20 (21), 6208–6236. 10.1016/j.bmc.2012.09.013. [DOI] [PubMed] [Google Scholar]
- Pennington L. D.; Moustakas D. T. The Necessary Nitrogen Atom: A Versatile High-Impact Design Element for Multiparameter Optimization. J. Med. Chem. 2017, 60 (9), 3552–3579. 10.1021/acs.jmedchem.6b01807. [DOI] [PubMed] [Google Scholar]
- Kütt A.; Selberg S.; Kaljurand I.; Tshepelevitsh S.; Heering A.; Darnell A.; Kaupmees K.; Piirsalu M.; Leito I. pKa Values in Organic Chemistry – Making Maximum Use of the Available Data. Tetrahedron Lett. 2018, 59 (42), 3738–3748. 10.1016/j.tetlet.2018.08.054. [DOI] [Google Scholar]
- Tshepelevitsh S.; Kütt A.; Lõkov M.; Kaljurand I.; Saame J.; Heering A.; Plieger P. G.; Vianello R.; Leito I. On the Basicity of Organic Bases in Different Media. Eur. J. Org. Chem. 2019, 2019 (40), 6735–6748. 10.1002/ejoc.201900956. [DOI] [Google Scholar]
- Heering A.; Stoica D.; Camões F.; Anes B.; Nagy D.; Nagyné Szilágyi Z.; Quendera R.; Ribeiro L.; Bastkowski F.; Born R.; Nerut J.; Saame J.; Lainela S.; Liv L.; Uysal E.; Roziková M.; Vičarová M.; Snedden A.; Deleebeeck L.; Radtke V.; Krossing I.; Leito I. Symmetric Potentiometric Cells for the Measurement of Unified pH Values. Symmetry 2020, 12 (7), 1150. 10.3390/sym12071150. [DOI] [Google Scholar]
- The abbreviation tBu4Box2CH2 denotes bis(4,6-tert-butyl-benzoxazol-2-yl)methane (CAS 2128672–84–2).
- NIST Chemistry WebBook, NIST Standard Reference Database 69; Linstrom P. J., Mallard W. G., Eds.; National Institute of Standards and Technology: Gaithersburg MD, 20899, 10.18434/T4D303. [DOI] [Google Scholar]
- Toomsalu E.; Koppel I. A.; Burk P. Critical Test of Some Computational Chemistry Methods for Prediction of Gas-Phase Acidities and Basicities. J. Chem. Theory Comput. 2013, 9 (9), 3947–3958. 10.1021/ct4003916. [DOI] [PubMed] [Google Scholar]
- Leito I.; Kaljurand I.; Piirsalu M.; Tshepelevitsh S.; Zheng J.; Roses M.; Gal J.-F.. Acid dissociation constants in selected dipolar non-hydrogen-bond-donor solvents. Zenodo, deposited December 16, 2024, version 1.1.0. 10.5281/zenodo.12608876 (accessed 2025–06–02). [DOI]
- Dolman D.; Stewart R. Strongly Basic Systems. VIII. The H– Function for Dimethyl Sulfoxide – Water – Tetramethylammonium Hydroxide. Can. J. Chem. 1967, 45 (9), 911–924. 10.1139/v67-155. [DOI] [Google Scholar]
- Yagil G. The Proton Dissociation Constant of Pyrrole, Indole and Related Compounds. Tetrahedron 1967, 23 (6), 2855–2861. 10.1016/0040-4020(67)85151-2. [DOI] [PubMed] [Google Scholar]
- Taft R. W.; Bordwell F. G. Structural and Solvent Effects Evaluated from Acidities Measured in Dimethyl Sulfoxide and in the Gas Phase. Acc. Chem. Res. 1988, 21 (12), 463–469. 10.1021/ar00156a005. [DOI] [Google Scholar]
- Meot-Ner M.; Liebman J. F.; Kafafi S. A. Carbon Acidities of Aromatic Compounds. 2. Ionic Probes of Aromaticity in Annelated Rings. J. Am. Chem. Soc. 1988, 110 (18), 5937–5941. 10.1021/ja00226a001. [DOI] [PubMed] [Google Scholar]
- Stewart R.The Proton: Applications to Organic Chemistry, 1st ed.; Academic Press: Orlando, 1985. [Google Scholar]
- Taft R. W.; Topsom R. D.. The Nature and Analysis of Substitutent Electronic Effects. In Prog. Phys. Org. Chem.; Taft R. W., Ed.; Wiley, 1987; Vol. 16, pp 1–83. 10.1002/9780470171950.ch1. [DOI] [Google Scholar]
- Chrisment J.; Delpuech J.-J. Thermodynamic and Kinetic Acidities in Dimethyl Sulfoxide. Part 2. Acetylenic Compounds. J. Chem. Soc., Perkin Trans. 2 1977, (4), 407. 10.1039/p29770000407. [DOI] [Google Scholar]
- Catalan J.; Elguero J. Basicity and Acidity of Azoles. Adv. Heterocycl. Chem. 1987, 41, 187–274. 10.1016/S0065-2725(08)60162-2. [DOI] [Google Scholar]
- Chen Y.; Rich R. L.; Gai F.; Petrich J. W. Fluorescent Species of 7-Azaindole and 7-Azatryptophan in Water. J. Phys. Chem. 1993, 97 (9), 1770–1780. 10.1021/j100111a011. [DOI] [Google Scholar]
- Carnerero J. M.; González-Benjumea A.; Carmona C.; Balón M. Spectroscopic Study of the Ground and Excited State Prototropic Equilibria of 4-Azaindole. Spectrochim. Acta, Part A 2012, 97, 1072–1078. 10.1016/j.saa.2012.07.099. [DOI] [PubMed] [Google Scholar]
- Filleux-Blanchard M. L.; Fieux J.; Halle J. C. Hindered Rotation about the N–Aryl Bond in Anilines. J. Chem. Soc. D 1971, 0 (15), 851–852. 10.1039/C29710000851. [DOI] [Google Scholar]
- Vlasov V. M.; Yakobson G. G.; Terekhova M. I.; Petrov E. S.; Shatenstein A. I.; Karpov L. Ya. Equilibrium NH-Acidity of Polyfluorinated Arylamines and Benzanilides. J. Fluorine Chem. 1980, 16 (6), 561–562. 10.1016/S0022-1139(00)84076-2. [DOI] [Google Scholar]
- Koppel I.; Koppel J.; Maria P.-C.; Gal J.-F.; Notario R.; Vlasov V. M.; Taft R. W. Comparison of Brönsted Acidities of Neutral NH-Acids in Gas Phase, Dimethyl Sulfoxide and Water. Int. J. Mass Spectrom. 1998, 175 (1–2), 61–69. 10.1016/S0168-1176(98)00113-X. [DOI] [Google Scholar]
- Catalan J.; Claramunt R. M.; Elguero J.; Laynez J.; Menendez M.; Anvia F.; Quian J. H.; Taagepera M.; Taft R. W. Basicity and Acidity of Azoles: The Annelation Effect in Azoles. J. Am. Chem. Soc. 1988, 110 (13), 4105–4111. 10.1021/ja00221a001. [DOI] [Google Scholar]
- Balón M.; Hidalgo J.; Guardado P.; Muñoz M. A.; Carmona C. Acid–Base and Spectral Properties of β-Carbolines. Part 2. Dehydro and Fully Aromatic β-Carbolines. J. Chem. Soc., Perkin Trans. 2 1993, (1), 99–104. 10.1039/P29930000099. [DOI] [Google Scholar]
- Stewart R.; O’Donnell J. P. Strongly basic systems: III. The H– function for various solvent systems. Can. J. Chem. 1964, 42 (7), 1681–1693. 10.1139/v64-251. [DOI] [Google Scholar]
- Terekhova M. I.; Os’kina I. A.; Sutula V. D.; Pogozhnikova O. Yu.; Tabatskaya A. A.; Petrov E. S.; Vlasov V. M. Effects of Substituents in the Equilibrium CH and NH Acidity of Substituted 1-Naphthylacetonitriles and 1-Naphthylamines. J. Org. Chem. USSR 1987, 23, 361–369. [Google Scholar]
- Yagil G. Effect of Ionic Hydration in Equilibria and Rates in Concentrated Electrolyte Solutions. III. The H- Scale in Concentrated Hydroxide Solutions. J. Phys. Chem. 1967, 71 (4), 1034–1044. 10.1021/j100863a040. [DOI] [Google Scholar]
- Albert A.; Goldacre R.; Phillips J. 455. The Strength of Heterocyclic Bases. J. Chem. Soc. 1948, 2240. 10.1039/jr9480002240. [DOI] [Google Scholar]
- Bartmess J. E.; Scott J. A.; McIver R. T. Scale of Acidities in the Gas Phase from Methanol to Phenol. J. Am. Chem. Soc. 1979, 101 (20), 6046–6056. 10.1021/ja00514a030. [DOI] [Google Scholar]
- Cumming J. B.; Kebarle P. Summary of Gas Phase Acidity Measurements Involving Acids AH. Entropy Changes in Proton Transfer Reactions Involving Negative Ions. Bond Dissociation Energies D (A—H) and Electron Affinities EA(A). Can. J. Chem. 1978, 56 (1), 1–9. 10.1139/v78-001. [DOI] [Google Scholar]
- Taft R. W.; Anvia F.; Taagepera M.; Catalan J.; Elguero J. Electrostatic Proximity Effects in the Relative Basicities and Acidities of Pyrazole, Imidazole, Pyridazine, and Pyrimidine. J. Am. Chem. Soc. 1986, 108 (12), 3237–3239. 10.1021/ja00272a013. [DOI] [Google Scholar]
- Hansch C.; Leo A.; Taft R. W. A Survey of Hammett Substituent Constants and Resonance and Field Parameters. Chem. Rev. 1991, 91 (2), 165–195. 10.1021/cr00002a004. [DOI] [Google Scholar]
- Terrier F.; Kizilian E.; Goumont R.; Faucher N.; Wakselman C. α-Sulfonyl Carbanions: Combined Kinetic, Thermodynamic, and NMR Approaches for the Study of the Ionization of Benzyltriflones in Me2SO and H2O–Me2SO Mixtures. J. Am. Chem. Soc. 1998, 120 (37), 9496–9503. 10.1021/ja9806074. [DOI] [Google Scholar]
- Goumont R.; Magnier E.; Kizilian E.; Terrier F. Acidity Inversions of α-NO2 and α-SO2CF3 Activated Carbon Acids as a Result of Contrasting Solvent Effects on Transfer from Water to Dimethyl Sulfoxide Solutions. J. Org. Chem. 2003, 68 (17), 6566–6570. 10.1021/jo034244o. [DOI] [PubMed] [Google Scholar]
- Saame J.; Rodima T.; Tshepelevitsh S.; Kütt A.; Kaljurand I.; Haljasorg T.; Koppel I. A.; Leito I. Experimental Basicities of Superbasic Phosphonium Ylides and Phosphazenes. J. Org. Chem. 2016, 81 (17), 7349–7361. 10.1021/acs.joc.6b00872. [DOI] [PubMed] [Google Scholar]
- Schwesinger R.; Schlemper H.; Hasenfratz C.; Willaredt J.; Dambacher T.; Breuer T.; Ottaway C.; Fletschinger M.; Boele J.; Fritz H.; Putzas D.; Rotter H. W.; Bordwell F. G.; Satish A. V.; Ji G.-Z.; Peters E.-M.; Peters K.; Schnering H. G. von; Walz L. Extremely Strong, Uncharged Auxiliary Bases; Monomeric and Polymer-Supported Polyaminophosphazenes (P2–P5). Liebigs Annalen 1996, 1996 (7), 1055–1081. 10.1002/jlac.199619960705. [DOI] [Google Scholar]
- Pikma M.-L.; Tshepelevitsh S.; Selberg S.; Kaljurand I.; Leito I.; Kütt A. P. KaH Values and θH Angles of Phosphanes to Predict Their Electronic and Steric Parameters. Dalton Trans. 2024, 53 (34), 14226–14236. 10.1039/D4DT01430H. [DOI] [PubMed] [Google Scholar]
- TURBOMOLE, V7.8, 2023, a Development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 1989–2007, TURBOMOLE GmbH, since 2007. https://www.turbomole.org (accessed 2025–02–05).
- TURBOMOLE, V7.7, 2022, a Development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 1989–2007, TURBOMOLE GmbH, since 2007. https://www.turbomole.org (accessed 2025–02–05).
- Helgaker T.; Klopper W.; Koch H.; Noga J. Basis-Set Convergence of Correlated Calculations on Water. J. Chem. Phys. 1997, 106 (23), 9639–9646. 10.1063/1.473863. [DOI] [Google Scholar]
- Riplinger C.; Sandhoefer B.; Hansen A.; Neese F. Natural Triple Excitations in Local Coupled Cluster Calculations with Pair Natural Orbitals. J. Chem. Phys. 2013, 139 (13), 134101 10.1063/1.4821834. [DOI] [PubMed] [Google Scholar]
- Neese F. Software Update: The ORCA Program System—Version 5.0. WIREs Comput. Mol. Sci. 2022, 12 (5), e1606 10.1002/wcms.1606. [DOI] [Google Scholar]
- Martin J. M. L. Ab Initio Total Atomization Energies of Small Molecules — towards the Basis Set Limit. Chem. Phys. Lett. 1996, 259 (5), 669–678. 10.1016/0009-2614(96)00898-6. [DOI] [Google Scholar]
- Vasilyev V. Online Complete Basis Set Limit Extrapolation Calculator. Comput. Theor. Chem. 2017, 1115, 1–3. 10.1016/j.comptc.2017.06.001. [DOI] [Google Scholar]
- Curtiss L. A.; Redfern P. C.; Raghavachari K. Gaussian-4 Theory Using Reduced Order Perturbation Theory. J. Chem. Phys. 2007, 127 (12), 124105 10.1063/1.2770701. [DOI] [PubMed] [Google Scholar]
- 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 16, Revision A.03, 2016.
- Kretsch J.; Koehne I.; Lõkov M.; Leito I.; Stalke D. Bis(Benzoxazol-2-Yl)Methanes Hounding NacNac: Varieties and Applications in Main Group Metal Coordination. Eur. J. Inorg. Chem. 2019, 2019 (28), 3258–3264. 10.1002/ejic.201900631. [DOI] [Google Scholar]
- Koppel I. A.; Taft R. W.; Anvia F.; Zhu S.-Z.; Hu L.-Q.; Sung K.-S.; DesMarteau D. D.; Yagupolskii L. M.; Yagupolskii Y. L. The Gas-Phase Acidities of Very Strong Neutral Bronsted Acids. J. Am. Chem. Soc. 1994, 116 (7), 3047–3057. 10.1021/ja00086a038. [DOI] [Google Scholar]
- Kütt A.; Leito I.; Kaljurand I.; Sooväli L.; Vlasov V. M.; Yagupolskii L. M.; Koppel I. A. A Comprehensive Self-Consistent Spectrophotometric Acidity Scale of Neutral Brønsted Acids in Acetonitrile. J. Org. Chem. 2006, 71 (7), 2829–2838. 10.1021/jo060031y. [DOI] [PubMed] [Google Scholar]
- Kaupmees K.; Järviste R.; Leito I. Basicity of Very Weak Bases in 1,2-Dichloroethane. Chem.—Eur. J. 2016, 22 (48), 17445–17449. 10.1002/chem.201602629. [DOI] [PubMed] [Google Scholar]
- Kaljurand I.; Kütt A.; Sooväli L.; Rodima T.; Mäemets V.; Leito I.; Koppel I. A. Extension of the Self-Consistent Spectrophotometric Basicity Scale in Acetonitrile to a Full Span of 28 pKa Units: Unification of Different Basicity Scales. J. Org. Chem. 2005, 70 (3), 1019–1028. 10.1021/jo048252w. [DOI] [PubMed] [Google Scholar]
- Haav K.; Kadam S. A.; Toom L.; Gale P. A.; Busschaert N.; Wenzel M.; Hiscock J. R.; Kirby I. L.; Haljasorg T.; Lokov M.; Leito I. Accurate Method To Quantify Binding in Supramolecular Chemistry. J. Org. Chem. 2013, 78 (16), 7796–7808. 10.1021/jo400626p. [DOI] [PubMed] [Google Scholar]
- Kolomeitsev A. A.; Koppel I. A.; Rodima T.; Barten J.; Lork E.; Röschenthaler G.-V.; Kaljurand I.; Kütt A.; Koppel I.; Mäemets V.; Leito I. Guanidinophosphazenes: Design, Synthesis, and Basicity in THF and in the Gas Phase. J. Am. Chem. Soc. 2005, 127 (50), 17656–17666. 10.1021/ja053543n. [DOI] [PubMed] [Google Scholar]
- University of Tartu . UT Rocket. 10.23673/PH6N-0144 (accessed 2025–02–05). [DOI]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The data underlying this study are available in the published article and its Supporting Information.



