Next Article in Journal
Lactoferrin Affects the Viability of Bacteria in a Biofilm and the Formation of a New Biofilm Cycle of Mannheimia haemolytica A2
Previous Article in Journal
Hints of Biological Activity of Xerosydryle: Preliminary Evidence on the Early Stages of Seedling Development
Previous Article in Special Issue
Strong Bases and beyond: The Prominent Contribution of Neutral Push–Pull Organic Molecules towards Superbases in the Gas Phase
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Strong Bases Design: Key Techniques and Stability Issues

by
Andrey V. Kulsha
1,
Oleg A. Ivashkevich
2,*,
Dmitry A. Lyakhov
3 and
Dominik Michels
3
1
Chemical Department, Belarusian State University, 14 Leningradskaya Str., 220006 Minsk, Belarus
2
Research Institute for Physical Chemical Problems, Belarusian State University, 14 Leningradskaya Str., 220006 Minsk, Belarus
3
Computer, Electrical and Mathematical Science and Engineering Division, 4700 King Abdullah University of Science and Technology, Thuwal 23955-6900, Saudi Arabia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(16), 8716; https://doi.org/10.3390/ijms25168716 (registering DOI)
Submission received: 24 July 2024 / Revised: 6 August 2024 / Accepted: 8 August 2024 / Published: 9 August 2024

Abstract

:
Theoretical design of molecular superbases has been attracting researchers for more than twenty years. General approaches were developed to make the bases potentially stronger, but less attention was paid to the stability of the predicted structures. Hence, only a small fraction of the theoretical research has led to positive experimental results. Possible stability issues of extremely strong bases are extensively studied in this work using quantum chemical calculations on a high level of theory. Several step-by-step design examples are discussed in detail, and general recommendations are given to avoid the most common stability problems. New potentially stable structures are theoretically studied to demonstrate the future prospects of molecular superbases design.

1. Introduction

The history of strong molecular organic bases goes back more than a century. After the synthesis of pentamethylguanidine (PMG) by Schenck [1] in 1912, progressively stronger molecular bases with nitrogen basicity centers such as 1,5,7-triazabicyclo[4.4.0]dec-5-ene [2] (TBD) or MeN=P(NMe2)3 [3] (Me-P1) have been discovered. However, that strength growth has been uneven: the jump of about 20 orders of magnitude was performed as soon as Reinhard Schwesinger applied the “homologization” concept to amidines [4,5] and phosphazenes [6,7] (Figure 1).
Since then, Schwesinger’s proton sponge (SPS) and polyphosphazenes from P2 to P5 [8] have come to be used as benchmarks to assess the strength of newly synthesized molecular bases. Numerous applications of strong molecular bases in organic synthesis [9,10,11] led to increased interest in various kinds of molecular bases [12,13], including carbenes [14,15], phosphines [16,17,18,19], and even borylenes [20]. However, none of them exceeded the strength of tBu-P5-pyrr 3 significantly (Scheme 1). If metal-free ionic bases are taken into account, then Schwesinger’s phosphazenium fluorides [21,22] expand the scale by several orders of magnitude, with [P(N=P(NMe2)3)4]+F as the absolute record holder.
Consequently, a challenge arose to design new molecular bases. To reach the high basicity of an arbitrary molecule B being designed, one should either stabilize its protonated form BH+ or destabilize its neutral form B. Key techniques to do this are summarized as a general five-step design algorithm.
Step 1
choosing the basicity center. That should be a negatively charged atom with a lone electron pair available for protonation. The trivalent nitrogen atom has been a common choice for more than a century; however, the generally weaker acidity of C–H bond compared to N–H bond makes carbene bases generally stronger. Nevertheless, a lot depends on further steps, making it possible to design strong bases with other types of basicity centers [16,17,18,19,20,21,22].
Step 2
steric loading of the basicity center by groups with lone electron pairs, which destabilize B to a greater extent than BH+. That is implemented, for example, in various classes of “proton sponges” [23,24,25,26,27], Verkade bases [28,29], cyclic aminopyridines [30,31], rotaxane or catenane superbases [32,33,34], and adamanzanes [34,35,36].
Step 3
“homologization principle”: the parent electron-rich unit such as amidine [4,5], guanidine [37], phosphazene [6,7,18], or cyclopropenimine [38] moiety is repeated several times to build a large conjugated structure that delocalizes the positive charge in BH+.
Step 4
adding donor groups such as OMe, NMe2, N(CH2)4, N=C(NMe2)2, N=P(NMe2)3 to the conjugated framework [12,13,39] to increase the electron density on the protonation site in B.
Step 5
adjusting the structure to additionally stabilize BH+ with hydrogen bonds or other intramolecular interactions [40,41].
Although the guide generally works for various types of bases, not all the steps may be necessary at the same time. For example, protonated carbene bases are usually bad hydrogen bond donors, which makes Step 5 almost useless for them. Another point is that every step has its own limitations. Thus, an excess of donor group introduction at Step 4 may lead to conjugation breaking, while too large steric hindrance provided at Step 3 may turn the base kinetically inhibited.
The availability of quantum chemical research methods has stimulated theoretical studies of various potentially superbasic structures in the last 25 years [42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84]. Although many of these predicted structures were theoretically expected to surpass the current basicity champions, none of them actually did. We have previously shown that these inconsistencies may be quite predictable by the same quantum chemistry methods as those used for the design [85]. The goals of this article are to develop general rules to avoid the most common problems while designing new superbases and to provide specific examples of such a design.
The strength of a given base B may be measured and computed in the gas phase [86,87,88,89,90] as well as in the solution [89,90,91,92,93,94,95]. In the latter case, pKa(BH+) values for solvents like THF, DMSO, or MeCN are commonly used. These values become formal for the strongest bases due to solvent degradation [8,22], but that is not the case for hexamethylphosphoramide (HMPA) because of its exceptional stability in highly basic media [85,96]. Indeed, HMPA withstands the presence of deprotonated toluene [97] and deprotonated THF [98], not to mention its ability to solvate electrons [99]. Although HMPA had been suspected of being carcinogenic [100], it was doubted later [101]. Thus, computed pKa(BH+) values in HMPA are used in this work to compare the basicity of the considered structures. According to our previous studies [85,96,102], good linear correlation with experimental pKa values in other solvents suggests that computed pKa values should be quite accurate with respect to each other, although some systematic bias may still appear. To give an idea of the basicity scale in HMPA, a few calculated pKa(BH+) anchor points are given in Table 1.

2. Results and Discussion

2.1. Flawed Design Examples

To illustrate the possible problems arising upon molecular superbase design, a few sample structures are considered.
Structure 5 (Scheme 2) might have been quite a strong carbene base with computed pKa(BH+) = 38.84. However, the obvious problem is its self-deprotonation leading to pyridine. The reason behind it is that the acidity of the NH group in 5 is too high to make it a stable, strong base. Surprisingly, a lot of structures proposed in the literature suffer from the same problem, and that can be shown with quantum chemical computations as it is done by us for the structure 5. For example, an unstable tautomer of 4-aminobenzamidine is proposed as a superbase in ref. [44]; certain “croissant” structures from ref. [58] contain too acidic imidazole moieties; proton sponges from ref. [59] could tautomerize due to the acidity of NH2 groups; allenes from ref. [66] contain CH-acidic cyclopentadiene moieties; allenes from ref. [84] irreversibly lose the allene moiety upon protonation. In azaphosphiridines from ref. [71], proton migration from endocyclic nitrogen atom to endocyclic carbon atom could lead to ring opening; similar ring-opening tautomerization is possible for the tetrahedrane scaffolds from ref. [74]. The tautomerization of dendritic allenes from ref. [72] is less obvious but has been addressed by us earlier [85].
On the other hand, some authors explicitly warn about the self-deprotonation problem [45,53], e.g., “peralkylation of potential superbases is strongly recommended in order to prevent intramolecular self-protonation of the most basic sites” [53]. Thus, one could propose structure 6 (Scheme 2) to avoid self-protonation. However, unfortunately, 6 can at least undergo either self-demethylation to 4-methylpyridine or dimerization to the neutral form of methylviologen [103]. Such inter- or intramolecular rearrangements should always be considered as other possible degradation ways for the structure being designed. Unlike self-deprotonation, they are generally less obvious, but the common routes are either dimerization or nucleophilic attack of positively charged Lewis acidic sites by negatively charged basicity centers. Carbenes [63,65], silylenes [60,69], and germylenes [67] are the most susceptible to these degradation routes due to their high reactivity [104,105,106,107]. Two explicit degradation ways for certain structures from ref. [63] have already been considered by us [85], while a similar process has been recently observed in the experiment [108].
Another illustrative example of intramolecular nucleophilic attack is the isomerization of 4,5-diarsaphenanthrene dioxide derivatives from ref. [68]. For structure 7 (Scheme 3), our calculations predict the isomerization to be exergonic by 73 kJ/mol in the HMPA solution and by 128 kJ/mol in the gas phase. Generally, the most basic molecules may become unstable due to the presence of even such weak Lewis acids as cyclopropene groups [64,65,73,75,80]; the corresponding degradation has already been confirmed experimentally [38].
Structure 8 (Scheme 4) is a product of applying the “homologization principle” to TBD. At first glance, 8 should be quite stable, and its mesoionic nature suggests high basicity of the outer nitrogen atoms. Indeed, our calculations give pKa(BH+) = 21.27 in HMPA for 8H+ and a gas-phase dipole moment of more than 12 Debye for 8. However, closer analysis reveals its possible tautomerization to 8a with the corresponding pKtaut = 0.2. Thus, self-deprotonation may be accompanied by intramolecular rearrangement. The experimental evidence of that was provided by Schwesinger: strongly basic media cause degradation of N(CH2)4 group to N(CH2)2CH=CH2 group [22], while NC(CH3)3 group decomposes to NH group and isobutylene [8]. The dimethylamino group seems to be somewhat more stable; however, Lewis acidity of the nearby phosphorus(V) atom may induce CH2=NCH3 elimination [14].
From the above, one can conclude that under extremely basic conditions, such common designing blocks as P(NMe2)3 donor groups, ethylene bridges, or even alkyl groups larger than methyl may become unstable. For example, our calculations show that hexaethylenetetramine 9 [35] could have been a strong base with pKa(BH+) = 21.05, but it appears to be unstable against self-deprotonation (Scheme 5).
Another case is substituted tetraaminoallene 10, the protonated form of which was synthesized by Schwesinger [4] back in 1987. It could have been a record-breaking base with pKa(BH+) = 45.19 if not for the degradation to 10a (Scheme 6).
One could try to modify 10, removing the ethylene bridges and making the seven-membered ring conjugated (Scheme 7). According to our calculations, the resulting structure 11 could still be record-breaking with pKa(BH+) = 42.71. It passes the self-deprotonation test, although the lowest-lying tautomer 11a is quite close (pKtaut = 0.5), but unlike the phosphazene story, the barrier for CH2=NCH3 elimination from 11a is reliably high (174 kJ/mol). The protonation site in 11 can be methylated by one of the adjacent NMe2 groups in 11H+, but the corresponding barrier for CH3+ migration is high enough (133 kJ/mol), suggesting years of lifetime in the worst case (both 11 and 11H+ concentrations are high). However, all these stability arguments are ruined by dimerization to 11b, especially because electron-rich 11b appears to have too high alkali-metal-like reactivity: the calculated sum of its 1st and 2nd ionization energies is just 9.7 eV. Thus, one should remember that even if dimerization or other structure rearrangement occurs to a minor extent, the high reactivity of the product may be of decisive importance for the original structure stability.
Looking at the structure of the strongest currently known carbene base 1 (Scheme 1), one could try to improve its basicity [77,79]. However, quite low stability of similar structures [109,110] suggests that dimerization of 1 or its derivatives is possible. Our calculations reveal an interesting situation: while the direct dimerization is not favorable, it becomes quite exergonic if accompanied by N–N bond cleavage, which is essentially the reduction of the hydrazine moiety (Scheme 8). Such a reduction is facilitated by the presence of donor substituents attached to the phenyl rings, that is why more basic derivatives of 1 are still unknown. The dimerization mechanism may involve an open-shell singlet state [111] of 1, giving an idea why it has not been isolated in pure form yet [15].
The ultimate basicity of Schwesinger’s phosphazenium fluorides is explained by the large proton affinity of the “naked” fluoride anion and by the ability of the proton to attach two fluoride ions simultaneously (general trends for C, N, O, F basicity centers in molecular and ionic structures have been discussed by us earlier [85]). To improve the basicity, one could build a cation with lower fluoride ion affinity [112] than Schwesinger’s [P(N=P(NMe2)3)4]+. According to the recent study [113], donor-substituted tetraphenylphosphonium is worth trying. Since the fluoride anion is small enough to reach the Lewis acidic phosphorus atom in PPh4+ [114], ortho-substituted phenyl groups are preferable to provide the corresponding steric hindrance. Thus, tetrakis[2,6-dimethoxy-4-(dimethylamino)phenyl]phosphonium cation 12+ (Scheme 9) could be suggested to build the potentially record-breaking new superbase 12+F.
Our calculations show that the fluoride ion affinity of 12+ is 39 kJ/mol lower than that of [P(N=P(NMe2)3)4]+, but the following problem arises: the methoxy group can undergo deprotonation and attack the nearby phosphorus atom, leading to the neutral molecule 12a. Indeed, the calculated Gibbs energy change for the reaction
12+F + 12+F = 12a + 12+HF2
is −18 kJ/mol, and the equilibrium might shift further to 12a because the 12+HF2 ion pair seems to dissociate much more than 12+F in the HMPA solution. Nevertheless, that could still be fine since 12a appears to be quite a strong molecular base with computed pKa(BH+) = 46.39 if the protonation back to 12+ is assumed. However, unfortunately, 12a protonation proceeds with 3,5-dimethoxy-N,N-dimethylaniline elimination due to steric overloading of the phosphorus atom (Scheme 9). Considering HCN molecule as a protonation probe, we calculated the protonation barriers to differ by 34 kJ/mol (gas phase) or 9 kJ/mol (HMPA solution) in favor of the degradation route.

2.2. Successful Design Examples

Despite a plethora of stability problems arising at a high basicity level, one could still design potentially stable superbases [42,43,46,47,48,49,50,51]. Key techniques to improve the stability of the superbase B being designed are summarized as another general five-step guide.
Step 6. 
Avoid the presence of groups in B that are either Brønsted or Lewis acidic, and also groups that tend to rearrange upon deprotonation.
Step 7. 
Check that both B protonation and BH+ deprotonation are favored at the same sites; identify possible tautomers.
Step 8. 
Perform a full conformational analysis of all tautomers of both B and BH+ in their ground and low-lying excited states to ensure their relaxation.
Step 9. 
Check that neither B nor BH+ tends to dimerization or other rearrangements.
Step 10. 
Look for as many degradation ways of B as possible, find the lowest-barrier one, and estimate the corresponding half-life under the conditions being proposed for practical usage, including the possible interaction with the solvent.
Our previously designed [85,96] bases 13, 14a, and 14b (Scheme 10) have passed all these tests. While 13 (pKa(BH+) = 33.74) is not stronger than tBu-P4 and therefore has fewer stability risks, bases 14a and especially 14b reach the practical limit of the basicity scale, going quite beyond the deadline of THF degradation. The basicity centers of both 14a and 14b are carbon atoms in the middle of CCC moieties, and the degree of their protonation varies from anion to trication depending on pH (Table 2). The closeness of pKa values of neutral and monoprotonated forms makes both bases autoionize in HMPA: the autoionization degree is 2% for 14a and 11% for 14b.
Methyl group acidity is enough to make both bases tautomerize in the neutral form as well as in the deprotonated anionic form (see [85] for the details on 14a). Although 14b is a stronger base than 14a, the tautomerization of 14b is less pronounced (pKtaut = 2.87 and 1.33 for neutral and anionic forms, respectively) and corresponds to deprotonation of the methyl groups attached to endocyclic nitrogen atoms. The acidity of dimethylamino groups is negligible: the corresponding pKtaut values for neutral 14b are 9.64 (peripheral NMe2 group) and 11.28 (NMe2 group closer to the boron atom), and the corresponding tautomers appear to be quite stable against CH2=NCH3 elimination (the kinetic barriers are 107 and 110 kJ/mol, respectively).
Gas-phase basicity of neutral 14b was calculated to be 1337 kJ/mol, while the same level of theory gives a lower value of 1332 kJ/mol for the Cs2O molecule [85]. Since the latter has been recently proposed as a threshold for hyperbasicity to clarify the meaning of the “hyperbase” term [88], base 14b appears to be the first molecular hyperbase that is presumably stable and synthetically achievable.
The lowest-barrier degradation path for 14b was found to be CH3+ migration from the endocyclic nitrogen atom in the protonated form to the basicity center in the neutral form, like it was found for structure 11. The calculated kinetic barrier of 127 kJ/mol in HMPA solution suggests a year-scale stability for 14b, as it was previously shown for 14a with a similar barrier of 126 kJ/mol [85].
Possible synthetic routes to 14a were addressed earlier [85]; the route to 14b might be similar except for the NMe2 groups introduction, likely requiring the basicity site protection. Protonated forms 14aH+ and 14bH+ are both D2d-symmetric and rigid, allowing crystallization of the corresponding salts for XRC structure validation. It should also be noted that 14bH+ has little strain and perfect steric hindrance of the boron atom, making it a very promising and thermally stable weak-coordinating cation. According to the calculations, it has 14 kJ/mol lower fluoride ion affinity than [P(N=P(NMe2)3)4]+, making 14b·HF a potentially record-breaking base that could be easier to synthesize than 14b itself. To track the deprotonation of 14aH+ and 14bH+ in the solution, a new pH indicator synthesizable from cumene and CCl4 (Scheme 11) has been proposed by us [102]. It is isoelectronic to crystal violet and possesses sharp color changes from orange-yellow neutral form to green anion (pKa1 = 45.60) and then to deep blue dianion (pKa2 = 50.62).
To add some variety to the protonation sites, we also introduce a new series of bases 15ad (Scheme 12) designed to have a metal atom as the basicity center. They look synthesizable because similar platinum complexes have been known since 1989 [115], and the ligand for 15a has been known since 1974 [116,117]. The acid–base properties of that kind of metal complexes have already received great attention from both theoretical [118] and experimental [119] points of view, but surprisingly neither of 15ad has been considered yet. Their exceptional basicity is explained by a specific, like in Verkade bases [29], coordination of the central atom that allows donor substituents to push out the excess electron density to the axial direction. Indeed, the related complex [HPt(dmpe)2]+, where dmpe = Me2PCH2CH2PMe2, has a much lower calculated pKa value of 23.21, which makes [Pt(dmpe)2] a bit weaker base than SPS (Table 1). The experimentally measured pKa values in acetonitrile (32.88 for protonated SPS [120] and 31.1 for [HPt(dmpe)2]+ [121]) agree with the calculations.
Base 15b seems to have greater applied significance: it is stronger than tBu-P4, has enough steric hindrance of phosphorus atoms, and seems to have no apparent problems with synthesis. Bases 15c and 15d are more of theoretical interest: additional steric loading of the basicity center with negatively charged oxygen atoms allows them to reach the level of the strongest known molecular bases (Scheme 1). It is interesting that the related complex Pt(PF3)4 requires one of the strongest superacids for the protonation [122], demonstrating quite a wide basicity range for the PtP4 moiety.
Compared to [Pt(dmpe)2], complexes 15ad are predicted to be less stable. The computed pK values in HMPA for the substitution reactions
[PtP(CH2CH2PR2)3] + 2dmpe = [Pt(dmpe)2] + P(CH2CH2PR2)3
are –8.00, –8.95, –14.52, and –15.16 for 15ad, respectively, showing a clear stability decrease trend. Calculations show that the Pt atom in 15b is attacked by a PF3 molecule with quite a low barrier of 65 kJ/mol. Thus, ligand substitution seems to be the main degradation route for 15ad in the presence of competing ligands. On the other hand, the chelate effect provides good stability against bulky molecules with low affinity to the Pt atom because four Pt–P bonds must be broken to eliminate the ligand completely. That provides some theoretical evidence of 15ad stability in the HMPA solution because the HMPA molecule is quite large and has much lower affinity to the Pt atom than phosphine-type ligands. Anyway, even if the vulnerability of 15ad to small nucleophilic species limits their use as superbases in real applications, it may not diminish the theoretical interest in them as unusual examples of strong base design.

3. Methods

The values of pKa(BH+) were computed as ΔG/RTln10 − 0.7557, where ΔG is the Gibbs energy change for the proton transfer from the protonated base BH+ to HMPA molecule in the solution, R is the molar gas constant (8.31446 J·mol–1·K–1), T is the temperature (298.15 K), and −0.7557 is the decimal logarithm of the numerical value of HMPA molar volume (dm3·mol–1) [123]. Gas-phase Gibbs energy of a given species was computed as the sum of gas-phase total energy on the DLPNO-CCSD(T) level of theory [124,125,126,127,128,129,130] and thermal correction on the PBE0 level of theory [131]. Solvation energy was computed on the PBE0 level of theory within the CPCM model [132] using UFF atomic radii [133] scaled by 1.1 and a dielectric constant of 31.6 for HMPA [134]. Geometry optimizations and vibrational frequency calculations were performed on the PBE0 level of theory. Gas-phase basicity calculations were adjusted for the basis set superposition error [135] (BSSE) using geometries of protonated structures. The details on basis sets and the corresponding effective core potentials (ECP) for heavy atoms are presented in Table 3. DLPNO-CCSD(T) calculations were performed within the ORCA 5.0 package [136], while the PBE0 calculations were performed within the Gaussian16 package [137].
The reliability of the chosen levels of theory and basis sets for the computations of basicity in the solution and in the gas phase was shown by us earlier [85,96,102]. Optimized gas phase geometries, DLPNO-CCSD(T) total energies, and Gibbs energies for all structures studied are available in the Supplementary Materials.

4. Conclusions

We have briefly reviewed the 25-year history of quantum chemical design of strong molecular bases. Although some of the theoretical predictions, like cyclopropenimines [42] from 1999, hydrogen-bonded guanidines [44] from 2002, or phosphazenyl phosphines [48] from 2006, have found excellent experimental confirmation [17,38,40], the vast majority of theoretically proposed structures remain far from real life. Unfortunately, this situation has led to a gradually growing discredit of quantum chemical methods of design.
Analysis of the existing approaches to strong base design allowed us to formulate a general five-step guide to superbasic molecule construction. More importantly, we have deeply analyzed the unsuccessful cases and identified the main instability reasons of theoretically predicted structures in superbasic media. Providing a few explicit examples of step-by-step design, we have illustrated possible stability problems with sample structures as well as with the examples from the literature. As a result, we have formulated another five-step guide to checking the stability of bases with quantum chemistry methods.
We have introduced new successful cases of strong base design. One of them, 15b, is predicted to be stronger than tBu-P4, while having a metal atom as the protonation site. Another proposed base, 14b, is predicted to be more basic than the Cs2O molecule in the gas phase, which makes 14b a hyperbase. In the solution, it must surpass the strongest known molecular bases by more than ten orders of magnitude, setting up a practically reachable basicity limit. Meanwhile, the monoprotonated form of 14b appears to be a robust, highly symmetric weak-coordinating cation.
In general, we can conclude that the lack of experimental confirmation of the theoretical strong base design is not a failure of quantum chemistry but a methodological flaw. Quantum chemistry does provide enough instruments to control the stability of the structures being designed; hence, it does remain a reliable assistant to govern the experiment in the effective direction. Thus, for example, the predicted route of 1 degradation could help to establish the conditions for its isolation in the pure form, while the elusive hexaethylenetetramine could probably be synthesized via derivatives of its predicted tautomer. Careful revising of previously proposed structures that turned out to be unstable might lead to stability improvements. Moreover, the design of stable superbases with new types of protonation sites could be a promising direction of future research.
Anyway, as the growth of available computing power makes quantum chemical research accessible literally to everyone, it becomes very important to follow an efficient methodology for such research. We sincerely hope that our contributions to this methodology will help to build a straight road from theoretical strong bases design to experiment—the criterion of truth.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms25168716/s1.

Author Contributions

Conceptualization, A.V.K.; methodology, A.V.K.; software, D.A.L. and D.M.; resources, D.A.L. and D.M.; writing—original draft preparation, A.V.K.; writing—review and editing, A.V.K. and O.A.I.; supervision, O.A.I.; funding acquisition, D.M. All authors have read and agreed to the published version of the manuscript.

Funding

The APC was funded by KAUST.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the Supplementary Material, further inquiries can be directed to the corresponding author.

Acknowledgments

All Gaussian16 computations were performed on KAUST’s Ibex HPC. We thank the KAUST Supercomputing Core Lab team for assistance with execution tasks on Skylake nodes. We thank Borislav Kovačević for useful comments and suggestions made during the article preparation.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Schenck, M. Zur Kenntnis der methylierten Guanidine. Hoppe-Seyler’s Z. Physiol. Chem. 1912, 77, 328–393. [Google Scholar] [CrossRef]
  2. McKay, A.F.; Kreling, M.-E. Preparation and Chemistry of Δ8-hexahydro-1,4,8-pyrimidazole, Δ9-1,5,9-triazabicyclo(4.4.0)decene, and Δ9-1,4,9-triazabicyclo(5.3.0)decene. Can. J. Chem. 1957, 35, 1438–1445. [Google Scholar] [CrossRef]
  3. Issleib, K.; Lischewski, M. Dimethylamino-Iminophosphorane. Synth. Inorg. Met.-Org. Chem. 1973, 3, 255–266. [Google Scholar] [CrossRef]
  4. Schwesinger, R. Tricyclic 2,4-Diaminovinamidines—Readily Accessible, Very Strong CHN Bases. Angew. Chem. Int. Ed. Engl. 1987, 26, 1164–1165. [Google Scholar] [CrossRef]
  5. Schwesinger, R.; Mißfeldt, M.; Peters, K.; von Schnering, H.G. Novel, Very Strongly Basic, Pentacyclic “Proton Sponges” with Vinamidine Structure. Angew. Chem. Int. Ed. Engl. 1987, 26, 1165–1167. [Google Scholar] [CrossRef]
  6. Schwesinger, R.; Schlemper, H. Peralkylated Polyaminophosphazenes—Extremely Strong, Neutral Nitrogen Bases. Angew. Chem. Int. Ed. Engl. 1987, 26, 1167–1169. [Google Scholar] [CrossRef]
  7. Schwesinger, R.; Hasenfratz, C.; Schlemper, H.; Walz, L.; Peters, E.-M.; Peters, K.; von Schnering, H.G. How Strong and How Hindered Can Uncharged Phosphazene Bases Be? Angew. Chem. Int. Ed. Engl. 1993, 32, 1361–1363. [Google Scholar] [CrossRef]
  8. Schwesinger, R.; Schlemper, H.; Hasenfratz, C.; Willaredt, J.; Dambacher, T.; Breuer, T.; Ottaway, C.; Fletschinger, M.; Boele, J.; Fritz, H.; et al. Extremely Strong, Uncharged Auxiliary Bases; Monomeric and Polymer-Supported Polyaminophosphazenes (P2–P5). Liebigs Ann. 1996, 1996, 1055–1081. [Google Scholar] [CrossRef]
  9. Wang, Y.-H.; Cao, Z.-Y.; Li, Q.-H.; Lin, G.-Q.; Zhou, J.; Tian, P. Activating Pronucleophiles with High pKa Values: Chiral Organo-Superbases. Angew. Chem. Int. Ed. 2020, 59, 8004–8014. [Google Scholar] [CrossRef]
  10. Puleo, T.R.; Sujansky, S.J.; Wright, S.E.; Bandar, J.S. Organic Superbases in Recent Synthetic Methodology Research. Chem. Eur. J. 2021, 27, 4216–4229. [Google Scholar] [CrossRef]
  11. Trofimov, B.A.; Schmidt, E.Y. Superbasis in Organic Synthesis. Chem. Probl. 2022, 20, 325–340. [Google Scholar] [CrossRef]
  12. Vazdar, K.; Margetić, D.; Kovačević, B.; Sundermeyer, J.; Leito, I.; Jahn, U. Design of Novel Uncharged Organic Superbases: Merging Basicity and Functionality. Acc. Chem. Res. 2021, 54, 3108–3123. [Google Scholar] [CrossRef] [PubMed]
  13. Pozharskii, A.F.; Ozeryanskii, V.A.; Filatova, E.A. Heterocyclic superbases: Retrospective and current trends. Chem. Heterocycl. Compd. 2012, 48, 200–219. [Google Scholar] [CrossRef]
  14. Ullrich, S.; Kovačević, B.; Koch, B.; Harms, K.; Sundermeyer, J. Design of non-ionic carbon superbases: Second generation carbodiphosphoranes. Chem. Sci. 2019, 10, 9483–9492. [Google Scholar] [CrossRef]
  15. Vermersch, F.; Yazdani, S.; Junor, G.P.; Grotjahn, D.B.; Jazzar, R.; Bertrand, G. Stable Singlet Carbenes as Organic Superbases. Angew. Chem. Int. Ed. 2021, 60, 27253–27257. [Google Scholar] [CrossRef] [PubMed]
  16. Mehlmann, P.; Mück-Lichtenfeld, C.; Tan, T.T.Y.; Dielmann, F. Tris(imidazolin-2-ylidenamino)phosphine: A crystalline phosphorus(III) superbase that splits carbon dioxide. Chem. Eur. J. 2017, 23, 5929–5933. [Google Scholar] [CrossRef] [PubMed]
  17. Ullrich, S.; Kovačević, B.; Xie, X.; Sundermeyer, J. Phosphazenyl Phosphines: The Most Electron-Rich Uncharged Phosphorus Brønsted and Lewis Bases. Angew. Chem. Int. Ed. 2019, 58, 10335–10339. [Google Scholar] [CrossRef]
  18. Weitkamp, R.F.; Neumann, B.; Stammler, H.-G.; Hoge, B. Phosphorus-Containing Superbases: Recent Progress in the Chemistry of Electron-Abundant Phosphines and Phosphazenes. Chem. Eur. J. 2021, 27, 10807–10825. [Google Scholar] [CrossRef] [PubMed]
  19. Buß, F.; Röthel, M.B.; Werra, J.A.; Rotering, P.; Wilm, L.F.B.; Daniliuc, C.G.; Löwe, P.; Dielmann, F. Tris(tetramethylguanidinyl)phosphine: The Simplest Non-ionic Phosphorus Superbase and Strongly Donating Phosphine Ligand. Chem. Eur. J. 2022, 28, e202104021. [Google Scholar] [CrossRef]
  20. Lv, W.; Dai, Y.; Guo, R.; Su, Y.; Ruiz, D.A.; Liu, L.L.; Tung, C.-H.; Kong, L. Geometrically Constrained Organoboron Species as Lewis Superacids and Organic Superbases. Angew. Chem. Int. Ed. 2023, 62, e202308467. [Google Scholar] [CrossRef]
  21. Schwesinger, R.; Link, R.; Thiele, G.; Rotter, H.; Honert, D.; Limbach, H.-H.; Männle, F. Stable Phosphazenium Ions in Synthesis—An Easily Accessible, Extremely Reactive “Naked” Fluoride Salt. Angew. Chem. Int. Ed. Engl. 1991, 30, 1372–1375. [Google Scholar] [CrossRef]
  22. Schwesinger, R.; Link, R.; Wenzl, P.; Kossek, S. Anhydrous Phosphazenium Fluorides as Sources for Extremely Reactive Fluoride Ions in Solution. Chem. Eur. J. 2005, 12, 438–445. [Google Scholar] [CrossRef] [PubMed]
  23. Kögel, J.F.; Oelkers, B.; Kovačević, B.; Sundermeyer, J. A New Synthetic Pathway to the Second and Third Generation of Superbasic Bisphosphazene Proton Sponges: The Run for the Best Chelating Ligand for a Proton. J. Am. Chem. Soc. 2013, 135, 17768–17774. [Google Scholar] [CrossRef] [PubMed]
  24. Belding, L.; Dudding, T. Synthesis and Theoretical Investigation of a 1,8-Bis(bis(diisopropylamino)cyclopropeniminyl)naphthalene Proton Sponge Derivative. Chem. Eur. J. 2014, 20, 1032–1037. [Google Scholar] [CrossRef] [PubMed]
  25. Kögel, J.F.; Xie, X.; Baal, E.; Gesevičius, D.; Oelkers, B.; Kovačević, B.; Sundermeyer, J. Superbasic Alkyl-Substituted Bisphosphazene Proton Sponges: Synthesis, Structural Features, Thermodynamic and Kinetic Basicity, Nucleophilicity and Coordination Chemistry. Chem. Eur. J. 2014, 20, 7670–7685. [Google Scholar] [CrossRef] [PubMed]
  26. Belding, L.; Stoyanov, P.; Dudding, T. Synthesis, Theoretical Analysis, and Experimental pKa Determination of a Fluorescent, Nonsymmetric, In–Out Proton Sponge. J. Org. Chem. 2016, 81, 6–13. [Google Scholar] [CrossRef] [PubMed]
  27. Kögel, J.F.; Margetić, D.; Xie, X.; Finger, L.H.; Sundermeyer, J. A Phosphorus Bisylide: Exploring a New Class of Superbases with Two Interacting Carbon Atoms as Basicity Centers. Angew. Chem. Int. Ed. 2017, 56, 3090–3093. [Google Scholar] [CrossRef] [PubMed]
  28. Kisanga, P.B.; Verkade, J.G.; Schwesinger, R. pKa Measurements of P(RNCH2CH3)3N. J. Org. Chem. 2000, 65, 5431–5432. [Google Scholar] [CrossRef] [PubMed]
  29. Kisanga, P.B.; Verkade, J.G. Synthesis of new proazaphosphatranes and their application in organic synthesis. Tetrahedron 2001, 57, 467–475. [Google Scholar] [CrossRef]
  30. Uchida, N.; Taketoshi, A.; Kuwabara, J.; Yamamoto, T.; Inoue, Y.; Watanabe, Y.; Kanbara, T. Synthesis, Characterization, and Catalytic Reactivity of a Highly Basic Macrotricyclic Aminopyridine. Org. Lett. 2010, 12, 5242–5245. [Google Scholar] [CrossRef]
  31. Uchida, N.; Kuwabara, J.; Taketoshi, A.; Kanbara, T. Molecular Design of Organic Superbases, Azacalix[3](2,6)pyridines: Catalysts for 1,2- and 1,4-Additions. J. Org. Chem. 2012, 77, 10631–10637. [Google Scholar] [CrossRef] [PubMed]
  32. Power, M.J.; Morris, D.T.J.; Vitorica-Yrezabal, I.J.; Leigh, D.A. Compact Rotaxane Superbases. J. Am. Chem. Soc. 2023, 145, 8583–8599. [Google Scholar] [CrossRef] [PubMed]
  33. Capocasa, G.; Frateloreto, F.; Valentini, M.; Di Stefano, S. Molecular entanglement can strongly increase basicity. Commun. Chem. 2024, 7, 116. [Google Scholar] [CrossRef] [PubMed]
  34. Chambron, J.-C.; Meyer, M. The ins and outs of proton complexation. Chem. Soc. Rev. 2009, 38, 1663–1673. [Google Scholar] [CrossRef] [PubMed]
  35. Miyahara, Y.; Tanaka, Y.; Amimoto, K.; Akazawa, T.; Sakuragi, T.; Kobayashi, H.; Kubota, K.; Suenaga, M.; Koyama, H.; Inazu, T. The Proton Cryptate of Hexaethylenetetramine. Angew. Chem. Int. Ed. 1999, 38, 956–959. [Google Scholar] [CrossRef]
  36. Springborg, J. Adamanzanes–Bi- and tricyclic tetraamines and their coordination compounds. Dalton Trans. 2003, 2003, 1653–1665. [Google Scholar] [CrossRef]
  37. Vazdar, K.; Kunetskiy, R.; Saame, J.; Kaupmees, K.; Leito, I.; Jahn, U. Very Strong Organosuperbases Formed by Combining Imidazole and Guanidine Bases: Synthesis, Structure, and Basicity. Angew. Chem. Int. Ed. 2014, 53, 1435–1438. [Google Scholar] [CrossRef] [PubMed]
  38. Nasca, E.D.; Lambert, T.H. Higher-Order Cyclopropenimine Superbases: Direct Neutral Brønsted Base Catalyzed Michael Reactions with α-Aryl Esters. J. Am. Chem. Soc. 2015, 137, 10246–10253. [Google Scholar] [CrossRef]
  39. Kunetskiy, R.A.; Polyakova, S.M.; Vavřík, J.; Císařová, I.; Saame, J.; Nerut, E.R.; Koppel, I.; Koppel, I.A.; Kütt, A.; Leito, I.; et al. A New Class of Organosuperbases, N-Alkyl- and N-Aryl-1,3-dialkyl-4,5-dimethylimidazol-2-ylidene Amines: Synthesis, Structure, pKBH+ Measurements, and Properties. Chem. Eur. J. 2012, 18, 3621–3630. [Google Scholar] [CrossRef]
  40. Glasovac, Z.; Kovačević, B.; Meštrović, E.; Eckert-Maksić, M. Synthesis and properties of novel guanidine bases. N,N′,N″-Tris(3-dimethylaminopropyl)-guanidine. Tetrahedron Lett. 2005, 46, 8733–8736. [Google Scholar] [CrossRef]
  41. Ullrich, S.; Barić, D.; Xie, X.; Kovačević, B.; Sundermeyer, J. Basicity Enhancement by Multiple Intramolecular Hydrogen Bonding in Organic Superbase N,N′,N″,N‴-Tetrakis(3-(dimethylamino)propyl)triaminophosphazene. Org. Lett. 2019, 21, 9142–9146. [Google Scholar] [CrossRef] [PubMed]
  42. Maksić, Z.B.; Kovačević, B. Spatial and Electronic Structure of Highly Basic Organic Molecules:  Cyclopropeneimines and Some Related Systems. J. Phys. Chem. A 1999, 103, 6678–6684. [Google Scholar] [CrossRef]
  43. Maksić, Z.B.; Kovačević, B. Absolute Proton Affinity of Some Polyguanides. J. Org. Chem. 2000, 65, 3303–3309. [Google Scholar] [CrossRef] [PubMed]
  44. Maksić, Z.B.; Vianello, R. Quest for the Origin of Basicity:  Initial vs Final State Effect in Neutral Nitrogen Bases. J. Phys. Chem. A 2002, 106, 419–430. [Google Scholar] [CrossRef]
  45. Maksić, Z.B.; Glasovac, Z.; Despotović, I. Predicted high proton affinity of poly-2,5-dihydropyrrolimines—The aromatic domino effect. J. Phys. Org. Chem. 2002, 15, 499–508. [Google Scholar] [CrossRef]
  46. Kovačević, B.; Glasovac, Z.; Maksić, Z.B. The intramolecular hydrogen bond and intrinsic proton affinity of neutral organic molecules: N,N′,N″-tris(3-aminopropyl)guanidine and some related systems. J. Phys. Org. Chem. 2002, 15, 765–774. [Google Scholar] [CrossRef]
  47. Kovačević, B.; Maksić, Z.B.; Vianello, R.; Primorac, M. Computer aided design of organic superbases: The role of intramolecular hydrogen bonding. New J. Chem. 2002, 26, 1329–1334. [Google Scholar] [CrossRef]
  48. Bucher, G. DFT Calculations on a New Class of C3-Symmetric Organic Bases: Highly Basic Proton Sponges and Ligands for Very Small Metal Cations. Angew. Chem. Int. Ed. 2003, 42, 4039–4042. [Google Scholar] [CrossRef]
  49. Alder, R.W. Design of C2-Chiral Diamines that Are Computationally Predicted to Be a Million-fold More Basic than the Original Proton Sponges. J. Am. Chem. Soc. 2005, 127, 7924–7931. [Google Scholar] [CrossRef]
  50. Kovačević, B.; Maksić, Z.B. High basicity of phosphorus—Proton affinity of tris-(tetramethylguanidinyl)phosphine and tris-(hexamethyltriaminophosphazenyl)phosphine by DFT calculations. Chem. Commun. 2006, 42, 1524–1526. [Google Scholar] [CrossRef]
  51. Kovačević, B.; Maksić, Z.B. High basicity of tris-(tetramethylguanidinyl)-phosphine imide in the gas phase and acetonitrile—A DFT study. Tetrahedron Lett. 2006, 47, 2553–2555. [Google Scholar] [CrossRef]
  52. Despotović, I.; Maksić, Z.B.; Vianello, R. Engineering Neutral Organic Bases and Superbases by Computational DFT Methods—Carbonyl Polyenes. Eur. J. Org. Chem. 2006, 2006, 5505–5514. [Google Scholar] [CrossRef]
  53. Despotović, I.; Maksić, Z.B.; Vianello, R. Computational design of Brønsted neutral organic superbases—[3]iminoradialenes and quinonimines are important synthetic targets. New J. Chem. 2007, 31, 52–62. [Google Scholar] [CrossRef]
  54. Despotović, I.; Kovačević, B.; Maksić, Z.B. Pyridine and s-triazine as building blocks of nonionic organic superbases—A density functional theory B3LYP study. New J. Chem. 2007, 31, 447–457. [Google Scholar] [CrossRef]
  55. Despotović, I.; Maksić, Z.B.; Vianello, R. Design of Brønsted Neutral Organic Bases and Superbases by Computational DFT Methods: Cyclic and Polycyclic Quinones and [3]Carbonylradialenes. Eur. J. Org. Chem. 2007, 2007, 3402–3413. [Google Scholar] [CrossRef]
  56. Despotović, I.; Kovačević, B.; Maksić, Z.B. Hyperstrong Neutral Organic Bases:  Phosphazeno Azacalix[3](2,6)pyridines. Org. Lett. 2007, 9, 4709–4712. [Google Scholar] [CrossRef]
  57. Margetić, D.; Trošelj, P.; Ishikawa, T.; Kumamoto, T. Design of New Scaffolds for Increased Superbasicity of Bisguanidine Proton Sponges. Bull. Chem. Soc. Jpn. 2010, 83, 1055–1057. [Google Scholar] [CrossRef]
  58. Maksić, Z.B.; Peran, N. Polycyclic croissant-like organic compounds are powerful superbases in the gas phase and acetonitrile—A DFT study. Chem. Commun. 2011, 47, 1327–1329. [Google Scholar] [CrossRef]
  59. Lo, R.; Singh, A.; Kesharwani, M.K.; Ganguly, B. Rational design of a new class of polycyclic organic bases bearing two superbasic sites and their applications in the CO2 capture and activation process. Chem. Commun. 2012, 48, 5865–5867. [Google Scholar] [CrossRef]
  60. Biswas, A.K.; Lo, R.; Ganguly, B. First Principles Studies toward the Design of Silylene Superbases: A Density Functional Theory Study. J. Phys. Chem. A 2013, 117, 3109–3117. [Google Scholar] [CrossRef]
  61. Despotović, I.; Vianello, R. Engineering exceptionally strong oxygen superbases with 1,8-diazanaphthalene di-N-oxides. Chem. Commun. 2014, 50, 10941–10944. [Google Scholar] [CrossRef] [PubMed]
  62. Barić, D.; Dragičević, I.; Kovačević, B. Cyclopropenimine as a hydrogen bond acceptor—Towards the strongest non-phosphorus superbases. Tetrahedron 2014, 70, 8571–8576. [Google Scholar] [CrossRef]
  63. Leito, I.; Koppel, I.A.; Koppel, I.; Kaupmees, K.; Tshepelevitsh, S.; Saame, J. Basicity Limits of Neutral Organic Superbases. Angew. Chem. Int. Ed. 2015, 54, 9262–9265. [Google Scholar] [CrossRef] [PubMed]
  64. Barić, D.; Kovačević, B. Designing a next generation of proton sponges: Cyclopropeniminophosphazenes as the strongest pincer ligands. Tetrahedron Lett. 2016, 57, 442–445. [Google Scholar] [CrossRef]
  65. Barić, D.; Kovačević, B. Cyclopropenimine as pincer ligand and strong electron donor in proton sponges. J. Phys. Org. Chem. 2016, 29, 750–758. [Google Scholar] [CrossRef]
  66. Margetić, D.; Antol, I. A DFT study of endocyclic allenes: Unprecedentedly superbasic hydrocarbons. New J. Chem. 2016, 40, 8191–8193. [Google Scholar] [CrossRef]
  67. Biswas, A.K.; Ganguly, B. Revealing Germylene Compounds to Attain Superbasicity with Sigma Donor Substituents: A Density Functional Theory Study. Chem. Eur. J. 2017, 23, 2700–2705. [Google Scholar] [CrossRef] [PubMed]
  68. Tandarić, T.; Vianello, R. Design of Exceptionally Strong Organic Superbases Based on Aromatic Pnictogen Oxides: Computational DFT Analysis of the Oxygen Basicity in the Gas Phase and Acetonitrile Solution. J. Phys. Chem. A 2018, 122, 1464–1471. [Google Scholar] [CrossRef] [PubMed]
  69. Biswas, A.K.; Si, M.K.; Ganguly, B. The effect of σ/π, σ and π donors on the basicity of silylene superbases: A density functional theory study. New J. Chem. 2018, 42, 11153–11159. [Google Scholar] [CrossRef]
  70. Saadat, K.; Shiri, A.; Kovačević, B. Substituted troponimines: When aromatization of the conjugate acid leads to very strong neutral organic superbases. New J. Chem. 2018, 42, 14568–14575. [Google Scholar] [CrossRef]
  71. Saeidian, H.; Barfinejad, E. Design of Exceptional Strong Organosuperbases Based on Iminophosphorane and Azaphosphiridine Derivatives: Harnessing Ring Strain and Aromaticity to Engineer Neutral Superbases. ChemistrySelect 2019, 4, 3088–3095. [Google Scholar] [CrossRef]
  72. Radić, N.; Maksić, Z.B. Carbon Atom as an Extremely Strong Nucleophilic and Electrophilic Center: Dendritic Allenes Are Powerful Organic Proton and Hydride Sponges. J. Org. Chem. 2019, 84, 2425–2438. [Google Scholar] [CrossRef] [PubMed]
  73. Barić, D. Utilizing the Azaazulene Scaffolds in the Design of New Organic Superbases. ACS Omega 2019, 4, 15197–15207. [Google Scholar] [CrossRef] [PubMed]
  74. Valadbeigi, Y.; Vianello, R. Is It Possible to Achieve Organic Superbases Beyond the Basicity Limit Using Tetrahedrane Scaffolds? ChemistrySelect 2020, 5, 5794–5798. [Google Scholar] [CrossRef]
  75. Saeidian, H.; Mirjafary, Z. Engineering non-ionic carbon super- and hyperbases by a computational DFT approach: Substituted allenes have unprecedented cation affinities. New J. Chem. 2020, 44, 12967–12977. [Google Scholar] [CrossRef]
  76. Saadat, K.; Shiri, A.; Kovačević, B. Step Forward to Stronger Neutral Organic Superbases: Fused Troponimines. J. Org. Chem. 2020, 85, 11375–11381. [Google Scholar] [CrossRef] [PubMed]
  77. Deljuie, F.; Rouhani, M.; Saeidian, H. Exceptional design of super/hyperbases based on spiro-alleneic structures in gas phase: A density functional theory study. J. Phys. Org. Chem. 2022, 35, e4423. [Google Scholar] [CrossRef]
  78. Koneshlou, T.; Rouhani, M.; Saeidian, H.; Aliabad, J.M. Super/hyperbasicity of novel diquinonimino derivatives of guanidine in gas phase. Chem. Phys. Lett. 2022, 804, 139915. [Google Scholar] [CrossRef]
  79. Jalezadeh, A.; Mirjafary, Z.; Rouhani, M.; Saeidian, H. Basicity of five-membered cyclic allenes: Proton and cation affinity evaluation using density functional theory calculations. Int. J. Mass Spectrom. 2022, 482, 116929. [Google Scholar] [CrossRef]
  80. Valadbeigi, Y.; Taheri, R. Superbasicity of imines with bicyclo[5.1.0]octa-1,3,5,7-tetraene scaffold due to electron delocalization in the conjugated acids. Comput. Theor. Chem. 2023, 1222, 114076. [Google Scholar] [CrossRef]
  81. Koneshlou, T.; Rouhani, M.; Saeidian, H.; Aliabad, J.M. Biguanide-dihydropyrimidine dual scaffolds with impressive basicities according to DFT calculations. Comput. Theor. Chem. 2023, 1225, 114178. [Google Scholar] [CrossRef]
  82. Saha, A.; Ganguly, B. Exploiting the (–C–H···C–) Interaction to Design Cage-Functionalized Organic Superbases and Hyperbases: A Computational Study. ACS Omega 2023, 8, 38546–38556. [Google Scholar] [CrossRef]
  83. Yarikordeh, S.; Rouhani, M.; Saeidian, H. Computationally design aspects of superbasic amidine-arsinine 1-oxide binary frameworks. Chem. Phys. Lett. 2023, 833, 140905. [Google Scholar] [CrossRef]
  84. Al-Husseini, S.; Rouhani, M.; Saeidian, H. A brief computational look at the basicity strength of some experimentally observed strained allenes. Chem. Phys. Lett. 2024, 842, 141221. [Google Scholar] [CrossRef]
  85. Kulsha, A.V.; Ragoyja, E.G.; Ivashkevich, O.A. Strong Bases Design: Predicted Limits of Basicity. J. Phys. Chem. A 2022, 126, 3642–3652. [Google Scholar] [CrossRef]
  86. Maksić, Z.B.; Kovačević, B.; Vianello, R. Advances in Determining the Absolute Proton Affinities of Neutral Organic Molecules in the Gas Phase and Their Interpretation: A Theoretical Account. Chem. Rev. 2012, 112, 5240–5270. [Google Scholar] [CrossRef]
  87. Raczyńska, E.D.; Gal, J.-F.; Maria, P.-C. Enhanced Basicity of Push–Pull Nitrogen Bases in the Gas Phase. Chem. Rev. 2016, 116, 13454–13511. [Google Scholar] [CrossRef]
  88. Raczyńska, E.D.; Gal, J.-F.; Maria, P.-C. Strong Bases and beyond: The Prominent Contribution of Neutral Push–Pull Organic Molecules towards Superbases in the Gas Phase. Int. J. Mol. Sci. 2024, 25, 5591. [Google Scholar] [CrossRef]
  89. Kaljurand, I.; Saame, J.; Rodima, T.; Koppel, I.; Koppel, I.A.; Kögel, J.F.; Sundermeyer, J.; Köhn, U.; Coles, M.P.; Leito, I. Experimental Basicities of Phosphazene, Guanidinophosphazene, and Proton Sponge Superbases in the Gas Phase and Solution. J. Phys. Chem. A 2016, 120, 2591–2604. [Google Scholar] [CrossRef]
  90. Lõkov, M.; Tshepelevitsh, S.; Heering, A.; Plieger, P.G.; Vianello, R.; Leito, I. On the Basicity of Conjugated Nitrogen Heterocycles in Different Media. Eur. J. Org. Chem. 2017, 2017, 4475–4489. [Google Scholar] [CrossRef]
  91. Glasovac, Z.; Eckert-Maksić, M.; Maksić, Z.B. Basicity of organic bases and superbases in acetonitrile by the polarized continuum model and DFT calculations. New J. Chem. 2009, 33, 588–597. [Google Scholar] [CrossRef]
  92. 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, 7349–7361. [Google Scholar] [CrossRef]
  93. Rossini, E.; Bochevarov, A.D.; Knapp, E.W. Empirical Conversion of pKa Values between Different Solvents and Interpretation of the Parameters: Application to Water, Acetonitrile, Dimethyl Sulfoxide, and Methanol. ACS Omega 2018, 3, 1653–1662. [Google Scholar] [CrossRef]
  94. 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, 6735–6748. [Google Scholar] [CrossRef]
  95. Glasovac, Z.; Kovačević, B. Modeling pKa of the Brønsted Bases as an Approach to the Gibbs Energy of the Proton in Acetonitrile. Int. J. Mol. Sci. 2022, 23, 10576. [Google Scholar] [CrossRef]
  96. Kulsha, A.V.; Ivashkevich, O.A. Quantum-chemical study of the stability of solvents with respect to strong organic bases. Dokl. Natl. Acad. Sci. Belarus 2023, 67, 380–387. [Google Scholar] [CrossRef]
  97. Ebel, H.F.; Schneider, R. Ionization of Benzylmagnesium Chloride. Angew. Chem. Int. Ed. Engl. 1965, 4, 878. [Google Scholar] [CrossRef]
  98. Clayden, J.; Yasin, S.A. Pathways for decomposition of THF by organolithiums: The role of HMPA. New J. Chem. 2002, 26, 191–192. [Google Scholar] [CrossRef]
  99. Gremmo, N.; Randles, J.E.B. Solvated electrons in hexamethylphosphoramide. Part 1.—Conductivity of solutions of alkali metals. J. Chem. Soc. Faraday Trans. 1 1974, 70, 1480–1487. [Google Scholar] [CrossRef]
  100. Lee, K.P.; Trochimowicz, H.J. Morphogenesis of Nasal Tumors in Rats Exposed to Hexamethylphosphoramide by Inhalation. Environ. Res. 1984, 33, 106–118. [Google Scholar] [CrossRef] [PubMed]
  101. Harman, A.E.; Voigt, J.M.; Frame, S.R.; Bogdanffy, M.S. Mitogenic responses of rat nasal epithelium to hexamethylphosphoramide inhalation exposure. Mutat. Res. Fundam. Mol. Mech. Mutagen. 1997, 380, 155–165. [Google Scholar] [CrossRef] [PubMed]
  102. Kulsha, A.V.; Ivashkevich, O.A. pH Indicators for Strong Molecular Bases: A Theoretical Approach. J. Phys. Chem. A 2024, 128, 4701–4704. [Google Scholar] [CrossRef] [PubMed]
  103. Bockman, T.M.; Kochi, J.K. Isolation and Oxidation-Reduction of Methylviologen Cation Radicals. Novel Disproportionation in Charge-Transfer Salts by X-ray Crystallography. J. Org. Chem. 1990, 55, 4127–4135. [Google Scholar] [CrossRef]
  104. Alder, R.W.; Blake, M.E.; Chaker, L.; Harvey, J.N.; Paolini, F.; Schütz, J. When and How Do Diaminocarbenes Dimerize? Angew. Chem. Int. Ed. 2004, 43, 5896–5911. [Google Scholar] [CrossRef] [PubMed]
  105. Protchenko, A.V.; Birjkumar, K.H.; Dange, D.; Schwarz, A.D.; Vidovic, D.; Jones, C.; Kaltsoyannis, N.; Mountford, P.; Aldridge, S. A Stable Two-Coordinate Acyclic Silylene. J. Am. Chem. Soc. 2012, 134, 6500–6503. [Google Scholar] [CrossRef]
  106. Lui, M.W.; Merten, C.; Ferguson, M.J.; McDonald, R.; Xu, Y.; Rivard, E. Contrasting Reactivities of Silicon and Germanium Complexes Supported by an N-Heterocyclic Guanidine Ligand. Inorg. Chem. 2015, 54, 2040–2049. [Google Scholar] [CrossRef] [PubMed]
  107. Johansen, M.A.L.; Ghosh, A. The curious chemistry of carbones. Nat. Chem. 2023, 15, 1042. [Google Scholar] [CrossRef] [PubMed]
  108. Loh, Y.K.; Melaimi, M.; Munz, D.; Bertrand, G. An Air-Stable “Masked” Bis(imino)carbene: A Carbon-Based Dual Ambiphile. J. Am. Chem. Soc. 2023, 145, 2064–2069. [Google Scholar] [CrossRef]
  109. Lavallo, V.; Dyker, C.A.; Donnadieu, B.; Bertrand, G. Synthesis and Ligand Properties of Stable Five-Membered-Ring Allenes Containing Only Second-Row Elements. Angew. Chem. Int. Ed. 2008, 47, 5411–5414. [Google Scholar] [CrossRef]
  110. Melaimi, M.; Parameswaran, P.; Donnadieu, B.; Frenking, G.; Bertrand, G. Synthesis and Ligand Properties of a Persistent, All-Carbon Four-Membered-Ring Allene. Angew. Chem. Int. Ed. 2009, 48, 4792–4795. [Google Scholar] [CrossRef]
  111. Ariai, J.; Ziegler, M.; Würtele, C.; Gellrich, U. An N-Heterocyclic Quinodimethane: A Strong Organic Lewis Base Exhibiting Diradical Reactivity. Angew. Chem. Int. Ed. 2024, 63, e202316720. [Google Scholar] [CrossRef]
  112. Erdmann, P.; Leitner, J.; Schwarz, J.; Greb, L. An Extensive Set of Accurate Fluoride Ion Affinities for p-Block Element Lewis Acids and Basic Design Principles for Strong Fluoride Ion Acceptors. ChemPhysChem 2020, 21, 987–994. [Google Scholar] [CrossRef]
  113. Dempsey, S.H.; Kass, S.R. Liberating the Anion: Evaluating Weakly Coordinating Cations. J. Org. Chem. 2022, 87, 15466–15482. [Google Scholar] [CrossRef]
  114. Brown, S.J.; Clark, J.H. Tetraphenylfluorophosphorane. J. Chem. Soc. Chem. Commun. 1983, 19, 1256–1257. [Google Scholar] [CrossRef]
  115. Brüggeller, P. Five-coordinate platinum(II) hydrides containing 1,1,4,7,10,10-hexaphenyl-1,4,7,10-tetraphosphadecane as a tetradentate monometallic ligand. Inorg. Chem. 1990, 29, 1742–1750. [Google Scholar] [CrossRef]
  116. King, R.B.; Cloyd, J.C., Jr. Poly(tertiary phosphines and arsines). X. Synthesis of methylated poly(tertiary phosphines). J. Am. Chem. Soc. 1975, 97, 53–60. [Google Scholar] [CrossRef]
  117. Bampos, N.; Field, L.D.; Messerle, B.A.; Smernik, R.J. Synthesis of new tetradentate oligophosphine ligands. Inorg. Chem. 1993, 32, 4084–4088. [Google Scholar] [CrossRef]
  118. Qi, X.-J.; Liu, L.; Fu, Y.; Guo, Q.-X. Ab Initio Calculations of pKa Values of Transition-Metal Hydrides in Acetonitrile. Organometallics 2006, 25, 5879–5886. [Google Scholar] [CrossRef]
  119. Morris, R.H. Brønsted–Lowry Acid Strength of Metal Hydride and Dihydrogen Complexes. Chem. Rev. 2016, 116, 8588–8654. [Google Scholar] [CrossRef]
  120. Schwesinger, R. Starke ungeladene Stickstoffbasen. Nachr. Chem. Tech. Lab. 1990, 38, 1214–1226. [Google Scholar] [CrossRef]
  121. Curtis, C.J.; Miedaner, A.; Ellis, W.W.; DuBois, D.L. Measurement of the Hydride Donor Abilities of [HM(diphosphine)2]+ Complexes (M = Ni, Pt) by Heterolytic Activation of Hydrogen. J. Am. Chem. Soc. 2002, 124, 1918–1925. [Google Scholar] [CrossRef] [PubMed]
  122. Drews, T.; Rusch, D.; Seidel, S.; Willemsen, S.; Seppelt, K. Systematic Reactions of [Pt(PF3)4]. Chem. Eur. J. 2008, 14, 4280–4286. [Google Scholar] [CrossRef] [PubMed]
  123. Ozari, Y.; Jagur-Grodzinski, J. Donor strength of N-substituted phosphoramides. J. Chem. Soc. Chem. Commun. 1974, 10, 295–296. [Google Scholar] [CrossRef]
  124. Liakos, D.G.; Sparta, M.; Kesharwani, M.K.; Martin, J.M.L.; Neese, F. Exploring the Accuracy Limits of Local Pair Natural Orbital Coupled-Cluster Theory. J. Chem. Theory Comput. 2015, 11, 1525–1539. [Google Scholar] [CrossRef]
  125. Liakos, D.G.; Neese, F. Is It Possible to Obtain Coupled Cluster Quality Energies at near Density Functional Theory Cost? Domain-Based Local Pair Natural Orbital Coupled Cluster vs Modern Density Functional Theory. J. Chem. Theory Comput. 2015, 11, 4054–4063. [Google Scholar] [CrossRef]
  126. Riplinger, C.; Pinski, P.; Becker, U.; Valeev, E.F.; Neese, F. Sparse maps—A systematic infrastructure for reduced-scaling electronic structure methods. II. Linear scaling domain based pair natural orbital coupled cluster theory. J. Chem. Phys. 2016, 144, 024109. [Google Scholar] [CrossRef]
  127. Saitow, M.; Becker, U.; Riplinger, C.; Valeev, E.F.; Neese, F. A new near-linear scaling, efficient and accurate, open-shell domain-based local pair natural orbital coupled cluster singles and doubles theory. J. Chem. Phys. 2017, 146, 164105. [Google Scholar] [CrossRef]
  128. Guo, Y.; Riplinger, C.; Becker, U.; Liakos, D.G.; Minenkov, Y.; Cavallo, L.; Neese, F. Communication: An improved linear scaling perturbative triples correction for the domain based local pair-natural orbital based singles and doubles coupled cluster method [DLPNO-CCSD(T)]. J. Chem. Phys. 2018, 148, 011101. [Google Scholar] [CrossRef]
  129. Mallick, S.; Roy, B.; Kumar, P. A comparison of DLPNO-CCSD(T) and CCSD(T) method for the determination of the energetics of hydrogen atom transfer reactions. Comput. Theor. Chem. 2020, 1187, 112934. [Google Scholar] [CrossRef]
  130. Sandler, I.; Chen, J.; Taylor, M.; Sharma, S.; Ho, J. Accuracy of DLPNO-CCSD(T): Effect of Basis Set and System Size. J. Phys. Chem. A 2021, 125, 1553–1563. [Google Scholar] [CrossRef]
  131. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  132. Barone, V.; Cossi, M. Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent Model. J. Phys. Chem. A 1998, 102, 1995–2001. [Google Scholar] [CrossRef]
  133. Rappé, A.K.; Casewit, C.J.; Colwell, K.S.; Goddard III, W.A.; Skiff, W.M. UFF, a full periodic table force field for molecular mechanics and molecular dynamics simulations. J. Am. Chem. Soc. 1992, 114, 10024–10035. [Google Scholar] [CrossRef]
  134. Mahajan, G.R.; Kumbharkhane, A.C. Dielectric relaxation study of hexamethylphosphoramide—1,4-dioxane mixtures using time domain reflectometry (TDR) technique. Phys. Chem. Liq. 2012, 50, 513–522. [Google Scholar] [CrossRef]
  135. Simon, S.; Duran, M.; Dannenberg, J.J. How does basis set superposition error change the potential surfaces for hydrogen-bonded dimers? J. Chem. Phys. 1996, 105, 11024–11031. [Google Scholar] [CrossRef]
  136. Neese, F. Software update: The ORCA program system—Version 5.0. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2022, 12, e1606. [Google Scholar] [CrossRef]
  137. 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.; et al. Gaussian 16, Revision C.02; Gaussian, Inc.: Wallingford, CT, USA, 2019. [Google Scholar]
  138. Dunning, T.H., Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  139. Weigend, F.; Köhn, A.; Hättig, C. Efficient use of the correlation consistent basis sets in resolution of the identity MP2 calculations. J. Chem. Phys. 2002, 116, 3175–3183. [Google Scholar] [CrossRef]
  140. Woon, D.E.; Dunning, T.H., Jr. Gaussian basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon. J. Chem. Phys. 1993, 98, 1358–1371. [Google Scholar] [CrossRef]
  141. Peterson, K.A. Systematically convergent basis sets with relativistic pseudopotentials. I. Correlation consistent basis sets for the post-d group 13–15 elements. J. Chem. Phys. 2003, 119, 11099–11112. [Google Scholar] [CrossRef]
  142. Metz, B.; Stoll, H.; Dolg, M. Small-core multiconfiguration-Dirac–Hartree–Fock-adjusted pseudopotentials for post-d main group elements: Application to PbH and PbO. J. Chem. Phys. 2000, 113, 2563–2569. [Google Scholar] [CrossRef]
  143. Stoychev, G.L.; Auer, A.A.; Neese, F. Automatic Generation of Auxiliary Basis Sets. J. Chem. Theory Comput. 2017, 13, 554–562. [Google Scholar] [CrossRef] [PubMed]
  144. Figgen, D.; Peterson, K.A.; Dolg, M.; Stoll, H. Energy-consistent pseudopotentials and correlation consistent basis sets for the 5d elements Hf–Pt. J. Chem. Phys. 2009, 130, 164108. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Discovery timeline of the strongest molecular bases with nitrogen basicity centers.
Figure 1. Discovery timeline of the strongest molecular bases with nitrogen basicity centers.
Ijms 25 08716 g001
Scheme 1. The most basic known carbene 1 [15], phosphine 2 [17], azene 3 [7], and borylene 4 [20]. Protonation sites are shown in bold.
Scheme 1. The most basic known carbene 1 [15], phosphine 2 [17], azene 3 [7], and borylene 4 [20]. Protonation sites are shown in bold.
Ijms 25 08716 sch001
Scheme 2. Sample superbasic structures and their degradation ways.
Scheme 2. Sample superbasic structures and their degradation ways.
Ijms 25 08716 sch002
Scheme 3. Degradation of the superbase 7 proposed in ref. [68].
Scheme 3. Degradation of the superbase 7 proposed in ref. [68].
Ijms 25 08716 sch003
Scheme 4. The proposed “homologue” of TBD and its degradation.
Scheme 4. The proposed “homologue” of TBD and its degradation.
Ijms 25 08716 sch004
Scheme 5. Degradation of hexaethylenetetramine 9 to its opened-cage isomer 9a.
Scheme 5. Degradation of hexaethylenetetramine 9 to its opened-cage isomer 9a.
Ijms 25 08716 sch005
Scheme 6. Degradation of the substituted tetraaminoallene 10.
Scheme 6. Degradation of the substituted tetraaminoallene 10.
Ijms 25 08716 sch006
Scheme 7. Another proposed superbase 11, its tautomer 11a, and dimer 11b.
Scheme 7. Another proposed superbase 11, its tautomer 11a, and dimer 11b.
Ijms 25 08716 sch007
Scheme 8. Predicted dimerization of the strongest known carbene superbase 1.
Scheme 8. Predicted dimerization of the strongest known carbene superbase 1.
Ijms 25 08716 sch008
Scheme 9. Proposed weakly coordinating cation 12+ and its degradation.
Scheme 9. Proposed weakly coordinating cation 12+ and its degradation.
Ijms 25 08716 sch009
Scheme 10. Presumably stable superbases designed by us [85,96].
Scheme 10. Presumably stable superbases designed by us [85,96].
Ijms 25 08716 sch010
Scheme 11. Proposed synthetic route of a new pH indicator for extremely basic media.
Scheme 11. Proposed synthetic route of a new pH indicator for extremely basic media.
Ijms 25 08716 sch011
Scheme 12. New presumably stable superbases 15ad and pKa values of their protonated forms.
Scheme 12. New presumably stable superbases 15ad and pKa values of their protonated forms.
Ijms 25 08716 sch012
Table 1. Calculated pKa(BH+) values in HMPA for a few selected bases [85,102].
Table 1. Calculated pKa(BH+) values in HMPA for a few selected bases [85,102].
Base BpKa(BH+)
pyridine1.94
Et3N8.88
TBD17.01
SPS25.78
tBu-P435.65
carbene 139.28
phosphine 240.59
azene 340.82
borylene 440.87
Table 2. Calculated pKa values for various forms of 14a and 14b in HMPA.
Table 2. Calculated pKa values for various forms of 14a and 14b in HMPA.
Protonation Degree14a14b
neutral base53.0254.67
monoprotonated49.0452.24
diprotonated10.3513.01
triprotonated5.039.48
Table 3. Basis sets and ECPs utilized for the calculations.
Table 3. Basis sets and ECPs utilized for the calculations.
AtomsBasis SetECPAuxiliary Basis Set
for DLPNO-CCSD(T)
H–Fcc-pVTZ [138]cc-pVTZ/C [139]
Pcc-pVTZ [140]cc-pVTZ/C [139]
Ascc-pVTZ-PP [141]Stuttgart-Köln [142]AutoAux [143]
Ptcc-pVTZ-PP [144]Stuttgart-Köln [144]AutoAux [143]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kulsha, A.V.; Ivashkevich, O.A.; Lyakhov, D.A.; Michels, D. Strong Bases Design: Key Techniques and Stability Issues. Int. J. Mol. Sci. 2024, 25, 8716. https://doi.org/10.3390/ijms25168716

AMA Style

Kulsha AV, Ivashkevich OA, Lyakhov DA, Michels D. Strong Bases Design: Key Techniques and Stability Issues. International Journal of Molecular Sciences. 2024; 25(16):8716. https://doi.org/10.3390/ijms25168716

Chicago/Turabian Style

Kulsha, Andrey V., Oleg A. Ivashkevich, Dmitry A. Lyakhov, and Dominik Michels. 2024. "Strong Bases Design: Key Techniques and Stability Issues" International Journal of Molecular Sciences 25, no. 16: 8716. https://doi.org/10.3390/ijms25168716

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop