Next Article in Journal
Cecropin-Loaded Zeolitic Imidazolate Framework Nanoparticles with High Biocompatibility and Cervical Cancer Cell Toxicity
Previous Article in Journal
Novel C60 Fullerenol-Gentamicin Conjugate–Physicochemical Characterization and Evaluation of Antibacterial and Cytotoxic Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Functionalized C3-Symmetric Building Blocks—The Chemistry of Triaminotrimesic Acid

1
Institute of Organic Chemistry, Karlsruhe Institute of Technology (KIT), Fritz-Haber-Weg 6, 76131 Karlsruhe, Germany
2
Department of Chemistry, University of Helsinki, P.O. Box 55 (A. I. Virtasen aukio 1), FIN-00014 Helsinki, Finland
3
Institute of Biological and Chemical Systems—FMS, Karlsruhe Institute of Technology (KIT), Hermann-Von-Helmholtz-Platz 1, 76344 Leopoldshafen, Germany
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(14), 4369; https://doi.org/10.3390/molecules27144369
Submission received: 30 May 2022 / Revised: 29 June 2022 / Accepted: 1 July 2022 / Published: 7 July 2022

Abstract

:
A series of C3-symmetric fully substituted benzenes were prepared based on alkyl triamino-benzene-tricarboxylates. Starting with a one step-synthesis, the alkyl triamino-benzene-tricarboxylates were synthesized using the corresponding cyanoacetates. The reactivity of these electronically sophisticated compounds was investigated by the formation of azides, the click reaction of the azides and a Sandmeyer-like reaction. Caused by the low stability of triaminobenzenes, direct N-alkylation was rarely reported. The use of the stable alkyl triamino-benzene-tricarboxylates allowed us total N-alkylation under standard alkylation conditions. The molecular structures of the C3-symmetric structures have been corroborated by an X-ray analysis.

1. Introduction

The symmetry of molecular building blocks plays a pivotal role in the overall geometry of the materials that are formed. Besides many examples of C2-symmetric [1] building blocks, the C3-symmetrical ones have found fewer applications, other than in life sciences [2] or material sciences. However, C3-symmetrical-based geometries can be found in star-shaped molecules, dendrimers, and molecular cages [3], thus allowing the formation of columnar structuring due to strong π-π interactions in the case of appropriately functionalized monomers [4] (for reviews, see Ref. [5]). The applications range from discotic liquid crystals [6,7]; mesogens [8]; OLED emitters featuring thermally-activated delayed fluorescence; [9,10,11] gels; [12] metal-organic frameworks (MOFs) (trimesic acid/BTC: MOF-177 [13]; MOF-199; 2D-covalent organic frameworks (COFs) [14,15]; hole-transporting materials for photovoltaics; [16] materials for second harmonic generation/non-linear optics by generating octopoles [17,18,19]; hydrogels [20,21] and many more [22]. Noteworthy in this context are functionalized truxenes [23,24] such as truxenones [25,26,27], or triazatruxenes [28,29] (e.g., D, Figure 1), which exhibit outstanding photophysical properties [30,31,32,33].
Fully substituted derivatives of type A (X, Y are non-hydrogen atoms, Figure 1) also exhibit additional properties due to steric crowding. The non-planarity of some derivatives leads to different functionalization of the side groups, e.g., carboxylic acids (Figure 1, Structure B) [34].
The synthesis of such molecules can either start from condensation reactions (mostly carbonyl compounds in aldol-like reactions), the cyclotrimerization reactions of alkynes (Reppe-type chemistry) [35], or by the functionalization of appropriately equipped building blocks of type A, such as triesters, triamines and trihalides [36] (Figure 1).
In particular, cores featuring nitrogen and/or carbon-based functionalized groups have been successfully used in many materials. An overview is given in Table S1.
Most of the syntheses for aminocarboxy-substituted arenes start with an internal acylation of triaminobenzene. However, the combination of nitrogen and oxidized carbon-based functionalized groups have not been reported in depth. Methyl triaminobenzenetricarboxylate—the only derivative known so far—was used in only one patent, wherein 2,4,6,8,10,12-hexacyano-1,3,5,7,9,11-hexa-azatriphenylene, a derivative of HAT-CN was prepared from trimethyl 2,4,6-triaminobenzene-1,3,5-tricarboxylate (2a) via thermal cyclocondensation with urea; halogenation with POCl3 in PhNMe2; and cyanolysis with KCN in MeCN [37].
Herein, it is our intention to report the syntheses, structures and reactivities of novel C3-symmetric fully substituted benzenes with the tandem amino/alkylcarboxylate groups, based on the easily manageable cyclotrimerization of alkyl cyanoacetates. These electronically sophisticated structures of alkyl triamino-benzene-tricarboxylates led to extremely challenging subsequent reactions of the amine group. However, azide formation and a Sandmeyer-like reaction were studied within this report. Due to the low stability of triaminobenzenes, direct N-alkylation was rarely reported. Using the stable alkyl triamino-benzene-tricarboxylates allowed us total N-alkylation. An overview of the synthesized building blocks is given in Scheme 1.

2. Results

2.1. Syntheses of C3-Symmetric Alkyl Triamino-Benzene-Tricarboxylates

According to the literature-known synthesis for 2a, starting with the methylcyanoacetate, methyl triamino-benzene-tricarboxylate was synthesized in a moderate yield [38]. Despite extensive experimentation, the yield of the methyl triamino-benzene-tricarboxylate 2a could not be improved in our hands. However, this atom-economic reaction is scalable and provides the required building block in multi-gram amounts. Besides the methyl triamino-benzene-tricarboxylate 2a, cyclotrimerization of the alkylcyanoacetates 1bg led to the derivatives 2bg in moderate yields (Table 1).
In comparison to 1,3,5-triaminobenzene and several other derivatives, which darken slowly after being left in air, the synthesized alkyl triamino-benzene-tricarboxylates do not show a color change after several months stored in air [39].

2.2. Diazotation and Azide Formation

The diazotization and conversion of triaminobenzenes have been reported [40]. Although 1,3,5-triaazidobenzenes are known and reported to be stable, most publications only deal with the theoretical calculations of these molecules [41,42,43,44,45,46,47,48], and there are only around 20 reported examples (including tetra/penta/hexa-azides) [40,43,44,49,50,51,52,53,54,55].
Herein, we report the syntheses of 1,3,5-triazidobenzenes, substituted with ester groups in a 2,4,6-position. The syntheses of triazides 3ag proceeded from the triamines 2ag under established conditions through a diazotization reaction. The triazides 3ag were obtained in yields of 36% to 60% (Table 2). The yields that were achieved for the triazides 3ag were in the range of other reported triazides that were synthesized from the corresponding amines [40,44]. A procedure in which the diazotization was carried out with tert-butyl nitrite and tosylic acid at 21 °C, followed by the addition of sodium azide, did not result in the formation of the triazide 3a. Triazides 3ac and 3g were stable at normal conditions (daylight included) for several weeks. In the case of the alkyl triamino-benzene-tricarboxylates 2d, 2e and 2f, the mono- and diazides that were also formed during the conversion to the triazide could not be separated from the corresponding triazides via column chromatography. Therefore, these structures are not listed in Table 2.
Cautionary note: azides with a C/N ratio of around 1:1 are potentially explosive [56,57].
In the following, the deprotection of the ester was successfully performed to give the triazidobenzene-tricarboxylic acid 4 in a moderate yield of 60% (Scheme 2). These structures might serve as interesting building blocks, e.g., for the generation of trinitrenes [58,59,60] or the synthesis of HKUST1-comparable MOFs.
Due to the ability to form the azide, we thought that reactions based on the diazonium salt, such as Sandmeyer reactions, should be possible. Nevertheless, Sandmeyer-like reactions using tert-butyl nitrite or sodium nitrite/hydrochloric acid and potassium iodide failed several times. Many attempts were necessary until conditions were found, which lead to a triple-halogenated compound 5. This method uses tert-butyl nitrite for diazotization and TMS-bromide for halogen transfer. Upon optimization, we find that the addition of the diazotization compound and TMS-bromide must take place alternately, leading to the tribromide 5 in a moderate yield of 34% (Scheme 3). This indicates that diazotization can only be performed at one amine group at a time. Compared to another synthesis route, which has five reaction steps from mesitylene to methyl tribromobenzene tricarboxylate 5, this route only needs two steps, starting with methylcyanoacetate [8,14].

2.3. Click Reactions

Tris-1,2,3-triazoles originating from triazides of type 3 are unknown, except for a single benzotriazole [61] and theoretical investigations [62]. In our hands, the click chemistry that was applied in the case of triazide 3a in a reaction with phenyl ethyne and p-bromophenyl ethyne gave the triazoles 6a-Ph and 6a-C6H4Br, respectively (Table 3). A click reaction with the triazide 3b, containing an ethyl ester instead of the methyl ester of 3a, resulted in triazole 6b-Ph. The yield of 6b-Ph is noticeably better than that of 6a-Ph, which can be explained by the better solubility of the ethyl ester 3b and intermediates on the way to compound 6b-Ph. Substituted alkynes enable different reactions of these molecules, e.g., network building via dialkynes or coupling reactions of the bromo-substituted tristriazole 6a-C6H4Br.

2.4. Alkylation

In the past, the direct N-alkylation of 1,3,5-triaminobenzene and many derivatives was not possible due to the low stability of the amines. Therefore, benzene-1,3,5-triol or 1,3,5-halogenated structures and secondary amines were commonly used to synthesize alkylated 1,3,5-triamino benzenes [63,64,65].
Using the stable methyl triamino-benzene-tricarboxylate 2a enabled the study of the total N-alkylation of 1,3,5- triamino benzenes. To investigate the reactivity of the methyl triamino-benzene-tricarboxylate 2a, we used different alkyl iodides. The previously unknown hexa-alkyl-triamines 7ac were successfully synthesized. Purification via column chromatography gave the alkylated structures 7ac in yields between 38% and 51% (Table 4). The reaction was performed at 100 °C and stirred for 16 h in case of 7a and 2d, in case of 7b and 7c. TLC of the crude reaction mixtures showed less alkylated fractions. Nevertheless, longer reaction times could not improve the yields.

2.5. Molecular Structures

The structures of the alkyl triamino-benzene-tricarboxylates 2ac were additionally confirmed by X-ray crystallography. The molecular structures show the possible formation of hydrogen bonds between the amino and ester groups, leading to the expected planar structure of 2. While 2a showed a planar propeller-like arrangement, the sterically more demanding ester groups of 2b and 2c twisted the functional groups marginally out of the plane (Figure 2).
Beside the triamines 2ac, the structure of 3a was also confirmed by the molecular structure, but due to the poor crystal quality we will not discuss this further (Figure 3).

3. Materials and Methods

3.1. General Procedure for Cyclotrimerizations

A pressure tube was charged with Cu(OAc)2*H2O (0.10 equiv.) and 1,4-dioxane. Alkyl cyanoacetate (1.00 equiv.) was added, and the mixture was bubbled with argon for 5 min. The mixture was heated to 130 °C for 72 h. After cooling to room temperature, the solid was filtered off, and the solvent was removed under reduced pressure. The product was purified by column chromatography (cyclohexane/ethyl acetate).

3.2. General Procedure for the Syntheses of Azides

Trialkyl 2,4,6-triaminobenzene-1,3,5-tricarboxylate (1.00 equiv.) was solved in THF and cooled to 0 °C. tert-Butyl nitrite (9.00 equiv.) was added dropwise. The mixture was stirred for 30 minutes, followed by the addition azido(trimethyl)silane (slow, 6.00 equiv.). The mixture was stirred for 72 h. The solvent was (carefully) removed under reduced pressure, and the residue was purified by column chromatography (cyclohexane/ethyl acetate).

3.3. Typical Procedure for the Click Reactions

3a (100 mg, 240 μmol, 1.00 equiv) and ethynylbenzene (85.7 mg, 839 μmol, 3.50 equiv.) were solved under argon in degassed DMSO (2.50 mL). Copper sulfate pentahydrate (8.97 mg, 35.9 μmol, 0.150 equiv) and sodium ascorbate (14.2 mg, 71.9 μmol, 0.300 equiv) were solved in degassed water (500 μL) and degassed DMSO (2.50 mL). The mixture was added dropwise to the ethynylbenzene-solution. The mixture was stirred at 50 °C for 3 d. After cooling to 25 °C, ethyl acetate (20 mL) and water (20 mL) were added, and the phases were separated. The organic layer was washed with brine (20 mL) and then dried over sodium sulfate. The solvent was removed under reduced pressure, and the residue was purified via flash chromatography (cyclohexane/ethyl acetate 20:1 to 4:1) to give the desired product 6a-Ph as a light-yellow solid.

3.4. Crystal Structure Determination

The single-crystal X-ray diffraction studies were carried out on a Bruker D8 Venture diffractometer with a PhotonII detector at 123(2) K; 173(2) K; or 298(2) K using Cu-Kα radiation (λ = 1.54178 Å). Dual space methods (SHELXT) [66] were used for the structure solution, and refinement was carried out using SHELXL (full-matrix least-squares on F2) [67]. Hydrogen atoms were localized by difference electron density determination and refined using a riding model (H(N, O) free). Semi-empirical absorption corrections were applied. CCDC 2,102,766 (2a); 2,102,767 (2b); and 2,102,768 (2c) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif (accessed on 12 August 2021). Due to the bad quality of the data of 3a (completeness approx. 82%), the data were not deposited with The Cambridge Crystallographic Data Centre).

3.5. NMR Measurements

The NMR spectra were recorded at 25 °C on an Bruker Avance 400 NMR instrument. More details on the NMR measurements can be found in the Supplementary Information.

4. Conclusions

Different C3-symmetric building blocks based on alkyl triamino-benzene-tricarboxylates have been reported in this manuscript. Despite only moderate yields, the simplicity of the syntheses allowed gram amounts of the alkyl triamino-benzene-tricarboxylates. Starting from the remarkably stable alkyl-triamino-benzene-tricarboxylates, we investigated azide formation and Sandmeyer-like reactions, as well as chemo-selective N-hexa-alkylation. More interestingly, click reactions were possible with the synthesized triazides, allowing further studies on the formation of porous organic polymers (POP). With the triazidobenzene-tricarboxylic acid, we have presented a building block that can be used, for example, in a similar way to trimesic acid in the synthesis of MOFs.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/molecules27144369/s1: synthetic procedure, crystallographic, and NMR data.

Author Contributions

Conceptualization, L.S.; methodology, L.S. and S.B.; validation, L.S.; formal analysis, L.S. and D.W.; investigation, L.S. and D.W.; resources, S.B.; data curation, L.S., D.W. and M.N.; writing—original draft preparation, L.S. and M.N.; writing—review and editing, S.B.; visualization, L.S. and M.N.; supervision, S.B.; project administration, S.B.; funding acquisition, S.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The procedures, as well as the analytical data, are available online in the Chemotion Repository (https://www.chemotion-repository.net/welcome (accessed on 28 June 2022)). Exact links are given in the Supplementary Information.

Acknowledgments

This work is supported by the Helmholtz Association Program at the Karlsruhe Institute of Technology. The Deutsche Forschungsgemeinschaft under Germany’s Excellence Strategy 3DMM2O—EXC-2082/1–390761711 is gratefully acknowledged. We thank Philip Stößel (Merck) for stimulating discussions.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not available.

References and Notes

  1. Through this manuscript, we refer to C3-symmetric building blocks: D3-symmetric ones are also included.
  2. Lin, Y.; Wu, X.; Feng, S.; Jiang, G.; Luo, J.; Zhou, S.; Vrijmoed, L.L.P.; Jones, E.B.G.; Krohn, K.; Steingroever, K.; et al. Five Unique Compounds: Xyloketals from Mangrove Fungus Xylaria sp. from the South China Sea Coast. J. Org. Chem. 2001, 66, 6252–6256. [Google Scholar] [CrossRef] [PubMed]
  3. Tamura, Y.; Takezawa, H.; Fujita, M. A Double-Walled Knotted Cage for Guest-Adaptive Molecular Recognition. J. Am. Chem. Soc. 2020, 142, 5504–5508. [Google Scholar] [CrossRef] [PubMed]
  4. Park, L.Y.; Hamilton, D.G.; McGehee, E.A.; McMenimen, K.A. Complementary C3-Symmetric Donor–Acceptor Components:  Cocrystal Structure and Control of Mesophase Stability. J. Am. Chem. Soc. 2003, 125, 10586–10590. [Google Scholar] [CrossRef] [PubMed]
  5. Wolff, J.J. Structures of benzene derivatives with symmetrical substitution by three donor and three acceptor groups. Adv. Strained Interesting Org. Mol. 1999, 7, 43–101. [Google Scholar] [CrossRef]
  6. Harada, Y.; Matsunaga, Y. New mesomorphic compounds: N,N′,N″-trialkanoyl-2,4,6-trimethyl-1,3,5-benzenetriamines. Bull. Chem. Soc. Jpn. 1988, 61, 2739–2741. [Google Scholar] [CrossRef] [Green Version]
  7. Omenat, A.; Barbera, J.; Serrano, J.L.; Houbrechts, S.; Persoons, A. Columnar liquid crystals with highly polar groups. Evaluation of the nonlinear optical properties. Adv. Mater. 1999, 11, 1292–1295. [Google Scholar] [CrossRef]
  8. Bushey, M.L.; Nguyen, T.-Q.; Nuckolls, C. Synthesis, Self-Assembly, and Switching of One-Dimensional Nanostructures from New Crowded Aromatics. J. Am. Chem. Soc. 2003, 125, 8264–8269. [Google Scholar] [CrossRef]
  9. Mamada, M.; Inada, K.; Komino, T.; Potscavage, W.J.; Nakanotani, H.; Adachi, C. Highly Efficient Thermally Activated Delayed Fluorescence from an Excited-State Intramolecular Proton Transfer System. ACS Cent. Sci. 2017, 3, 769–777. [Google Scholar] [CrossRef]
  10. Long, Y.; Mamada, M.; Li, C.; dos Santos, P.L.; Colella, M.; Danos, A.; Adachi, C.; Monkman, A. Excited State Dynamics of Thermally Activated Delayed Fluorescence from an Excited State Intramolecular Proton Transfer System. J. Phys. Chem. Lett. 2020, 11, 3305–3312. [Google Scholar] [CrossRef] [Green Version]
  11. Jiang, G.; Li, F.; Fan, J.; Song, Y.; Wang, C.-K.; Lin, L. Theoretical perspective for luminescent mechanism of thermally activated delayed fluorescence emitter with excited-state intramolecular proton transfer. J. Mater. Chem. C 2020, 8, 98–108. [Google Scholar] [CrossRef]
  12. Tang, L.M.; Wang, Y.J. Highly stable supramolecular hydrogels formed from 1,3,5-benzenetricarboxylic acid and hydroxyl pyridines. Chin. Chem. Lett. 2009, 20, 1259–1262. [Google Scholar] [CrossRef]
  13. Chae, H.K.; Siberio-Perez, D.Y.; Kim, J.; Go, Y.B.; Eddaoudi, M.; Matzger, A.J.; O’Keeffe, M.; Yaghi, O.M. A route to high surface area, porosity and inclusion of large molecules in crystals. Nature 2004, 427, 523–527. [Google Scholar] [CrossRef] [PubMed]
  14. Alahakoon, S.; Tan, K.; Pandey, H.; Diwakara, S.D.; McCandless, G.T.; Grinffiel, D.I.; Durand-Silva, A.; Thonhauser, T.; Smaldone, R.A. 2D-Covalent Organic Frameworks with Interlayer Hydrogen Bonding Oriented through Designed Nonplanarity. J. Am. Chem. Soc. 2020, 142, 12987–12994. [Google Scholar] [CrossRef]
  15. Wei, H.; Ning, J.; Cao, X.; Li, X.; Hao, L. Benzotrithiophene-Based Covalent Organic Frameworks: Construction and Structure Transformation under Ionothermal Condition. J. Am. Chem. Soc. 2018, 140, 11618–11622. [Google Scholar] [CrossRef] [PubMed]
  16. Molina-Ontoria, A.; Zimmermann, I.; Garcia-Benito, I.; Gratia, P.; Roldan-Carmona, C.; Aghazada, S.; Graetzel, M.; Nazeeruddin, M.K.; Martin, N. Benzotrithiophene-Based Hole-Transporting Materials for 18.2% Perovskite Solar Cells. Angew. Chem. Int. Ed. 2016, 55, 6270–6274. [Google Scholar] [CrossRef] [Green Version]
  17. Wolff, J.J.; Siegler, F.; Matschiner, R.; Wortmann, R. Optimized two-dimensional NLO chromophores with a threefold symmetry axis. Angew. Chem. Int. Ed. 2000, 39, 1436–1439. [Google Scholar] [CrossRef]
  18. Ledoux, I.; Zyss, J.; Siegel, J.S.; Brienne, J.; Lehn, J.M. Second-harmonic generation from nondipolar noncentrosymmetric aromatic charge-transfer molecules. Chem. Phys. Lett. 1990, 172, 440–444. [Google Scholar] [CrossRef]
  19. Cho, B.R.; Lee, S.J.; Lee, S.H.; Son, K.H.; Kim, Y.H.; Doo, J.-Y.; Lee, G.J.; Kang, T.I.; Lee, Y.K.; Cho, M.; et al. Octupolar Crystals for Nonlinear Optics: 1,3,5-Trinitro-2,4,6-tris(styryl)benzene Derivatives. Chem. Mater. 2001, 13, 1438–1440. [Google Scholar] [CrossRef]
  20. Sukul, P.K.; Das, P.; Dhakar, G.L.; Das, L.; Malik, S. Effect of Tricarboxylic Acids on the Formation of Hydrogels with Melem or Melamine: Morphological, Structural and Rheological Investigations. Gels 2022, 8, 51. [Google Scholar] [CrossRef]
  21. Sukul, P.K.; Malik, S. Removal of toxic dyes from aqueous medium using adenine based bicomponent hydrogel. RSC Adv. 2013, 3, 1902–1915. [Google Scholar] [CrossRef]
  22. Zhu, X.; Liu, B.; Cui, P.; Kilina, S.; Sun, W. Multinuclear 2-(Quinolin-2-yl)quinoxaline-Coordinated Iridium(III) Complexes Tethered by Carbazole Derivatives: Synthesis and Photophysics. Inorg. Chem. 2020, 59, 17096–17108. [Google Scholar] [CrossRef]
  23. Huang, C.; Fu, W.; Li, C.-Z.; Zhang, Z.; Qiu, W.; Shi, M.; Heremans, P.; Jen, A.K.Y.; Chen, H. Dopant-Free Hole-Transporting Material with a C3h Symmetrical Truxene Core for Highly Efficient Perovskite Solar Cells. J. Am. Chem. Soc. 2016, 138, 2528–2531. [Google Scholar] [CrossRef] [PubMed]
  24. Truxene after Trujillo Coke. Truxillense (Trujillo or Peruvian leaf) which is the variety used to flavor beverages such as Coca Cola. The names stems from “Turris Iulia”, the old name fro Trujillo.
  25. Wang, J.-Y.; Yan, J.; Ding, L.; Ma, Y.; Pei, J. One-dimensional microwires formed by the co-assembly of complementary aromatic donors and acceptors. Adv. Funct. Mater. 2009, 19, 1746–1752. [Google Scholar] [CrossRef]
  26. Luo, J.; Zhou, Y.; Niu, Z.-Q.; Zhou, Q.-F.; Ma, Y.; Pei, J. Three-dimensional architectures for highly stable pure blue emission. J. Am. Chem. Soc. 2007, 129, 11314–11315. [Google Scholar] [CrossRef] [PubMed]
  27. Kirschbaum, T.; Rominger, F.; Mastalerz, M. An Isosteric Triaza Analogue of a Polycyclic Aromatic Hydrocarbon Monkey Saddle. Chem. A Eur. J. 2020, 26, 14560–14564. [Google Scholar] [CrossRef]
  28. Ramos, F.J.; Rakstys, K.; Kazim, S.; Grätzel, M.; Nazeeruddin, M.K.; Ahmad, S. Rational design of triazatruxene-based hole conductors for perovskite solar cells. RSC Adv. 2015, 5, 53426–53432. [Google Scholar] [CrossRef]
  29. Connell, A.; Wang, Z.; Lin, Y.-H.; Greenwood, P.C.; Wiles, A.A.; Jones, E.W.; Furnell, L.; Anthony, R.; Kershaw, C.P.; Cooke, G.; et al. Low cost triazatruxene hole transporting material for >20% efficiency perovskite solar cells. J. Mater. Chem. C 2019, 7, 5235–5243. [Google Scholar] [CrossRef] [Green Version]
  30. Ahn, D.H.; Kim, S.W.; Lee, H.; Ko, I.J.; Karthik, D.; Lee, J.Y.; Kwon, J.H. Highly efficient blue thermally activated delayed fluorescence emitters based on symmetrical and rigid oxygen-bridged boron acceptors. Nat. Photonics 2019, 13, 540–546. [Google Scholar] [CrossRef]
  31. Lai, W.-Y.; He, Q.-Y.; Zhu, R.; Chen, Q.-Q.; Huang, W. Kinked star-shaped fluorene/triazatruxene co-oligomer hybrids with enhanced functional properties for high-performance, solution-processed, blue organic light-emitting diodes. Adv. Funct. Mater. 2008, 18, 265–276. [Google Scholar] [CrossRef]
  32. Liu, X.; Xu, Y.; Jiang, D. Conjugated Microporous Polymers as Molecular Sensing Devices: Microporous Architecture Enables Rapid Response and Enhances Sensitivity in Fluorescence-On and Fluorescence-Off Sensing. J. Am. Chem. Soc. 2012, 134, 8738–8741. [Google Scholar] [CrossRef]
  33. Rakstys, K.; Abate, A.; Dar, M.I.; Gao, P.; Jankauskas, V.; Jacopin, G.; Kamarauskas, E.; Kazim, S.; Ahmad, S.; Gratzel, M.; et al. Triazatruxene-Based Hole Transporting Materials for Highly Efficient Perovskite Solar Cells. J. Am. Chem. Soc. 2015, 137, 16172–16178. [Google Scholar] [CrossRef] [PubMed]
  34. Reynaerts, R.; Minoia, A.; Gali, S.M.; Daukiya, L.; Van Velthoven, N.; De Vos, D.; Lazzaroni, R.; Mali, K.S.; De Feyter, S. Coplanar versus Noncoplanar Carboxyl Groups: The Influence of Sterically Enforced Noncoplanarity on the 2D Mixing Behavior of Benzene Tricarboxylic Acids. J. Phys. Chem. C 2020, 124, 24874–24882. [Google Scholar] [CrossRef]
  35. Oakdale, J.S.; Sit, R.K.; Fokin, V.V. Ruthenium-Catalyzed Cycloadditions of 1-Haloalkynes with Nitrile Oxides and Organic Azides: Synthesis of 4-Haloisoxazoles and 5-Halotriazoles. Chem. Eur. J. 2014, 20, 11101–11110. [Google Scholar] [CrossRef] [PubMed]
  36. Rieser, J.; Ismail, N.; Abou-Elenien, G.; Wallenfels, K. Mono-, bis- and trishydrazo, and -azo compounds in the tricyanobenzene series. I. Preparation and properties of the title compounds. Liebigs Ann. Chem. 1981, 12, 1586–1597. [Google Scholar] [CrossRef]
  37. Stoessel, P.; Buesing, A.; Breuning, E.; Pflumm, C.; Parham, A.H.; Eberle, T.; Mujica-Fernaud, T. Heteroaromatic Compounds for Organic Electroluminescent Devices. WO 2012095143A1, 19 July 2012. [Google Scholar]
  38. Desimoni, G.; Gamba Invernizzi, A.; Quadrelli, P.; Righetti, P.P. Copper(II) in organic synthesis. IX. The copper(II)-catalyzed Michael reaction as a route to polysubstituted benzene derivatives. Gazz. Chim. Ital. 1991, 121, 483–485. [Google Scholar]
  39. Gill, J.E.; MacGillivray, R.; Munro, J. The preparation of symmetrical aromatic triamines and triisocyanates. J. Chem. Soc. 1949, 374, 1753–1754. [Google Scholar] [CrossRef]
  40. Chapyshev, S.V.; Misochko, E.Y.; Akimov, A.V.; Dorokhov, V.G.; Neuhaus, P.; Grote, D.; Sander, W. Molecular Structure and Magnetic Parameters of Septet 2,4,6-Trinitrenotoluene. J. Org. Chem. 2009, 74, 7238–7244. [Google Scholar] [CrossRef]
  41. Breslow, D.S.; Marcantonio, A.F. Polyazide Crosslinking Agents. US 3297674, 10 January 1967. [Google Scholar]
  42. Dorofeeva, O.V.; Ryzhova, O.N.; Suntsova, M.A. Accurate Prediction of Enthalpies of Formation of Organic Azides by Combining G4 Theory Calculations with an Isodesmic Reaction Scheme. J. Phys. Chem. A 2013, 117, 6835–6845. [Google Scholar] [CrossRef]
  43. Du, H.; Xu, X.; Liu, Y.; Liu, H.; Wang, F.; Zhang, J.; Gong, X. Theoretical studies on the nitro and azido derivatives of benzene. Huaxue Xuebao 2011, 69, 269–276. [Google Scholar]
  44. Hoogboom, J.; Juricek, M.; Lauko, J.; Rehak, J.; Woltinge, T.; van de Rult, K.; Flipse, K.F.J.; Rowan, A.E. Click chemistry approach to graphitic structures. PMSE Prepr. 2007, 97, 110–111. [Google Scholar]
  45. Kim, H.J.; Lee, W.S. Anion Receptor Comprising Aromatic Amines Substituted with Electron Withdrawing Groups and Electrolyte Using the Same for Alkali Metal Batteries. WO 2007126262A1, 8 November 2007. [Google Scholar]
  46. Kugler, T. Organic Electroluminescent Diode. GB2462433A, 10 February 2010. [Google Scholar]
  47. Liu, X.-F.; Xu, W.-G.; Lu, S.-X. DFT theoretical study on nitrogen-rich compounds C6H6−n(N3)n(n = 1–6). Gaodeng Xuexiao Huaxue Xuebao 2009, 30, 1406–1409. [Google Scholar]
  48. Liu, X.-J.; He, W.-P.; Huang, J. QSPR study on thermodynamic properties of the nitro and azido derivatives of benzene. Yuanzi Yu Fenzi Wuli Xuebao 2015, 32, 754–762. [Google Scholar]
  49. Boiko, V.N.; Gogoman, I.V.; Shchupak, G.M.; Yagupol’skii, L.M. Nucleophilic substitution in aromatic compounds with fluorine-containing substituents. X. Reactions of 2,4,6-trichloro- and 2,6-dichloro-3,5-bis[(trifluoromethyl)sulfonyl]nitrobenzenes with nucleophiles. Zh. Org. Khim. 1987, 23, 2586–2591. [Google Scholar]
  50. Chapyshev, S.V. Synthesis of 2,4,6-triazidotoluene for Use as Photoaffinic and Crosslinking Agent. RU 2430080C2, 27 September 2011. [Google Scholar]
  51. Chapyshev, S.V.; Chernyak, A.V. Triazidation of 2,4,6-trifluorobenzenes. J. Fluorine Chem. 2013, 156, 303–306. [Google Scholar] [CrossRef]
  52. Misochko, E.Y.; Masitov, A.A.; Akimov, A.V.; Korchagin, D.V.; Chapyshev, S.V. Heavy Atom Effect on Magnetic Anisotropy of Matrix-Isolated Monobromine Substituted Septet Trinitrene. J. Phys. Chem. A 2015, 119, 2413–2419. [Google Scholar] [CrossRef] [PubMed]
  53. Yang, W.; Huang, J.; Fan, G.; Ma, Q.; Qu, Y. Synthesis Method of Energetic Compound 1-(3,5-dinitro-2,4,6-triaminobenzene)-1,2,4-triazole and Its Derivatives. CN 106588792A, 26 April 2017. [Google Scholar]
  54. Yang, W.; Lu, H.; Liao, L.; Fan, G.; Ma, Q.; Huang, J. Synthesis, and single crystal structure of fully-substituted polynitrobenzene derivatives for high-energy materials. RSC Adv. 2018, 8, 2203–2208. [Google Scholar] [CrossRef] [Green Version]
  55. Yang, X.; Lin, X.; Yang, L.; Zhang, T. A novel method to synthesize stable nitrogen-rich polynitrobenzenes with π-stacking for high-energy-density energetic materials. Chem. Commun. 2018, 54, 10296–10299. [Google Scholar] [CrossRef]
  56. Brase, S.; Gil, C.; Knepper, K.; Zimmermann, V. Organic azides. An exploding diversity of a unique class of compounds. Angew. Chem. Int. Ed. 2005, 44, 5188–5240. [Google Scholar] [CrossRef]
  57. Keicher, T.; Loebbecke, S. Lab-Scale Synthesis of Azido Compounds: Safety Measures and Analysis; John Wiley & Sons Ltd.: Hoboken, NJ, USA, 2010; pp. 3–27. [Google Scholar]
  58. Chapyshev, S.V.; Korchagin, D.V.; Grote, D.; Sander, W. EPR spectroscopy of multicomponent, multispin molecular system obtained by the photolysis of 2,4,6-triazido-3-cyano-5-fluoropyridine in solid argon. Magn. Reson. Chem. 2019, 57, 472–478. [Google Scholar] [CrossRef]
  59. Chapyshev, S.V.; Mendez-Vega, E.; Sander, W. Molecular Magnets: The Synthesis and Characterization of High-Spin Nitrenes. Chem.-Eur. J. 2021, 27, 1258–1269. [Google Scholar] [CrossRef]
  60. Mendez-Vega, E.; Mieres-Perez, J.; Chapyshev, S.V.; Sander, W. Persistent Organic High-Spin Trinitrenes. Angew. Chem. Int. Ed. 2019, 58, 12994–12998. [Google Scholar] [CrossRef] [PubMed]
  61. Stoessel, P.; Buesing, A.; Voges, F.; Pflumm, C.; Parham, A.H.; Eberle, T.; Mujica-Fernaud, T. Benzene Derivatives with Nitrogen-Containing Substituents for Organic Electronic Devices and Their Preparation and the Devices. WO 2012143079A1, 26 October 2012. [Google Scholar]
  62. Lu, P.; Wu, Y.; Kang, H.; Wei, H.; Liu, H.; Fang, M. What can pKa and NBO charges of the ligands tell us about the water and thermal stability of metal organic frameworks? J. Mater. Chem. A 2014, 2, 16250–16267. [Google Scholar] [CrossRef]
  63. Effenberger, F.; Niess, R. Aminobenzole, I.V. N-Persubstituierte 3.5-Diamino-phenole und 1.3.5-Triamino-benzole. Chem. Ber. 1968, 101, 3787–3793. [Google Scholar] [CrossRef]
  64. Beller, M.; Breindl, C.; Riermeier, T.H.; Tillack, A. Synthesis of 2,3-Dihydroindoles, Indoles, and Anilines by Transition Metal-Free Amination of Aryl Chlorides. J. Org. Chem. 2001, 66, 1403–1412. [Google Scholar] [CrossRef]
  65. Witulski, B.; Senft, S.; Thum, A. Palladium-Catalyzed Multiple Aryl Aminations of Polybromobenzenes. Synlett 1998, 1998, 504–506. [Google Scholar] [CrossRef]
  66. Sheldrick, G. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  67. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef]
Figure 1. Generic structure of C3-symmetric benzenes (A), 2,4,6-triiodobenzene-1,3,5-tricarboxylic acid (B), [34] two prominent C3h-symmetric disk-like structures triquinolonobenzene (TQB) (C) [9] and 5,10,15-trihexyl-10,15-dihydro-5H-diindolo [3,2-a:3′,2′-c] carbazole (D); precursor dipyrimido [4,5-f:4′,5′-h] quinazoline-2,4,6,8,10,12-(1H,3H,5H,7H,-9H,11H)-hexone (E) [35].
Figure 1. Generic structure of C3-symmetric benzenes (A), 2,4,6-triiodobenzene-1,3,5-tricarboxylic acid (B), [34] two prominent C3h-symmetric disk-like structures triquinolonobenzene (TQB) (C) [9] and 5,10,15-trihexyl-10,15-dihydro-5H-diindolo [3,2-a:3′,2′-c] carbazole (D); precursor dipyrimido [4,5-f:4′,5′-h] quinazoline-2,4,6,8,10,12-(1H,3H,5H,7H,-9H,11H)-hexone (E) [35].
Molecules 27 04369 g001
Scheme 1. Overview of the herein reported building blocks.
Scheme 1. Overview of the herein reported building blocks.
Molecules 27 04369 sch001
Scheme 2. Synthesis of triazidobenzene-tricarboxylic acid 4.
Scheme 2. Synthesis of triazidobenzene-tricarboxylic acid 4.
Molecules 27 04369 sch002
Scheme 3. Synthesis of methyl tribromobenzene tricarboxylate 5.
Scheme 3. Synthesis of methyl tribromobenzene tricarboxylate 5.
Molecules 27 04369 sch003
Figure 2. Molecular structures of 2a, 2b and 2c, intramolecular hydrogen bonds drawn at dashed lines, displacement parameters are drawn at 50% probability level (for details, see supporting information and cif-file).
Figure 2. Molecular structures of 2a, 2b and 2c, intramolecular hydrogen bonds drawn at dashed lines, displacement parameters are drawn at 50% probability level (for details, see supporting information and cif-file).
Molecules 27 04369 g002
Figure 3. Molecular structure of 3a, displacement parameters are drawn at 30% probability level (for details, see supporting information and cif-file).
Figure 3. Molecular structure of 3a, displacement parameters are drawn at 30% probability level (for details, see supporting information and cif-file).
Molecules 27 04369 g003
Table 1. Scope of the various alkyl triamino-benzene-tricarboxylates 2ag.
Table 1. Scope of the various alkyl triamino-benzene-tricarboxylates 2ag.
Molecules 27 04369 i001
Starting MaterialRProductTemp. [°C]Yields [%]
1aMe 2a13035 a
1bEt2b13028
1ci Pr2c13018
1dt Bu2d10015
1ei Bu2e10012
1fn Pent2f10022
1gBn2g13023
Table 2. Syntheses of triazides 3ac, g.
Table 2. Syntheses of triazides 3ac, g.
Molecules 27 04369 i002
Starting MaterialRProductYield [%]
2aMe3a60
2bEt3b47
2ci Pr3c50
2gBn3g36
Table 3. Syntheses of tris-1,2,3-triazoles 6.
Table 3. Syntheses of tris-1,2,3-triazoles 6.
Molecules 27 04369 i003
Starting MaterialR1ProductR2Yield
3aMe6a-PhPh19
3aMe6a-C6H4BrC6H4Br18
3bEt6b-PhPh27
Table 4. Chemo-selective hexa-N-alkylation of methyl triamino-benzene-tricarboxylate 2a.
Table 4. Chemo-selective hexa-N-alkylation of methyl triamino-benzene-tricarboxylate 2a.
Molecules 27 04369 i004
ProductRYield [%]
7aMe48
7bEt38
7cBn51
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Schmidt, L.; Wagner, D.; Nieger, M.; Bräse, S. Functionalized C3-Symmetric Building Blocks—The Chemistry of Triaminotrimesic Acid. Molecules 2022, 27, 4369. https://doi.org/10.3390/molecules27144369

AMA Style

Schmidt L, Wagner D, Nieger M, Bräse S. Functionalized C3-Symmetric Building Blocks—The Chemistry of Triaminotrimesic Acid. Molecules. 2022; 27(14):4369. https://doi.org/10.3390/molecules27144369

Chicago/Turabian Style

Schmidt, Lisa, Danny Wagner, Martin Nieger, and Stefan Bräse. 2022. "Functionalized C3-Symmetric Building Blocks—The Chemistry of Triaminotrimesic Acid" Molecules 27, no. 14: 4369. https://doi.org/10.3390/molecules27144369

Article Metrics

Back to TopTop