Next Article in Journal
Investigating New Methods to Increase Adventitious Root Formation
Previous Article in Journal
Suitability Evaluation of Underutilized Crops Under Future Climate Change Using Ecocrop Model: A Case of Bambara Groundnut in Nigeria
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Synthesis of New bis 1-Substituted 1H-Tetrazoles via Efficient Heterocyclizations from Symmetric Dianilines, Methyl Orthoester, and Sodium Azide †

by
Julio C. Flores-Reyes
1,
Roberto E. Blanco-Carapia
1,
Alfredo López-Olvera
2,
Perla Islas-Jácome
1,
Yizrell Medina-Martínez
1,
Mónica A. Rincón-Guevara
3,
Ilich A. Ibarra
2,
Leticia Lomas-Romero
1,*,
Eduardo González-Zamora
1,* and
Alejandro Islas-Jácome
1,*
1
Departamento de Química, Universidad Autónoma Metropolitana-Iztapalapa, San Rafael Atlixco 186, Col. Vicentina, C.P. Iztapalapa, 09340 Ciudad de México, Mexico
2
Laboratorio de Fisicoquímica y Reactividad de Superficies, Instituto de Investigaciones en Materiales, Universidad Nacional Autónoma de México, Circuito Exterior S/N, Ciudad Universitaria, C.P. Coyoacán, 04510 Ciudad de México, Mexico
3
Departamento de Biotecnología, Universidad Autónoma Metropolitana-Iztapalapa, San Rafael Atlixco 186, Col. Vicentina, C.P. Iztapalapa, 09340 Ciudad de México, Mexico
*
Authors to whom correspondence should be addressed.
Presented at the 23rd International Electronic Conference on Synthetic Organic Chemistry, 15 November 2019–15 December 2019; Available online: https://ecsoc-23.sciforum.net/.
Proceedings 2019, 41(1), 26; https://doi.org/10.3390/ecsoc-23-06521
Published: 14 November 2019

Abstract

:
The synthesis and characterization of three new bis 1-substituted 1H-tetrazoles are described. Two products were synthesized in a single step via a direct heterocyclization of primary amines, methyl orthoester (trymethyl orthoformate), and sodium azide in 30% and 91% yields, respectively. Besides, another one was prepared via a three-step synthetic strategy: SNAr (32%), nitro-group reduction (66%), and primary amine heterocyclization (83%), yielding 18%, overall. The aim behind the synthesis of new tetrazole-containing products is to construct novel MOF-like structures to evaluate their gas capture properties (CO2, CO, and SO2) under relative humidity conditions.

1. Introduction

Metal organic frameworks (MOFs) are hybrid structures consisting in metal cations (nodes) connected through organic moieties (ligands) with different lengths that propagate into extended networks, which produce crystalline porous materials in 1-, 2-, or 3-dimensions [1]. The use of MOFs has been reported in a wide variety of fields of science and technology, for instance, as catalysts [2], luminescent materials [3], drug carriers/delivers [4], and as chemosensors [5]. However, their use as materials for gas capture, separation, and storage remains one of their most sought-after applications [6]. One of the two key components to fabricate MOFs is the metal cation. This component acts as a node and usually belongs to transition metals group [7]; although there are some reports of MOFs constructed with lanthanoid elements [8]. The other key component is the ligand, since it is responsible for most of the material’s structural properties such as the flexibility, porosity, and topology [9]. Regarding the ligands used to assemble MOFs, most of them contain carboxylates as the electron-donating functional groups due to the increased stability that these ‘teeth’ offers when coordinating to the metal cation, since they do it in a charge-compensating way [10]. However, there are some advantages when ligands containing azole-based donating groups are used. Tetrazoles are basic in their nature due to the four electron-withdrawing N-atoms in their structures. This property increases directly with the number of N-atoms present. Furthermore, the coordination between a transition metal cation (soft Lewis acid) and the azole (soft Lewis base) is expected to result in a stable bond, since their covalent nature grow up [11]. Furthermore, tetrazoles are an important class of heterocyclic compounds with many applications, for example, in medicinal chemistry [12], materials and polymer sciences [13], and as corrosion inhibitors [14]. In the same context, their salts have been used as energetic materials [15], among other interesting applications. The use of tetrazoles as ligands to fabricate complex MOFs was documented first by Franke et al. [16]. It has been reported that tetrazoles can act as a monodentate ligands through the N-4 donor sites [17]. In addition to the previously mentioned properties of tetrazoles, the ligands presented in this work may exhibit angular and flexible conformations, which could give rise to diverse types of MOF topologies. Also, the aromatic rings of the ligands could affect the packing of crystal structures through non-covalent interactions [18].
There are two main methods to synthesize 1-substituted 1H-tetrazoles: (i) through a Huisgen-type [3+2] acid-catalyzed cycloadditions between an azide ion source and isocyanides [19], and (ii) via heterocyclizations of primary amines, an orthoester, and the sodium azide [20] (Scheme 1). The second method has some advantages over the first one, since primary amines are inexpensive and readily available reagents, as well as the orthoesters. This reaction has a wide scope and can be used with numerous substrates to effectively introduce the 1-substituted 1H-tetrazole moiety into polyfunctional molecules [21]. As an example, Muttenthaler and co-workers [22] synthesized a series of bis 1-substitued 1H-tetrazoles from primary amines making use of this methodology obtaining good to excellent yields. Even to present days, this remains the most used methodology towards the synthesis of 1-substituted 1H-tetrazoles, despite it was first reported in a patent in 1973 [23].
The aim of this work is to increase the reserve of ligands with donor groups (different to carboxylates), which can be used to fabricate new MOFs. Thus, for the present work, the synthesis and characterization of three new bis 1-substituted 1H-tetrazoles with potential application to construct new MOFs is described. One of them is linked by a disulfane-bridge and the other by a biphenyl moiety, both synthesized in a single step via a direct heterocyclization of primary amines. The third one is linked by an aromatic secondary amine group. This one was obtained via a three-step synthetic strategy. It is noteworthy that these ligands have not previously been synthesized nor isolated.

2. Results and Discussion

2.1. Synthesis of 1,2-bis(4-(1H-tetrazol-1-yl)phenyl)disulfane (2)

The 4,4’-disulfanediyldianiline (1) reacted with an excess of both, sodium azide (3.0 equiv.) and trimethyl orthoformate (22.8 equiv.) in acetic acid as the solvent for 24 h to give the target molecule 1,2-bis(4-(1H-tetrazol-1-yl)phenyl)disulfane (2) in 30% yield via a primary amine heterocyclization (Scheme 2).
The bis-tetrazole compound 2 was characterized by its physicochemical properties, as well as by classic spectroscopic techniques. Figure 1 shows the 1H and 13C-NMR spectra of compound 2. There is a characteristic singlet at 10.05 ppm (Figure 1), which is attributed to the H-atom from the tetrazole, as well as a couple of doublets for the A2B2 system from their aromatic rings. Figure 1b shows a key peak at 168.4 ppm corresponding to the C-atom from the tetrazole (C-5).

2.2. Synthesis of 4,4’-bis(4-(1H-tetrazol-1-yl)phenoxy)-1,1’-biphenyl (4)

The 4,4′-(1,1′-biphenyl-4,4′-diyldioxy)dianiline (3) was combined sequentially with sodium azide and trimethyl orthoformate in acetic acid as the solvent to furnish the target compound 4,4’-bis(4-(1H-tetrazol-1-yl)phenoxy)-1,1’-biphenyl (4) in 91% yield via a primary amine heterocyclization. The yield for this reaction is remarkably good considering the structural complexity of the ligand 4 (Scheme 3). It is noteworthy also that this yield (91%) is considerable higher than that one for the product 2 (30%) probably because the disulfane bridge is sensitive to higher temperatures and the acidic conditions as a result of the AcOH media.
The bis-tetrazole 4 was characterized by its physicochemical properties, as well as by spectroscopic techniques. Figure 2 shows the 1H and 13C-NMR spectra of the compound 4. The 1H-NMR spectrum (Figure 2a) shows the typical singlet at 10.06 ppm, which is attributed to the H-atom from the tetrazole ring. The structure of 4 contains four symmetric para-substituted aromatic rings, so four doublets corresponding to non-equivalent A2B2 systems are expected. The 13C-NMR spectrum (Figure 2b) shows the expected nine signals for the ligand, the key one being the peak at 142.2 ppm, which is characteristic for the C-atom of the 1-substituted tetrazole ring.

2.3. Synthesis of bis(4-(1H-tetrazol-1-yl)phenyl)amine (9)

The first step to synthesize the bis(4-(1H-tetrazol-1-yl)phenyl)amine (9) was a SNAr reaction between the para-nitroaniline (5) and the 1-chloro-4-nitrobenzene (6) to furnish the precursor bis(4-nitrophenyl)amine (7) in 32% yield. This reaction proceeded due to the strong activation of C-1 from 6 by the nitro group (Scheme 4).
The precursor 7 was characterized successfully by its physicochemical properties, as well as by spectroscopic techniques. Figure 3 shows the 1H and 13C-NMR spectra of precursor 7. The 1H-NMR spectrum (Figure 3a) shows a key singlet at 9.96 ppm, corresponding to the NH-proton, as well as a pair of doublets for A2B2 systems from both aromatic rings. The 13C-NMR spectrum (Figure 3b) shows the expected four signals for the compound with a key peak at 147.5 ppm, corresponding to the more unshielded C-1 and C-8 due to bonds to the amine group, as well as to the electron-withdrawing effect coming from the nitro groups.
Having the bis(4-nitrophenyl)amine (7) in hand, the next step was to hydrogenate both nitro groups using the metallic iron/iron(II) sulfate heptahydrate reductive system in water as the solvent to furnish the N1-(4-aminophenyl)benzene-1,4-diamine (8) in 66% yield (Scheme 5). Unfortunately, the precursor 8 underwent a quick carbonatation with CO2 from air, which did not allow its characterization by spectroscopic techniques. Thus, the compound 8 was used as soon as prepared (detected by TLC) for the next step, a heterocyclization with trimethyl orthoformate and sodium azide in acetic acid as the solvent, affording the new bis(4-(1H-tetrazol-1-yl)phenyl)amine (9) in 83% yield.
Figure 4 shows the 1H and 13C-NMR spectra of the compound 9. The 1H-NMR spectrum (Figure 4a) shows two key singlets, one at 9.96 ppm, corresponding to the H-atom from the tetrazole ring and the other at 8.99 ppm, which belongs to the NH-proton. Two doublets for the A2B2 systems is present as well. For the 13C-NMR spectrum (Figure 4b), the expected five signals are present, with a characteristic peak at 141.9 ppm, which corresponds to the C-atom of the 1-substituted 1H-tetrazole moiety (C-5).

3. Conclusions

The heterocyclization towards the disulfane-bridged ligand occurred with low yields because this functional group is very sensitive to high temperatures. The dioxy-biphenyl-bridged analogue was prepared in excellent yields considering its structural complexity. The amino-linked ligand required three steps for its synthesis. However, it would be interesting to prepare MOFs with an additional coordinating group. First attempts using these new ligands to try constructing novel porous coordination polymers are being conducted.

4. Experimental Section

4.1. General Information, Instrumentation, and Chemicals

1H and 13C NMR spectra were acquired on a Bruker Advance III (500 MHz) spectrometer. The solvent was deuterated dimethyl sulfoxide (d6-DMSO). Chemical shifts are reported in parts per million (/ppm). The internal reference for NMR spectra is with respect to tetramethyl silane (TMS) at 0.0 ppm. Coupling constants are reported in Hertz (J/Hz). Multiplicities of the signals are reported using the standard abbreviations: singlet (s), doublet (d), triplet (t), quartet (q), and multiplet (m). NMR data were treated using the MestReNova software (12.0.0–20080). Reaction progress was monitored by thin layer chromatography (TLC) on precoated kieselgel 60 F254 plates and the spots were visualized under UV light (254 or 365 nm). Melting points were determined on a Fisher-Johns apparatus and are uncorrected. Structural drawings were created using the ChemDraw professional software (15.0.0.106). All starting materials were purchased from Sigma-Aldrich and were used without further purification. The solvents were distilled and dried according to standard procedures.

4.2. Synthesis of 1,2-bis(4-(1H-tetrazol-1-yl)phenyl)disulfane (2)

In a 50-mL round-bottomed flask equipped with a magnetic stirring bar, were added 1.0 g (1.0 equiv.) of 4,4’-disulfanediyldianiline (1) and 0.78 g (3.0 equiv.) of sodium azide. The mixture was dissolved in 5 mL of acetic acid and then 10 mL of trimethyl orthoformate (22.0 equiv.) was added dropwise at RT. The reaction vessel was flushed with argon, it was placed in an oil bath at 100 °C and kept at reflux for 24 h. After this time, the crude of the reaction was filtered, washed with an acetone/ethanol (1:1, v/v) mixture, and dried under vacuum to afford 0.43 g of a grey solid with a 30% yield; Rf = 0.13 (Hex-AcOEt = 2:3, v/v); mp = 168–170 °C; 1H NMR (500 MHz, d6-DMSO): δ 10.05 (s, 2H), 7.58 (d, J = 8.7 Hz, 4H, H-8, H-10, H-15, H-19), 7.41 (d, J = 8.7 Hz, 4H, H-7, H-11, H-16, H-18); 13C NMR (126 MHz, d6-DMSO): δ 168.4 (C-5, C-24), 139.5 (C-9, C-14), 130.1 (C-6, C-7, C-11, C-16, C-17, C-18), 119.7 (C-8, C-10, C-15, C-19); HRMS [M+H]+ calculated for C14H11N8S2+ = 355.0503, found = 355.0517.

4.3. Synthesis of 4,4’-bis(4-(1H-tetrazol-1-yl)phenoxy)-1,1’-biphenyl (4)

A total of 1.08 g (1.0 equiv.) of 4,4′-(1,1′-biphenyl-4,4′-diyldioxy)dianiline (3), 1.5 g (3.0 equiv.) of sodium azide, and 5.5 mL (10.0 equiv.) of trimethyl orthoformate were sequentially added to a 50-mL round-bottomed flask equipped with a magnetic stirring bar. The flask was placed on an oil bath at 90 °C and 10 mL of acetic acid was added to the mixture. The reaction progress was monitored by TLC and was stopped after 3 h stirring. Then, the reaction mixture was poured into 200 mL of distilled water. The formed precipitate was collected by filtration. The solid was washed with cold distilled water and ethanol. Then, it was let to dry at RT, furnishing 2.10 g of a white powder with 91% yield; Rf = 0.61 (Hex-AcOEt = 1:3, v/v); mp = 219–221 °C; 1H-NMR (500 MHz, d6-DMSO): δ 10.06 (s, 2H, H-5, H-36), 7.94 (d, J = 9.0 Hz, 4H, H-15, H-17, H-20, H-24), 7.75 (d, J = 8.8 Hz, 4H, H-7, H-11, H-28, H-30), 7.32 (d, J = 9.0 Hz, 4H, H-14, H-18, H-21, H-23), 8.21 (d, J = 8.7 Hz, 4H, H-8, H-10, H-27, H-31); 13C NMR (126 MHz, d6-DMSO): δ 157.6 (C-9, C-26), 155.4 (C-13, C-22), 142.3 (C-5, C-36), 135.4 (C-16, C-19), 129.1 (C-6, C-29), 128.3 (C-15, C-17, C-20, C-24), 123.3 (C-7, C-11, C-28, C-30), 119.6 (C-14, C-18, C-21, C-23), 119.5 (C-8, C-10, C-27, C-31). HRMS [M+H]+ calculated for C26H19N8O2+ = 475.1586, found = 475.1598.

4.4. Synthesis of bis(4-nitrophenyl)amine (7)

In a 50-mL round-bottomed flask equipped with a magnetic stirring bar, were dissolved 2.06 g (2.5 equiv.) of p-nitroaniline (5), 0.96 (1.0 equiv.) of 1-chloro-4-nitrobenzene (6), and 3.35 g (3.9 equiv.) of potassium carbonate in 18 mL of N,N-dimethylformamide. The flask was connected to a reflux condenser and placed in an oil bath at 145 °C. The reaction progress was monitored by TLC and was stopped after 24 h stirring. After this time, the mixture was poured into 500 mL of distilled water, and then, concentrated hydrochloric acid was added until an acidic pH reached (pH = 2) and then, the mixture was filtered and washed with distilled water. The filtrate was crystallized from acetonitrile to furnish 0.55 g of a bright orange needles with 32% yield; Rf = 0.61 (Hex-AcOEt = 1:1, v/v); mp = 166–168 °C; 1H-NMR (500 MHz, d6-DMSO): δ 9.96 (s, 1H, H-7), 8.20 (d, J = 9.3 Hz, 4H, H-3, H-5, H-10, H-12), 7.36 (d, J = 9.3 Hz, 4H, H-2, H-6, H-9, H-13). 13C NMR (126 MHz, d6-DMSO): 147.5 (C-1, C-8), 140.5 (C-4, C-11), 125.7 (C-3, C-5, C-10, C-12), 117.0 (C-2, C-6, C-9, C-13). HRMS [M+H]+ calculated for C12H10N3O4+ = 260.0627, found = 260.0669.

4.5. Synthesis of N1-(4-aminophenyl)benzene-1,4-diamine (8)

In a 50-mL two-necked round-bottomed flask equipped with a magnetic stirring bar, were added 0.26 g (1.0 equiv.) of bis(4-nitrophenyl)amine (7), 1.12 g (20.1 equiv.) of metallic iron, and 0.41 g (1.5 equiv.) of iron(II) sulfate heptahydrate. The flask was connected to a reflux condenser, flushed with argon and then 10 mL of distilled water was added. The flask was placed in an oil bath and was refluxed for 4 h. After this time, the reaction mixture was cooled to RT, filtered, and extracted with 3 × 15 mL of ethyl acetate, the organic layer was dried with anhydrous sodium sulfate and evaporated to dryness furnishing 0.13 g of a light brown solid in 66% yield. This product undergoes a rapid carbonatation. Thus, it was used immediately for the subsequent step without spectrometric characterization; Rf = 0.48 (Hex-AcOEt = 1:3, v/v); mp = 172–174 °C.

4.6. Synthesis of bis(4-(1H-tetrazol-1-yl)phenyl)amine (9)

In a 50-mL round-bottomed flask equipped with a magnetic stirring bar, were added 0.11 g (1.0 equiv.) of N1-(4-aminophenyl)benzene-1,4-diamine (8), 0.23 g (6.1 equiv.) of sodium azide, and 1.0 mL (16.0 equiv.) of trimethyl orthoformate. The flask was placed on an oil bath at 90 °C and 2 mL of acetic acid were added to the mixture. The reaction progress was monitored by TLC and was stopped after 3 h stirring. Then, the reaction mixture was poured into 100 mL of distilled water. The formed precipitate was collected by filtration. The solid was washed with cold distilled water and ethanol and purified by column chromatography with ethyl acetate furnishing 0.14 g of a light brown crystalline solid with 83% yield; Rf = 0.40 (Hex-AcOEt = 1:3, v/v); mp = 224–226 °C; 1H-NMR (500 MHz, d6-DMSO): δ 9.96 (s, 2H, H-5, H-23), 8.99 (s, 1H, H-12), 7.79 (d, J = 8.9 Hz, 4H, H-7, H-11, H-15, H-17), 7.36 (d, J = 8.9 Hz, 4H, H-8, H-10, H-14, H-18). 13C NMR (126 MHz, d6-DMSO): δ 143.8 (C-9, C-13), 141.9 (C-5, C-23), 126.4 (C-6, C-16), 122.7 (C-7, C-11, C-15, C-17), 117.6 (C-8, C-10, C-14, C-18). HRMS [M+H]+ calculated for C14H12N9+ = 306.1171, found = 306.1187.

Author Contributions

Synthesis and characterization, J.C.F.-R., R.E.B.-C., A.L.-O., P.I.-J. and Y.M.-M.; HRMS and NMR analyses, M.A.R.-G.; resources, I.A.I. and L.L.-R.; writing—review and editing, E.G.-Z. and A.I.-J. All authors have read and agreed to the published version of the manuscript.

Acknowledgments

I.A.I. thanks CONACyT (1789) and PAPIIT UNAM (IN101517) for financial support. L.L.-R. acknowledges PRODEP-SEP (511-6/18-8354) for financial support. E.G.-Z. thanks CONACyT (236879) for financial support. A.I.-J. acknowledges CONACyT-SEP CB-2017-2018 (A1-S-32582) for financial support.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Izatt, R.M. Macrocyclic and Supramolecular Chemistry; Wiley: Chichester, UK, 2016. [Google Scholar] [CrossRef]
  2. Lee, J.; Farha, O.K.; Roberts, J.; Scheidt, K.A.; Nguyen, S.T.; Hupp, J.T. Metal-organic framework materials as catalysts. Chem. Soc. Rev. 2009, 38, 1450–1459. [Google Scholar] [CrossRef] [PubMed]
  3. Allendorf, M.D.; Bauer, C.A.; Bhakta, R.K.; Houk, R.J. Luminescent metal–organic frameworks. Chem. Soc. Rev. 2009, 38, 1330–1352. [Google Scholar] [CrossRef] [PubMed]
  4. Orellana-Tavra, C.; Baxter, E.F.; Tian, T.; Bennett, T.D.; Slater, N.K.; Cheetham, A.K.; Fairen-Jimenez, D. Amorphous metal–organic frameworks for drug delivery. Chem. Commun. 2015, 51, 13878–13881. [Google Scholar] [CrossRef] [PubMed]
  5. Kreno, L.E.; Leong, K.; Farha, O.K.; Allendorf, M.; Van Duyne, R.P.; Hupp, J.T. Metal–organic framework materials as chemical sensors. Chem. Rev. 2012, 112, 1105–1125. [Google Scholar] [CrossRef] [PubMed]
  6. Li, J.R.; Kuppler, R.J.; Zhou, H.C. Selective gas adsorption and separation in metal–organic frameworks. Chem. Soc. Rev. 2009, 38, 1477–1504. [Google Scholar] [CrossRef] [PubMed]
  7. Batten, S.R.; Neville, S.M.; Turner, D.R. Coordination Polymers: Design, Analysis and Application; Royal Society of Chemistry: Cambridge, UK, 2009. [Google Scholar] [CrossRef]
  8. Liu, X.; Akerboom, S.; Jong, M.D.; Mutikainen, I.; Tanase, S.; Meijerink, A.; Bouwman, E. Mixed-lanthanoid metal–organic framework for ratiometric cryogenic temperature sensing. Inorg. Chem. 2015, 54, 11323–11329. [Google Scholar] [CrossRef] [PubMed]
  9. Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O’Keeffe, M.; Yaghi, O.M. Systematic design of pore size and functionality in isoreticular MOFs and their application in methane storage. Science 2002, 295, 469–472. [Google Scholar] [CrossRef] [PubMed]
  10. Deria, P.; Bury, W.; Hod, I.; Kung, C.W.; Karagiaridi, O.; Hupp, J.T.; Farha, O.K. MOF functionalization via solvent-assisted ligand incorporation: Phosphonates vs. carboxylates. Inorg. Chem. 2015, 54, 2185–2192. [Google Scholar] [CrossRef] [PubMed]
  11. Zhang, J.P.; Zhang, Y.B.; Lin, J.B.; Chen, X.M. Metal azolate frameworks: From crystal engineering to functional materials. Chem. Rev. 2012, 112, 1001–1033. [Google Scholar] [CrossRef] [PubMed]
  12. Herr, R.J. 5-Substituted-1H-tetrazoles as carboxylic acid isosteres: Medicinal chemistry and synthetic methods. Bioorg. Med. Chem. 2002, 10, 3379–3393. [Google Scholar] [CrossRef] [PubMed]
  13. Zhao, H.; Qu, Z.R.; Ye, H.Y.; Xiong, R.G. In situ hydrothermal synthesis of tetrazole coordination polymers with interesting physical properties. Chem. Soc. Rev. 2008, 37, 84–100. [Google Scholar] [CrossRef]
  14. Kuznetsov, Y.I.; Kazansky, L.P. Physicochemical aspects of metal protection by azoles as corrosion inhibitors. Russ. Chem. Rev. 2008, 77, 219. [Google Scholar] [CrossRef]
  15. Gao, H.; Shreeve, J.N.M. Azole-based energetic salts. Chem. Rev. 2011, 111, 7377–7436. [Google Scholar] [CrossRef] [PubMed]
  16. Franke, P.L.; Haasnoot, J.G.; Zuur, A.P. Tetrazoles as ligands. Part IV. Iron (II) complexes of monofunctional tetrazole ligands, showing high-spin (p5T2g) ⇋ low-spin transitions. Inorg. Chim. Acta 1982, 59, 5–9. [Google Scholar] [CrossRef]
  17. Aromí, G.; Barrios, L.A.; Roubeau, O.; Gamez, P. Triazoles and tetrazoles: Prime ligands to generate remarkable coordination materials. Coord. Chem. Rev. 2011, 255, 485–546. [Google Scholar] [CrossRef]
  18. Huang, Z.; Song, H.B.; Du, M.; Chen, S.T.; Bu, X.H.; Ribas, J. Coordination polymers assembled from angular dipyridyl ligands and CuII, CdII, CoII salts: Crystal structures and properties. Inorg. Chem. 2004, 43, 931–944. [Google Scholar] [CrossRef]
  19. Jin, T.; Kamijo, S.; Yamamoto, Y. Synthesis of 1-substituted tetrazoles via the acid-catalyzed [3+2] cycloaddition between isocyanides and trimethylsilyl azide. Tetrahedron Lett. 2004, 45, 9435–9437. [Google Scholar] [CrossRef]
  20. Gaponik, P.N.; Karavai, V.P.; Grigor’ev, Y.V. Synthesis of 1-substituted tetrazoles by heterocyclization of primary amines, orthoformic ester, and sodium azide. Chem. Heterocycl. Compd. 1985, 21, 1255–1258. [Google Scholar] [CrossRef]
  21. Grigoriev, Y.V.; Voitekhovich, S.V.; Karavai, V.P.; Ivashkevich, O.A. Synthesis of tetrazole and its derivatives by heterocyclization reaction involving primary amines, orthoesters, and azides. Chem. Heterocycl. Compd. 2017, 53, 670–681. [Google Scholar] [CrossRef]
  22. Muttenthaler, M.; Bartel, M.; Weinberger, P.; Hilscher, G.; Linert, W. Synthesis and characterisation of new ditetrazole-ligands as more rigid building blocks of envisaged iron (II) spin-crossover coordination polymers. J. Mol. Struct. 2005, 741, 159–169. [Google Scholar] [CrossRef]
  23. Kamiya, T.; Saito, Y. Process for Preparing 1H-tetrazole Compounds. U.S. Patent 3,767,667; U.S. Patent and Trademark Office, Washington, DC, USA, 23 October 1973. [Google Scholar]
Scheme 1. Scheme 1. Synthesis of 1-substituted 1H-tetrazoles from primary amines, orthoesters, and the azide anion.
Scheme 1. Scheme 1. Synthesis of 1-substituted 1H-tetrazoles from primary amines, orthoesters, and the azide anion.
Proceedings 41 00026 sch001
Scheme 2. Scheme 2. Synthesis of the new disulfane-bridged bis-tetrazole ligand 2.
Scheme 2. Scheme 2. Synthesis of the new disulfane-bridged bis-tetrazole ligand 2.
Proceedings 41 00026 sch002
Figure 1. (a) 1H-NMR spectrum of the disulfane-bridged bis-tetrazole 2. (b) 13C-NMR spectrum of the disulfane-bridged bis-tetrazole 2.
Figure 1. (a) 1H-NMR spectrum of the disulfane-bridged bis-tetrazole 2. (b) 13C-NMR spectrum of the disulfane-bridged bis-tetrazole 2.
Proceedings 41 00026 g001
Scheme 3. Scheme 3. Synthesis of new biphenyldioxy-bridged bis-tetrazole ligand.
Scheme 3. Scheme 3. Synthesis of new biphenyldioxy-bridged bis-tetrazole ligand.
Proceedings 41 00026 sch003
Figure 2. (a) 1H-NMR spectrum of the biphenyldioxy-bridged bis-tetrazole 4. (b) 13C-NMR spectrum of the biphenyldioxy-bridged bis-tetrazole 4.
Figure 2. (a) 1H-NMR spectrum of the biphenyldioxy-bridged bis-tetrazole 4. (b) 13C-NMR spectrum of the biphenyldioxy-bridged bis-tetrazole 4.
Proceedings 41 00026 g002
Scheme 4. Scheme 4. Synthesis of the precursor bis(4-nitrophenyl)amine 7.
Scheme 4. Scheme 4. Synthesis of the precursor bis(4-nitrophenyl)amine 7.
Proceedings 41 00026 sch004
Figure 3. (a) 1H-NMR spectrum of the bis(4-nitrophenyl)amine 7. (b) 13C-NMR spectrum of the bis(4-nitrophenyl)amine 7.
Figure 3. (a) 1H-NMR spectrum of the bis(4-nitrophenyl)amine 7. (b) 13C-NMR spectrum of the bis(4-nitrophenyl)amine 7.
Proceedings 41 00026 g003
Scheme 5. Nitro groups reduction and primary amine heterocyclization toward the new bis-tetrazole 9.
Scheme 5. Nitro groups reduction and primary amine heterocyclization toward the new bis-tetrazole 9.
Proceedings 41 00026 sch005
Figure 4. (a) 1H-NMR spectrum of the bis(4-(1H-tetrazol-1-yl)phenyl)amine 9. (b) 13C-NMR spectrum of the bis(4-(1H-tetrazol-1-yl)phenyl)amine 9.
Figure 4. (a) 1H-NMR spectrum of the bis(4-(1H-tetrazol-1-yl)phenyl)amine 9. (b) 13C-NMR spectrum of the bis(4-(1H-tetrazol-1-yl)phenyl)amine 9.
Proceedings 41 00026 g004
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

Flores-Reyes, J.C.; Blanco-Carapia, R.E.; López-Olvera, A.; Islas-Jácome, P.; Medina-Martínez, Y.; Rincón-Guevara, M.A.; Ibarra, I.A.; Lomas-Romero, L.; González-Zamora, E.; Islas-Jácome, A. Synthesis of New bis 1-Substituted 1H-Tetrazoles via Efficient Heterocyclizations from Symmetric Dianilines, Methyl Orthoester, and Sodium Azide. Proceedings 2019, 41, 26. https://doi.org/10.3390/ecsoc-23-06521

AMA Style

Flores-Reyes JC, Blanco-Carapia RE, López-Olvera A, Islas-Jácome P, Medina-Martínez Y, Rincón-Guevara MA, Ibarra IA, Lomas-Romero L, González-Zamora E, Islas-Jácome A. Synthesis of New bis 1-Substituted 1H-Tetrazoles via Efficient Heterocyclizations from Symmetric Dianilines, Methyl Orthoester, and Sodium Azide. Proceedings. 2019; 41(1):26. https://doi.org/10.3390/ecsoc-23-06521

Chicago/Turabian Style

Flores-Reyes, Julio C., Roberto E. Blanco-Carapia, Alfredo López-Olvera, Perla Islas-Jácome, Yizrell Medina-Martínez, Mónica A. Rincón-Guevara, Ilich A. Ibarra, Leticia Lomas-Romero, Eduardo González-Zamora, and Alejandro Islas-Jácome. 2019. "Synthesis of New bis 1-Substituted 1H-Tetrazoles via Efficient Heterocyclizations from Symmetric Dianilines, Methyl Orthoester, and Sodium Azide" Proceedings 41, no. 1: 26. https://doi.org/10.3390/ecsoc-23-06521

Article Metrics

Back to TopTop