Next Article in Journal
Hydrogels Based on Polysaccharides Grafted Ferulic Acid: A Biomimetic Approach
Previous Article in Journal
Development and Validation of a Multi-Level Computational Protocol for Drug Repurposing in the Treatment of Bacterial Infections
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Synthesis of a Symmetrical tris-Tetrazole as Isostere of a Tricarboxylic Acid: Behind New Tridentate Ligands for MOFs †

by
Julio C. Flores-Reyes
1,
Perla Islas-Jácome
1,
Atilano Gutiérrez-Carrillo
1,
Mónica A. Rincón-Guevara
2,
Galdina V. Suárez-Moreno
3,
Óscar Vázquez-Vera
3,
Leticia Lomas-Romero
1,*,
Eduardo González-Zamora
1,* and
Alejandro Islas-Jácome
1,*
1
Departamento de Química, Universidad Autónoma Metropolitana-Iztapalapa, Avenida Ferrocarril San Rafael Atlixco, Número 186, Colonia Leyes de Reforma 1A Sección, Alcaldía Iztapalapa, Ciudad de México 09310, Mexico
2
Departamento de Biotecnología, Universidad Autónoma Metropolitana-Iztapalapa, Avenida Ferrocarril San Rafael Atlixco, Número 186, Colonia Leyes de Reforma 1A Sección, Alcaldía Iztapalapa, Ciudad de México 09310, Mexico
3
Instituto Politécnico Nacional, Unidad Profesional Interdisciplinaria de Biotecnología, Av. Acueducto S/N, Barrio la Laguna Ticomán, Ciudad de México 07340, Mexico
*
Authors to whom correspondence should be addressed.
Presented at the 25th International Electronic Conference on Synthetic Organic Chemistry, 15–30 November 2021; Available online: https://ecsoc-25.sciforum.net.
Chem. Proc. 2022, 8(1), 25; https://doi.org/10.3390/ecsoc-25-11751
Published: 14 November 2021

Abstract

:
The synthesis and characterization of three tridentate ligands for new MOFs are described. A triple aminolysis of the 1,3,5-tribenzoyl chloride with p-aminobenzoic acid gave the tricarboxylic acid 3 in 90% yield. Moreover, the same reaction, also from the 1,3,5-tribenzoyl chloride, but using p-aminobenzonitrile gave the new tris-benzonitrile 5 in 85% yield. Finally, this later one was treated with sodium azide and a Lewis acid to synthesize the new tris-tetrazole-based ligand 7 in 72% yield through a [3 + 2] azide-nitrile cycloaddition. It is noteworthy that the isosterism between carboxylic acids and tetrazoles may be considered to design and fabricate new MOFs with similar properties.

1. Introduction

Metal-organic frameworks (MOFs), also known as porous coordination polymers (PCPs), are formed by the union of single metal cations (primary building unit, PBU) or metal clusters (secondary building unit, SBU) with organic ligands through coordination bonds functionalized with donor groups, whose main characteristics are crystallinity, and permanent porosity [1]. These materials form ordered networks that spread in 1D, 2D and 3D architectures and have found ample use in diverse applications, for example, in catalysis [2], biomedicine [3], sensing and luminescence [4], and as potential tools in environmental remediation as platforms for hazardous gas capture, for example, H2S [5], SO2 [6], and CO2 [7], just to name a few.
While the choice of metal ion to construct MOFs is important, the variability in the material’s structure and properties, for example pore size, will come from the ligand design. In this sense, the most used ligands for MOF synthesis are either neutral or anionic. Carboxylates are among the most common anionic coordinating groups and are often preferred due to the strong binding produced with metal ions through the formation of SBUs, which confer the network with great stability and promote its self-assembly. As for neutral ligands, the most used are pyrazine and 4,4′-bipyridine [8].
On the other hand, azoles, which are aromatic five-membered N-containing heterocycles, can be deprotonated to form the corresponding azolate, which allows them to bind metal ions in a similar way as pyridines [9]. Tetrazoles contain four N-atoms, and due to their electron-withdrawing effect, this heterocycle is turned slightly acidic, having pKa values like those of carboxylic acids (pKa = ~4.5–4.9), although tetrazoles are slightly larger. Due to these features, it is said that 5-substiuted tetrazoles are isosteres of carboxylic acids [10], and thus, they are also capable of coordinating to metal ions.
Carboxylates have various coordination modes [11] and can bind metal ions in a monodentate or a bidentate manner; furthermore, each oxygen donor can coordinate with up to three different metal ions [9]. However, bidentate binding produces a stronger interaction because it leads to the formation of SBUs [12]. 5-substituted tetrazoles, on the other hand, show up to ten different coordination modes, depending on whether they are in the neutral form or as a tetrazolates [13], making them very versatile ligands for MOF synthesis. Lastly, nitrile ligands generally exhibit a single linear coordination mode. However, they also show other less common side-on coordination to metal ions [14], as shown in Figure 1.
The classic method to synthesize 5-substituted tetrazoles consists of a Huisgen-type [3 + 2] dipolar cycloaddition between nitriles and a source of the azide anion, catalyzed by a Zn(II) salt. This methodology is operationally simple, minimizes the release of highly toxic hydrazoic acid, uses environmentally benign solvents, and often generates the desired compound in good yields, especially if the nitrile is bound to an electron-deficient aromatic ring [15]. With this communication, we aim to provide the synthesis and characterization of two new tridentate ligands and thus increase the number of available ligands that can be used in the synthesis of novel MOF-type materials. The tris-nitrile ligand was synthesized via a triple nucleophilic acyl substitution (triple aminolysis) and was later used to generate a tris-tetrazole-based ligand. We also present the synthesis of a previously reported tris-carboxylic acid but in a higher yield. These syntheses were conducted in environmentally benign solvents, and the products were obtained in good yields after a simple aqueous workup.

2. Results and Discussion

2.1. Synthesis of 4,4′,4″-((Benzene-1,3,5-tricarbonyl)tris(azanediyl))tribenzoic Acid (3)

The tridentate ligand 3 was synthesized, as reported by M. S. Lah and co-workers [16], with a slight modification. The process involves a triple nucleophilic acyl substitution by the reaction of 1,3,5-benzenetricarbonyl trichloride (1) with three equivalents of p-aminobenzoic acid (2) in the presence of triethylamine in acetone as a solvent, as shown in Scheme 1. The target ligand was obtained in 90% yield, which is consistent with several reports in the literature [17,18,19].
Figure 2 shows the NMR characterization of ligand 3. The 1H spectrum (Figure 2a) depicts a singlet at around 11.0 ppm, which can be attributed to the acidic protons of the carboxylic acid moieties. Next, there is another singlet at 8.80 ppm that accounts for the NH protons of the amide groups, and finally, a multiplet at 7.99 ppm integrates the fifteen aromatic protons. The two small doublets at around 7.60 and 6.60 ppm are attributed to unreacted p-aminobenzoic acid that was leftover. The 13C NMR spectrum shown in Figure 2b shows the expected eight signals, the key ones being those corresponding to the three carboxylic acid carbonyls at 167.2 ppm and the three amide carbonyls at 164.8 ppm, respectively.

2.2. Synthesis of N1,N3-Bis(4-(1H-tetrazol-5-yl)phenyl)-N5-(4-(2H-tetrazol-5-yl)phenyl)benzene-1,3,5-tricarboxamide (7)

The ligand 7 was synthesized considering that 5-substituted tetrazoles are isosteres of carboxylic acids, meaning that these compounds have a similar electronic structure, as well as near steric and physicochemical properties [10]. This ligand was obtained via a two-step sequence consisting of a triple nucleophilic acyl substitution followed by a Zn-catalyzed azide-nitrile [3+2] intermolecular cycloaddition. Thus, for the first step of the sequence, 1,3,5-benzenetricarbonyl trichloride (1) was reacted with three equivalents of p-aminobenzonitrile (4) utilizing the conditions described for the previous ligand, affording the tris-nitrile 5 in 85% yield. This later compound is of high interest because it could potentially be used as a ligand for constructing new MOFs due to the presence of three cyanide groups. As was mentioned in the introduction, the cyanide group acts as a pseudohalide ligand and, when combined with transition metals, is prone to forming crystalline materials [12]. Nonetheless, ligand 5 was taken as an intermediate for the second step of the reaction scheme, and it was reacted with nine equivalents of sodium azide and Zn(II) bromide in tert-butanol with microwave heating to afford the target tris-tetrazolic ligand 7 in 72% yield, as shown in Scheme 2.
Nine equivalents of both sodium azide and Zn(II) bromide seem to be excessive to perform a [3 + 2] dipolar cycloaddition. However, they were the necessary amounts to synthesize the tris-tetrazole 7 in a reasonable yield (72%), which can be considered a good yield, especially because the synthesis of this ligand had never been reported, anywhere.
The 1H NMR spectrum for compound 5 is shown in Figure 3a; it shows a singlet at 11.01 ppm corresponding to the NH protons, another singlet at 8.77 ppm that integrates the three aromatic protons of the central phenyl ring, and finally, a set of two doublets at 8.04 and 7.87 ppm, respectively, which integrate for six protons each and are attributed to the three para-substituted phenyl rings. Figure 3b shows the 13C NMR spectrum for compound 5, which contains the expected signals for the eight magnetically different carbons. There is a key signal at 164.9 ppm that corresponds to the three amide carbonyls and another signal at 118.9 that belongs to the cyanide groups.
The purification of compound 7 was difficult due to the tetrazole’s reduced solubility in most organic solvents, but some characteristic signals can be distinguished in the NMR spectra. The 1H NMR spectrum shown in Figure 4a depicts a singlet at 11.19 ppm corresponding to the three NH protons, a multiplet at 8.87 ppm corresponding to the three aromatic protons of the center ring, and finally, two sets of signals at 8.14 and 7.89 ppm, respectively, which correspond to the twelve aromatic protons of the three para-substituted aromatic rings. On the 13C NMR spectrum depicted in Figure 4b, the eight expected signals for compound 7 are shown. The key signal for the amide carbonyls is found at 164.8 ppm, and another key signal corresponding to the tetrazole carbons is located at 162.3 ppm. As can be seen in both spectrums, the impurity of tris-tetrazole 7 corresponds just to its precursor 5 with a relationship near to 4:1.

3. Conclusions

Tridentate ligand 3 was synthesized in an excellent 90% yield, which is consistent with literature reports. It has been extensively used with Zn salts for MOF synthesis, but there are reports of its use with several rare earth salts as well [20], which opens the possibility of trying new combinations of metal salts with this ligand. The novel tris-nitrile 5 and tris-tetrazole 7 were obtained in 85% and 72% yields, respectively, although the latter one containing traces of its precursor 5. These compounds could potentially be used with Cu(I) or Ag(I) salts for the construction of new MOFs. These experiments are underway.

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 MestReNova software (12.0.0–20080). The 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). Structural drawings were created using ChemDraw professional software (15.0.0.106). All starting materials were purchased from Sigma-Aldrich and were used without further purification or dehydration. The solvents were distilled and dried according to standard procedures.

4.2. Synthesis of 4,4′,4″-((Benzene-1,3,5-tricarbonyl)tris(azanediyl))tribenzoic Acid (3)

In a 1 L round-bottomed flask, a solution was prepared by dissolving 2.06 g (15.02 mmol, 5 equiv.) of 4-aminobenzoic acid in 80 mL of acetone, and 1.34 mL of triethylamine was added (9.51 mmol, 3 equiv.). To this solution was added 0.55 mL (3.01 mmol, 1 equiv.) of 1,3,5-benzenetricarbonyl trichloride under inert atmosphere, and the resulting mixture was stirred for 24 h. After this time, 500 mL of distilled water was added, and the precipitate that formed was filtered and washed with diethyl ether affording 1.54 g of an off-white solid in 90% yield. 1H NMR (500 MHz, d6-DMSO): δ 10.99 (s, 3H, H-11, H-38, H-40), 8.80 (s, 3H, H-7, H-23, H-30), 8.01–7.96 (m, 15H, H-13, H-15, H-17, H-2, H-3, H-5, H-6, H-25, H-26, H-28, H-29, H-32, H-33, H-35, H-36); 13C NMR (126 MHz, d6-DMSO): δ 167.2 (C-10, C-37, C-39), 164.8 (C-8, C-19, C-21), 142.8 (C-1, C-24, C-31), 135.3 (C-12, C-14, C-16), 130.3 (C-3, C-5, C-26, C-28, C-33, C-35), 130.2 (C-13, C-15, C-17), 126.6 (C-4, C-27, C-34), 119.7 (C-2, C-6, C-25, C-29, C-32, C-36).

4.3. Synthesis of N1,N3,N5-Tris(4-cyanophenyl)benzene-1,3,5-tricarboxamide (5)

In a 100 mL round-bottomed flask equipped with a magnetic stir bar, a solution was prepared by dissolving 0.28 g of 4-aminobenzonitrile (2.36 mmol, 3.3 equiv.) in 14 mL of acetone to which 0.32 mL of triethylamine (2.30 mmol, 3.3 equiv.) was added. To this solution was added 0.13 mL (0.74 mmol, 1 equiv.) of 1,3,5-benzenetricarbonyl trichloride under inert atmosphere, and the resulting mixture was stirred for 24 h. After this time, 50 mL of distilled water was added, and the precipitate that formed was filtered and washed with diethyl ether affording 0.32 g of an off-white solid in 85% yield. 1H NMR (500 MHz, d6-DMSO): δ 10.98 (s, 3H, H-7, H-21, H-28), 8.75 (s, 3H, H-12, H-14, H-16), 8.04 (d, J = 8.8 Hz, 6H, H-2, H-6, H-23, H-27, H-30, H-34), 7.87 (d, J = 8.8 Hz, 6H, H-3, H-5, H-24, H-26, H-31, H-33); 13C NMR (126 MHz, d6-DMSO): δ 164.9 (C-8, C-17, C-19), 143.1 (C-1, C-22, C-29), 135.0 (C-11, C-13, C-15), 133.2 (C-3, C-5, C-24, C-26, C-31, C-33), 130.5 (C-12, C-14, C-16), 120.3 (C-2, C-6, C-23, C-27, C-30, C-34), 118.9 (C-10, C-35, C-36), 105.8 (C-4, C-25, C-32).

4.4. Synthesis of N1,N3-Bis(4-(1H-tetrazol-5-yl)phenyl)-N5-(4-(2H-tetrazol-5-yl)phenyl)benzene-1,3,5-tricarboxamide (7)

In a microwave reactor 10 mL vial, 0.079 g (0.16 mmol, 1.0 equiv.) of compound 5, 0.091 g of sodium azide (1.40 mmol, 9.0 equiv.), and 0.32 g of ZnBr2 (1.40 mmol, 9.0 equiv.) were added to 1 mL of t-BuOH. It was placed in the microwave reactor for 60 min at 100 °C and a power of 300 W. After this time, the contents of the vial were poured into 10% HCl, and the resulting solution was stirred for 20 min. The precipitate that was formed was filtered and washed with distilled water, ethanol, and diethyl ether, affording 73.6 mg of a light brown solid in 72% yield. 1H NMR (500 MHz, d6-DMSO): δ 11.19 (s, 3H, H-12, H-25, H-32), 8.87–8.85 (m, 3H, H-16, H-18, H-20), 8.14–8.12 (m, 6H, H-7, H-11, H-28, H-30, H-35, H-37), 7.89–7.86 (m, 6H, H-8 H-10, H-27, H-31, H-34, H-38). 13C NMR (126 MHz, d6-DMSO): δ 164.8 (C-13, C-21, C-23), 162.3 (C-4, C-39, C-44), 143.2 (C-9, C-26, C-33), 134.7 (C-15, C-17, C-19), 133.1 (C-8, C-10, C-27, C-31, C-34, C-38), 127.6 (C-16, C-18, C-20), 127.1 (C-6, C-29, C-36), 120.3 (C-7, C-11, C-28, C-30, C-35, C-37).

Author Contributions

All authors contributed equally to this work (J.C.F.-R., P.I.-J., A.G.-C., M.A.R.-G., G.V.S.-M., Ó.V.-V., L.L.-R., E.G.-Z. and A.I.-J.). All authors have read and agreed to the published version of the manuscript.

Funding

A.I.-J. acknowledges “Proyecto Apoyado por el Fondo Sectorial de Investigación para la Educación CONACyT-SEP CB-2017-2018 (A1-S-32582)” for financial support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

J.C.F.-R. (886026) and P.I.-J. (1098528) thank CONACyT-México for their scholarships.

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. Lu, W.; Wei, Z.; Gu, Z.G.; Liu, T.L.; Park, J.; Park, J.; Tian, J.; Zhang, M.; Zhang, Q.; Gentle, T., III; et al. Tuning the structure and function of metal–organic frameworks via linker design. Chem. Soc. Rev. 2014, 43, 5561–5593. [Google Scholar] [CrossRef] [PubMed]
  2. Zhu, L.; Liu, X.-Q.; Jiang, H.-L.; Sun, L.-B. Metal-Organic Frameworks for Heterogeneous Basic Catalysis. Chem. Rev. 2017, 117, 8129–8176. [Google Scholar] [CrossRef] [PubMed]
  3. Giménez-Marquéz, M.; Hidalgo, T.; Serre, C.; Horcajada, P. Nanostructured metal-organic frameworks and their bio-related applications. Coord. Chem. Rev. 2016, 307, 342–360. [Google Scholar] [CrossRef]
  4. Müller-Buschbaum, K.; Beuerle, F.; Feldman, C. MOF based luminescence tuning and chemical/physical sensing. Microporous Mesoporous Mater. 2015, 216, 171–199. [Google Scholar] [CrossRef]
  5. López-Olvera, A.; Flores, J.G.; Aguilar-Pliego, J.; Brozek, C.K.; Gutiérrez-Alejandre, A.; Ibarra, I.A. Chemical Transformation of H2S within the Pores of Metal–Organic Frameworks: Formation of Polysulfides. Chem. Mater. 2021, 33, 6269–6276. [Google Scholar] [CrossRef]
  6. Martínez-Ahumada, E.; He, D.; Berryman, V.; López-Olvera, A.; Hernandez, M.; Jancik, V.; Martis, V.; Vera, M.A.; Lima, E.; Parker, D.J.; et al. SO2 Capture Using Porous Organic Cages. Angew. Chem. Int. Ed. 2021, 60, 17556–17563. [Google Scholar] [CrossRef] [PubMed]
  7. Cotlame-Salinas, V.; López-Olvera, A.; Islas-Jácome, A.; González-Zamora, E.; Ibarra, I.A. CO2 capture enhancement in MOFs via the confinement of molecules. React. Chem. Eng. 2021, 6, 441–453. [Google Scholar] [CrossRef]
  8. Kitagawa, S.; Kitaura, R.; Noro, S. Functional porous coordination polymers. Angew. Chem. Int. Ed. 2004, 43, 2334–2375. [Google Scholar] [CrossRef] [PubMed]
  9. Jie-Peng, Z.; Yue-Biao, Z.; Jian-Bin, L.; Xiao-Ming, C. Metal Azolate Frameworks: From Crystal Engineering to Functional Materials. Chem. Rev. 2012, 112, 1001–1033. [Google Scholar]
  10. Ballatore, C.; Huryn, D.; Smith III, A. Carboxylic acid (bio)isosteres in drug design. ChemMedChem 2013, 8, 385–395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Dojer, B.; Pevec, A.; Belaj, F.; Kristl, M. Two New Zinc(II) Acetates with 3- and 4-Aminopyridine: Syntheses and Structural Properties. Acta Chim. Slov. 2015, 62, 312–318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Batten, S.R.; Neville, S.M.; Turner, D.R. Coordination Polymers: Design Analysis and Application, 6th ed.; Royal Society of Chemistry: Cambridge, UK, 2009; pp. 172–178. [Google Scholar]
  13. Aromí, G.; Barrios, L.A.; Roubeau, O.; Gámez, P. Triazoles and tetrazoles: Prime ligands to generate remarkable coordination materials. Coord. Chem. Rev. 2011, 255, 485–546. [Google Scholar] [CrossRef]
  14. Lennartson, A.; McKenzie, C.J. Bridging nitrile groups in a metal–organic framework. J. Coord. Chem. 2012, 65, 4194–4202. [Google Scholar] [CrossRef]
  15. Demko, Z.P.; Sharpless, K.B. Preparation of 5-Substituted 1H-Tetrazoles from Nitriles in Water. J. Org. Chem. 2001, 66, 7945–7950. [Google Scholar] [CrossRef] [PubMed]
  16. Zou, Y.; Park, M.; Hong, S.; Lah, M.S. A designed metal–organic framework based on a metal–organic polyhedron. Chem. Commun. 2008, 20, 2340–2342. [Google Scholar] [CrossRef] [PubMed]
  17. Zeng, W.; Wang, G.; Zheng, B.; Wang, Z.; Bai, J. A porous amide-functionalized pto-type MOF exhibiting selective capture and separation of cationic MB dye. J. Coord. Chem. 2021, 74, 241–251. [Google Scholar] [CrossRef]
  18. Song, X.; Zou, Y.; Liu, X.; Oh, M.; Lah, M.S. A two-fold interpenetrated (3,6)-connected metal–organic framework with rutile topology showing a large solvent cavity. New J. Chem. 2010, 34, 2396–2399. [Google Scholar] [CrossRef] [Green Version]
  19. Howe, R.; Smalley, A.; Guttenplan, A.; Doggett, M.; Eddleston, M.; Tan, J.; Lloyd, G. A family of simple benzene 1,3,5-tricarboxamide (BTA) aromatic carboxylic acid hydrogels. Chem. Commun. 2013, 49, 4268–4270. [Google Scholar] [CrossRef] [PubMed]
  20. Duan, J.; Higuchi, M.; Foo, M.; Horike, S.; Rao, K.; Kitagawa, S. A Family of Rare Earth Porous Coordination Polymers with Different Flexibility for CO2/C2H4 and CO2/C2H6 Separation. Inorg. Chem. 2013, 52, 8244–8249. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Most common coordination modes of (a) carboxylates, (b) 5-substituted tetrazoles, (c) 5-substituted tetrazolates, and (d) nitriles.
Figure 1. Most common coordination modes of (a) carboxylates, (b) 5-substituted tetrazoles, (c) 5-substituted tetrazolates, and (d) nitriles.
Chemproc 08 00025 g001
Scheme 1. Synthesis of ligand 3.
Scheme 1. Synthesis of ligand 3.
Chemproc 08 00025 sch001
Figure 2. (a) 1H NMR spectrum and (b) 13C NMR spectrum of ligand 3.
Figure 2. (a) 1H NMR spectrum and (b) 13C NMR spectrum of ligand 3.
Chemproc 08 00025 g002aChemproc 08 00025 g002b
Scheme 2. Synthesis of ligands 5 and 7.
Scheme 2. Synthesis of ligands 5 and 7.
Chemproc 08 00025 sch002
Figure 3. (a) 1H NMR spectrum and (b) 13C NMR spectrum of intermediate 5.
Figure 3. (a) 1H NMR spectrum and (b) 13C NMR spectrum of intermediate 5.
Chemproc 08 00025 g003
Figure 4. (a) 1H NMR spectrum and (b) 13C NMR spectrum of ligand 7.
Figure 4. (a) 1H NMR spectrum and (b) 13C NMR spectrum of ligand 7.
Chemproc 08 00025 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Flores-Reyes, J.C.; Islas-Jácome, P.; Gutiérrez-Carrillo, A.; Rincón-Guevara, M.A.; Suárez-Moreno, G.V.; Vázquez-Vera, Ó.; Lomas-Romero, L.; González-Zamora, E.; Islas-Jácome, A. Synthesis of a Symmetrical tris-Tetrazole as Isostere of a Tricarboxylic Acid: Behind New Tridentate Ligands for MOFs. Chem. Proc. 2022, 8, 25. https://doi.org/10.3390/ecsoc-25-11751

AMA Style

Flores-Reyes JC, Islas-Jácome P, Gutiérrez-Carrillo A, Rincón-Guevara MA, Suárez-Moreno GV, Vázquez-Vera Ó, Lomas-Romero L, González-Zamora E, Islas-Jácome A. Synthesis of a Symmetrical tris-Tetrazole as Isostere of a Tricarboxylic Acid: Behind New Tridentate Ligands for MOFs. Chemistry Proceedings. 2022; 8(1):25. https://doi.org/10.3390/ecsoc-25-11751

Chicago/Turabian Style

Flores-Reyes, Julio C., Perla Islas-Jácome, Atilano Gutiérrez-Carrillo, Mónica A. Rincón-Guevara, Galdina V. Suárez-Moreno, Óscar Vázquez-Vera, Leticia Lomas-Romero, Eduardo González-Zamora, and Alejandro Islas-Jácome. 2022. "Synthesis of a Symmetrical tris-Tetrazole as Isostere of a Tricarboxylic Acid: Behind New Tridentate Ligands for MOFs" Chemistry Proceedings 8, no. 1: 25. https://doi.org/10.3390/ecsoc-25-11751

Article Metrics

Back to TopTop