Next Article in Journal
Does Soil Organic Matter Affect the Impact of the Ionic Liquid Ethylammonium Nitrate in the Pure State and as Mixture with Lithium Salt on Basal Soil Respiration?
Previous Article in Journal
Influence of Nonpolar Medium on Antioxidant Capacity of Bergaptol and Xanthotoxol—Kinetic DFT Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Green Synthesis of Symmetric Dimaleamic Acids from Dianilines and Maleic Anhydride: Behind New Bidentate Ligands for MOFs †

by
Julio C. Flores-Reyes
1,
José L. Sosa-Juárez
1,
Mayra Sánchez-Serratos
1,
Perla Islas-Jácome
1,
Atilano Gutiérrez-Carrillo
1,
Francisco Méndez
1,2,3,
Galdina V. Suárez-Moreno
4,*,
Alejandro Islas-Jácome
1,* and
Eduardo González-Zamora
1,*
1
Departamento de Química, Universidad Autónoma Metropolitana-Iztapalapa, San Rafael Atlixco 186, Col. Vicentina, Iztapalapa, Ciudad de México C.P. 09340, Mexico
2
Le Studium, Loire Valley Institute for Advanced Studies, 4500 Orleans & Tours, France;
3
Haute Température et Irradiation (CEMHTI), UPR3079 CNRS, CEMHTI 1, Avenue de la Recherche Scientifique, 45071 Orleans, France
4
Instituto Politécnico Nacional, Unidad Profesional Interdisciplinaria de Biotecnología, Av. Acueducto S/N, Barrio la Laguna Ticomán, CDMX C.P. 07340, Mexico
*
Authors to whom correspondence should be addressed.
Presented at the 24th International Electronic Conference on Synthetic Organic Chemistry, 15 November–15 December 2020; Available online: https://sciforum.net/conference/ecsoc-24.
Published: 14 November 2020

Abstract

:
We herein report the synthesis and characterization of six α,β-unsaturated dicarboxylic acid ligands with different phenyl spacers, and two ligands with a biphenyl and anthraquinone spacers. All these dimaleamic acids were synthesized in 16 to 99% yields via a base-catalyzed maleimide ring opening in water (ligand 2), or by a di-N-acylation from the corresponding diamines and maleic anhydride in acetic acid (ligands 4, 6, 8, 10, 12, 14 and 16). These reactions were performed using green solvents, while requiring minimal work up procedures, making them suitable alternatives to access these types of bidentate ligands quickly, which can be used to fabricate new metal-organic frameworks (MOFs).

1. Introduction

Metal-organic frameworks (MOFs) are hybrid structures composed by the union, via coordination bonds, of metal ions or metal clusters to organic ligands with donor atoms, producing crystalline and often highly porous materials through a repetitive network that propagates into one, two or three dimensions [1]. Given that the amount of ligands and metal salts that can be combined for MOF synthesis is practically unlimited, MOFs can be tailored according to their intended use, and have therefore been used in several applications of economic, technological and environmental importance such as luminescence and sensing [2], electrochemistry [3], catalysis [4], gas capture, storage and separation [5] and biomedicine [6]. Most of the ligands used to construct MOFs are neutral or anionic. The drawback of using neutral ligands is the requirement of counter ions due to the positive charge generated by the formation of the coordination bond, while anionic ligands do not have this disadvantage because they bind metal atoms via a charge compensation [7]. Pyridine and pyrazine are among the most common donor groups present on neutral ligands, while for anionic ligands, carboxylates are the most common ones [8]. The carboxylate group can bind to a metal atom in a monodentate or bidentate manner, the latter of which produces the strongest binding. When bidentate binding occurs, it causes the in-situ formation of inorganic clusters called “secondary building units” (SBUs), which confer great rigidity to the framework being constructed and facilitate its self-assembly [9]. The spacers used in the ligand’s backbone also play a key role in MOF synthesis. As a result, most ligands have bound or fused aromatic rings as central or extending units because their rigidity can benefit the crystal packing and arrangement, for instance, via π–π stacking-type interactions [10].
Maleamic acid is a nitrogen-containing analog of maleic acid. N-substituted maleamic acids are highly conjugated, and their use as polydentate ligands has been well documented with lanthanide ions such as La3+ [11], Eu3+, Tb3+ and Yb3+ [12], and transition metals such as Cu2+ [13,14]. The most common method for maleamic acid synthesis involves the reaction between a primary amine and maleic anhydride under mild reaction conditions (Scheme 1), in a wide variety of solvents, for instance, dichloromethane [15], diethyl ether [16], toluene [17], tetrahydrofuran [18] or xylene [19], which may pose health and environmental hazards.
The main objective of this work is to provide environmentally benign procedures for the synthesis of ligands that can be used to construct novel MOF structures. To do this, we present the synthesis and spectroscopic characterization of eight bis-maleamic acid ligands, the first of which (ligand 2) was synthesized in water as the solvent from a bis-maleimide, and the rest in acetic acid from the corresponding diamines and maleic anhydride. Also, ligands 12 and 16 have not been previously reported, and their structures show that this procedure can be applicable to fabricate ligands with a more complex backbone.

2. Results and Discussion

2.1. Synthesis of (2E,2′E)-4,4′-(1,4-phenylenebis(azanediyl))bis(4-oxobut-2-enoic acid) (2)

For the synthesis of the ligand 2, the N,N′-(1,4-phenylene)dimaleimide 1 was reacted in an aqueous solution [0.1 M] of sodium hydroxide for 3 h to afford the target molecule (2E,2′E)-4,4′-(1,4-phenylenebis(azanediyl))bis(4-oxobut-2-enoic acid) in an almost quantitative yield (99%) via a base-catalyzed maleimide ring opening reaction (Scheme 2). This methodology has been utilized to obtain maleamic acids [20,21,22], but, to the best of our knowledge, the hydrolysis of a bis-maleimide such as 1 to produce a bis-maleamic acid has not been reported yet.
The bis-carboxylic acid 2 was characterized by its physicochemical properties, as well as by spectroscopic techniques. Figure 1 shows the 1H spectrum of compound 2. There is a broad singlet at 13.24 ppm, which is characteristic of the acidic protons from the carboxylic acid moiety, as well as a key singlet at 10.46 ppm, which corresponds to NH (amide protons). There is also a singlet at 7.60 ppm corresponding to the four H atoms from the aromatic ring, and a couple of doublets at 6.47 and 6.31 ppm, respectively, which correspond to the alkene moiety.

2.2. Synthesis of (2Z,2′Z)-4,4′-((methylenebis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (4)

4,4′-methylenedianiline (3) was reacted with a slight excess of maleic anhydride for 1 h in acetic acid as solvent to generate the desired compound (2Z,2′Z)-4,4′-((methylenebis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (4) in 62% yield (Scheme 3).
Compound 4 was characterized by its spectroscopic properties. Figure 2a shows the 1H NMR spectra of compound 4. The broad singlet corresponding to the carboxylic acid protons can be found at 13.07 ppm, followed by the singlet attributed to the NH protons at 10.43 ppm. Next, there are two doublets at 7.54 and 7.17 ppm corresponding to the A2B2 system of the aromatic rings, and finally another set of two doublets that correspond to the alkene protons. Figure 2b shows the 13C NMR spectrum. There are two key peaks at 166.7 and 163.1 ppm, respectively, which correspond to both carbonyl carbons, as well as another peak at 40 ppm belonging to the methylene carbon.

2.3. Synthesis of (2Z,2′Z)-4,4′-((oxybis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (6)

4,4′-oxydianiline (5) was reacted with two equivalents of maleic anhydride for 1 h in acetic acid as solvent to generate the desired compound (2Z,2′Z)-4,4′-((oxybis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (6) via an N-acylation reaction, in 50% yield (Scheme 4).
The ligand 6 was characterized by its spectroscopic properties. Figure 3a shows the 1H NMR spectra of compound 6. There is a key broad singlet corresponding to the carboxylic acid protons at 13.12 ppm followed by the singlet attributed to the NH protons at 10.43 ppm, the set of two doublets at 7.62 and 6.98 ppm corresponding to the A2B2 system of the aromatic rings and finally another set of two doublets corresponding to the vinylic protons. The 13C NMR spectrum shows the expected eight signals, with the key peaks at 166.6 and 163.0 ppm corresponding to both pairs of carbonyl carbons (Figure 3b).

2.4. Synthesis of (2Z,2′Z)-4,4′-((thiobis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (8)

4,4′-thiodianiline (7) reacted with two equivalents of maleic anhydride for 1 h in acetic acid as the solvent furnishing the desired compound (2Z,2′Z)-4,4′-((thiobis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (8) via an N-acylation reaction, in 55% yield (Scheme 5).
Ligand 8 was characterized by spectroscopic techniques. Figure 4a shows the 1H NMR spectrum, which shows the key broad singlet at 12.97 ppm that belongs to the carboxylic acid protons and another singlet for the NH protons at 10.47 ppm. There are two sets of two doublets, the first at 7.63 and 7.29 ppm corresponding to the A2B2 system of the aromatic rings, and the following set belonging to the alkene protons. Figure 4b shows the 13C NMR spectrum, which shows the expected eight signals, the two key peaks being at 166.8 and 163.3 ppm and corresponding to the carbonyl carbons.

2.5. Synthesis of (2Z,2′Z)-4,4′-((sulfonylbis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (10)

The 4,4′-sulfonyldianiline (9) reacted with two equivalents of maleic anhydride for 1 h in acetic acid as the solvent furnishing the target compound (2Z,2′Z)-4,4′-((sulfonylbis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (10) via an N-acylation reaction, in 39% yield (Scheme 6).
Thus, ligand 10 was characterized by spectroscopic techniques. Figure 5a shows the 1H NMR spectrum, which shows the key broad singlet at 12.86 ppm that corresponds to the carboxylic acid protons and another singlet for the NH protons at 10.68 ppm, and two sets of two doublets: the first at 7.88 and 7.82 ppm corresponding to the A2B2 system of the aromatic rings, and the following set belonging to the alkene protons. Figure 5b shows the 13C NMR spectrum, which shows the expected eight signals, the two key peaks being at 166.8 and 163.8 ppm, and corresponding to the carbonyl carbons.

2.6. Synthesis of (2Z,2′Z)-4,4′-((9,10-dioxo-9,10-dihydroanthracene-2,6-diyl)bis(azanediyl))bis(4-oxobut-2-enoic acid) (12)

2,6-diaminoanthracene-9,10-dione (11) reacted with two equivalents of maleic anhydride for 1 h in acetic acid as the solvent furnishing the target compound (2Z,2′Z)-4,4′-((9,10-dioxo-9,10-dihydroanthracene-2,6-diyl)bis(azanediyl))bis(4-oxobut-2-enoic acid) (12) via an N-acylation reaction, in 22% yield (Scheme 7).
Ligand 12 was characterized by spectroscopic techniques. Figure 6a shows the 1H NMR spectrum, which shows the key broad singlet at 12.88 ppm that corresponds to the carboxylic acid protons, and another singlet for the NH protons at 10.93 ppm. Three doublets at 8.47, 8.17 and 8.07 ppm correspond to the trisubstituted phenyl rings and two doublets of the alkene protons. Figure 6b shows the 13C NMR spectrum, which shows the expected 11 signals, the two key peaks being at 166.9 and 164.0 ppm, corresponding to the carbonyl from the amide and carboxylic acid, and another peak at 181.2 ppm, which accounts for both carbonyls of the anthraquinone core.

2.7. Synthesis of (2Z,2′Z)-4,4′-((disulfanediylbis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (14)

For the synthesis of ligand 14, 4,4′-disulfanediyldianiline (13) was reacted with two equivalents of maleic anhydride for 1 h in acetic acid as the solvent furnishing the target compound (2Z,2′Z)-4,4′-((disulfanediylbis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (14) via an N-acylation reaction, in 34% yield (Scheme 8).
Ligand 14 was characterized by spectroscopic techniques. Figure 7a shows the 1H NMR spectrum, which shows the key broad singlet at 12.94 ppm that corresponds to the carboxylic acid protons and another singlet for the NH protons at 10.50 ppm. The two doublets corresponding to the A2B2 system of the phenyl rings are present at 7.64 and 7.47 ppm, followed by the two doublets of the alkene protons at 6.46 and 6.30 ppm. Figure 7b shows the 13C NMR spectrum, which shows the expected seven signals, the two key peaks being at 166.8 and 163.3 ppm, and corresponding to the carbonyl carbons.

2.8. Synthesis of (2Z,2′Z)-4,4′-((([1,1′-biphenyl]-4,4′-diylbis(oxy))bis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (16)

For the synthesis of ligand 16, 4,4′-([1,1′-biphenyl]-4,4′-diylbis(oxy))dianiline (15) was reacted with two equivalents of maleic anhydride for 1 h in acetic acid as the solvent furnishing the target compound (2Z,2′Z)-4,4′-((([1,1′-biphenyl]-4,4′-diylbis(oxy))bis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) via an N-acylation reaction, in 16% yield (Scheme 9).
Ligand 16 was characterized by spectroscopic techniques. Figure 8a shows the 1H NMR spectrum, which shows the key broad singlet at 13.13 ppm that corresponds to the carboxylic acid protons and another singlet for the NH protons at 10.46 ppm. In this case, there are two sets of two doublets that correspond to both non-equivalent (anisochrone) A2B2 system of the four p-substituted phenyl rings at 7.66, 7.63, 7.06 and 7.05 ppm followed by the two doublets of the alkene protons at 6.48 and 6.31 ppm. Figure 8b shows the 13C NMR spectrum, which shows 11 signals, the two key peaks being at 166.7 and 163.1 ppm, and corresponding to the carbonyl carbons.

3. Conclusions

The ligand 2 was synthesized via a base catalyzed bis-maleimide ring opening in water in near quantitative yield (>99%). Ligands 4, 6, 8, 10, 12, 14 and 16 were synthesized via a reaction between the corresponding diamines and maleic anhydride in acetic acid as solvent in good to moderate yields (16–62%). Furthermore, ligands 12 and 16 have not been previously reported. The described processes use green solvents, mild reaction conditions, and a minimal work-up procedure, which makes them attractive alternatives for the synthesis of ligands with potential application for fabricating new MOFs.

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 (2E,2′E)-4,4′-(1,4-phenylenebis(azanediyl))bis(4-oxobut-2-enoic acid) (2)

In a 50 mL round-bottomed flask were added 0.27 g (1.0 equiv.) of 1,1′-(1,4-phenylene)bis(1H-pyrrole-2,5-dione) (1) into 20 mL of distilled water to obtain a turbid solution. Then, an aqueous solution of sodium hydroxide (10 mL, 0.5 M) was poured slowly into the mixture, avoiding surpassing a pH of 12 at room temperature. After 3 h of stirring, a yellow solution was obtained. This was acidified with [3 M] HCl until a pH of 2 was reached. This prompted a yellow precipitate to appear, which was collected by filtration, washed with distilled water and dried at room temperature to afford 0.30 g of a yellow powder in 99% yield. 1H NMR (500 MHz, d6-DMSO): δ 13.24 (s, 2H, H-16, H-20), 10.46 (s, 2H, H-7, H-10), 7.60 (s, 4H, H-2, H-3, H-5, H-6), 6.47 (d, J = 12.1 Hz, 2H, H-13, H-17), 6.31 (d, J = 12.1 Hz, 2H, H-14, H-18).

4.3. Synthesis of (2Z,2′Z)-4,4′-((methylenebis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (4)

In a 50 mL two-neck round-bottomed flask, a mixture of 0.20 g (2.0 equiv.) of maleic anhydride and 5 mL of acetic acid were heated until reflux. To this mixture, a solution of 0.20 g (1.0 equiv.) of 4,4′-methylendianiline in 5 mL of acetic acid was added dropwise. After the addition was completed, the mixture was kept at reflux until a precipitate was formed. The solution was let to cool at RT (room temperature) and 30 mL of water were added, the product was filtered and washed with an 8:2 water:ethanol mixture and was let to dry, affording 0.24 g of a white powder in 62% yield. 1H NMR (500 MHz, d6-DMSO): δ 13.07 (s, 2H, H-20, H-28), 10.39 (s, 2H, H-7, H-21), 7.54 (d, J = 8.6 Hz, 4H, H-2, H-6, H-13, H-15), 7.17 (d, J = 8.5 Hz, 4H, H-3, H-5, H-12, H-16), 6.46 (d, J = 12.1 Hz, 2H, H-17, H-25), 6.29 (d, J = 12.1 Hz, 2H, H-18, H-26), 3.86 (s, 1H, H-10); 13C NMR (126 MHz, d6-DMSO): δ 166.7 (C-8, C-22), 163.1 (C-19, C-27), 137.0 (C-1, C-14), 136.4 (C-4, C-11), 131.6 (C-17, C-25), 130. 7 (C-18, C-26), 129.0 (C-2, C-6, C-13, C-15), 119.8 (C-3, C-5, C-12, C-16), 40.0 (C-10).

4.4. Synthesis of (2Z,2′Z)-4,4′-((oxybis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (6)

In a 50 mL two-neck round-bottomed flask, a mixture of 0.20 g (2.0 equiv.) of maleic anhydride and 5 mL of acetic acid were heated until reflux. To this mixture, a solution of 0.20 g (1.0 equiv.) of 4,4′-oxydianiline in 5 mL of acetic acid was added dropwise. After the addition was completed, the mixture was kept at reflux until a precipitate was formed. The solution was let to cool at RT and 30 mL of water were added, the product was filtered and washed with an 8:2 water:ethanol mixture and was let to dry affording 0.20 g of a white powder in 50% yield. 1H NMR (500 MHz, d6-DMSO): δ 13.12 (s, 2H, H-23, H-27), 10.43 (s, 2H, H-14, H-17), 7.62 (d, J = 9.0 Hz, 4H, H-3, H-5, H-10, H-12), 6.98 (d, J = 9.0 Hz, 4H, H-2, H-6, H-9, H-13), 6.47 (d, J = 12.1 Hz, 2H, H-20, H-24), 6.29 (d, J = 12.1 Hz, 2H, H-21, H-25); 13C NMR (126 MHz, d6-DMSO) δ 166.6 (C-15, C-18), 163.0 (C-22, C-26), 152.9 (C-1, C-8), 134.0 (C-4, C-11), 131.8 (C-20, C-24), 130.5 (C-21, C-25), 121.3 (C-3, C-5, C-10, C-12), 118.8 (C-2, C-6, C-9, C-13).

4.5. Synthesis of (2Z,2′Z)-4,4′-((thiobis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (8)

In a 50 mL two-neck round-bottomed flask, a mixture of 0.20 g (2.0 equiv.) of maleic anhydride and 5 mL of acetic acid were heated until reflux. To this mixture, a solution of 0.22 g (1.0 equiv.) of 4,4′-thiodianiline in 5 mL of acetic acid was added dropwise. After the addition was completed, the mixture was kept at reflux until a precipitate was formed. The solution was let to cool at RT and 30 mL of water was added; the product was filtered and washed with an 8:2 water:ethanol mixture and was let to dry, affording 0.23 g of a grey powder in 55% yield. 1H NMR (500 MHz, d6-DMSO): δ 12.97 (s, 2H, H-23, H-27), 10.47 (2H, H-14, H-17), 7.63 (d, J = 8.7 Hz, 4H, H-3, H-5, H-10, H-12), 7.29 (d, J = 8.7 Hz, 4H, H-2, H-6, H-9, H-13), 6.46 (d, J = 12.0 Hz, 2H, H-20, H-24), 6.30 (d, J = 12.0 Hz, 2H, H-21, H-25); 13C NMR (126 MHz, d6-DMSO): δ 166.8 (C-15, C-18), 163.3 (C-22, C-26), 138.0 (C-4, C-11), 131.6 (C-20, C-24), 131.4 (C-3, C-5, C-10, C-12), 130.3 (C-21, C-25), 129.6 (C-1, C-8), 120.4 (C-2, C-6, C-9, C-13).

4.6. Synthesis of (2Z,2′Z)-4,4′-((sulfonylbis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (10)

In a 50 mL two-neck round-bottomed flask, a mixture of 0.20 g (2.0 equiv.) of maleic anhydride and 5 mL of acetic acid were heated until reflux. To this mixture, a solution of 0.25 g (1.0 equiv.) of 4,4′-sulfonyldianiline in 5 mL of acetic acid was added dropwise. After the addition was completed, the mixture was kept at reflux until a precipitate was formed. The solution was let to cool at RT and 30 mL of water was added; the product was filtered and washed with an 8:2 water:ethanol mixture and was let to dry, affording 0.17 g of a white powder in 39% yield. 1H NMR (500 MHz, d6-DMSO): δ 12.86 (s, 2H, H-22, H-30), 10.68 (s, 2H, H-9, H-23), 7.88 (d, J = 9.0 Hz, 4H, H-2, H-6, H-14, H-18), 7.82 (d, J = 9.0 Hz, 4H, H-3, H-5, H-15, H-17), 6.47 (d, J = 12.0 Hz, 2H, H-19, H-27), 6.31 (d, J = 12.0 Hz, 2H, H-20, H-28); 13C NMR (126 MHz, d6-DMSO): δ 166.8 (C-21, C-29), 163.8 (C-10, C-24), 143.1 (C-19, C-27), 135.6 (C-4, C-16), 131.6 (C-20, C-28), 130.2 (C-1, C-13), 128.4 (C-2, C-6, C-14, C-18), 119.4 (C-3, C-5, C-15, C-17).

4.7. Synthesis of (2Z,2′Z)-4,4′-((9,10-dioxo-9,10-dihydroanthracene-2,6-diyl)bis(azanediyl))bis(4-oxobut-2-enoic acid) (12)

In a 50 mL two-neck round-bottomed flask, a mixture of 0.20 g (2.0 equiv.) of maleic anhydride and 5 mL of acetic acid were heated until reflux. To this mixture, a solution of 0.24 g (1.0 equiv.) of 2,6-diaminoanthracene-9,10-dione in 5 mL of acetic acid was added dropwise. After the addition was completed, the mixture was kept at reflux until a precipitate was formed. The solution was let to cool at RT and 30 mL of water was added; the product was filtered and washed with an 8:2 water:ethanol mixture and was let to dry, affording 0.01 g of a red powder in 22% yield. 1H NMR (500 MHz, d6-DMSO): δ 12.88 (s, 2H, H-26, H-30), 10.93 (s, 2H, H-17, H-20), 8.47 (d, J = 2.4 Hz, 2H, H-10, H-11) 8.17 (d, J = 8.5 Hz, 2H, H-7, H-14), 8.07 (dd, J = 8.6, 2.3 Hz, 2H, H-8, H-13), 6.53 (d, J = 12.0 Hz, 2H, H-23, H-27), 6.37 (d, J = 11.9 Hz, 2H, H-24, H-28); 13C NMR (126 MHz, d6-DMSO): δ 181.2 (C-1, C-4), 166.9 (C-18, C-21), 164.0 (C-25, C-29), 144.3 (C-9, C-12), 134.3 (C-23, C-27), 131.5 (C-24, C-28), 130.5 (C-2, C-5), 128.5 (C-3, C-6), 128.3 (C-7, C-14), 123.7 (C-8, C-13), 116.1 (C-10, C-11).

4.8. Synthesis of (2Z,2′Z)-4,4′-((disulfanediylbis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (14)

In a 50 mL two-neck round-bottomed flask, a mixture of 0.20 g (2.0 equiv.) of maleic anhydride and 5 mL of acetic acid were heated until reflux. To this mixture, a solution of 0.25 g (1.0 equiv.) of 4,4′-disulfanediyldianiline in 5 mL of acetic acid was added dropwise. After the addition was completed, the mixture was kept at reflux until a precipitate was formed. The solution was let to cool at RT and 30 mL of water was added; the product was filtered and washed with an 8:2 water:ethanol mixture and was let to dry affording 0.15 g of a grey powder in 34% yield. 1H NMR (500 MHz, d6-DMSO): δ 12.94 (s, 2H, H-24, H-28), 10.48 (s, 2H, H-15, H-18), 7.64 (d, J = 8.8 Hz, 4H, H-3, H-5, H-11, H-13), 7.47 (d, J = 8.9 Hz, 4H, H-2, H-6, H-10, H-14), 6.46 (d, J = 12.1 Hz, 2H, H-21, H-25), 6.30 (d, J = 12.0 Hz, 2H, H-22, H-26); 13C NMR (126 MHz, d6-DMSO): δ 166.8 (C-16, C-19), 163.3 (C-23, C-27), 138.7 (C-4, C-12), 131.6 (C-21, C-25), 130.3 (C-22, C-26), 129.8 (C-2, C-6, C-10, C-14), 120.2 (C-3, C-5, C-11, C-13).

4.9. Synthesis of (2Z,2′Z)-4,4′-((([1,1′-biphenyl]-4,4′-diylbis(oxy))bis(4,1-phenylene))bis(azanediyl))bis(4-oxobut-2-enoic acid) (16)

In a 50 mL two-neck round-bottomed flask, a mixture of 0.20 g (2.0 equiv.) of maleic anhydride and 5 mL of acetic acid were heated until reflux. To this mixture, a solution of 0.37 g (1.0 equiv.) of 4,4′-([1,1′-biphenyl]-4,4′-diylbis(oxy))dianiline in 5 mL of acetic acid was added dropwise. After the addition was completed, the mixture was kept at reflux until a precipitate was formed. The solution was let to cool at RT and 30 mL of water was added; the product was filtered and washed with an 8:2 water:ethanol mixture and was let to dry, affording 0.09 g of a white powder in 16% yield. 1H NMR (500 MHz, d6-DMSO): δ 13.13 (s, 2H, H-26, H-41), 10.46 (s, 2H, H-14, H-35), 7.66 (d, J = 9.0 Hz, 4H, H-10, H-12, H-18, H-22), 7.63 (d, J = 8.7 Hz, 4H, H-3, H-5, H-30, H-32), 7.06 (d, J = 7.3 Hz, 4H, H-9, H-13, H-19, H-21), 7.05 (d, J = 6.9 Hz, 4H, H-2, H-6, H-29, H-33), 6.48 (d, J = 12.1 Hz, 2H, H-23, H-38), 6.31 (d, J = 12.0 Hz, 2H, H-24, H-39).; 13C NMR (126 MHz, d6-DMSO) δ 166.7 (C-15, C-36), 163.1 (C-25, C-40), 156.6 (C-8, C-20), 152.2 (C-1, C-28), 134.4 (C-4, C-31, C-11, C-17), 131.8 (C-23, C-38), 130.4 (C-24, C-39), 128.0 (C-10, C-12, C-18, C-22), 121.3 (C-3, C-5, C-30, C-32), 119.5 (C-9, C-13, C-19, C-21), 118.4 (C-2, C-6, C-29, C-33).

Author Contributions

All authors contributed equally to this work. 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. E.G.-Z. thanks CBI-UAM-I for financial support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

All authors thank Mónica Rincón-Guevara (CBS-UAM-I) for supporting the spectroscopic characterization.

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. Tabăcăru, A.; Pettinari, C.; Galli, S. Coordination polymers and metal-organic frameworks built up with poly(tetrazolate) ligands. Coord. Chem. Rev. 2018, 372, 1–30. [Google Scholar] [CrossRef]
  2. Lustig, W.P.; Mukherjee, S.; Rudd, N.D.; Desai, A.V.; Li, J.; Ghosh, S.K. Metal-organic frameworks: Functional luminescent and photonic materials for sensing applications. Chem. Soc. Rev. 2017, 46, 3242–3285. [Google Scholar] [CrossRef] [PubMed]
  3. Lytvynenko, A.S.; Kolotilov, S.V. Electrochemically active coordination polymers: A review. Theor. Exp. Chem. 2016, 52, 197–211. [Google Scholar] [CrossRef]
  4. Liu, J.; Chen, L.; Cui, H.; Zhang, J.; Zhang, L.; Su, C.-Y. Applications of metal–organic frameworks in heterogeneous supramolecular catalysis. Chem. Soc. Rev. 2014, 43, 6011–6061. [Google Scholar] [CrossRef]
  5. 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]
  6. 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]
  7. Robin, A.Y.; Fromm, K.M. Coordination polymer networks with O- and N-donors: What they are, why and how they are made. Coord. Chem. Rev. 2006, 250, 2127–2157. [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]
  9. 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. 172–173. [Google Scholar]
  10. Huang, Z.; Song, H.-B.; Du, M.; Chen, S.-T.; Bu, X.-H. 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]
  11. Niu, L.-E.; Yang, Y.; Qi, X.-H.; Liang, X. Synthesis and spectroscopic studies of Lanthanide(III) complexes with (Z)-4-oxo-4-(phenylamino) but-2-enoic acid. Synth. React. Inorg. Met. Org. Chem. 2009, 39, 1–5. [Google Scholar] [CrossRef]
  12. Li, M.; Liang, Y.-D.; Wu, Y.-X.; Li, K.-S. Synergistic effect of complexes of ethylenediamine double maleamic acid radical and lanthanum (III) with pentaerythritol on the thermal stability of poly(vinyl chloride). Polym. Degrad. Stabil. 2017, 140, 176–193. [Google Scholar] [CrossRef]
  13. Lazarou, K.N.; Psycharis, V.; Perlepes, S.P.; Raptopoulou, C.P. Complexes derived from the copper(II) perchlorate/maleamic acid/2,20-bipyridine and copper(II) perchlorate/maleic acid/2,20-bipyridine reaction systems: Synthetic, reactivity, structural and spectroscopic studies. Polyhedron 2009, 28, 1085–1096. [Google Scholar] [CrossRef]
  14. Lazarou, K.N.; Boudalis, A.K.; Perlepes, S.P.; Terzis, A.; Raptopoulou, C.P. Maleamate(–1) and maleate(–2) copper(II)–2,2′-bipyridine complexes: Synthesis, reactivity and structural and physical studies Eur. J. Inorg. Chem. 2009, 4554–4563. [Google Scholar] [CrossRef]
  15. Li, B.; Shen, Y.; Wu, H.; Wu, X.; Yuan, L.; Ji, Q. Design, synthesis and biological evaluation of novel 3,4-dihydro- 2(1H)-quinolinone derivatives as potential chitin synthase inhibitors and antifungal agents. Eur. J. Med. Chem. 2020, 195, 112278. [Google Scholar] [CrossRef] [PubMed]
  16. Sun, Y.; Zhang, G. Photoinduced decarboxylative amino-fluoroalkylation of maleic anhydride. Chem. Eur. J. 2020, 26, 419–422. [Google Scholar] [CrossRef] [PubMed]
  17. Badru, R.; Singh, B. Triethylamine-Catalyzed Synthesis of Oxazepine from Maleamic Acids. J. Heterocycl. Chem. 2015, 52, 635–640. [Google Scholar] [CrossRef]
  18. Guevara, J.A.; Trujillo, J.G.; Quintana, D.; Jiménez, H.A.; Arellano, M.G.; Bahena, J.R.; Tamay, F.; Ciprés, F.J. Acetylcholinesterase inhibition by products generated in situ from the transformation of N-arylisomaleimides. Med. Chem. Res. 2018, 27, 989–1003. [Google Scholar] [CrossRef]
  19. Long, Y.L.; Jiang, W.; Wu, X.; Wang, Q. Grafted copolymerization of N-phenylmaleimide and styrene in porous polyvinyl chloride particles suspended in aqueous solution. Des. Monomers Polym. 2019, 22, 66–78. [Google Scholar] [CrossRef]
  20. Kumar, P.P.; Devi, B.R.; Dubey, P.K.; Mohiuddin, S.M.G. PEG-600 mediated simple, efficient and eco-friendly synthesis of N-substituted imides and chemoselective C=C reduction. Green Chem. Lett. Rev. 2011, 4, 341–348. [Google Scholar] [CrossRef]
  21. Chen, Y.; Tssao, K.; De Francesco, E.; Keillor, J.W. Ring substituent effects on the thiol addition and hydrolysis reactions of N-arylmaleimides. J. Org. Chem. 2015, 80, 12182–12192. [Google Scholar] [CrossRef]
  22. Kiselev, V.D.; Kashaeva, H.A.; Potapova, L.N.; Kornilov, D.A.; Latypova, L.I.; Konovalov, A.I. Hydrophobic acceleration in the Diels—Alder reaction of 9-hydroxymethylanthracene with N-phenylmaleimide. Russ. Chem. Bull. Int. Ed. 2016, 65, 2202–2205. [Google Scholar] [CrossRef]
Scheme 1. General reaction scheme for the synthesis of maleamic acids.
Scheme 1. General reaction scheme for the synthesis of maleamic acids.
Chemproc 03 00092 sch001
Scheme 2. Synthesis of the ligand 2.
Scheme 2. Synthesis of the ligand 2.
Chemproc 03 00092 sch002
Figure 1. 1H NMR spectrum of compound 2.
Figure 1. 1H NMR spectrum of compound 2.
Chemproc 03 00092 g001
Scheme 3. Synthesis of ligand 4.
Scheme 3. Synthesis of ligand 4.
Chemproc 03 00092 sch003
Figure 2. (a) 1H NMR spectrum of ligand 4. (b) 13C NMR spectrum of ligand 4.
Figure 2. (a) 1H NMR spectrum of ligand 4. (b) 13C NMR spectrum of ligand 4.
Chemproc 03 00092 g002
Scheme 4. Synthesis of ligand 6.
Scheme 4. Synthesis of ligand 6.
Chemproc 03 00092 sch004
Figure 3. (a) 1H NMR spectrum of ligand 6. (b) 13C NMR spectrum of ligand 6.
Figure 3. (a) 1H NMR spectrum of ligand 6. (b) 13C NMR spectrum of ligand 6.
Chemproc 03 00092 g003
Scheme 5. Synthesis of ligand 8.
Scheme 5. Synthesis of ligand 8.
Chemproc 03 00092 sch005
Figure 4. (a)1H NMR spectrum of ligand 6. (b) 13C NMR spectrum of ligand 8.
Figure 4. (a)1H NMR spectrum of ligand 6. (b) 13C NMR spectrum of ligand 8.
Chemproc 03 00092 g004
Scheme 6. Synthesis of compound 10.
Scheme 6. Synthesis of compound 10.
Chemproc 03 00092 sch006
Figure 5. (a) 1H NMR spectrum of ligand 10. (b) 13C NMR spectrum of ligand 10.
Figure 5. (a) 1H NMR spectrum of ligand 10. (b) 13C NMR spectrum of ligand 10.
Chemproc 03 00092 g005
Scheme 7. Synthesis of ligand 12.
Scheme 7. Synthesis of ligand 12.
Chemproc 03 00092 sch007
Figure 6. (a) 1H NMR spectrum of ligand 12. (b) 13C NMR spectrum of ligand 12.
Figure 6. (a) 1H NMR spectrum of ligand 12. (b) 13C NMR spectrum of ligand 12.
Chemproc 03 00092 g006
Scheme 8. Synthesis of ligand disulfane bridged 14.
Scheme 8. Synthesis of ligand disulfane bridged 14.
Chemproc 03 00092 sch008
Figure 7. (a) 1H NMR spectrum of disulfane bridged ligand 14. (b) 13C NMR spectrum of disulfane bridged ligand 14.
Figure 7. (a) 1H NMR spectrum of disulfane bridged ligand 14. (b) 13C NMR spectrum of disulfane bridged ligand 14.
Chemproc 03 00092 g007
Scheme 9. Synthesis of ligand 16.
Scheme 9. Synthesis of ligand 16.
Chemproc 03 00092 sch009
Figure 8. (a) 1H NMR spectrum of ligand 16. (b) 13C NMR spectrum of ligand 16.
Figure 8. (a) 1H NMR spectrum of ligand 16. (b) 13C NMR spectrum of ligand 16.
Chemproc 03 00092 g008
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.; Sosa-Juárez, J.L.; Sánchez-Serratos, M.; Islas-Jácome, P.; Gutiérrez-Carrillo, A.; Méndez, F.; Suárez-Moreno, G.V.; Islas-Jácome, A.; González-Zamora, E. Green Synthesis of Symmetric Dimaleamic Acids from Dianilines and Maleic Anhydride: Behind New Bidentate Ligands for MOFs. Chem. Proc. 2021, 3, 92. https://0-doi-org.brum.beds.ac.uk/10.3390/ecsoc-24-08379

AMA Style

Flores-Reyes JC, Sosa-Juárez JL, Sánchez-Serratos M, Islas-Jácome P, Gutiérrez-Carrillo A, Méndez F, Suárez-Moreno GV, Islas-Jácome A, González-Zamora E. Green Synthesis of Symmetric Dimaleamic Acids from Dianilines and Maleic Anhydride: Behind New Bidentate Ligands for MOFs. Chemistry Proceedings. 2021; 3(1):92. https://0-doi-org.brum.beds.ac.uk/10.3390/ecsoc-24-08379

Chicago/Turabian Style

Flores-Reyes, Julio C., José L. Sosa-Juárez, Mayra Sánchez-Serratos, Perla Islas-Jácome, Atilano Gutiérrez-Carrillo, Francisco Méndez, Galdina V. Suárez-Moreno, Alejandro Islas-Jácome, and Eduardo González-Zamora. 2021. "Green Synthesis of Symmetric Dimaleamic Acids from Dianilines and Maleic Anhydride: Behind New Bidentate Ligands for MOFs" Chemistry Proceedings 3, no. 1: 92. https://0-doi-org.brum.beds.ac.uk/10.3390/ecsoc-24-08379

Article Metrics

Back to TopTop