Next Article in Journal
The Preparation of a Highly Efficient Ag3PO4/Ag/Bi2O2CO3 Photo-Catalyst and the Study of Its Photo-Catalytic Organic Synthesis Reaction Driven by Visible Light
Next Article in Special Issue
Terpyridine-Containing Imine-Rich Graphene for the Oxygen Reduction Reaction
Previous Article in Journal
Improving the Indigo Carmine Decolorization Ability of a Bacillus amyloliquefaciens Laccase by Site-Directed Mutagenesis
Previous Article in Special Issue
SnSx (x = 1, 2) Nanocrystals as Effective Catalysts for Photoelectrochemical Water Splitting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrocarboxylation of Dichlorobenzenes on a Silver Electrode in DMF

1
Shandong Provincial Key Laboratory of Chemical Energy Storage and Novel Cell Technology, College of Chemistry and Chemical Engineering, Liaocheng University, Liaocheng 252059, China
2
Shanghai Key Laboratory of Green Chemistry and Chemical Processes, School of Chemistry and Molecular Engineering, East China Normal University, Shanghai 200062, China
*
Authors to whom correspondence should be addressed.
Catalysts 2017, 7(9), 274; https://doi.org/10.3390/catal7090274
Submission received: 25 August 2017 / Revised: 12 September 2017 / Accepted: 13 September 2017 / Published: 15 September 2017
(This article belongs to the Special Issue Advances in Electrocatalysis)

Abstract

:
Carbon dioxide (CO2) is the largest contributor to the greenhouse effect, and fixing and using this greenhouse gas in a facile manner is crucial. This work investigates the electrocarboxylation of dichlorobenzenes with the atmospheric pressure of CO2 in an undivided cell with an Ag cathode and an Mg sacrificial anode. The corresponding carboxylic acids and their derivatives, which are important industrial and fine chemicals, are obtained. To deeply understand this reaction, we investigate the influence of various reaction conditions, such as supporting electrolyte, current density, electric charge, and reaction temperature, on the electrocarboxylation yield by using 1,4-dichlorobenzene as the model compound. The electrochemical behavior of dichlorobenzenes is studied through cyclic voltammetry. The relation among the distinct electronic effects of dichlorobenzenes, the electrochemical characteristics of their reduction, and the distribution law of target products is also established.

Graphical Abstract

1. Introduction

The growing carbon dioxide (CO2) content in the atmosphere is the major contributor to the greenhouse effect and global warming [1]. Measures to address the greenhouse effect can be implemented in two ways. The first one is to create a low-carbon economic development model that conserves energy and controls energy from the source. The second method is to convert CO2 into useful organic chemicals [2,3,4]. Continuous emission of CO2 is expected to occur for a long time due to human dependence on fossil fuel. Therefore, identifying effective means to convert and utilize CO2 has become a pressing issue in this modern society; moreover, the fixation of CO2 into useful products of industrial significance has elicited considerable attention [5,6,7,8,9].
Although CO2 is the main gas that causes the greenhouse effect, CO2 is an abundant, cheap, non-toxic, recyclable C1 building block for organic synthesis. However, CO2 conversion is a challenge to chemists because of the high chemical and thermodynamic stability of this gas. Electrochemical techniques [10] that deal with a clean reagent, the electron, can realize CO2 activation at normal temperature and pressure [11,12,13]. Therefore, electrocarboxylation is a facile method of CO2 fixation from ecological and economic viewpoints. Many substrate types, alcohols [14], epoxides [15,16], imines [17], organic halides [13,18,19,20,21,22,23,24,25,26,27,28], ketones [11,12,29,30,31,32], aldehydes [33], alkenes [34,35,36,37,38], and alkynes [39] have been used in previous studies.
Carboxylic acids and their derivatives are an important class of compounds, and several of them are fine chemicals of industrial interest, especially in the production of anti-inflammatory drugs. Electrocarboxylation of organic halides is a potential facile means to synthesize carboxylated products. Homogeneous [40,41,42] and heterogeneous [19,21,22,43,44] catalytic methods have been adopted to optimize this reaction because of the negative reduction potential of organic chlorides. A good catalytic electrode material can effectively reduce the reduction potential and improve the reaction. Hg has been the main material choice for a long time. However, Hg is discarded for environmental reasons. Meanwhile, studies have indicated that silver (Ag) exhibits extraordinary electrocatalytic activities toward the reduction of organic halides [13,18,19,20,21,22,23,43,44]. Although many studies have investigated electroreduction and electrocarboxylation of organic chlorides, most of them focused on monochloride compounds [19,21,22,40,41,42,43,44]. Only a few studies have examined the electroreduction of polyhalidecompounds [45,46,47,48,49,50,51,52,53,54,55,56,57,58,59], such as dichloromethane, trichloromethane, tetrachloromethane, and hexachlorocyclohexane.
In this study, electroreduction of dichlorobenzenes on an Ag electrode in the presence of CO2 under mild conditions is described (Scheme 1). The aim of this investigation is to set up an alternative methodology for electrocarboxylation of dichlorobenzenes with CO2 to the corresponding carboxylic acid, thus providing an environmentally friendly method for the electrochemical fixation of CO2 with dichlorobenzenes possessing aromatic carboxylic acids and their derivatives, which are useful for pharmaceutical or industrial application.

2. Results and Discussion

2.1. Electroanalytical Measurement of 1,4-Dichlorobenzene

Figure 1 presents the voltammetric characteristic of 1,4-dichlorobenzene (1a) in N,N-dimethylformamide (DMF) containing 0.1 M tetraethylammonium tetrafluoroborate (TEABF4) on glassy carbon (GC), and Ag electrodes at a scan rate of 0.1 V/s. As shown in curve a, no redox peak on the GC electrode existed within the scan range, which revealed a favorable potential window in the region of −0.6 to −2.5 V. Two successive irreversible cathodic peaks were detected at −2.04 V and −2.30 V (curve b) after 1a was added to the solution. The first reduction peak currents varied linearly with v1/2 (Figure S1, Supplementary Materials), which confirmed the diffusion control for the electroreduction process. The voltammetric characteristic of chlorobenzene on the GC electrode was also recorded for comparison. Curve c displayed an irreversible reduction peak at −2.30 V corresponding to the two-electron transfer of the C–Cl bond of chlorobenzene, similar to the second reduction peak of 1a on the GC electrode [50]. This result implies that the first cathodic peak was caused by the two-electron reductive cleavage of one of the C–Cl bonds, and the second one was caused by the reduction of another C–Cl bond. The gas chromatography–mass spectrometry (GC–MS) analysis also confirmed the synthesis of chlorobenzene through potentiostatic electrolysis at the first reduction peak in the presence of N2. When the voltammetric characteristic of 1a was recorded in DMF saturated with CO2 (0.2 M) [51], a different behavior was observed. As shown in curve d of Figure 1, the first reduction peak current increased, with its potential appreciably shifting to a more positive place, indicating that a chemical reaction may have occurred between the electroreduced intermediate and CO2. Meanwhile, the second reduction peak disappeared. The voltammetric characteristic of 1a was checked on the Ag electrode for comparison. As shown in curves b and e, the two reduction peaks shifted to a more positive potential because of the electrocatalytic properties of the Ag electrode. The electrocatalytic activities of electrode materials for the reduction of organic halides have been the subject of many studies [18,19,21,22,52,53,54]. These studies have shown that Ag exhibits powerful electrocatalytic activities toward the reduction of various organic halides. Therefore, the electroreduction of dichlorobenzenes on the Ag cathode was investigated herein.

2.2. Electrocarboxylation of Dichlorobenzene

Considering that electrolysis carried out under galvanostatic conditions is a much more convenient procedure than that implemented under potentiostatic conditions, the reaction conditions for the galvanostatic electrolysis of 1a were examined and are summarized in Table 1. The yields of mono- and di-carboxylation products were affected by the choice of supporting electrolyte (SE), current density (J), electric charge (Q), and reaction temperature (T).
SE strongly determined the effects of the electrocarboxylation of 1a. Tetrabutyl ammonium chloride (TBACl) was the best SE, and resulted in 36.7% total electrocarboxylation yield (Table 1, entry 5). Notably, with the same tetraalkylammonium cation, the total carboxylation yield dependences on the anion were in the following order: Cl−1 > Br−1 > I−1 (Table 1, entries 1–3, 5–7). Although negatively charged in the Ag electrode, inorganic halide anions have been reported to remain powerfully absorbed in the Ag electrode in the sequence of Cl−1 ≤ Br−1 ≤ I−1 and are thus accorded with many well-known halide properties, such as the solubility of products with Ag ions [18,55]. In addition, with the same anion of Cl−1, Br−1, and I−1, the synthesis carried out with the TBA+ cation provided better yields than those of the TEA+ cation cases (Table 1, entries 1–3, 5–7), indicating that the cations exerted some influence on the electrolysis.
The electrocarboxylation yields were also strongly affected by J (Table 1, entries 5, 8–12). The highest total electrocarboxylation yield of 43.6% was achieved with 14 mA/cm2 (Table 1, entry 10). The total electrocarboxylation yield decreased regardless of whether the current density was higher or lower than 14 mA/cm2. On the one hand, the electrode potential became negative when the current density was high. Hence a series of side reactions, such as the reduction of Mg2+ or CO2 to C1 or C2 compounds (carbonate, oxalate, and CO), occurred on the electrode surface and reduced the total electrocarboxylation yield [56]. On the other hand, the electrode potential became positive when current density was low. Therefore, the proportion of the Faradaic current was reduced, thus decreasing the carboxylation yield.
The influence of Q was also investigated (Table 1, entries 10, 13–17). The Faradaic efficiency did not reach 100% due to the existence of non-Faradaic current. Accordingly, an amount that is more than the theoretic charge was consumed to improve the yield in general. The best total yield of 45.3% was obtained when the passed electric charge reached 4.2 F/mol (Table 1, entry 15). Before the point of 4.2 F/mol, the total yield increased rapidly with the increase in the passed charge. Afterward, the total yield decreased to 30.5% for 5 F/mol.
T also influenced the electrolysis reaction and mainly affected the thermodynamics and kinetics of the electrocarboxylation reaction and the solubility of CO2 in the solvent [51]. In general, a low T increases the solubility of CO2 in DMF, and is conducive to the reaction. A low T also reduces the rate of mass transport and increases the activation energy of the reaction, both of which are dangerous to the electrocarboxylation of 1a. To investigate the effect of T, a set of electrolysis experiments were carried out at different T values (Table 1, entries 15, 18–22). Increasing T from −10 °C to 5 °C favored the reaction, but further increasing T resulted in a low total carboxylation yield of 1a. The best total carboxylation product with 46.3% yield was achieved at 5 °C (Table 1, entry 20).

2.3. Electrocarboxylation of Other Dichlorobenzenes

To investigate the effect of the C–Cl bond position on the benzene ring, cyclic voltammograms and electrolysis were extended to 1,3-dichlorobenzene (1b) and 1,2-dichlorobenzene (1c) under the optimized reaction conditions.
As shown in Figure 2, two successive irreversible cathodic peaks corresponding to the two-electron transfer reaction of the C–Cl bond were detected for all the three dichlorobenzenes on the Ag electrode. The cathodic peak potentials of the three dichlorobenzenes were very close. Electrolysis of 1b and 1c was conducted under the optimized conditions. As shown in Table 2, halogenated monocarboxylation product 2 was obtained for all the three dichlorobenzenes, and the yield of 2 decreased from 43.7% to 38.4% in the following order: 2c > 2b > 2a. The yield correlated with the first peak current of 1 (Figure 2): 1c (0.103 mA) > 1b (0.097 mA) > 1a (0.079 mA). High yields were obtained when the first peak current of 1 was large. In addition, dehalogenated monocarboxylation product 3 was detected for all the three dichlorobenzenes. Surprisingly, dicarboxylation products 4b and 4c were not detected. To understand this phenomenon, the proposed pathways for the electrocarboxylation of dichlorobenzene were studied, and energy calculations were performed using the Gaussian09W program (Revision A.1, Gaussian, Inc., Wallingford, CT, USA, 2009).
For monoaromatic halide compounds, carboxylated products were normally obtained by the reaction of CO2 with the anions that were generated by the two-electron transfer of monoaromatic halides [20,25,57,58]. Therefore, the possible electrocarboxylation pathway of dichlorobenzene is Scheme 2. The first electron reduction of 1 results in the carbon radical 5. After the second electron transfer, 5 is reduced to carbon anion 6, which reacts with CO2 to form halogenated monocarboxylation product 2. The phenomenon in which two successive irreversible cathodic peaks were detected for all the three dichlorobenzenes (Figure 2) is indicative of two two-electron transfer reactions for all three dichlorobenzenes. Therefore, 2 may undergo further two-electron transfer to the corresponding carbon anion 8, which would react with another CO2 to form decarboxylation product 4. Given that a small amount of water is still present in the solvent, the protonation process competes with carboxylation [13,58,59]. Then, dehalogenated monocarboxylation product 3 is formed by either carboxylation following protonation or protonation following carboxylation.
On the basis of the electrolysis result (Table 2) and the proposed electrocarboxylation pathway, dicarboxylation products 4b and 4c were not obtained from 1b and 1c, respectively. This finding indicates that several steps may have been blocked from 2b to 4b and from 2c to 4c. To determine the reason, the Gaussian09W program was used to perform gas-phase geometry optimizations and energy calculations using the density functional theory (DFT) B3LYP/6-31G* method [60]. The optimized structures with the net charges on the carbon radicals are shown in Scheme 2. The charge numbers of the carbon radical of 7a, 7b, and 7c were −0.041, −0.047, and −0.053, respectively. Electrophilicity decreased in the following order: 7a > 7b > 7c. 8 was difficult to obtain with weak electrophilicity. Generally, strong electron donation of the substrate goes against electrocarboxylation [29]. When 1b or 1c was used as a substrate, the weak electrophilicity of the carbon radical of 7b or 7c might have significantly prevented 7b or 7c from undergoing further one-electron transfer to the corresponding carbon anion 8b or 8c. Scheme 2 indicates that the charge numbers of the carbon anion of 8a, 8b, and 8c were −0.294, −0.288, and −0.283, respectively. Nucleophilicity decreased in the following order: 8a > 8b > 8c. When the electrocarboxylation was carried out using 1b or 1c as a substrate, the weak nucleophilicity of the carbon anion of 8b or 8c might have prevented 8b or 8c from attacking CO2 in a nucleophilic manner even when a small amount of 8b or 8c was generated [18]. Furthermore, based on the energy calculations, the energy differences of 4b8b and 4c8c were surprisingly higher than that of 4a8a (Scheme 2). Accordingly, the gas-phase geometry optimizations and energy calculations corroborated the phenomenon that the corresponding dicarboxylation products, 4b and 4c, were not obtained.
Our investigation of the electrocarboxylation of dichlorobenzenes on an Ag electrode showed that the initial stage was the formation of carbon anions that could attack CO2 in a nucleophilic manner to form halogenated monocarboxylation product 2 or dehalogenated monocarboxylation 3. Afterward, 2 was further reduced by another two-electron transfer followed by a reaction with CO2 to form dicarboxylation product 4, or by protonation to 3. The electronic effect of different C–Cl bond positions on the benzene ring played a decisive role in reaction efficiency and product distribution. When 1a was used as a substrate, three carboxylation products were obtained. By contrast, no dicarboxylation product was available for 1b or 1c when the electrophilicity of the corresponding carbon radical, and the nucleophilicity of the corresponding carbon anion, were weak.

3. Materials and Methods

3.1. Materials and Product Measurement

All dichlorobenzenes were commercially available from J&K Chemical Co. (Beijing, China). DMF was kept over 4 Å molecular sieves. All other reagents were used as received. The cyclic voltammograms were measured with a CHI 760E electrochemical station (Shanghai Chenhua Instruments Co., Shanghai, China) at 25 °C in an undivided glass cell equipped with a gas inlet and outlet. Galvanostatic electrolysis was performed with a DC-regulated power supply HY5003M (Hangzhou Huayi Electronics Industry Co., Ltd., Hangzhou, China) in an undivided glass cell. The product yields were determined through GC-2014C gas chromatography (GC) (SHIMADZU, Tokyo, Japan).

3.2. General Electroanalytical Procedure

The electroanalytical experiments were performed in a DMF solution containing 0.1 M TEABF4 as the SE. The working electrodes used for the electroanalytical experiments were GC with a diameter of 2 mm and Ag (d = 2 mm). The auxiliary and reference electrodes were a platinum (Pt) spiral and Ag/AgI/0.1 M TBAI in DMF, respectively. All experiments were performed at atmospheric pressure.

3.3. General Electrolysis Procedure

Galvanostatic electrolysis was conducted in a mixture of dichlorobenzenes (0.1 M) and SE (0.1 M) in 10 mL DMF solvent saturated with CO2 in an undivided electrochemical cell equipped with an Ag cathode (8 cm2) and a sacrificial Mg anode. At the end of each set of experiments, the reaction mixture was esterified in DMF by adding anhydrous K2CO3 (2 mmol) and MeI (9 mmol) at 55 °C for 5 h. Then, the solution was hydrolyzed and extracted with diethyl ether, and the combined organic layers were dried over anhydrous MgSO4, filtered, and evaporated. The yields of methyl esters corresponding to acids were determined via GC with decane as the internal standard.

4. Conclusions

We demonstrated a simple and efficient electrochemical means to achieve the electroreduction of dichlorobenzenes in the presence of CO2 on an Ag electrode. The impact of SE, J, Q, and T on the total electrocarboxylation yield was studied with 1,4-dichlorobenzene as the model compound. The electronic effect of different C–Cl bond positions on the benzene ring played a decisive role in reaction efficiency and product distribution. This work provides an environmentally friendly method for the electrochemical fixation of CO2 with dichlorobenzenes possessing aromatic carboxylic acids and their derivatives, which are useful for pharmaceutical or industrial use. Thus, the applications of electrochemical methodology are expanded.

Supplementary Materials

The following are available online at www.mdpi.com/2073-4344/7/9/274/s1, Figure S1: Cyclic voltammograms of 5 mM 1,4-dichlorobenzene in DMF containing 0.1 M TEABF4 on GC electrode at (a) v = 0.1 V/s; (b) 0.2 V/s; (c) 0.3 V/s; (d) 0.4 V/s; (e) 0.5 V/s; (f) 0.6 V/s; (g) 0.7 V/s; (h) 0.8 V/s; (i) 0.9 V/s and (j) 1.0 V/s.

Acknowledgments

This work was financially supported by the funds from the National Natural Science Foundation of China (21503104, 21601078, 21401094), Natural Science Foundation of Shandong Province (ZR2014BQ010, ZR2016BQ20, ZR2016BQ21), Science and Technology Development Project Fund of Liaocheng (2014GJH08), Research Fund for the Doctoral Program of Liaocheng University (31805), and Colleges and Universities in Shandong Province Science and Technology Projects (J16LC05, J17KA097).

Author Contributions

Pei-Pei Luo, Ying-Tian Zhang and Bao-Li Chen conceived of and designed the experiments. Pei-Pei Luo, Ying-Tian Zhang, Bao-Li Chen, Shu-Xian Yu, Hua-Wei Zhou, Kong-Gang Qu, Yu-Xia Kong, Xian-Qiang Huang and Xian-Xi Zhang performed the experiments. Pei-Pei Luo and Bao-Li Chen analyzed the data. Pei-Pei Luo, Bao-Li Chen, Xian-Xi Zhang and Jia-Xing Lu wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Karl, T.R.; Trenberth, K.E. Modern global climate change. Science 2003, 302, 1719. [Google Scholar] [CrossRef] [PubMed]
  2. Sakakura, T.; Choi, J.C.; Yasuda, H. Transformation of carbon dioxide. Chem. Rev. 2007, 107, 2365–2387. [Google Scholar] [CrossRef] [PubMed]
  3. Habisreutinger, S.N.; Schmidt-Mende, L.; Stolarczyk, J.K. Photocatalytic reduction of CO2 on TiO2 and other semiconductors. Angew. Chem. Int. Ed. 2013, 52, 7372–7408. [Google Scholar] [CrossRef] [PubMed]
  4. Zhao, Y.; Yu, B.; Yang, Z.; Zhang, H.; Hao, L.; Gao, X.; Liu, Z. A protic ionic liquid catalyzes Co(2) conversion at atmospheric pressure and room temperature: Synthesis of quinazoline-2,4(1 h,3 h)-diones. Angew. Chem. Int. Ed. 2014, 53, 5922–5925. [Google Scholar] [CrossRef] [PubMed]
  5. Min, X.; Kanan, M.W. Pd-catalyzed electrohydrogenation of carbon dioxide to formate: High mass activity at low overpotential and identification of the deactivation pathway. J. Am. Chem. Soc. 2015, 137, 4701–4708. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, S.; Kang, P.; Meyer, T.J. Nanostructured tin catalysts for selective electrochemical reduction of carbon dioxide to formate. J. Am. Chem. Soc. 2014, 136, 1734–1737. [Google Scholar] [CrossRef] [PubMed]
  7. Yuan, J.; Liu, L.; Guo, R.R.; Zeng, S.; Wang, H.; Lu, J.X. Electroreduction of CO2 into ethanol over an active catalyst: Copper supported on titania. Catalysts 2017, 7, 220. [Google Scholar] [CrossRef]
  8. Liu, Q.; Wu, L.; Jackstell, R.; Beller, M. Using carbon dioxide as a building block in organic synthesis. Nat. Commun. 2015, 6, 5933. [Google Scholar] [CrossRef] [PubMed]
  9. Gao, S.; Lin, Y.; Jiao, X.; Sun, Y.; Luo, Q.; Zhang, W.; Li, D.; Yang, J.; Xie, Y. Partially oxidized atomic cobalt layers for carbon dioxide electroreduction to liquid fuel. Nature 2016, 529, 68–71. [Google Scholar] [CrossRef] [PubMed]
  10. Chen, B.L.; Xiao, Y.; Xu, X.M.; Yang, H.P.; Wang, H.; Lu, J.X. Alkaloid induced enantioselective electroreduction of acetophenone. Electrochim. Acta 2013, 107, 320–326. [Google Scholar] [CrossRef]
  11. Chen, B.L.; Tu, Z.Y.; Zhu, H.W.; Sun, W.W.; Wang, H.; Lu, J.X. CO2 as a C1-organic building block: Enantioselective electrocarboxylation of aromatic ketones with CO2 catalyzed by cinchona alkaloids under mild conditions. Electrochim. Acta 2014, 116, 475–483. [Google Scholar] [CrossRef]
  12. Chen, B.L.; Zhu, H.W.; Xiao, Y.; Sun, Q.L.; Wang, H.; Lu, J.X. Asymmetric electrocarboxylation of 1-phenylethyl chloride catalyzed by electrogenerated chiral [CoI(salen)] complex. Electrochem. Commun. 2014, 42, 55–59. [Google Scholar] [CrossRef]
  13. Lan, Y.C.; Wang, H.; Wu, L.X.; Zhao, S.F.; Gu, Y.Q.; Lu, J.X. Electroreduction of dibromobenzenes on silver electrode in the presence of CO2. J. Electroanal. Chem. 2012, 664, 33–38. [Google Scholar] [CrossRef]
  14. Wu, L.X.; Wang, H.; Xiao, Y.; Tu, Z.Y.; Ding, B.B.; Lu, J.X. Synthesis of dialkyl carbonates from CO2 and alcohols via electrogenerated N-heterocyclic carbenes. Electrochem. Commun. 2012, 25, 116–118. [Google Scholar] [CrossRef]
  15. Xiao, Y.; Chen, B.L.; Yang, H.P.; Wang, H.; Lu, J.X. Electrosynthesis of enantiomerically pure cyclic carbonates from CO2 and chiral epoxides. Electrochem. Commun. 2014, 43, 71–74. [Google Scholar] [CrossRef]
  16. Khoshro, H.; Zare, H.R.; Namazian, M.; Jafari, A.A.; Gorji, A. Synthesis of cyclic carbonates through cycloaddition of electrocatalytic activated CO2 to epoxides under mild conditions. Electrochim. Acta 2013, 113, 263–268. [Google Scholar] [CrossRef]
  17. Feroci, M.; Inesi, A.; Rossi, L. The reaction of amines with an electrogenerated base. Improved synthesis of arylcarbamic esters. Tetrahedron Lett. 2000, 41, 963–966. [Google Scholar] [CrossRef]
  18. Niu, D.F.; Xiao, L.P.; Zhang, A.J.; Zhang, G.R.; Tan, Q.Y.; Lu, J.X. Electrocatalytic carboxylation of aliphatic halides at silver cathode in acetonitrile. Tetrahedron 2008, 64, 10517–10520. [Google Scholar] [CrossRef]
  19. Isse, A.A.; Gennaro, A. Electrocatalytic carboxylation of benzyl chlorides at silver cathodes in acetonitrile. Chem. Commun. 2002, 2798–2799. [Google Scholar] [CrossRef]
  20. Isse, A.A.; Durante, C.; Gennaro, A. One-pot synthesis of benzoic acid by electrocatalytic reduction of bromobenzene in the presence of CO2. Electrochem. Commun. 2011, 13, 810–813. [Google Scholar] [CrossRef]
  21. Isse, A.A.; De Giusti, A.; Gennaro, A.; Falciola, L.; Mussini, P.R. Electrochemical reduction of benzyl halides at a silver electrode. Electrochim. Acta 2006, 51, 4956–4964. [Google Scholar] [CrossRef]
  22. Scialdone, O.; Galia, A.; Errante, G.; Isse, A.A.; Gennaro, A.; Filardo, G. Electrocarboxylation of benzyl chlorides at silver cathode at the preparative scale level. Electrochim. Acta 2008, 53, 2514–2528. [Google Scholar] [CrossRef]
  23. Chen, D.; Fabre, P.L.; Reynes, O. Electrocarboxylation of chloroacetonitrile by a cobalt(I) complex of terpyridine. Electrochim. Acta 2011, 56, 8603–8610. [Google Scholar] [CrossRef] [Green Version]
  24. Feng, Q.; Huang, K.; Liu, S.; Wang, X. Electrocatalytic carboxylation of 2-amino-5-bromopyridine with CO2 in ionic liquid 1-butyl-3-methyllimidazoliumtetrafluoborate to 6-aminonicotinic acid. Electrochim. Acta 2010, 55, 5741–5745. [Google Scholar] [CrossRef]
  25. Tateno, H.; Nakabayashi, K.; Kashiwagi, T.; Senboku, H.; Atobe, M. Electrochemical fixation of CO2 to organohalides in room-temperature ionic liquids under supercritical CO2. Electrochim. Acta 2015, 161, 212–218. [Google Scholar] [CrossRef]
  26. Wang, H.M.; Sui, G.J.; Wu, D.; Feng, Q.; Wang, H.; Lu, J.X. Selective electrocarboxylation of bromostyrene at silver cathode in dmf. Tetrahedron 2016, 72, 968–972. [Google Scholar] [CrossRef]
  27. Zhao, S.F.; Wang, H.; Lan, Y.C.; Liu, X.; Lu, J.X.; Zhang, J. Influences of the operative parameters and the nature of the substrate on the electrocarboxylation of benzophenones. J. Electroanal. Chem. 2012, 664, 105–110. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, H.; He, L.; Sui, G.J.; Lu, J.X. Electrocatalytic reduction of PhCH2Br on a Ag–Y zeolite modified electrode. RSC Adv. 2015, 5, 42663–42665. [Google Scholar] [CrossRef]
  29. Sui, G.J.; Sun, Q.L.; Wu, D.; Meng, W.J.; Wang, H.; Lu, J.X. Electrocatalytic reduction of PhCH2Cl on Ag-zsm-5 zeolite modified electrode. RSC Adv. 2016, 6, 63493–63496. [Google Scholar] [CrossRef]
  30. Scialdone, O.; Sabatino, M.A.; Belfiore, C.; Galia, A.; Paternostro, M.P.; Filardo, G. An unexpected ring carboxylation in the electrocarboxylation of aromatic ketones. Electrochim. Acta 2006, 51, 3500–3505. [Google Scholar] [CrossRef]
  31. Feng, Q.; Huang, K.; Liu, S.; Yu, J.; Liu, F. Electrocatalytic carboxylation of aromatic ketones with carbon dioxide in ionic liquid 1-butyl-3-methylimidazoliumtetrafluoborate to α-hydroxy-carboxylic acid methyl ester. Electrochim. Acta 2011, 56, 5137–5141. [Google Scholar] [CrossRef]
  32. Saravanan, K.R.; Chandrasekaran, M.; Suryanarayanan, V. Efficient electrocarboxylation of benzophenone on silver nanoparticles deposited boron doped diamond electrode. J. Electroanal. Chem. 2015, 757, 18–22. [Google Scholar] [CrossRef]
  33. Doherty, A.P. Electrochemical reduction of butyraldehyde in the presence of CO2. Electrochim. Acta 2002, 47, 2963–2967. [Google Scholar] [CrossRef]
  34. Wang, H.; Zhang, G.; Liu, Y.; Luo, Y.; Lu, J. Electrocarboxylation of activated olefins in ionic liquid BMIMBF4. Electrochem. Commun. 2007, 9, 2235–2239. [Google Scholar] [CrossRef]
  35. Zhang, K.; Xiao, Y.; Lan, Y.; Zhu, M.; Wang, H.; Lu, J. Electrochemical reduction of aliphatic conjugated dienes in the presence of carbon dioxide. Electrochem. Commun. 2010, 12, 1698–1702. [Google Scholar] [CrossRef]
  36. Li, J.; Inagi, S.; Fuchigami, T.; Hosono, H.; Ito, S. Selective monocarboxylation of olefins at 12CaO·7Al2O3 electride cathode. Electrochem. Commun. 2014, 44, 45–48. [Google Scholar] [CrossRef]
  37. Yuan, G.Q.; Jiang, H.F.; Lin, C.; Liao, S.J. Efficient electrochemical synthesis of 2-arylsuccinic acids from CO2 and aryl-substituted alkenes with nickel as the cathode. Electrochim. Acta 2008, 53, 2170–2176. [Google Scholar] [CrossRef]
  38. Steinmann, S.N.; Michel, C.; Schwiedernoch, R.; Wu, M.; Sautet, P. Electro-carboxylation of butadiene and ethene over Pt and Ni catalysts. J. Catal. 2016, 343, 240–247. [Google Scholar] [CrossRef]
  39. Yuan, G.Q.; Jiang, H.F.; Lin, C. Efficient electrochemical dicarboxylations of arylacetylenes with carbon dioxide using nickel as the cathode. Tetrahedron 2008, 64, 5866–5872. [Google Scholar] [CrossRef]
  40. Damodar, J.; Krishna Mohan, S.R.; Jayarama Reddy, S.R. Synthesis of 2-arylpropionic acids by electrocarboxylation of benzylchlorides catalysed by PdCl2(PPh3)2. Electrochem. Commun. 2001, 3, 762–766. [Google Scholar] [CrossRef]
  41. Fabre, P.L.; Reynes, O. Electrocarboxylation of chloroacetonitrile mediated by electrogenerated cobalt(I) phenanthroline. Electrochem. Commun. 2010, 12, 1360–1362. [Google Scholar] [CrossRef] [Green Version]
  42. Bessel, C.A.; Rolison, D.R. Electrocatalytic reactivity of zeolite-encapsulated Co(salen) with benzyl chloride. J. Am. Chem. Soc. 1997, 119, 12673–12674. [Google Scholar] [CrossRef]
  43. Isse, A.A.; Falciola, L.; Mussini, P.R.; Gennaro, A. Relevance of electron transfer mechanism in electrocatalysis: The reduction of organic halides at silver electrodes. Chem. Commun. 2006, 344–346. [Google Scholar] [CrossRef] [PubMed]
  44. Scialdone, O.; Galia, A.; Filardo, G.; Isse, A.A.; Gennaro, A. Electrocatalytic carboxylation of chloroacetonitrile at a silver cathode for the synthesis of cyanoacetic acid. Electrochim. Acta 2008, 54, 634–642. [Google Scholar] [CrossRef]
  45. Fiori, G.; Rondinini, S.; Sello, G.; Vertova, A.; Cirja, M.; Conti, L. Electroreduction of volatile organic halides on activated silver cathodes. J. Appl. Electrochem. 2005, 35, 363–368. [Google Scholar] [CrossRef]
  46. Isse, A.A.; Sandonà, G.; Durante, C.; Gennaro, A. Voltammetric investigation of the dissociative electron transfer to polychloromethanes at catalytic and non-catalytic electrodes. Electrochim. Acta 2009, 54, 3235–3243. [Google Scholar] [CrossRef]
  47. Durante, C.; Isse, A.A.; Sandonà, G.; Gennaro, A. Electrochemical hydrodehalogenation of polychloromethanes at silver and carbon electrodes. Appl. Catal. B Environ. 2009, 88, 479–489. [Google Scholar] [CrossRef]
  48. Huang, B.; Isse, A.A.; Durante, C.; Wei, C.; Gennaro, A. Electrocatalytic properties of transition metals toward reductive dechlorination of polychloroethanes. Electrochim. Acta 2012, 70, 50–61. [Google Scholar] [CrossRef]
  49. Peverly, A.A.; Karty, J.A.; Peters, D.G. Electrochemical reduction of (1R,2R,3S,4R,5R,6S)-hexachlorocyclohexane (lindane) at silver cathodes in organic and aqueous–organic media. J. Electroanal. Chem. 2013, 692, 66–71. [Google Scholar] [CrossRef]
  50. Mubarak, M.S.; Peters, D.G. Electrochemical reduction of di-, tri-, and tetrahalobenzenes at carbon cathodes in dimethylformamide evidence for a halogen dance during the electrolysis of 1,2,4,5-tetrabromobenzene. J. Electroanal. Chem. 1997, 435, 47–53. [Google Scholar] [CrossRef]
  51. Gennaro, A.; Isse, A.A.; Vianello, E. Solubility and electrochemical determination of CO2 in some dipolar aprotic solvents. J. Electroanal. Chem. 1990, 289, 203–215. [Google Scholar] [CrossRef]
  52. Rondinini, S.; Mussini, P.R.; Muttini, P.; Sello, G. Silver as a powerful electrocatalyst for organic halide reduction: The critical role of molecular structure. Electrochim. Acta 2001, 46, 3245–3258. [Google Scholar] [CrossRef]
  53. Bellomunno, C.; Bonanomi, D.; Falciola, L.; Longhi, M.; Mussini, P.R.; Doubova, L.M.; Di Silvestro, G. Building up an electrocatalytic activity scale of cathode materials for organic halide reductions. Electrochim. Acta 2005, 50, 2331–2341. [Google Scholar] [CrossRef]
  54. Fedurco, M.; Coppex, L.; Augustynski, J. Ab initio and electrochemical studies on the reductive bond dissociation in haloethanols. J. Phys. Chem. B 2002, 106, 2625–2633. [Google Scholar] [CrossRef]
  55. Rondinini, S.B.; Mussini, P.R.; Crippa, F.; Sello, G. Electrocatalytic potentialities of silver as a cathode for organic halide reductions. Electrochem. Commun. 2000, 2, 491–496. [Google Scholar] [CrossRef]
  56. Zhang, K.; Wang, H.; Zhao, S.F.; Niu, D.F.; Lu, J.X. Asymmetric electrochemical carboxylation of prochiral acetophenone: An efficient route to optically active atrolactic acid via selective fixation of carbon dioxide. J. Electroanal. Chem. 2009, 630, 35–41. [Google Scholar] [CrossRef]
  57. Aishah, A.J.; Hartini, M.A.; Normala, S.; Norhuda, A.M.; Hanis, H.H.N.; Razif, H.M.; Sugeng, T. Carbon dioxide fixation method for electrosynthesis of benzoic acid from chlorobenzene. J. Nat. Gas Chem. 2007, 16, 273–277. [Google Scholar] [CrossRef]
  58. Isse, A.A.; Berzi, G.; Falciola, L.; Rossi, M.; Mussini, P.R.; Gennaro, A. Electrocatalysis and electron transfer mechanisms in the reduction of organic halides at ag. J. Appl. Electrochem. 2009, 39, 2217–2225. [Google Scholar] [CrossRef]
  59. Zhao, S.F.; Wu, L.X.; Wang, H.; Lu, J.X.; Bond, A.M.; Zhang, J. A unique proton coupled electron transfer pathway for electrochemical reduction of acetophenone in the ionic liquid [BMIM][BF4] under a carbon dioxide atmosphere. Green Chem. 2011, 13, 3461–3468. [Google Scholar] [CrossRef]
  60. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision C.01, Gaussian, Inc.: Wallingford, CT, USA, 2010.
Scheme 1. Electrocarboxylation of dichlorobenzenes.
Scheme 1. Electrocarboxylation of dichlorobenzenes.
Catalysts 07 00274 sch001
Figure 1. Cyclic voltammograms obtained with the scan rate of 0.1 V/s in DMF + 0.1 M TEABF4-: (a) blank cyclic voltammetry on GC electrode; (b) as (a) +5 mM 1,4-dichlorobenzene; (c) as (a) +5 mM chlorobenzene; (d) as (b) in the presence of CO2; (e) cyclic voltammetry of 5 mM 1,4-dichlorobenzene on Ag electrode.
Figure 1. Cyclic voltammograms obtained with the scan rate of 0.1 V/s in DMF + 0.1 M TEABF4-: (a) blank cyclic voltammetry on GC electrode; (b) as (a) +5 mM 1,4-dichlorobenzene; (c) as (a) +5 mM chlorobenzene; (d) as (b) in the presence of CO2; (e) cyclic voltammetry of 5 mM 1,4-dichlorobenzene on Ag electrode.
Catalysts 07 00274 g001
Figure 2. Cyclic voltammograms of 5 mM dichlorobenzene obtained on Ag electrode with the scan rate of 0.1 V/s in DMF + 0.1 M TEABF4-: (a) 1,4-dichlorobenzene; (b) 1,3-dichlorobenzene; (c) 1,2-dichlorobenzene.
Figure 2. Cyclic voltammograms of 5 mM dichlorobenzene obtained on Ag electrode with the scan rate of 0.1 V/s in DMF + 0.1 M TEABF4-: (a) 1,4-dichlorobenzene; (b) 1,3-dichlorobenzene; (c) 1,2-dichlorobenzene.
Catalysts 07 00274 g002
Scheme 2. Proposed pathways for electrocarboxylation of dichlorobenzenes (Insertion: Optimized structures of expected carboxylated intermediate for dichlorobenzenes, grey—C atom, red—O atom, white—H atom).
Scheme 2. Proposed pathways for electrocarboxylation of dichlorobenzenes (Insertion: Optimized structures of expected carboxylated intermediate for dichlorobenzenes, grey—C atom, red—O atom, white—H atom).
Catalysts 07 00274 sch002
Table 1. Electrocarboxylation of 1,4-dichlorobenzene under various synthetic parameters a.
Table 1. Electrocarboxylation of 1,4-dichlorobenzene under various synthetic parameters a.
EntrySEJ (mA/cm2)Q (F/mol)T (°C)Carboxylation Product Yields (%) bTotal Yield (%) b
234
1TEACl10403.1trace1.04.1
2TEABr10402.1trace1.03.1
3TEAI10401.2trace1.02.2
4TEABF410402.3tracetrace2.3
5TBACl104031.23.22.336.7
6TBABr104024.33.42.430.1
7TBAI104025.3trace1.026.3
8TBACl84016.14.22.222.5
9TBACl134035.74.32.942.9
10TBACl144037.63.32.743.6
11TBACl154034.44.93.342.6
12TBACl164026.82.62.231.6
13TBACl142014.92.61.719.2
14TBACl143023.02.61.627.2
15TBACl144.2037.24.73.445.3
16TBACl144.5026.83.92.333.0
17TBACl145024.83.52.230.5
18TBACl144.2−1041.61.01.043.6
19TBACl144.2−540.71.91.844.4
20TBACl144.2538.44.63.346.3
21TBACl144.21038.64.13.145.8
22TBACl144.22524.26.73.134.0
a General conditions: DMF = 10 mL, SE = 0.1 M, 1,4-dichlorobenzene = 0.1 M, Ag as cathode, Mg as anode, PCO2 = 1 atm; b The yield based on starting substrate is determined by GC.
Table 2. Electrocarboxylation of dichlorobenzenes a.
Table 2. Electrocarboxylation of dichlorobenzenes a.
EntrySubstrateCarboxylation Product Yields (%) bTotal Yield (%) b
234
11,4-dichlorobenzene (1a)38.44.63.346.3
21,3-dichlorobenzene (1b)41.47.8049.2
31,2-dichlorobenzene (1c)43.75.8049.5
a General conditions: DMF = 10 mL, TBACl = 0.1 M, haloacetophenone = 0.1 M, Q = 4.2 F/mol, J = 14 mA/cm2, T = 5 °C, PCO2 = 1 atm, Ag as cathode, Mg as anode; b The yield based on starting substrate is determined by GC.

Share and Cite

MDPI and ACS Style

Luo, P.-P.; Zhang, Y.-T.; Chen, B.-L.; Yu, S.-X.; Zhou, H.-W.; Qu, K.-G.; Kong, Y.-X.; Huang, X.-Q.; Zhang, X.-X.; Lu, J.-X. Electrocarboxylation of Dichlorobenzenes on a Silver Electrode in DMF. Catalysts 2017, 7, 274. https://doi.org/10.3390/catal7090274

AMA Style

Luo P-P, Zhang Y-T, Chen B-L, Yu S-X, Zhou H-W, Qu K-G, Kong Y-X, Huang X-Q, Zhang X-X, Lu J-X. Electrocarboxylation of Dichlorobenzenes on a Silver Electrode in DMF. Catalysts. 2017; 7(9):274. https://doi.org/10.3390/catal7090274

Chicago/Turabian Style

Luo, Pei-Pei, Ying-Tian Zhang, Bao-Li Chen, Shu-Xian Yu, Hua-Wei Zhou, Kong-Gang Qu, Yu-Xia Kong, Xian-Qiang Huang, Xian-Xi Zhang, and Jia-Xing Lu. 2017. "Electrocarboxylation of Dichlorobenzenes on a Silver Electrode in DMF" Catalysts 7, no. 9: 274. https://doi.org/10.3390/catal7090274

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

Article Metrics

Back to TopTop