Skip to main content
ACS Omega logoLink to ACS Omega
. 2022 Nov 15;7(47):42828–42834. doi: 10.1021/acsomega.2c04748

Highly Selective Electrochemical Synthesis of Urea Derivatives Initiated from Oxygen Reduction in Ionic Liquids

Erin Witherspoon , Pinghua Ling , William Winchester , Qi Zhao §, Ahmad Ibrahim , Kevin E Riley , Zhe Wang †,*
PMCID: PMC9713781  PMID: 36467946

Abstract

graphic file with name ao2c04748_0008.jpg

The development of more efficient and sustainable methods for synthesizing substituted urea compounds and directly utilizing CO2 has long been a major focus of synthetic organic chemistry as these compounds serve critical environmental and industrial roles. Herein, we report a green approach to forming the urea compounds directly from CO2 gas and primary amines, triggered by oxygen electroreduction in ionic liquids (ILs). These reactions were carried out under mild conditions, at very low potentials, and achieved high conversion rates. The fact that O2 gas was utilized as the sole catalyst in this electrochemical loop, without additional reagents, is a significant milestone for eco-friendly syntheses of C–N compounds and establishes an effective and green CO2 scavenging method.

Introduction

Currently, about 3.54 billion people globally are fed by synthetic nitrogen fertilizers, over 70% of which are composed of urea and its derivatives. Demand is expected to increase even further as the population continues to grow.1,2 Moreover, as an important class of carbonyl compounds and valuable chemical intermediates, urea derivatives are widely used in pharmaceuticals, textiles, medicine, and other agricultural chemicals.35 They are also used as dyes, energy carriers, antioxidants, corrosion inhibitors, and additives in plastics.69 Presently, the conventional production of urea is accomplished through the reaction of NH3/RNH2 and CO2.10,11 This synthesis is generally described in Scheme 1.

Scheme 1. The Conventional Production of Urea through the Reaction of NH3/RNH2 and CO2.

Scheme 1

The standard Gibbs free energy changes for reactions (1) and (2) are −23.8 and +16.7 kJ, respectively, at 25 °C and 1 bar. As a kinetically slow process, reaction (2) is thermodynamically infeasible at the ambiance; thus, significant external energy and catalysts are needed.12 This reaction is initiated from the carbamate salt form, at which point the amine group is required to nucleophilically attack the carbonyl group. Strong bases such as CsOH, Cs2CO3, KOH, and [BMIM]OH are utilized in the chemical process to generate negatively charged R-NH or R-NH* radical. However, their reactions with CO2 limit the production efficiency when using CO2 gas directly. Thus, current methods are energy-intensive (150–200 °C, 150–250 bar), with production costs being ∼2% of annual global energy.13 Dangerous reagents such as phosgene and isocyanates and multicycle processes are also required to improve the conversion efficiency.14

Using renewable electrical energy to synthesize urea derivatives is considered as a promising green alternative for next-generation industrial production.15,16 The C–N structure has previously been generated by activating the linear form of CO2 to CO2 by applying overpotentials of about −1.9 V, but temperatures of 100–200 °C were still required.1724 Further, Xia’s group used −15 kV of the negative corona to ionize NH3 and CO2 gases to form urea at 1 bar and 20 °C, achieving 82% CO2 conversion and 51% urea selectivity.25 In aqueous solutions, Wang and Li’s group investigated the coupling CO2 and N2 electrochemical reduction on metal-based catalyst surfaces. A faradaic efficiency (FE) of 8.92% was reported on PdCu and TiO2 catalysts.14 Both mechanisms involved COOH* as a key electrochemical intermediate from CO2. To avoid the high dissociation energy of N2 triple bonds (941 kJ/mol) and the limited solubility of N2 in H2O (0.02 v/v, 25 °C, 1 bar), nitrite solutions were used as feedstock with efficiencies of about 25–55% on different gas-diffusion metal electrodes.26,27 However, conventional precious metal-based alloy catalysts suffer from ambiguous active sites and easy corrosion.15,28,29 In the electrochemical approach, the co-reduction reactions achieved on metal-based catalysts could increase the conversion through a possible coupling pathway of NH2* and COOH* intermediates for urea formation. Accordingly, the solubility of ammonia and CO2 play a key role in conversion efficiency.30 Other limiting factors include the high overpotential of the CO2 reduction and the hydrogen evolution reaction involved in aqueous solutions.

We believe that this synthesis could be improved further by using green and readily available materials under milder conditions. Ionic liquids (ILs) are novel green solvents used in many chemical and electrochemical reactions.31,32 Here, hydrophobic ILs are used as electrolytes/solvents for urea formation directly from CO2 and amines since they are inert toward all compounds involved in the reaction and CO2 is highly soluble in the solvents. Our group has completed extensive work on electrochemical reactions in ILs.3335 As reported, strong nucleophiles, superoxide and peroxide ions, generated through the electrochemical reduction of oxygen, could be even more stable in ILs than in traditional aprotic solvents (i.e., DMSO and acetonitrile) and participate in several organic reactions in completely green processes.36,37

In this work, we hypothesize that active oxygen ions (or reactive oxygen species (ROS)) could activate various amines and further enable the nucleophilic attack of carbamate salts. It is well known that at room temperature and even at low pressure, CO2 quickly reacts with amines in the first step of the chemical synthesis process, reaction (1). The linear structure of CO2 becomes bent to carbamate acid or carbamate anion, and the exposed nucleophilic site of carbon could effectively react with R-NH or R-NH* nucleophiles. Thus, the recovered ammonium salts could be directly converted into urea with less additional energy input, which is a more practical industrial approach and an efficient ammonia-recovery method. Herein, the reactions were evaluated under ambient temperature and pressure in both protic and aprotic ILs, without catalysts, to provide potential adaptability and universality of this method. Cyclic voltammetry (CV), theoretical density functional theory (DFT) calculations, and in situ Fourier transform infrared (FTIR) spectroscopy were conducted to understand the reaction mechanisms and possible pathways.

Materials and Methods

Materials

Carbon dioxide (99.99%) and oxygen (99.99%) were purchased from Airgas, Inc. All amines, deuterated solvents, and sodium peroxide (95%) were obtained from Fisher Scientific. ILs were from IoLiTec Ionic Liquids Technologies GmbH. All chemicals were used as received without further purification. The nuclear magnetic resonance (NMR) spectra were recorded using a Bruker Avance 300 spectrometer and were calibrated using residual undeuterated solvent as an internal reference (CDCl3, 77.16 ppm 13C NMR; C2D6OS, 35.92 ppm 13C, 2.50 ppm 1H NMR). The FTIR absorption spectra were recorded on a Varian 3100 Excalibur Series spectrometer in the region of 400–2000 cm–1 using two NaCl disks also from Fisher Scientific for non-in situ measurements. The electrochemical tests were performed with a CHI660E electrochemical analyzer with platinum wire as the reference and counter electrodes and a gold plate as the working electrode.

Preparation of C–N Organic Compounds

700 μL of 1-butyl-1-methylimidazolium bis(trifluoromethylsulfonyl)-imide ([BMIM] [NTf2]) or 1-butyl-1-methylpyrrolidinium bis(trifluoromethylsulfonyl)-imide ([BMPyrr] [NTf2]) (Figure S1) was loaded in a Teflon cell then CO2 (50 sccm) and O2 (20 sccm) were bubbled in for 30 min. The mixture was reacted at −1.9 V for another 30 min. Finally, 300 μL of the selected amine was added into the cell, and the reaction continued for another 5 h at −1.9 V. For reactions using [BMPyrr] [NTf2], −1.7 V was applied. For generation using Na2O2, 700 μL of [BMPyrr] [NTf2] was added to 300 μL of CHA, and then CO2 was purged through the mixture for ∼1 min. Next, ∼0.250 g of Na2O2 was added, and the mixture was shaken for ∼1 min, and subsequently, CO2 was purged for ∼1 min more. The final reaction mixtures were vacuum filtered and washed with acetone.

Cyclic Voltammetry Measurements

CV measurements were completed using the same electrochemical cell and setup; however, only 50 μL of CHA was added in this case to avoid buildup of the solid product, and rather than constant potential, the potential was swept between 0.5 and −2.0 V. In addition, step 1, the salt formation step, took place outside the cell, and a specific mass (0.5 g) was added subsequently. CVs were scanned at 200 mV/s for 8 cycles.

In Situ Fourier Transform Infrared Spectroscopy

In situ EC-FTIR measurements were recorded using a Bio-Logic Thin Layer Spectroelectrochemical Cell with a gold gauze working electrode, a platinum gauze counter electrode, and a platinum wire reference electrode, with [BMIM] [NTf2] as the electrolyte. A constant potential of −1.9 V was applied continuously to the cell for over 7 h; O2 gas was purged for the same duration. The spectra were recorded every 10 min with 20 scans and a sensitivity of 1.

13C Isotope Nuclear Magnetic Resonance Spectroscopy Characterizations

A Bruker 200 MHz NMR spectrometer was used to record 13C isotope products using 13C NMR. Both the 13C salt product and 13C urea product were prepared using the same technique as 12C products but substituting 12CO2 with 13CO2, supplied from Sigma-Aldrich (99.0%). In this case, a magnetic stir bar was added to the reaction vessel since the 13CO2 gas was a low-pressure gas incapable of bubbling into the mixture. The reaction was stirred at room temperature for 15 min before collecting the NMR spectrum.

Computations

DFT computations were carried out at the PW6B95/def2-TZVP level of theory and utilized the SMD implicit solvation model (ε = 10) to mimic the IL environment. All calculations were carried out using the Gaussian G16 suite of molecular electronic structure programs. All structural optimizations were carried out using “tight” convergence criteria. Thermal corrections used to compute Gibbs energies for products, reactants, and intermediates were computed using harmonic vibrational analysis.

Results and Discussion

To investigate the reaction between activated O2 and the carbamate salt formed in step 1 (Scheme 1), a three-electrode system was used to study the electrochemical process between CO2/cyclohexylamine (CHA) salt and ROS in [BMIM] [NTf2]. Figure 1 (black curve) shows a typical CV curve of the two-step redox process of O2 in the IL, which was in agreement with previous reports. Superoxide, generated by gaining one electron at −1.2 V,35,3842 and peroxide anion, generated by gaining an additional electron at −1.7 V,4244 are quasi-reversible redox couples, as shown in Scheme 2. After adding 0.5 g of CHA/CO2 salt into the cell (red curve), the peak current (ip) decreased. The addition of salt may increase the viscosity of the solution, and thus, mass transport rates also decrease. Based on the Cottrell equation (1),45 the redox current dropped consequently, as previously reported.46,47 From the red curve of Figure 1, it is clear that the O22– oxidation peak, around −0.3 V, disappeared, but the superoxide peak remained after salt was added. This suggested that O22– was consumed in the reaction and may activate salt, while O2•– does not.

Figure 1.

Figure 1

CV curves of pure O2 gas in [BMIM] [NTf2] (black curve) and with 0.5 g of the salt product from CHA and CO2 added (red curve).

Scheme 2. Oxygen Electroreduction Reactions.

Scheme 2

Based on the CV results, constant potential experiments were also performed in [BMIM] [NTf2] and additionally in aprotic 1-butyl-1-methylpyrrolidinium bis(trifluoromethylsulfonyl)-imide ([BMPyrr] [NTf2]) (Figure S1). The applied potential was optimized, based on Figure 1, between −0.8 and 1.9 V. In [BMPyrr] [NTf2], no urea was formed at potentials of −1.2 V or lower, while the highest production rate occurred around −1.6 V (Figure 2). This implied the two-electron oxygen reduction product, peroxide, rather than superoxide, may trigger reaction (2) (Scheme 1), converting the salt product into urea. In [BMIM] [NTf2], a low production rate was found at −1.2 V. This may have been due to a disproportionation reaction between superoxide and the active proton in the solvent, producing O2 and H2O2 as reported in other studies.4851 The production rates between −1.4 and −2.0 V increased accordingly with the formation of peroxide in [BMIM] [NTf2].

Figure 2.

Figure 2

The production rate versus the applied potential in [BMIM] [NTf2] and [BMPyrr] [NTf2] at a conventional Au electrode for 5 h.

Furthermore, we have characterized the compositions of salts and subsequent electrochemical products by NMR spectroscopy and FTIR spectroscopy (see Figures S2–S34). The 13C NMR spectrum of the salt formed from CHA and CO2 in [BMIM] [NTf2] (CDCl3, 25 °C, 200 MHz) is shown in Figure S4. Based on the chemical environment, the signal peak at 162 ppm was assigned to the characteristic peak of the carbamate carbonyl carbon (C=O). The C–N carbon of the ammonium and the carbamate ions in C6H11-NH3+:C6H11-NH-CO2 were responsible for the peaks at 34.53 and 33.96 ppm, respectively. After introducing oxygen gas and electrochemical potential at −1.9 V for 5 h, only one peak at 34.39 ppm was observed (Figure S2). This indicated that a product with a symmetrical structure, (R-NH)2CO, was produced. Negligible double peaks from the salt were present in the NMR spectra of the final products, which implied that a high yield of salt–urea conversion was achieved. The FTIR spectrum (Figure S32) of the final products using the same starting reagents displayed the characteristic peak of urea at 1628 cm–1 and was attributable to the C=O stretching frequency in the N–CO–N structure (blue curve). An identical peak was also observed in an authentic dicyclohexylurea (DCU)-saturated [BMIM] [NTf2] solution (green curve). This peak was negligible in the carbamate salt structure (red curve). The results confirmed the formation of DCU synthesized through an electrochemical process starting with CO2 and O2 gases. Similar results were obtained from cyclopentylamine (CPA). It should also be emphasized that no detectable byproducts were observed in the NMR and FTIR measurements in either IL.

Additional control experiments were performed using sodium peroxide (Na2O2) with CHA and CO2 in [BMPyrr] [NTf2]. The product and salt structures were characterized using 13C NMR and FTIR (Figures S33–S36). The results were identical to those obtained using the same amine and IL with O2 and CO2 gases and applied potential. The production rate for this reaction method was determined to be 13.7%. The low production was most likely due to the extremely low solubility of Na2O2 in the IL,52,53 which could hinder the catalyst effect. Further, the low solubility made extracting pure urea product from the mixture nearly impossible. For these reasons, activating the reaction electrochemically may be considered the superior method since a high conversion yield was achieved (the peroxide anion was completely soluble in this case), and product removal was easy.

The same procedure was repeated by replacing CO2 gas with 13CO2 gas to further confirm the reaction pathway and the formation of the urea structure. The NMR spectra were obtained in the [BMIM] [NTf2] system (see Figures S37 and S38a,b). The 13C urea product was observed at 159.23 ppm, while the 12C product peak was present at 159.30 ppm. According to Lambert and Greifenstein, an isotopic shift of 0.05–0.1 ppm for 12C → 13C is typically seen.54,55 Here, hydrogen-bonding and increasing bond length and angle (118° for the N–C–N angle in 1,3-dicyclohexylurea) may all contribute to an increased downfield shift, while the direction of the shift may also depend on the compound itself and the solvent used.56,57 These NMR results confirmed that the carbonyl carbon was derived from CO2.

Additionally, a control reaction was performed using the standard procedure, but excess CO2 was eliminated prior to bubbling O2 into the mixture. No urea production was observed. This indicated that excess CO2 might have another role in this reaction. It may react with the peroxide anion to form peroxide carbonate derivatives that would be protonated by the ammonium ions from the carbamate salts, which would generate hydroperoxide carbonate. However, no related product peaks were found in the NMR or FTIR spectra. Considering that our results agree well with previous work, we could conclude that the carbonyl carbon is in fact provided by CO2, rather than an amine.58

To investigate the mechanism even further, we have performed DFT calculations of reactions relevant to this process. Figure 3 gives the proposed mechanism, along with calculated electronic and Gibbs energies for steps along the path in the formation of DCU (and water) from CHA and CO2. Here, the reduction of oxygen gas to peroxide anion on the cathode, along with re-oxidation steps, resulting in the restoration of oxygen gas on the anode, is also included. It is seen that steps involving the reduction of dissolved O2 and those associated with re-oxidation of HO2 to HO2 and O2 to O2 represent the largest shifts in energy (either free or electronic) with the reduction being favorable (ΔG = −52.70 kcal/mol) and oxidation having an energy penalty (ΔG = +135.61 kcal/mol for steps 6 and 8 combined). For steps 2–5, representing reactions directly involved with the conversion of the C6H11-NH3+:C6H11-NH-CO2 salt to (C6H11-NH)2CO, electronic energy shifts are favorable for each step (the overall energy change is ΔE = −66.50 kcal/mol). The overall change in the Gibbs energy for steps 2–5 is favorable (ΔG = −70.63 kcal/mol). However, one of these steps, step 4, is associated with a relatively small energy penalty of 7.64 kcal/mol. It is likely that steps 4 and 5 in this mechanism are concerted steps, with the hydroxyl group leaving the cyclohexylcarbamic acid, while the C–N bond, resulting in urea formation, is established. The Gibbs energy change for the combination of steps 4 and 5 is −12.10 kcal/mol, which may vary with the steric effect of different amines.

Figure 3.

Figure 3

Gibbs energy diagram for steps involved in the reaction R-NH3+ + R-NH-CO2- → R-NH-CO-NH-R + H2O (top left). Electronic and Gibbs energies for each step in the mechanism (top right). Molecular structure diagrams for reaction steps.

To study how generally this method can be applied, the same procedure was repeated with various amines in [BMIM] [NTf2] (Figures S5–S24) and [BMPyrr] [NTf2] (Figures S25–S31). The NMR results, as shown in Figures S2–S19, indicate that CHA, CPA, and linear amines all form urea structures in [BMIM] [NTf2]. Branched amines yield little or no product; aniline also failed to generate the product (Figures S20–S24). The results were consistent with previous work.59 However, the reasons for the diminished reactivities are not known but are thought to be related to modified electronic properties of NH2 groups relative to linear amines, CHA, and CPA. Reduced reactivity might be attributable either to reduced participation in the urea formation reactions themselves or to variations in the solubilities of the R-NH3+:R-NH-COO salts. Interestingly, both dipropylamine and dihexylamine reacted to form significant yields of (R2-N)2CO ureas when [BMPyrr] [NTf2] was the solvent. The percent conversion and turnover frequencies (TOF) were determined without CO2 recycling and are given in Table 1. The results were largely consistent with the NMR (and FTIR) results discussed above. Reaction yields in the two ILs are seen to vary substantially for different amine reactants. It showed more than 90% conversion rate for CHA, CPA, and pentylamine in [BMPyrr] [NTf2] and decreased with increasing carbon chain length, which may be due to the steric effect mentioned previously. For amines with more than eight carbons, conversion rates dropped to lower than 22.5%. However, amines demonstrating high urea yields through this reaction covered the most common ureas available. The reasons for this are not completely understood. One factor could be the presence of a protic or aprotic cation in the IL. The [BMPyrr] cation has a tendency to promote the stability of the peroxide dianion to a greater extent than the protic [BMIM] cation.60,61 It is also possible that the IL cation directly affects the solubility of the R-NH3+:R-NH-COO salt, indirectly affecting reaction yields.

Table 1. Percent Conversion and Turnover Frequencies of Various Amines.

amines % conv. in [BMIM] [NTf2] TOF (s–1) in [BMIM] [NTf2] % conv. in [BMPyrr] [NTf2] TOF (s–1) in [BMPyrr] [NTf2]
cyclohexylamine 88 78.411 93 108.44
CHA/sodium peroxide N/A N/A 13.7 N/A
cyclopentylamine 86 36.21 90 35.98
pentylamine 87 208.91 90 6.26
hexylamine 65 88.00 76 16.28
heptylamine 46 12.86 49 147.48
octylamine 14 10.03 22.5 73.46
dihexylamine 7 86.73 30 218.25
dipropylamine 5 N/A 43 N/A
piperidine 0 N/A 0 N/A

A spectroelectrochemical cell from Bio-Logic was used to measure the formation of urea in situ as reported in other systems.58,62 The spectra were recorded every 10 min starting from pure [BMIM] [NTf2] (Figure 4). After 20 min, 10 μL of CHA was added with CO2 purging for 5 min. The signals of the spectrum became lower since the solid salt instantly formed while being suspended in the IL. Then, O2 gas was purged into the system with a constant potential of −1.9 V applied for the entirety of the EC experiment. A new peak was observed at 3386 cm–1, which was attributed to the N–H stretching in 1,3-dicyclohexylurea.63,64 With time, the peak stopped increasing, which implied quick saturation due to the low solubility of urea. Moreover, the stability of [BMIM] [NTf2] as the sole electrolyte and solvent was investigated. The EC reaction was monitored over 7 h with a continuously applied potential of −1.9 V. From our results, the IL peaks remained unchanged throughout the experiments, demonstrating its excellent stability. The structural stability was also confirmed by 1H NMR measurements of the [BMIM] [NTf2] before and after the reaction (see Figure S39a,b).

Figure 4.

Figure 4

In situ EC-FTIR measurements. The spectra were recorded every 10 min. The first labeled spectrum resulted from pure [BMIM] [NTf2]. CHA and CO2 were added at 20 min (third curve). The fourth spectrum was recorded after 10 min of applying O2 gas and −1.9 V.

In conclusion, we have developed a highly selective, efficient, safe, and inexpensive method for synthesizing C–N compounds, which can be performed at ambiance within a short time. Unlike previously reported techniques, no external catalysts nor reagents other than oxygen gas are needed, and special modification or preparation of the electrode surface is not required. The most accessible amines on the market can be used to generate short carbon chain urea with more than 92% production. The concept of such a catalyst system could be expanded to other fields and exploited industrially.

Acknowledgments

This publication was made possible by the support from the Army Research Office under grant number W911NF-18-1-0458 and the National Science Foundation (CHE-1832167, HRD-1700429, and HRD-1505219).

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.2c04748.

  • Experimental details, computations, 13C NMR spectra of all compounds, and FTIR spectra of selected compounds (PDF)

The authors declare no competing financial interest.

Supplementary Material

ao2c04748_si_001.pdf (2.9MB, pdf)

References

  1. Houlton B. Z.; Almaraz M.; Aneja V.; Austin A. T.; Bai E.; Cassman K. G.; Compton J. E.; Davidson E. A.; Erisman J. W.; Galloway J. N.; Gu B.; Yao G.; Martinelli L. A.; Scow K.; Schlesinger W. H.; Tomich T. P.; Wang C.; Zhang X. A World of Cobenefits: Solving the Global Nitrogen Challenge. Earth’s Future 2019, 7, 865–872. 10.1029/2019EF001222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Nations, F. a. A. O. o. t. U ., World fertilizer trends and outlook to 2022.; Food and Agriculture Organization of the United Nations: 2019. [Google Scholar]
  3. Ghosh A. K.; Brindisi M. Urea Derivatives in Modern Drug Discovery and Medicinal Chemistry. J. Med. Chem. 2020, 63, 2751–2788. 10.1021/acs.jmedchem.9b01541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Shirvan A. R.; Shakeri M.; Bashari A., Recent advances in application of chitosan and its derivatives in functional finishing of textiles. in The impact and prospects of green chemistry for textile technology; Woodhead Publishing, 2019, 107–133. [Google Scholar]
  5. Atashkar B.; Zolfigol M. A.; Mallakpour S. Applications of biological urea-based catalysts in chemical processes. Mol. Cata. 2018, 452, 192–246. 10.1016/j.mcat.2018.03.009. [DOI] [Google Scholar]
  6. Li H.-Q.; Lv P.-C.; Yan T.; Zhu H.-L. Urea derivatives as anticancer agents. Anti-Cancer Agents Med. Chem. 2009, 9, 471–480. 10.2174/1871520610909040471. [DOI] [PubMed] [Google Scholar]
  7. Yokoya M.; Kimura S.; Yamanaka M. Urea derivatives as functional molecules: supramolecular capsules, supramolecular polymers, supramolecular gels, artificial hosts, and catalysts. Chem. - Eur. J. 2021, 27, 5601–5614. 10.1002/chem.202004367. [DOI] [PubMed] [Google Scholar]
  8. Amendola V.; Fabbrizzi L.; Mosca L. Anion recognition by hydrogen bonding: urea-based receptors. Chem. Soc. Rev. 2010, 39, 3889–3915. 10.1039/b822552b. [DOI] [PubMed] [Google Scholar]
  9. Song E. Y.; Kaur N.; Park M.-Y.; Jin Y.; Lee K.; Kim G.; Lee K. Y.; Yang J. S.; Shin J. H.; Nam K.-Y.; No K. T.; Han G. Synthesis of amide and urea derivatives of benzothiazole as Raf-1 inhibitor. Eur. J. Med. Chem. 2008, 43, 1519–1524. 10.1016/j.ejmech.2007.10.008. [DOI] [PubMed] [Google Scholar]
  10. Meessen J. Urea synthesis. Chem. Ing. Tech. 2014, 86, 2180–2189. 10.1002/cite.201400064. [DOI] [Google Scholar]
  11. Wu C.; Cheng H.; Liu R.; Wang Q.; Hao Y.; Yu Y.; Zhao F. Synthesis of urea derivatives from amines and CO2 in the absence of catalyst and solvent. Green Chem. 2010, 12, 1811–1816. 10.1039/c0gc00059k. [DOI] [Google Scholar]
  12. Xu M.; Jupp A. R.; Ong M. S.; Burton K. I.; Chitnis S. S.; Stephan D. W. Synthesis of urea derivatives from CO2 and silylamines. Am. Ethnol. 2019, 131, 5763–5767. 10.1002/ange.201900058. [DOI] [PubMed] [Google Scholar]
  13. Lv C.; Zhong L.; Liu H.; Fang Z.; Yan C.; Chen M.; Kong Y.; Lee C.; Liu D.; Li S.; Liu J.; Song L.; Chen G.; Yan Q.; Yu G. Selective electrocatalytic synthesis of urea with nitrate and carbon dioxide. Nat. Sustainable 2021, 4, 868–876. 10.1038/s41893-021-00741-3. [DOI] [Google Scholar]
  14. Wang H.; Xin Z.; Li Y., Synthesis of Ureas from CO2. Chemical Transformations of Carbon Dioxide; Springer; 2017, 177–202, 10.1007/978-3-319-77757-3_5. [DOI] [Google Scholar]
  15. Zhu X.; Zhou X.; Jing Y.; Li Y. Electrochemical synthesis of urea on MBenes. Nat. Commun. 2021, 12, 4080. 10.1038/s41467-021-24400-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Yuan M.; Chen J.; Bai Y.; Liu Z.; Zhang J.; Zhao T.; Shi Q.; Li S.; Wang X.; Zhang G. Electrochemical C–N coupling with perovskite hybrids toward efficient urea synthesis. Chem. Sci. 2021, 12, 6048–6058. 10.1039/D1SC01467F. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Centi G.; Quadrelli E. A.; Perathoner S. Catalysis for CO 2 conversion: a key technology for rapid introduction of renewable energy in the value chain of chemical industries. Energy Environ. Sci. 2013, 6, 1711–1731. 10.1039/c3ee00056g. [DOI] [Google Scholar]
  18. Sakakura T.; Choi J.-C.; Yasuda H. Transformation of carbon dioxide. Chem. Rev. 2007, 107, 2365–2387. 10.1021/cr068357u. [DOI] [PubMed] [Google Scholar]
  19. Rakowski Dubois M.; Dubois D. L. Development of molecular electrocatalysts for CO2 reduction and H2 production/oxidation. Acc. Chem. Res. 2009, 42, 1974–1982. 10.1021/ar900110c. [DOI] [PubMed] [Google Scholar]
  20. Aresta M.; Dibenedetto A.; Angelini A. Catalysis for the valorization of exhaust carbon: from CO2 to chemicals, materials, and fuels. Technological use of CO2. Chem. Rev. 2014, 114, 1709–1742. 10.1021/cr4002758. [DOI] [PubMed] [Google Scholar]
  21. Benson E. E.; Kubiak C. P.; Sathrum A. J.; Smieja J. M. Electrocatalytic and homogeneous approaches to conversion of CO2 to liquid fuels. Chem. Soc. Rev. 2009, 38, 89–99. 10.1039/B804323J. [DOI] [PubMed] [Google Scholar]
  22. Peterson A. A.; Abild-Pedersen F.; Studt F.; Rossmeisl J.; Nørskov J. K. How copper catalyzes the electroreduction of carbon dioxide into hydrocarbon fuels. Energy Environ. Sci. 2010, 3, 1311–1315. 10.1039/c0ee00071j. [DOI] [Google Scholar]
  23. Marcos M. L.; González-Velasco J.; Bolzán A. E.; Arvia A. J. Comparative electrochemical behaviour of CO2 on Pt and Rh electrodes in acid solution. J. Electroanal. Chem. 1995, 395, 91–98. 10.1016/0022-0728(95)04098-9. [DOI] [Google Scholar]
  24. Lim R. J.; Xie M.; Sk M. A.; Lee J.-M.; Fisher A.; Wang X.; Lim K. H. A review on the electrochemical reduction of CO2 in fuel cells, metal electrodes and molecular catalysts. Catal. Today 2014, 233, 169–180. 10.1016/j.cattod.2013.11.037. [DOI] [Google Scholar]
  25. Ma F.-X.; Wang J.; Wang F.-B.; Xia X.-H. The room temperature electrochemical synthesis of N-doped graphene and its electrocatalytic activity for oxygen reduction. Chem. Commun. 2015, 51, 1198–1201. 10.1039/C4CC07402E. [DOI] [PubMed] [Google Scholar]
  26. Shibata M.; Furuya N. Electrochemical synthesis of urea at gas-diffusion electrodes: Part VI. Simultaneous reduction of carbon dioxide and nitrite ions with various metallophthalocyanine catalysts. J. Electroanal. Chem. 2001, 507, 177–184. 10.1016/S0022-0728(01)00363-1. [DOI] [Google Scholar]
  27. Shibata M.; Yoshida K.; Furuya N. Electrochemical Synthesis of Urea at Gas-Diffusion Electrodes: IV. Simultaneous Reduction of Carbon Dioxide and Nitrate Ions with Various Metal Catalysts. J. Electroanal. Chem. 1998, 145, 2348. 10.1149/1.1838641. [DOI] [Google Scholar]
  28. da Silva Freitas W.; D’Epifanio A.; Mecheri B. Electrocatalytic CO2 reduction on nanostructured metal-based materials: Challenges and constraints for a sustainable pathway to decarbonization. J. CO2 Util. 2021, 50, 101579 10.1016/j.jcou.2021.101579. [DOI] [Google Scholar]
  29. Phan Q. T.; Poon K. C.; Sato H. A review on amorphous noble-metal-based electrocatalysts for fuel cells: Synthesis, characterization, performance, and future perspective. Int. J. Hydrogen Energy 2021, 46, 14190–14211. 10.1016/j.ijhydene.2021.02.006. [DOI] [Google Scholar]
  30. Manaka Y.; Nagatsuka Y.; Motokura K. Organic bases catalyze the synthesis of urea from ammonium salts derived from recovered environmental ammonia. Sci. Rep. 2020, 10, 2834. 10.1038/s41598-020-59795-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Ohno H., Electrochemical aspects of ionic liquids; John Wiley & Sons: 2005. [Google Scholar]
  32. Armand M.; Endres F.; MacFarlane D. R.; Ohno H.; Scrosati B., Ionic-liquid materials for the electrochemical challenges of the future. Materials for sustainable energy: a collection of peer-reviewed research and review articles from Nature Publishing Group; World Scientific: 2011, 129–137. [Google Scholar]
  33. Wang Z.; Kumar A.; Sevilla M. D.; Zeng X. Anaerobic Oxidation of Methane to Methyl Radical in NTf2-Based Ionic Liquids. J. Phys. Chem. C 2016, 120, 13466–13473. 10.1021/acs.jpcc.6b00153. [DOI] [Google Scholar]
  34. Tang Y.; Wang Z.; Chi X.; Sevilla M. D.; Zeng X. In Situ Generated Platinum Catalyst for Methanol Oxidation via Electrochemical Oxidation of Bis(trifluoromethylsulfonyl)imide Anion in Ionic Liquids at Anaerobic Condition. J. Phys. Chem. C 2016, 120, 1004–1012. 10.1021/acs.jpcc.5b09777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Wang Z.; Lin P.; Baker G. A.; Stetter J.; Zeng X. Ionic liquids as electrolytes for the development of a robust amperometric oxygen sensor. Anal. Chem. 2011, 83, 7066–7073. 10.1021/ac201235w. [DOI] [PubMed] [Google Scholar]
  36. Villagrán C.; Aldous L.; Lagunas M. C.; Compton R. G.; Hardacre C. Electrochemistry of phenol in bis {(trifluoromethyl) sulfonyl} amide ([NTf2]−) based ionic liquids. J. Electroanal. Chem. 2006, 588, 27–31. 10.1016/j.jelechem.2005.11.023. [DOI] [Google Scholar]
  37. Gondosiswanto R.; Hibbert D. B.; Fang Y.; Zhao C. Ionic liquid microstrips impregnated with magnetic nanostirrers for sensitive gas sensors. ACS Appl. Mater. Interfaces 2017, 9, 43377–43385. 10.1021/acsami.7b14657. [DOI] [PubMed] [Google Scholar]
  38. Wang Z.; Zeng X. Bis (trifluoromethylsulfonyl) imide (NTf2)-based ionic liquids for facile methane electro-oxidation on Pt. J. Electroanal. Chem. 2013, 160, H604. 10.1149/2.039309jes. [DOI] [Google Scholar]
  39. Hayyan M.; Hashim M. A.; AlNashef I. M. Superoxide ion: generation and chemical implications. Chem. Rev. 2016, 116, 3029–3085. 10.1021/acs.chemrev.5b00407. [DOI] [PubMed] [Google Scholar]
  40. Hayyan M.; Mjalli F. S.; AlNashef I. M.; Hashim M. A. Chemical and electrochemical generation of superoxide ion in 1-butyl-1-methylpyrrolidinium bis (trifluoromethylsulfonyl) imide. Int. J. Electrochem. Sci. 2012, 7, 8116–8127. [Google Scholar]
  41. Khan A.; Gunawan C. A.; Zhao C. Oxygen reduction reaction in ionic liquids: fundamentals and applications in energy and sensors. ACS Sustainable Chem. Eng. 2017, 5, 3698–3715. 10.1021/acssuschemeng.7b00388. [DOI] [Google Scholar]
  42. Pozo-Gonzalo C., Oxygen reduction reaction in ionic liquids: an overview. In: Torriero A. (eds) Electrochemistry in Ionic Liquids; Springer: Cham. 2015, 507–529. [Google Scholar]
  43. Zhang Y.; Pozo-Gonzalo C. Variations and applications of the oxygen reduction reaction in ionic liquids. Chem. Commun. 2018, 54, 3800–3810. 10.1039/c8cc00595h. [DOI] [PubMed] [Google Scholar]
  44. De Giorgio F.; Soavi F.; Mastragostino M. Effect of lithium ions on oxygen reduction in ionic liquid-based electrolytes. Electrochem. Commun. 2011, 13, 1090–1093. 10.1016/j.elecom.2011.07.004. [DOI] [Google Scholar]
  45. Bard A. J.; Faulkner L. R., Electrochemical Methods: Fundamentals and Applications, 2nd Edition; John Wiley & Sons, Incorporated: 2000. [Google Scholar]
  46. Tomcsányi L.; De Battisti A.; Hirschberg G.; Varga K.; Liszi J. The study of the electrooxidation of chloride at RuO2/TiO2 electrode using CV and radiotracer techniques and evaluating by electrochemical kinetic simulation methods. Electrochim. Acta 1999, 44, 2463–2472. 10.1016/S0013-4686(98)00381-8. [DOI] [Google Scholar]
  47. Bott A. W. Characterization of chemical reactions coupled to electron transfer reactions using cyclic voltammetry. Curr. Sep. 1999, 18, 9–16. [Google Scholar]
  48. Khan A.; Lu X.; Aldous L.; Zhao C. Oxygen reduction reaction in room temperature protic ionic liquids. J. Phys. Chem. C 2013, 117, 18334–18342. 10.1021/jp405759j. [DOI] [Google Scholar]
  49. Switzer E. E.; Zeller R.; Chen Q.; Sieradzki K.; Buttry D. A.; Friesen C. Oxygen reduction reaction in ionic liquids: The addition of protic species. J. Phys. Chem. C 2013, 117, 8683–8690. 10.1021/jp400845u. [DOI] [Google Scholar]
  50. Pozo-Gonzalo C.; Torriero A. A. J.; Forsyth M.; MacFarlane D. R.; Howlett P. C. Redox chemistry of the superoxide ion in a phosphonium-based ionic liquid in the presence of water. J. Phys. Chem. Lett. 2013, 4, 1834–1837. 10.1021/jz400715r. [DOI] [PubMed] [Google Scholar]
  51. Sawyer D. T., In Oxygen chemistry; Oxford university press: 1991; Vol. 26. [Google Scholar]
  52. DeLand F. H., In CRC Handbook of Chemistry and Physics; CRC Press, Inc., RC, West; Astle MJ; Beyer WH, Eds. Boca Raton, Florida, 1983, 2386 [Google Scholar]
  53. Budavari S.; O’Neil M. J.; Smith A.; Heckelman P. E., In The merck index.; Merck Rahway: NJ: 1989; Vol. 11. [Google Scholar]
  54. Lambert J. B.; Greifenstein L. G. Origin of the chemical-shift isotope effect. Stereochemical evidence. J. Am. Chem. Soc. 1974, 96, 5120–5124. 10.1021/ja00823a018. [DOI] [Google Scholar]
  55. Hansen P. E., Isotope effects on nuclear shielding. In Annual reports on NMR spectroscopy; Elsevier: 1984; Vol. 15, pp. 105–234, 10.1016/S0066-4103(08)60208-2. [DOI] [Google Scholar]
  56. Hansen P. E. Isotope effects on chemical shifts in the study of intramolecular hydrogen bonds. Molecules 2015, 20, 2405–2424. 10.3390/molecules20022405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Vaughan P.; Donohue J. The structure of urea. Interatomic distances and resonance in urea and related compounds. Acta Crystallogr. 1952, 5, 530–535. 10.1107/S0365110X52001477. [DOI] [Google Scholar]
  58. Wang J. IR spectroelectrochemical investigations of solvent roles in deactivation of poly (3-methylthiophene) films. Electrochim. Acta 1997, 42, 2545–2554. 10.1016/S0013-4686(96)00447-1. [DOI] [Google Scholar]
  59. Mindrup E. M.; Schneider W. F. Computational comparison of the reactions of substituted amines with CO2. ChemSusChem 2010, 3, 931–938. 10.1002/cssc.201000060. [DOI] [PubMed] [Google Scholar]
  60. Monaco S.; Arangio A. M.; Soavi F.; Mastragostino M.; Paillard E.; Passerini S. An electrochemical study of oxygen reduction in pyrrolidinium-based ionic liquids for lithium/oxygen batteries. Electrochim. Acta 2012, 83, 94–104. 10.1016/j.electacta.2012.08.001. [DOI] [Google Scholar]
  61. Hayyan M.; Mjalli F. S.; Hashim M. A.; AlNashef I. M.; Al-Zahrani S. M.; Chooi K. L. Long term stability of superoxide ion in piperidinium, pyrrolidinium and phosphonium cations-based ionic liquids and its utilization in the destruction of chlorobenzenes. J. Electroanal. Chem. 2012, 664, 26–32. 10.1016/j.jelechem.2011.10.008. [DOI] [Google Scholar]
  62. Hartl F.; Luyten H.; Nieuwenhuis H. A.; Schoemaker G. C. Versatile cryostated optically transparent thin-layer electrochemical (OTTLE) cell for variable-temperature UV-Vis/IR spectroelectrochemical studies. Appl. Spectrosc. 1994, 48, 1522–1528. 10.1366/0003702944027787. [DOI] [Google Scholar]
  63. SpectraBase; Wiley: 1989, 1990–2021. [Google Scholar]
  64. Rehder S.; Sakmann A.; Rades T.; Leopold C. S. Thermal degradation of amorphous glibenclamide. Eur. J. Pharm. Biopharm. 2012, 80, 203–208. 10.1016/j.ejpb.2011.07.009. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao2c04748_si_001.pdf (2.9MB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES