Next Article in Journal
Purification and Characterization of A New Cold-Adapted and Thermo-Tolerant Chitosanase from Marine Bacterium Pseudoalteromonas sp. SY39
Next Article in Special Issue
Functionalized Periodic Mesoporous Organosilicas: Tunable Hydrophobic Solid Acids for Biomass Conversion
Previous Article in Journal
Cardiomyocyte Progenitor Cells as a Functional Gene Delivery Vehicle for Long-Term Biological Pacing
Previous Article in Special Issue
Metal Organic Frameworks Based Materials for Heterogeneous Photocatalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Catalytic Conversion of Carbon Dioxide through C-N Bond Formation

State Key Laboratory of Coal Conversion, Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, China
*
Author to whom correspondence should be addressed.
Submission received: 1 December 2018 / Revised: 19 December 2018 / Accepted: 4 January 2019 / Published: 5 January 2019
(This article belongs to the Special Issue Catalysis: Providing Sustainable Solutions to Global Challenges)

Abstract

:
From the viewpoint of green chemistry and sustainable development, it is of great significance to synthesize chemicals from CO2 as C1 source through C-N bond formation. During the past several decade years, many studies on C-N bond formation reaction were involved, and many efforts have been made on the theory. Nevertheless, several great challenges such as thermodynamic limitation, low catalytic efficiency and selectivity, and high pressure etc. are still suffered. Herein, recent advances are highlighted on the development of catalytic methods for chemical fixation of CO2 to various chemicals through C-N bond formation. Meanwhile, the catalytic systems (metal and metal-free catalysis), strategies and catalytic mechanism are summarized and discussed in detail. Besides, this review also covers some novel synthetic strategies to urethanes based on amines and CO2. Finally, the regulatory strategies on functionalization of CO2 for N-methylation/N-formylation of amines with phenylsilane and heterogeneous catalysis N-methylation of amines with CO2 and H2 are emphasized.

1. Introduction

With the development of green chemistry, sustainable raw materials are urgently required, and growing attention is being paid to green resources such as biomass [1], carbon dioxide (CO2) [2,3,4] etc. due to their good features such as abundance, renewability and sustainability. In the past several decades, utilization of CO2 as an ideal C1 building block become a hot and promising field in both academic and industrial respects. A large number of advantages promoted the wide and deep research on catalytic conversion of CO2 into chemicals (C-O, C-N and C-C bond construction) and fuels [5,6,7]. Though great breakthroughs have been achieved during these years, however, CO2 conversion still faces many challenges due to its thermodynamic stability and kinetic inertness. Therefore, many strategies have been employed for effective CO2 conversion, including CO2 and substrate activation and active reagents (such as hydrogen, alkene, alkyne, epoxides, amines, etc.) [8,9,10,11] with high free-energy. Among the active substrates, amines were demonstrated to be some of the most high-efficiency candidates for chemical fixation of CO2 through C-N bond formation. The earliest and most successful example is the production of urea from ammonia and CO2, which accounts for most (>90%) of all of the industrial chemicals produced using CO2 as a chemical feedstock so far. Besides, many other N-contained chemicals such as urea derivatives, carbamates, oxazolinones, quinazoline-2,4(1H,3H)-diones, formamides, methylamines, etc. could also be synthesized through C-N bond formation from CO2 and various amines or their derivatives [9].
In 2011, the He group comprehensively reviewed the development of CO2 conversion through C-N bond formation with the production of oxazolidinones, quinazolines, carbamates, isocyanates and polyurethanes, etc. [6] In the subsequent years, several short reviews describing special fields were also reviewed. For example, catalytic strategy insights were gained on the reaction of CO2 and aziridines (a 100% atom economical route) to prepare oxazolidinones and polyurethanes in 2012 [12] and 2017 (heterogeneous catalysis schemes) [13]. Subsequently in 2013, Ghosh et al. summarized various synthetic methods to produce oxazolidinones from CO2 with a variety of aziridines, propargylamines and 2-aminoalcohols [14]. In addition, part of a mini-review from the Tomishige group focuses on the direct conversion of CO2 with aminoalcohols and diamines to give cyclic carbamates and cyclic ureas via heterogeneous catalysts [15]. Recently, Li et al. focused on catalytic synthesis of urea derivatives using CO2 and amines [16]. Moreover, a book chapter from the He group mainly focuses on the synthesis of carbonyl-containing chemicals, including 2-oxazolidinones and quinazoline-2,4(1H,3H)-diones, through C-N bond formation [17].
Methodology research will provide a solid theory for CO2 conversion into value-added products through C-N bond formation, and novel strategies have sprung up in the past few years. Herein, we would like to highlight recent (2014–2018) developments in catalytic methods for chemical fixation of CO2 to form various chemicals through C-N bond formation (Scheme 1). Notably, the work described in the previous reviews will not be discussed in this review, unless specially required in individual cases.

2. Chemical Fixation of CO2 through C-N Bond Formation

2.1. Sequential Carboxylation and Cyclization with C-O/C-N Bond Formation

2.1.1. Synthesis of Oxazolidinone Derivatives

Oxazolidinones, a class of five-membered cyclic carbamates, display biological activities and good antibacterial properties. Substituted oxazolidinones are not only used as substrates in organic synthesis, but also as chiral auxiliaries in asymmetric synthesis [18]. In particular, some oxazolidinones can be used as the core unit of antibacterial agents, such as toloxatone, linezolid and tedizolid [19]. Oxazolidinones can be prepared through the intramolecular nucleophilic cyclization of propynyl carbamates as well as chemical fixation of CO2 with several different reagents such as aziridines, aminoalcohols and propargylic amines [20]. However, from an availability and economic viability viewpoint, the three-component reaction of propargylic alcohols, primary amines and CO2 represents the most attractive method for the synthesis of oxazolidinone derivatives [19].
In the three-component reaction, the carboxylative cyclization of the propargyl alcohol with CO2 affords an α-alkylene cyclic carbonate, which then undergoes a nucleophilic attack by a primary amine followed by an intramolecular nucleophilic attack and dehydration to generate 4-alkylene-2-oxazolidinones or 2(3H)-oxazolones (See Scheme 2, Path A) [21]. It is worth noting that the formation of products is strongly associated with the substituents on the propargylic alcohols.
This reaction was catalyzed by copper salts including CuCl [22] and CuI [23], with the assistance of ionic liquids (ILs) [20] to obtain 4-methylene-2-oxazolidinones or 2(3H)-oxazolones. Alternatively, metal-free catalysts, i.e., bicyclic guanidines [24] and [DMIm]BF4 [25] are effective in the cycloaddition reactions of CO2 with propargylic alcohols and primary amines under relatively mild conditions, although the product yields are not high enough. Moreover, silver salts are suitable catalysts showing high efficiency in the three-component reaction (Scheme 3) [26].
Gratifyingly, He et al. obtained corresponding 2-oxazolidinones in moderate to excellent yields by using a bifunctional Ag2WO4/Ph3P system (Scheme 3a and Figure 1) [21]. The reaction was carried out under 0.5 MPa pressure of CO2 and 50 °C without using any solvent. As shown in Scheme 3a, the high catalytic activity was ascribed to the dual activation capacity of Ag2WO4. The silver ion, acting as a π-Lewis acid, could activate the C-C triple bond. Meanwhile, both CO2 and the O-H bond of the propargylic alcohol were activated by tungstate anion.
Recently, a noninterpenetrated primitive-cubic (pcu)-type sulfonate-based metal-organic framework (MOF) was synthesized by Fei et al. [27]. Due to the unusual linker-defective nature of the MOF, Ag(I) active sites were incorporated. The synergistic effect of high CO2 affinity and alkyne activation nature enabled Ag(I)-decorated sulfonate-MOF successfully catalyze the cyclic carboxylation of propargyl alcohol into oxazolidinones under atmospheric pressure of CO2 in the product yield of up to 100% (Scheme 3b). Notably, a fast mass transport resulting from the high surface area and the robust nature of the pcu topology result in a high catalytic efficiency.
An alternative catalyst to obtain 2-oxazolidinone derivatives or 2(3H)-oxazolones was a simple organocatalyst, i.e., 2,2′,2′′-terpyridine in the absence of any solvent (Scheme 3c) [18]. Nevertheless, high temperature and CO2 pressure was indispensable. Besides, this catalytic system was not suitable for anilines or internal propargyl alcohols.

2.1.2. Synthesis of Carbamates

One of the most important acyclic carbamates, i.e., β-oxopropylcarbamate, has attracted increased attention owing to its growing applications in agriculture and pharmacology [28]. It also acts as a useful intermediate in organic synthesis as well as an amine protective group in peptide chemistry [29]. A one-pot reaction of a propargylic alcohol, secondary amine and CO2 is an environmentally benign strategy to produce β-oxopropylcarbamate with high atom economy and easy handling [19]. Unlike the formation of 2-oxazolidinone derivatives, β-oxopropylcarbamates can be obtained via nucleophilic ring-opening of α-alkylidene cyclic carbonates and subsequent tautomerization (See Scheme 2, Path B) [21].
There are various well-established protocols with metal complexes including ruthenium(II) [30], iron complexes based on 1,1′-bis(diphenylphosphino)-ferrocene [31] and silver complexes [29] etc. that were developed before 2012. Additionally, bicyclic guanidines [24] and noncatalytic systems [28] were also demonstrated effective in the coupling of propargylic alcohols, secondary amines and CO2. Unfortunately, these approaches generally have severe limitations such as high pressure (≥2 MPa) and additional energy requirements.
In recent years, silver-containing catalytic systems have showed excellent performance in the production of β-oxopropylcarbamates through the three-component reactions, as shown in Table 1.
In 2014, He and co-workers reported that silver complex [(PPh3)2Ag]2CO3, which could be formed in situ from Ag2CO3 and PPh3, promoted the reaction even under atmospheric CO2 pressure [32]. According to the results of 1H and 13C nuclear magnetic resonance (NMR) spectroscopy, CO2 and propargylic alcohol were presumably simultaneously activated by [(PPh3)2Ag]2CO3, which greatly facilitated the reaction (Figure 2). Additionally, the scope of the reaction system was demonstrated to be wide (Figure 3). Furthermore, the combinations of Ag2WO4 and Ag2O with PPh3, respectively, were also used as catalysts to synthetize β-oxopropylcarbamates under mild reaction conditions. Ag2WO4/PPh3 showed good catalytic activity under the solvent-free condition primarily because of the reactant activation through [Ag(PPh3)]+ and [WO4]2− (Scheme 3a) [21]. Nevertheless, the aromatic amines could not afford the desired products in this system. In contrast, the employment of Ag2O/PPh3 successfully accomplished the reaction of propargylic alcohols and ammonium carbamate (amine-CO2) to achieve the quantitative fixation of CO2 [33]. Herein, ammonium carbamate formed from amine and CO2 was used as a CO2 surrogate. As shown in Scheme 4, propargylic carbonate intermediate A is initially generated from the reaction of propargylic alcohol with ammonium carbamate. Subsequently, the intermediate A is converted into silver propargylic carbonate B in presence of silver(I) catalyst. Then, after the activation of C≡C bond by the silver species, intermediate C is obtained and follows by the formation of α-alkylidene cyclic carbonate. Finally, in the presence of a secondary amine, the corresponding β-oxopropylcarbamate can be obtained via tautomerization.
In addition, Song et al. also reported that Ag2CO3/(p-MeOC6H4)3P could catalyze the coupling reaction of terminal propargylic alcohols, CO2, and secondary amines with a high turnover number (turnover frequency) of up to 3350 (35 h−1) only using 0.01 mol% silver catalyst [34]. More recently, 13Ccarbonyl-labeled reactions, 13C-NMR spectrum and high-resolution mass spectrometer (HRMS) characterization were performed by Zhang and Hao et al. in the AgCl/Et4NCl system. The results further confirmed that carbon atom in -N-C- skeleton of product derived from CO2 and the evidence of the previously reported reaction routes [35].

2.2. Direct Carboxylation/Cyclization with C-N Bond Formation

2.2.1. Carboxylative Cyclization of Propargyl Amines with Carbon Dioxide

Propargylic amines, as some of the important propargylic compounds, are not only easily available but also suitable for further transformations because of their two functional sites (an amino group and a triple bond) [36]. To date, great efforts have been made to develop the carboxylative cyclization of propargylic amines with CO2 to obtain 2-oxazolidinones. As described in Scheme 5, CO2 initially reacts with propargylic amine to afford intermediate A. Then, an intramolecular cyclization occurs with the generation of product. Notably, the second step is considered to be the rate-determining step [37].
Taking into account that activation of the carbon-carbon triple bond is of importance to promote the carboxylative cyclization, coinage metals (Au, Ag and Cu) have been extensively employed as catalysts due to π-coordination with the alkyne (Scheme 6). Among them, N-heterocyclic carbene (NHC)-Au complexes obviously paved the way to chemical fixation of CO2 (1 and 2, Scheme 6a). Originally, Ikariya et al. found that AuCl(IPr) (1) was capable of catalyzing the synthesis of (Z)-5-alkylidene-1,3-oxazolidin-2-ones in methanol under mild conditions [38]. Further investigation of the substrate scope revealed that the catalytic activity was weakened by the introduction of aromatic substituents at the alkyne terminus [39]. Metallodendrimers have a kind of functional or catalytic site at the core, and their solubility and physical properties can be modified via peripheral structures [40]. Accordingly, amphiphilic dendritic Au/G1[PEG] (2) also exhibited good catalytic activity in aqueous media under atmospheric pressure of CO2 [41].
Moreover, Sadeghzadeh developed various fibrous nanosilica (KCC-1)/metal nanoparticles (NPs) to catalyze the incorporation of CO2 into propargylic amines (3, 4 and 5, Scheme 6a). KCC-1 has a unique fibrous morphology bearing with high surface area, tunable pore size and pore volume, controllable particle size, and improved stability. High polyglycerol (HPG)@KCC-1/PPh2/Au (3) and Fe3O4/KCC-1/tetrazolylidene/Au (4) catalyst not only catalyzed and gave 2-oxazolidinones in good to excellent yields under 0.5 MPa CO2 pressure at room temperature with water as solvent, but also could be recovered and reused several times without appreciable loss of activity [42,43]. Especially, the turnover number of the latter system was up to 1200. In addition, KCC-1/Salen/Ru(II) NPs (5) was also efficient in production of oxazolidinones, but high reaction temperature was required [44].
Recently, a gold catalyst containing an Z-type ligand composed of a diphosphine-borane (DPB), namely [Au(DPBF)]SbF6 (6, Scheme 6a) for the carboxylation reaction was developed [45]. The desired products were formed in air (ca. CO2 300–700 ppm) at room temperature in the presence of 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU). However, the yields of the vast majority of corresponding products were around 40%.
As shown in Scheme 6b, silver compounds also displayed excellent performance in the carboxylative cyclization. In 2012, Yoshida et al. reported that AgNO3/DBU (7) catalytic system was used for CO2 capture and transformation at 60 °C and 0.1 MPa CO2 in dimethyl sulfoxide (DMSO) [46]. During the process, atmospheric CO2 was initially trapped by DBU to give DBU-CO2 complex subsequently, which could propel the reaction.
Additionally, a simple and low cost CuI/DBU (8, Scheme 6c) system was explored to synthesize 2-oxazolidinones through the coupling reaction of propargylic amines and CO2 by Wang et al. [47]. The DFT results suggested that CuI and DBU have an excellent synergistic effect in promoting the reactions. DBU could capture and transfer protons; meanwhile, CuI activated the C≡C triple bond of propargylic amine. Similarly, CuI/[DBUH][TFE] (9, Scheme 6c) system also afforded 2-oxazolidinones in outstanding yields under mild conditions [48]. Moreover, He and co-workers exploited a series of metal-substituted polyoxometalate (POM)-based ILs including Cu(II), Co(II), Fe(II), Ni(II), Zn(II) and Mn(II) [49]. Among them, [(nC7H15)4N]6[α-SiW11O39Cu] (10, Scheme 6c) displayed higher catalytic activity than that of other metal-substituted POM-based ILs. It was mainly ascribed to simultaneous activation of propargylic amines and CO2 by the single-component catalyst. Notably, this procedure featured as environmentally benign and low energy-input manner.
In recent years, considerable attention has also been paid to Zn-based catalysts in CO2 fixation because of their tunable Lewis acidity, natural abundance and eco-friendly properties. In 2016, He et al. found that the mononuclear ZnII complex ZnCl2(TBD)2 (TBD = 1,5,7-triazabicyclo[4.4.0]dec-5-ene) could fix CO2 by propargylic amines to deliver corresponding 2-oxazolidinones without any solvent (Scheme 7a) [50]. Nevertheless, the internal propargylic amine was unreactive with CO2 in this catalytic system. In contrast, the terminal and internal secondary propargylamines were smoothly transformed into oxazolidinone derivatives by Indenediide-based Pd complex catalysts under mild conditions in DMSO (Scheme 7b) [51]. In particular, six-membered cyclic carbamates and N-aryl oxazolidinones were obtained (Scheme 7c). In addition, primary propargylamines, which were prone to occurring side reaction due to the NH2 group, also led to the desired oxazolidinones in good yields.
As can be seen from the above examples, the reported metal catalytic systems generally employed expensive noble metals or complicated catalytic components. At the same time, metal-free catalytic processes can reduce the cost and avoid the pollution caused by metals. In recent years, there are well-established protocols with vigorous metal-free catalytic systems such as TBA2[WO4] (TBA = tetra-n-butylammonium) [52], ILs [53], NHCs (e.g., 1,3-di-t-butylimidazol-2-ylidene) [54] Bu4NF [55,56] and so on.
In 2012, Minakata reported an effective catalytic system being composed of t-BuOI and CH3CN for the carboxylative cyclization of propargyl amine and CO2, and 2-oxazolidinone bearing iodomethyl group could be generated under mild conditions (Scheme 8a) [57]. Afterwards, with two equivalents of a protic ionic liquid namely 1,8-diazabicyclo-[5.4.0]-7-undecenium 2-methyl- imidazolide ([DBUH][MIm]), the desired products were obtained in good to excellent yields (Scheme 8b) [37]. Gratifyingly, this cheaper and greener catalyst could be easily recycled and reused at least five times. Han et al. also found that [DBUH][MIm] promoted both the CO2 electrophilic attack and the intramolecular cyclization step by capturing and transferring proton through theoretical studies, respectively.
Very recently, the reaction of propargylic amines and CO2 were carried out in the presence of a cyanuric acid (H3CYA)-based organocatalyst (Scheme 8c) [58]. [Et4N]3[CYA] was the main component of the most effective catalyst among these organocatalysts. Notably, this was the first reported example of direct organocatalytic conversion of low-concentration CO2 in air (0.04% v/v) for 2-oxazolidinone synthesis. It could be ascribed to the better activation of CO2 in the presence of [H3-nCYA]n− (n = 1−3) than that of [Phth] and [2-PyO] anions (Scheme 8e). Besides, triethanolamine (TEOA) as a low-cost and biodegradable alkanolamine showed excellent performance on affording 2-oxazolidinones (Scheme 8d) [59]. As can be seen from Scheme 8f, TEOA activated CO2 to form a ring-shaped carbonate intermediate, which could be beneficial to the reaction procedure.
It is worth mentioning that allylamines are less active than propargyl amines in the reaction with CO2 because C≡C bonds are more prone to being attacked by carbamates than C=C bonds. However, the carboxylative cyclization of allylamines and CO2 would significantly expand the toolbox of 2-oxazolidinones synthesis. Most recently, Yu and co-workers summarized recent progress on this reaction by homogeneous catalysis in detail [60]. In addition, they also reviewed various cyclization reactions of allenic amines with CO2.

2.2.2. Direct/Indirect Condensation of Amino Alcohols and Carbon Dioxide

In 2013, there were two excellent reviews dealing with the synthesis of 2-oxazolidinones through carbonylation of corresponding amino alcohols with CO2 (before 2013) by Pulla’s [14] and Tomishige’s [15] groups. Recently, Fekri et al. provided a review of recent developments (before 2017) in the synthesis of 2-oxazolidinone derivatives via dehydrative condensation of β-amino alcohols and CO2 with a discussion on the mechanistic aspects of the reactions in detail [61]. Based on the above works, this review will briefly discuss only the most recent literature (2017–2018) on direct/indirect condensation of amino alcohols and CO2.
Despite efficient processes, the direct condensation of amino alcohols and CO2 were limited by harsh conditions (high temperature and pressure) and side reactions [62,63,64]. This could be attributed to the thermodynamic limitation of the reaction with water as by-product (Scheme 9a). Hence, organic bases, electrophiles and CeO2 were used for removing water from the system in most cases to improve the yield of products [65,66].
In 2018, Repo et al. developed an efficient one-pot approach using an external base and p-toluenesulfonyl chloride (TsCl) as a hydroxyl group activating reagent to synthesize cyclic carbamates from CO2 and amino alcohols (Scheme 9b) [67]. This stoichiometric reaction system allowed 2-oxazolidinones with the 3-aryl-5-alkyl substitution pattern to be obtained under mild conditions in good yields and high enantiomeric excess. In the reaction (Scheme 9c), a carbamate salt is formed from CO2 and amino alcohol, followed by the tosylation of the -OH to improve its leaving group character. Finally, ideal product is obtained via a ring-closing process. However, the carbamate species must be formed firstly and stabilized to avoid the competition of N-tosylation reaction. Unlike the above base-promoted work, a series of K-La-MgO with different K loadings (1, 3, 5, and 7 wt%) synthesized by combustion were screened for carbonylation of diethanolamine in a green and clean route [68]. Among them, 5% K-La-MgO was found to show the best performance. The catalytic process with 5% K-La-MgO offered 72% conversion of diethanol amine and 100% selectivity of 3-(2-hydroxyethyl)-1,3-oxazolidin-2-one at 150 °C and 2.0 MPa CO2 pressure.
Additionally, a thermodynamically favourable one-pot three-component approach of propargylic alcohols, CO2 and 2-aminoethanols was described by He et al. in 2016 at first, offering 2-oxazolidinones with equal amount of α-hydroxyl ketone (Scheme 10a). As shown in Table 2, organic solvent, high CO2 pressure and excess propargylic alcohols were indispensable in the Ag-based catalytic system to achieve excellent yields [69,70], while in the CuI/1,10-phen-catalyzed three-component reaction, the desired products in good to excellent yields were obtained under mild reaction conditions without solvent. It was worth mentioning that cheap, commercially available copper catalysts, and the ability of C≡C activation have been widely utilized in CO2 transformations. The reaction consists of two steps. Firstly, α-alkylidene cyclic carbonate intermediate M1 is formed via the carboxylative cyclization of propargylic alcohol and CO2. Then, M1 undertakes a nucleophilic addition by a 2-aminoalcohol to form the corresponding β-oxopropylcarbamate species M2, which simultaneously affords the 2-oxazolidinone and an α-hydroxyketone with a subsequent intramolecular cyclization (Scheme 10b). Notably, copper complex Cu2I2(phen)2 in situ formed from CuI and 1,10-phen could activate the C≡C through coordination. Most recently, 1,5,7-triazabicylo[4.4.0]dec-5-ene ([TBDH][TFE]) trifluoroethanol was used to catalyze the synthesis of oxazolidinones from propargylic alcohols, 2-aminoethanols and CO2 under atmospheric pressure [71]. The catalysts could be recovered and reused at least five times without obvious loss of activity.

2.3. Quinazoline-2,4(1H,3H)-Diones through Carbonylation of 2-aminobenzonitrile and CO2

Quinazoline-2,4(1H,3H)-diones have wide applications in the synthesis of pharmaceuticals, such as Zenarestat, Prazosin, Doxazosin, and Bunazosin [72]. The synthesis of quinazoline-2,4(1H,3H)-dione derivatives through carbonylation of 2-aminobenzonitrile and CO2 is an environmentally benign process with 100% atom efficiency. This reaction was highlighted by Soleimani-Amiri et al. [72]. They classified the reactions based on the type of catalysts to review the works before 2017. Thus, we continued to summarise the published work in recent two years.
According to the previous literature [73,74,75], amino group of 2-aminobenzonitrile is initially activated by basic catalysts, which facilitates the nucleophilic attack of 2-aminobenzonitrile at CO2 to generate intermediate A. Then, nucleophilic cyclization of A produces B. Finally, B converts into quinazoline-2,4(1H,3H)-dione with the regeneration of the catalysts (Scheme 11).
Undeniably, numerous IL-based systems [73,74,76,77,78] exhibit excellent performance in the transformation of 2-aminobenzonitrile and CO2 [79]. However, almost all the reported IL systems involve high usage amounts of dual catalyst-solvent media (>1.0 equiv.). There were no intensive and systematic research figuring out the role of the IL, such as the effects of cation and anion until Dyson’s work [80] and Wang’s work [81] was published (Scheme 12).
Dyson et al. found that the cation from the IL did not play a direct role in the activation of 2-aminobenzonitrile, CO2, or the intermediates in the catalytic cycle [80]. On the contrary, the basicity of the anion had an inseparable relationship with the reaction rate. Notably, the acidity of quinazoline-2,4(1H,3H)-dione resulted in its deprotonation in basic catalytic conditions with the formation of quinazolide anion. Furthermore, the reaction was limited due to the neutralization of the basic IL catalyst and quinazolide anion.
On the other hand, according to quantum-chemical calculations, NMR spectroscopic investigations and controlled experiments, Wang et al. proposed that the basicity of cation affects the catalytic activity of ILs dramatically (Figure 4) [81]. In addition, the cation with more OH groups exhibited poorer catalytic activity. Therefore, a hydroxyl functionalized aprotic IL [Ch][Im] was designed and exhibited high catalytic activity (Scheme 13b).
Inspired by Dyson’s work, Fujita et al. found that nBu4NF was an effective quaternary ammonium salt for catalyzing the cyclization of 2-aminobenzonitriles when a catalyst loading of 1 mol% was used at 110 °C under CO2 pressure of 2 MPa for 24 h [56].
As introduced in Section 2.2.1 of this review, cyanuric acid-based organocatalystw could be good activators for CO2 capture and utilization [58]. [Et4N]3[CYA] was also an efficient organocatalyst in converting atmospheric pressure of CO2 into quinazoline-2,4(1H,3H)-diones without any external base or promoter.
In recent years, great efforts have been made to develop heterogeneous catalysts for a greener procedure to quinazoline-2,4(1H,3H)-dione. Both a graphitic carbon nitride catalyst [82] and a functionalized basic nanocrystalline zeolite [83] prepared by Srivastava et al. were demonstrated to be effective in the synthesis of quinazoline-2,4(1H,3H)-dione from the reaction of 2-aminobenzonitrile and CO2 (Scheme 13).
In recent years, much attention was paid to the synthesis of quinazoline-2,4(1H,3H)-diones via the multicomponent reactions involving CO2 (Scheme 14). In 2017, Cheng et al. reported palladium-catalyzed multicomponent reactions of O-alkynylanlines, aryl iodides, and atmospheric pressure of CO2, generating 3,3-diaryl 2,4-quinolinediones [84]. This reaction provided a one pot strategy to incorporate CO2 into heterocycles bearing a quaternary carbon center (Scheme 14a). As shown in Scheme 14c, the carboxylation of amine with CO2 and the trans-oxopalladation of C≡C bond by ArPdIIX species sequential occur to afford intermediate A. Subsequently, intermediate A undergoes an reductive elimination, deprotonation and rearrangements to generate the product. Moreover, palladium-catalyzed cyclization reactions of O-haloanilines, CO2 and isocyanides were employed to prepare N3-substituted quinazoline-2,4(1H,3H)-dione with moderate to excellent yields (Scheme 14b) [85].

2.4. Urea Derivatives from Amine Derivatives and CO2

It is well known that urea derivatives have a wide range of applications in pharmaceutical chemistry, organic synthesis as well as analytical chemistry [86,87]. Traditionally, urea derivatives were synthesized by the stoichiometric reaction of amines and isocyanates [88]. Moreover, the reaction of amines and highly toxic phosgene/phosgene analogs was performed to afford urea derivatives [89]. Besides, catalytic oxidative carbonylation of amines in the presence of CO was also developed to obtain ureas [90]. Notably, alternative synthetic routes of CO2 as a source with amines have attracted considerable attention in recent years.
In the last decades, bicyclic guanidines, as important organic bases, have shown to be active catalysts in chemical fixation of CO2. For example, N,N′-dialkylureas were synthesized via the carbonylation of amines with CO2 using TBD as catalyst (Scheme 15) [91]. Among three TBD-CO2 species (ac) shown in Scheme 15, the formation of bicarbonate-guanidinium salt c is ascribed to trace of water. The reaction proceeds with a sequential electrophilic attack of amine on c, elimination of water from M1 and reaction between M2 and amine leading to N,N′-dialkylureas.
The use of metal oxides and metal salts of oxalates significantly paved the way for N,N’-dialkylurea synthesis in recent years (Table 3). CeO2 showed high activity for the direct synthesis of 1,3-dibutylurea from CO2 and n-butylamine [92]. The catalytic system was applicable to various amines such as linear primary alkylamines and branched primary alkylamines (Figure 5). Meanwhile, secondary amines and aniline, which were unreactive using CeO2 in N-methyl-2-pirrolidinone (NMP), were also transformed into the corresponding ureas when 2-cyanopyridine was combined with CeO2 in NMP.
Y2(C2O4)3 and Y0.08Zr0.92O1.96 mixed oxide were independently used to catalyze carbonylation of aliphatic primary amines with CO2. Y2(C2O4)3 provided the highest yield of N,N′-dialkylurea among various metal salts of oxalates (Zr(C2O4)2, Ce2(C2O4)3, Mn(C2O4), Na2(C2O4), Ni(C2O4)) [93]. Y0.08Zr0.92O1.96 mixed oxide as the heterogeneous catalyst showed higher catalytic activity than Y2(C2O4)3 [94]. The presence of oxygen vacancies in the mixed oxide was of great importance for adsorption and activation of CO2 in the carbonylation process. It should be ascribed to the additional reduction potential derived from the oxygen vacancies for the reduction of CO2 to CO and/or surface carbonaceous species (an important intermediate).
Indium-catalyzed transformations of CO2 with aliphatic and aromatic silylamines have been exploited by Stephan et al. [95]. The indium compound In(N(SiMe3)2)Cl2·(THF)n (THF = tetrahydrofuran) was able to efficiently catalyze the reaction and afford a wide range of aryl and alkyl ureas with 0.05–5 mol% catalyst loadings. Notably, the available main group metal-catalyzed transformations of CO2 were improved.
Polyureas containing urea linkages and connected by hydrogen bonds are a new kind of polymer. In 2016, good to excellent yields of various polyureas were achieved with different diamines and CO2 by hexyltributylphosphonium aminotriazole (P4,4,4,6ATriz, ATriz = 3-amino-1H-1,2,4-triazole) (1, Scheme 16) IL catalyst [96]. It was found that the catalytic performance was essentially consistent with the basicity of ILs described by Deng et al. Recently, some functional polyureas were successfully synthesized via the polymerization of CO2 with diamines using amino triazole alkali salts i.e., MATriz as catalysts (2, Scheme 16), and M was Li, Na, K [97]. More importantly, the catalysts could be reused for several times without obvious deactivation.
Additionally, Arai et al. found that DBU (3, Scheme 16) was the most active among multiple organic and inorganic base catalysts in addition of CO2 to 4,7,10-trioxa-1,13-tridecanediamine (TOTDDA) [98]. DBU could activate both CO2 and -NH2 group of TOTDDA, which was demonstrated by in situ high-pressure attenuated total reflectance Fourier transform infrared spectroscopy (ATR-FTIR), resulting in a high catalytic performance (Figure 6).

2.5. Novel Synthetic Strategies of Carbamate Derivatives Based on Amines and CO2

2.5.1. N-Tosylhydrazones as a Building Block to Construct C-N Bond

N-Tosylhydrazones have been extensively employed as useful building blocks to construct complicated molecules. In 2015, Jiang et al. reported the first approach for the synthesis of organic carbamates by a three-component coupling reaction of CO2, amines, and N-tosylhydrazones (Scheme 17a) [99]. Notably, in the presence of K2CO3 and a mixed CH3CN/H2O solvent, the reaction system tolerates a variety of valuable functional groups on the products, which provided ample potential for further organic syntheses. Subsequently, Chung et al. modified the procedure performed under mild conditions with desired carbamates (Scheme 17b) [100]. Nevertheless, poor yields were observed compared with Jiang’s work.
At present, however, there is insufficient evidence for the proposed mechanism [99]. Presumably, for example, N-tosylhydrazone decomposes in the presence of K2CO3, and generates diazo compound, which can be followed by a protonation and a thermodynamically favorable liberation of dinitrogen to give allylic carbocation species M1 (a resonance hybrid of A and B) (Scheme 17d). Finally, the carbamate anion of M2 as a nucleophile attacks to M1 generating products of C and D.
Notably, Cheng et al. developed a strategy towards 4-aryl-2-quinolinones by a three-component coupling reaction of N-tosylhydrazones, 2-iodoanilines and CO2 catalyzed by PdII/phosphine system, which was benefited from the N-tosylhydrazone chemistry (Scheme 17c) [101]. Gratifyingly, four novel bonds: two C-C, one C=C and one C-N formed simultaneously in the heterocycles.

2.5.2. O-Aryl Carbamates from Three-Component Reactions Containing Amines and CO2

A straightforward route to O-aryl carbamates was achieved through a three-component reaction of CO2, amines and diaryliodonium salts with an organic base (DBU) as the promoter (Scheme 18a) [102]. In this phosgene- and metal-free protocol, diaryliodonium salts was used as electrophilic arylating agents to trap the in situ-generated carbamate anions. However, this transformation generated a stoichiometric amount of by-product, i.e., aryl iodides. In addition, the preformation of diaryliodonium salts was required.
Additionally, a more convenient route to O-aryl carbamates by a copper(I)-catalyzed oxidative coupling reaction between arylboronic acids, amines and CO2 was developed (Scheme 18b) [103]. At the same time, BF3·OEt2 was employed as the promoter and O2 as the oxidant. This approach tolerated a wide range of functionalized O-aryl carbamates.
Ribas and Company et al. reported a one-pot procedure of a model pincer-like arene, amines and CO2 to synthesize O-arylcarbamates (Scheme 18c) [104]. This reaction performed smoothly at room temperature by an aerobic Cu(II)-based catalyst and a nucleophile of carbamic ammonium salts in situ generated from amines and CO2 gas.

2.5.3. N-Arylcarbamate from Three-Component Reactions Containing Amines and CO2

N-Arylcarbamate synthesis directly from CO2, amines and organic halides often has negative environmental effects [105]. Choi et al. reported the first successful example of a halogen-free process for producing N-phenylcarbamate from CO2, aniline and dibutyltin dialkoxide (Scheme 19a) [106]. Gratifyingly, the corresponding methyl N-phenylcarbamate was formed in a yield of 80% along with N′-diphenylurea as by-product. Recently, they found that silicate ester (i.e., Si(OR)4) could be incorporated with aliphatic or aromatic amines and CO2 to synthesize N-arylcarbamate (Scheme 19b) [107]. Notably, the corresponding carbamates of aromatic amines could be obtained in a yield of up to 96%. Besides, the reaction was chemoselective toward amine activation. The use of Zn(OAc)2/1,10-phenanthroline (phen) catalysts offered the best performance because of the carboxylate-assisted proton activation, as shown in Figure 7.
Alternatively, adding the corresponding alcohols to the system including CO2 and amines was also an efficient route to obtain N-arylcarbamates. In 2018, the combination of CeO2 and 2-cyanopyridine was used in the direct synthesis of N-arylcarbamates from CO2, amines and alcohols (Scheme 19c) [108]. Various linear and branched alcohols were transformed into the corresponding carbamates in high selectivities.

2.5.4. Allyl Carbamates from Amines and CO2

As seen from Scheme 20a, Zhao et al. developed a three-component reaction of unsymmetrical allylic chlorides, CO2 and amines to afford allyl carbamates by Pd catalysis [109]. This method tolerated various allyl carbamates in moderate yields but with excellent regioselectivity (99/1).
A facile and efficient methodology for the synthesis of new acetal-type O-allyl carbamates by regioselective coupling reaction of CO2, amines and aryloxyallenes was explored in 2018 (Scheme 20b) [110]. A wide range of desired products were obtained in moderate to excellent yields using the catalytic system of AgOAc and TBD. Jiang et al. found that a cationic species generated from Ag(I) salt and TBD was the key active complex in the transformation, which was indicated by 1H-NMR studies.

3. C-N Bond Formation through CO2 Hydrogenation

For the past fifteen years, reductive methylation using N-site-based nucleophiles and CO2 to give the corresponding methyl amines or formamide, which represents one of the most economical and environmentally friendly routes for C-N bond formation using CO2 as C1 building block has been investigated intensively. Recently, in this hot field, several groups such as Cantat (2015, reductive functionalization of CO2 with amines) [111,112], Yan (2015, transition metal-catalyzed methylation) [113], Beller (2017, catalytic methylation) [114], He (2018, transition-metal-free catalysis) [115], Kühn (2018, an update of CO2 catalytic conversion) [116], Fernández-Alvarez and Oro (2018, homogeneous catalytic reduction of CO2 with silicon-hydrides) [117], Motokura (2018, organocatalysis with silanes) [118], have reviewed the recent developments in CO2 hydrogenation based on amine substrates from different viewpoints. On the basis of the previous reviews and the integrity of this article, here, the introduction of the development of CO2 conversion based on amines will be emphasized in a special view. The contents of this part include the regulatory strategies for functionalization of CO2 for N-methylation and N-formylation of amines with phenylsilane and heterogeneous catalysis N-methylation of amines with CO2 and H2.

3.1. Regulatory Functionalization of CO2 with Amines and Phenylsilane: N-Methylation and N-Formylation of Amines

In 2013, the methylation of amines was achieved using hydrosilanes as reductants through zinc complex catalysis by the Cantat group (Scheme 21) [119]. A mechanism going through a formamide intermediate was proposed. In recent years, many researchers undertook the study on selective catalysis of N-methylation and N-formylation of amines from amines, CO2, and hydrosilane.
Subsequently, they described the first iron catalysts able to promote the reductive functionalization of CO2 with amines using hydrosilanes as reductants to formamide and methylamine derivatives under mild reaction conditions (Scheme 22a) [120]. At room temperature, the chemoselectivity of formamide was almost 100%. After increasing the temperature to 100 °C, the selectivity of N-methylation was greatly improved and the majority methylamine derivatives were obtained under the elevated loading of catalyst. Similarly, iron-rich natural mineral Gibeon meteorite was used as an efficient heterogeneous catalyst for the N-formylation or N-methylation of amines with CO2 and hydrosilanes recently (Scheme 22b) [121]. In the work, a wide range of amines was converted into their corresponding formamides in high yields at room temperature. By simply varying the conditions e.g., increasing the temperature (25→100 °C) and catalytic loading (1→10 mol%), the selectivity was switched from N-formylated to N-methylated products. Moreover, the Gibeon meteorite catalyst was recycled and reused at least five times without appreciable activity loss.
In 2015, the Dyson group reported an effective thiazolium carbine-based catalyst for the N-formylation of amines, using polymethylhydrosiloxane (PMHS) as a reducing reagent under ambient pressure [122]. A board range of primary amines could be converted into the corresponding formamides with PMHS (200–300 mL) at 50 °C under atmospheric pressure of CO2. Notably, a interesting discovery, namely that the variation of reaction temperature could change the reaction products, was also made. For example, the N-formylation could be turned into N-methylation by changing the reaction temperature from 50 to 100 °C with PMHS (200–300 mL), and methylamine was obtained in high yield (Scheme 23).
Subsequently, the He group [123], Fu group and Lin group [124] reported the selective synthesis of formamides or methylamines from amines, CO2, and hydrosilane under the different catalytic conditions in the same year. In He’s work, butylammonium fluoride (TBAF) was used for the reductive functionalization of CO2 with amines to selectively afford formamides or methylamines by employing different hydrosilanes (Scheme 24). Formamides were obtained with triethoxysilane as reductant, and methylamines with phenylsilane in excellent yield under atmospheric pressure of CO2 at 30 and 50 °C, respectively. The mechanism of formation of the key silyl formate intermediate in the formylation step and fluoride-promoted hydride transfer from the hydrosilane to CO2/formamide was proposed.
The Fu group and Lin group developed a simple and effective alkali-metal carbonate, i.e., cesium carbonate, that catalyzed both the formylation and methylation reactions under mild conditions (Scheme 25). Through varying the reaction temperature and loading of silane, the formylation/methylation selectivity could be conveniently controlled. By means of experimental and computational studies, they revealed the possible mechanism including: (i) activation of Si−H by Cs2CO3, (ii) insertion of CO2 into Si−H, (iii) formylation of amines by silyl formate, and (iv) reduction of formamides to methylamines.
In 2007, the He group developed a glycine betaine-catalyzed strategy for the reaction of CO2 with amines and diphenylsilane (Scheme 26a) [125]. Through tuning the amount of CO2 and the reaction temperature, three kinds of products, i.e., formamide, methylamine, and aminal, were successively obtained. This scheme firstly achieved the hierarchical reduction of CO2 with amine and hydrosilane using organocatalysis with wide substrate scope. Almost at the same time, the Han group also reported the glycine betaine-catalyzed transformation of CO2 with amines to synthesize N-substituted compounds (Scheme 26b) [126]. Similarly, the selectivity to the reductive products could be controlled by the molar ratio of reactants (i.e., CO2, amines, and PhSiH3) and the temperature. Subsequently, they used lecithin as an organocatalyst for the formylation and methylation of various amines with CO2 to corresponding formamides and methylamines via using PhSiH3 as the hydrogen source [127]. Meanwhile, the product selectivity could also be easily controlled by the molar ratio of reactants and reaction temperature.
Recently, Chen and Xia reported an efficient DBU catalytic system for the selective N-methylation and N-formylation of amines with CO2 and PMHS (Scheme 27) [128]. The N-methylation products in high yields are obtained with 1 mol% DBU at 100 °C, and selective N-formylation of amines is realized with 10 mol% DBU at a lower temperature.
In previous studes, the selective N-methylation and N-formylation of amines with CO2 was mainly dependent on the reaction temperature because higher reaction temperatures favour methylamine formation. Recently, a tungstate catalyst was employed for selective catalysis reductive functionalization of CO2 with amines and phenylsilane (Scheme 28) [129]. By varying the CO2 pressure, 2-electron or 6-electron reduction of CO2 was achieved to give formamides or methyl- amines, respectively. Meanwhile, the N-formylation proceeding through the silyl formate intermediate and the N-methylation through an aminal intermediate were demonstrated via control experiments and NMR studies.
Afterwards, Li and He studied the ligand-controlled copper complex-catalyzed selective transformations of amines and CO2 to formamides and methylamines with phenylsilane as reductant (Scheme 29) [130]. Using the same reaction conditions and copper compound, selective catalysis was accurately regulated through simply changing the phosphine ligand.
Initially, the coordination of copper with phosphine ligand Ph2CyP or DPPB led to the copper complex and the subsequent formation of the active Cu–H species, followed by CO2 insertion with the generation of the copper formate (Cycle I). Then, copper formate reacts with amine and hydrosilane to give the formamide along with the formation of the active Cu–H species (Cycle II). The catalytic ability of copper complex determines the final products through the catalytic cycle I or cycle II. The Cu(Ph2CyP) complex gave the final formamide product, and Cu(DPPB) could further catalyze the formamide reduction to the methylamine product.

The Mechanism of Metal and Organocatalysis

The previously proposed catalytic pathways for the reactions of N-methylation and N-formylation are summarized in Scheme 30 [131]. Among, the mechanisms for the synthesis of aminal from amine, CO2 and hydrosilane (Cantat’s TBD-catalyzed method) is similar to the methylation (Scheme 31) [132]. The aminal product is obtained via a [Si]OCH2O[Si] intermediate, and HC(=O)O[Si] intermediate does not react with amine to form formamide. Therefore, the formation of aminal does not undergo the formamide intermediate process. Besides, other organocatalysts including glycine betaine [125,126], DBU [128], K2WO4 [129], also supported the mechanism of aminal intermediate for the methylation of amines with CO2 using hydrosilane as reductant through detailed control experiments (Scheme 32, equations 1–3) and DFT study [133]. In addition, in some metal-based systems such as Fe [120], Cs [124], Cu [130], and a few organocatalysts (e.g., TBAF [123], lecithin [127], etc.), the experimental evidence supports the mechanism of further formamide hydrogenation to the corresponding the methylation under the corresponding conditions (Scheme 32, equations 4–8). In these works, control experiments, e.g., using the product formamide as reaction material, were carried out to demonstrate the pathway (Scheme 30).
Recently, a mechanistic investigation was carried out by Nguyen group on the guanidine catalyzed reductive amination of CO2, using a combination of experimental and analysis method (1H, 29Si-NMR, FT-IR, MS, and GC profiling) (Scheme 33) [131]. The results showed that reduction of CO2 to formamide, aminal, and then N-methylamine is not sequential. Formamide was obtained dominantly at a lower temperature (23 °C), but an obvious increase of aminal, and then N-methylamine was detected at the higher temperature (60 °C).
Increasing the reaction temperature enables a competitive and higher-energy pathway to aminal and N-methylamine, which requires direct reduction of CO2 with PhSiH3 to formoxysilane PhSiH2OOCH intermediate. The theory provided a new evidence for the pathway to formamide from aminal (Scheme 33, pathway 9).

3.2. Heterogeneous Catalysis Methylation of Amines with CO2 and H2

The use of hydrosilane could lead to the formation of large amount of siloxanes as by-products which is not beneficial to the development of sustainable chemistry. Therefore, in place of hydrosilane with benign H2 would make the hydrogenation process cleaner. In addition, the advantages of heterogeneous catalyst such as easily recovery and recycle are obvious. Therefore, herein, heterogeneous catalysis on methylation of amines with CO2 and H2 was highlighted.
In 1995, the Baiker group reported the first example on the synthesis of methylamines from ammonia, H2 and CO2 through Cu/Al2O3 catalysts in a fixed-bed microreactor (Scheme 34) [134], but the efficiency was not high enough. Subsequently, they tested various types of metal–alumina catalysts (Cu, Ag, Pt, Ni, Co, Fe) to further improve the catalytic activities and selectivities [135,136,137]. The best amine production rates were obtained with Cu–Mg–Al mixed oxides at 280 °C, and over 79% selectivity of methylamine was obtained.
During the past decade years, many research groups continuously carried out the study on heterogeneous catalysis (Cu [138,139], Pd [140,141], Pt [142,143], Au [144,145], Re [146]) on methylation of amines with CO2 and H2, and a significant progress had been made totally (Table 4). For example, Shi and co-workers described effective heterogeneous catalyst systems CuAlOx [138] and Pd/CuZrOx [140] for the methylation of amines with CO2 and H2 as sources for the methyl group, respectively. Primary and second aromatic and aliphatic amines were converted to methylamines under the identical reactions (170 °C, H2/CO2: 6–7/3 MPa). In the same year, Shimizu et al. [142] reported a Pt-MoOx/TiO2 catalyst in the methylation of secondary amines under solvent-free conditions (0.5 MPa CO2 and 5 MPa H2). Besides the good performance in the wide scope of secondary amines, the catalyst was reused in methylation of N-methyl aniline with CO2 and H2 for at least ten cycles with no significant loss in its activity. Notably, the total TON during the successive 10 runs reached 433 higher than TONs of the homogeneous Ru catalysts. Recently, they also studied TiO2-supported Re catalyst (Re/TiO2) promoted the N-methylation of amines by using H2 and CO2 [146]. Re/TiO2 efficiently catalyzed the reaction of various amines including substituted anilines, tetrahydroquinolines, aliphatic secondary amines in 66–99% yield.
In 2015, Su and Wang [144] demonstrated a supported Au catalyst (Au deposited on alumina) which was highly effective for the methylation of both aromatic and aliphatic amines using CO2/H2 to N-methylated products (Scheme 35). The average TOF based on surface Au atoms was 45 h−1 in the methylation of aniline with CO2 and H2. Moreover, in their work, a variety of amines including aromatic, aliphatic, secondary, and primary amines were demonstrated with good performance (40–99% yield). Notably, the catalytic scheme could make the one-pot method of primary amines, aldehydes, and CO2 with H2 to unsymmetrical tertiary amines. In addition, to extend the work, Du et al. revealed that with the decrease of the Au nanoparticle size from 8.3 to 1.8 nm, the TOF values for methylation of N-methylaniline with CO2/H2 increased [145].
Zhao et al. demonstrated that PdGa bimetallic alloy nanoparticles were highly dispersed on the TiO2 support, and the CO2 could be activated through the interaction between the electron-deficient Pd and Ga, which was confirmed by several detailed characterization studies (TEM, TPR, XPS, CO-adsorption IR and high-pressure in situ FTIR). The generation of a formic acid intermediate was proposed for the methylation (Scheme 36) [141]. As a result, 98% conversion and 94% selectivity for N-methylaniline was obtained under the identical conditions (180 °C, 5 MPa H2, 5 MPa CO2, 10 h).
The reductive N-formylation of amines using CO2 and hydrogen is also a promising means of fixing CO2 into value-added chemicals, and heterogeneous catalysis schemes were also well developed recently (Table 5). Lately, the Shi group were the first to report a highly active supported palladium nanoparticle catalyst for the N-formylation of various aliphatic amines with CO2 and H2 (Scheme 37) [147]. During the preparation, Pd(NH3)xCly was initially adsorbed onto the carbon support, and then the active nano-Pd particles were generated by in situ reduction. Additionally, the activity of the Pd/C catalysts can be tuned by -OH groups through modulating the hydrophilic/hydrophobic properties of the carbon surface which further potentially promote the adsorption of CO2 and amines near the Pd sites.
At the same time, a bimetallic Pd−Au catalyst (Pd/Au molar ratio of 1:1) prepared by depositing Pd−Au alloy nanoparticles (around 3.0 nm) on polyaniline-functionalized carbon nanotubes (PANI-CNT) was demonstrated the excellent catalytic activity for the N-formylation of pyrrolidine using CO2/H2 [148]. Ju et al. revealed that Pd atoms are the true active sites for the hydrogenation reaction, and the interaction between Pd atoms and Au atoms on bimetallic Pd−Au/PANI-CNT is beneficial to enhance the catalytic performance through changing the electronic properties of the formed bimetallic Pd−Au nanoparticles. In the work, a variety of aliphatic secondary amines showed excellent reactivity and gave high yield (67.2–98.3%). The aliphatic primary amines afforded a lower yield (up to 62.2%), and the tert-butylamine nearly performed no reactivity (trace product). Meanwhile, aniline was also not suitable under the reaction condition (0.1% yield).
Recently, a Pd nanoparticle (average size <2.0 nm) catalyst supported on N-doped carbons (NCs) was prepared by Liu group and applied for effective N-formylation of amines with CO2 and H2 in ethanol [149]. The investigation revealed that the interaction between the Pd nanoparticles and nitrogen in the NC support was responsible for the good performance of the catalyst. The catalyst was reused for three times, and the product yield decreased (from 93% to 81%). TEM observations indicated the reason may be the aggregation of Pd nanoparticle (from 2 to 3.5 nm) under the experimental conditions. In addition, a serial of aliphatic primary and secondary amines showed good reactivity and afforded the target products in high yields (51–99%). However, the aromatic amines such as aniline and N-methylaniline gave the corresponding products in low yield (11%, 11%) under the same conditions.

4. Summary and Outlook

In conclusion, C-N bond formation reaction through CO2 as C1 source represents a big stride forward for the development of sustainable and green chemistry. In this work, we have summarized recent the advances of CO2 conversion to valuable chemicals from various N-contained substrates through metal or organocatalysis strategies. C-N bond formation has been achieved involving direct carboxylation/cyclization of CO2 with various N-contained substrates (such as propargyl amines, amino alcohols, allylic amines, allenic amines, etc.), sequential carboxylation and cyclization of propargyl alcohols, amines and CO2, as well as N-methylation/N-formylation of amine via CO2 hydrogenation. Among them, from the availability and the economic viability viewpoint, the multi-component cascade reaction of amines, CO2 and other nucleophilic candidates is especially receiving increased attention in recent years. Beside the advantages such as the thermodynamically favourable procedure, nontoxic and easily accessible material, the cascade approach was accomplished under mild reaction conditions with high efficiency. The insights on CO2 conversion through C-N bond formation reactions would be beneficial for the exploration of new strategies and directions on effective incorporation of CO2 into high value-added products.
Gratifyingly, encouraging advances on the development of novel strategies to construct organic molecules through C-N bond formation have been made, however, many problems and challenges still remain. Firstly, in most transformations involving CO2 conversion, the catalytic performance in transition metal systems is much better than that of organocatalysis strategies. Generally, the metal complexes easily forming active sites are more beneficial for activating the substrates such as alkene, alkyne and H2, etc. As a result, the catalytic efficiency and selectivity show much higher values, together with the lower catalytic loading, faster reaction rates, shorter times, and lower pressures and temperatures. Despite the superiority of metal complexes, their preparation processes are stricter (they are generally sensitive to water and air) and their stability is lower than that of organocatalysts. Also, the catalytic ability of organocatalysts has improved greatly. In addition, almost all of the metal-based catalysts were focused on the transition metals. Accordingly, the number of examples exploring main group metal-catalyzed transformations of CO2 is inadequate. Secondarily, the model products could be generally obtained in excellent yield and selectivity, but the examples of products being directly used as functional target molecules for medicines, natural products, or bioactive polymers, etc. are limited. Thus, the compatibility of catalytic methods in larger molecules bearing with much more complicated functional groups needs more exploration. Furthermore, for CO2 hydrogenation with amines, high yields and selectivities are achieved much easier using hydrosilanes than H2 as hydrogen source. However, the use of hydrosilanes will be lead to a large amount of siloxane by-products. Therefore, from the point of view of sustainable chemistry, the development of benign H2 process would have great significance and the exploration of robust methods for selective and effective catalysis is also essential despite being challenging. Besides, the efficiency of the noble transition metal-based homogenous schemes and metallic oxide-based heterogeneous catalytic systems must be further improved. Last but not least, it is challenging and imperative to discover more available methods and routes to transform CO2 into valuable chemicals with C-N bonds.

Funding

This research was funded by the National Natural Science Foundation of China (21602232) and the Natural Science Foundation of Shanxi Province (201701D221057).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Carbon dioxide (CO2); ionic liquids (ILs); primitive-cubic (pcu); metal-organic framework (MOF); nuclear magnetic resonance (NMR); gas chromatography (GC); high-resolution mass spectrometry (HRMS); N-heterocyclic carbene (NHC); nanoparticles (NPs); diphosphine-borane (DPB); 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU); dimethyl sulfoxide (DMSO); density functional theory (DFT); polyoxometalate (POM); 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD); tetra-n-butylammonium (TBA); cyanuric acid (H3CYA); triethanolamine (TEOA); p-toluenesulfonyl chloride (TsCl); 1,5,7-triazabicylo[4.4.0]dec-5-ene ([TBDH][TFE]); N-methyl-2-pirrolidinone (NMP); 3-amino-1H-1,2,4-triazole (ATriz); 4,7,10-trioxa-1,13- tridecanediamine (TOTDDA); attenuated total reflectance fourier transform infrared spectroscopy (ATR-FTIR); polymethylhydrosiloxane (PMHS); butylammonium fluoride (TBAF); turnover frequency (TOF); polyaniline-functionalized carbon nanotubes (PANI-CNT); N-doped carbons (NCs)

References

  1. Tuck, C.O.; Pérez, E.; Horváth, I.T.; Sheldon, R.A.; Poliakoff, M. Valorization of biomass: Deriving more value from waste. Science 2012, 337, 695–699. [Google Scholar] [CrossRef] [PubMed]
  2. He, M.; Sun, Y.; Han, B. Green carbon science: Scientific basis for integrating carbon resource processing, utilization, and recycling. Angew. Chem. Int. Ed. 2013, 52, 9620–9633. [Google Scholar] [CrossRef] [PubMed]
  3. 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. [Google Scholar] [CrossRef] [PubMed]
  4. Artz, J.; Müller, T.E.; Thenert, K.; Kleinekorte, J.; Meys, R.; Sternberg, A.; Bardow, A.; Leitner, W. Sustainable conversion of carbon dioxide: An integrated review of catalysis and life cycle assessment. Chem. Rev. 2018, 118, 434–504. [Google Scholar] [CrossRef] [PubMed]
  5. Huang, K.; Sun, C.-L.; Shi, Z.-J. Transition-metal-catalyzed C–C bond formation through the fixation of carbon dioxide. Chem. Soc. Rev. 2011, 40, 2435–2452. [Google Scholar] [CrossRef] [PubMed]
  6. Yang, Z.-Z.; He, L.-N.; Gao, J.; Liu, A.-H.; Yu, B. Carbon dioxide utilization with C–N bond formation: Carbon dioxide capture and subsequent conversion. Energy Environ. Sci. 2012, 5, 6602–6639. [Google Scholar] [CrossRef]
  7. Cokoja, M.; Bruckmeier, C.; Rieger, B.; Herrmann, W.A.; Kuhn, F.E. Transformation of carbon dioxide with homogeneous transition-metal catalysts: A molecular solution to a global challenge? Angew. Chem. Int. Ed. 2011, 50, 8510–8537. [Google Scholar] [CrossRef]
  8. Sakakura, T.; Choi, J.-C.; Yasuda, H. Transformation of carbon dioxide. Chem. Rev. 2007, 107, 2365–2387. [Google Scholar] [CrossRef]
  9. Liu, Q.; Wu, L.; Jackstell, R.; Beller, M. Using carbon dioxide as a building block in organic synthesis. Nat. Commun. 2015, 6, 5933–5948. [Google Scholar] [CrossRef]
  10. Klankermayer, J.; Wesselbaum, S.; Beydoun, K.; Leitner, W. Selective catalytic synthesis using the combination of carbon dioxide and hydrogen: Catalytic chess at the interface of energy and chemistry. Angew. Chem. Int. Ed. 2016, 55, 7296–7343. [Google Scholar] [CrossRef]
  11. Song, Q.-W.; Zhou, Z.-H.; He, L.-N. Efficient, selective and sustainable catalysis of carbon dioxide. Green Chem. 2017, 19, 3707–3728. [Google Scholar] [CrossRef]
  12. Song, Q.-W.; Zhao, Y.-N.; He, L.-N.; Gao, J.; Yang, Z.-Z. Synthesis of oxazolidinones/polyurethanes from aziridines and CO2. Curr. Catal. 2012, 1, 107–124. [Google Scholar] [CrossRef]
  13. Liu, X.-F.; Wang, M.-Y.; He, L.-N. Heterogeneous catalysis for oxazolidinone synthesis from aziridines and CO2. Curr. Org. Chem. 2017, 21, 698–707. [Google Scholar] [CrossRef]
  14. Pulla, S.; Felton, C.M.; Ramidi, P.; Gartia, Y.; Ali, N.; Nasini, U.B.; Ghosh, A. Advancements in oxazolidinone synthesis utilizing carbon dioxide as a C1 source. J. CO2 Util. 2013, 2, 49–57. [Google Scholar] [CrossRef]
  15. Tamura, M.; Honda, M.; Nakagawa, Y.; Tomishige, K. Direct conversion of CO2 with diols, aminoalcohols and diamines to cyclic carbonates, cyclic carbamates and cyclic ureas using heterogeneous catalysts. J. Chem. Technol. Biotechnol. 2014, 89, 19–33. [Google Scholar] [CrossRef]
  16. Wang, H.; Xin, Z.; Li, Y. Synthesis of ureas from CO2. Top Curr. Chem. 2017, 375, 1–26. [Google Scholar]
  17. Song, Q.-W.; He, L.-N. Heterocyclic synthesis through C-N bond formation with carbon dioxide in chemistry beyond chlorine. In Chemistry Beyond Chlorine; Springer International Publishing: Basel, Switzerland, 2016; Chapter 16; pp. 435–453. [Google Scholar]
  18. Liu, H.; Hua, R. Conversion of carbon dioxide into 2-oxazolidinones and 2(3H)-oxazolones catalyzed by 2,2′,2′′-terpyridine. Tetrahedron 2016, 72, 1200–1204. [Google Scholar] [CrossRef]
  19. Arshadi, S.; Vessally, E.; Hosseinian, A.; Soleimani-amiri, S.; Edjlali, L. Three-component coupling of CO2, propargyl alcohols, and amines: An environmentally benign access to cyclic and acyclic carbamates. J. CO2 Util. 2017, 21, 108–118. [Google Scholar] [CrossRef]
  20. Gu, Y.; Zhang, Q.; Duan, Z.; Zhang, J.; Zhang, S.; Deng, Y. Ionic liquid as an efficient promoting medium for fixation of carbon dioxide:  A clean method for the synthesis of 5-methylene-1,3-oxazolidin-2-ones from propargylic alcohols, amines, and carbon dioxide catalyzed by Cu(I) under mild conditions. J. Org. Chem. 2005, 70, 7376–7380. [Google Scholar] [CrossRef]
  21. Song, Q.-W.; Yu, B.; Li, X.-D.; Ma, R.; Diao, Z.-F.; Li, R.-G.; Li, W.; He, L.-N. Efficient chemical fixation of CO2 promoted by a bifunctional Ag2WO4/Ph3P system. Green Chem. 2014, 16, 1633–1638. [Google Scholar] [CrossRef]
  22. Xu, J.; Zhao, J.; Jia, Z.; Zhang, J. Facile and mild process for chemical fixation of CO2 to 4-methylene-1,3-oxazolidin-2-ones under solvent-free conditions. Synth. Commun. 2011, 41, 858–863. [Google Scholar] [CrossRef]
  23. Jiang, H.; Zhao, J.; Wang, A. An efficient and eco-friendly process for the conversion of carbon dioxide into oxazolones and oxazolidinones under supercritical conditions. Synthesis 2008, 2008, 763–769. [Google Scholar] [CrossRef]
  24. Ca, N.D.; Gabriele, B.; Ruffolo, G.; Veltri, L.; Zanetta, T.; Costa, M. Effective guanidine-catalyzed synthesis of carbonate and carbamate derivatives from propargyl alcohols in supercritical carbon dioxide. Adv. Synth. Catal. 2011, 353, 133–146. [Google Scholar] [CrossRef]
  25. Zhang, Q.; Shi, F.; Gu, Y.; Yang, J.; Deng, Y. Efficient and eco-friendly process for the synthesis of N-substituted 4-methylene-2-oxazolidinones in ionic liquids. Tetrahedron Lett. 2005, 46, 5907–5911. [Google Scholar] [CrossRef]
  26. Jiang, H.-F.; Zhao, J.-W. Silver-catalyzed activation of internal propargylic alcohols in supercritical carbon dioxide: Efficient and eco-friendly synthesis of 4-alkylidene-1,3-oxazolidin-2-ones. Tetrahedron Lett. 2009, 50, 60–62. [Google Scholar] [CrossRef]
  27. Zhang, G.; Yang, H.; Fei, H. Unusual missing linkers in an organosulfonate-based primitive-cubic (pcu)-type metal-organic framework for CO2 capture and conversion under ambient conditions. ACS Catal. 2018, 8, 2519–2525. [Google Scholar] [CrossRef]
  28. Qi, C.-R.; Jiang, H.-F. Efficient synthesis of β-oxopropylcarbamates in compressed CO2 without any additional catalyst and solvent. Green Chem. 2007, 9, 1284–1286. [Google Scholar] [CrossRef]
  29. Qi, C.; Huang, L.; Jiang, H. Efficient synthesis of β-oxoalkyl carbamates from carbon dioxide, internal propargylic alcohols, and secondary amines catalyzed by silver salts and DBU. Synthesis 2010, 1433–1440. [Google Scholar] [CrossRef]
  30. Bruneau, C.; Dixncuf, P.H. Catalytic synthesis of O-β-oxoalkylcarbamates. Tetrahedron Lett. 1987, 28, 2005–2008. [Google Scholar] [CrossRef]
  31. Kim, T.-J.; Kwon, K.-H.; Kwon, S.-C.; Baeg, J.-O.; Shim, S.-C.; Lee, D.-H. Iron complexes of 1,1’-bis(diphenylphosphino)ferrocene (BPPF) as efficient catalysts in the synthesis of carbamates. X-ray crystal structure of (BPPF)Fe(CO)3. J. Organomet. Chem. 1990, 389, 205–217. [Google Scholar] [CrossRef]
  32. Song, Q.-W.; Chen, W.-Q.; Ma, R.; Yu, A.; Li, Q.-Y.; Chang, Y.; He, L.-N. Bifunctional silver(I) complex-catalyzed CO2 conversion at ambient conditions: Synthesis of α-methylene cyclic carbonates and derivatives. ChemSusChem 2015, 8, 821–827. [Google Scholar] [CrossRef] [PubMed]
  33. Song, Q.-W.; Zhou, Z.-H.; Yin, H.; He, L.-N. Silver(I)-catalyzed synthesis of β-oxopropylcarbamates from propargylic alcohols and CO2 surrogate: A gas-free process. ChemSusChem 2015, 8, 3967–3972. [Google Scholar] [CrossRef] [PubMed]
  34. Song, Q.-W.; Liu, P.; Han, L.-H.; Zhang, K.; He, L.-N. Upgrading CO2 by incorporation into urethanes through silver-catalyzed one-pot stepwise amidation reaction. Chin. J. Chem. 2017, 36, 147–152. [Google Scholar] [CrossRef]
  35. Zhao, Q.-N.; Song, Q.-W.; Liu, P.; Zhang, K.; Hao, J. Ag(I)/(C2H5)4NCl cooperation catalysis for fixing CO2 or its derivatives into β-oxopropylcarbamates. Chem. Select 2018, 3, 6897–6901. [Google Scholar]
  36. Peshkov, V.A.; Pereshivko, O.P.; Nechaev, A.A.; Peshkov, A.A.; Van der Eycken, E.V. Reactions of secondary propargylamines with heteroallenes for the synthesis of diverse heterocycles. Chem. Soc. Rev. 2018, 47, 3861–3898. [Google Scholar] [CrossRef] [PubMed]
  37. Hu, J.; Ma, J.; Zhu, Q.; Zhang, Z.; Wu, C.; Han, B. Transformation of atmospheric CO2 catalyzed by protic ionic liquids: Efficient synthesis of 2-oxazolidinones. Angew. Chem. Int. Ed. 2015, 54, 5399–5403. [Google Scholar] [CrossRef] [PubMed]
  38. Hase, S.; Kayaki, Y.; Ikariya, T. NHC–gold(I) complexes as effective catalysts for the carboxylative cyclization of propargylamines with carbon dioxide. Organometallics 2013, 32, 5285–5288. [Google Scholar] [CrossRef]
  39. Hase, S.; Kayaki, Y.; Ikariya, T. Mechanistic aspects of the carboxylative cyclization of propargylamines and carbon dioxide catalyzed by gold(I) complexes bearing an N-heterocyclic carbene ligand. ACS Catal. 2015, 5, 5135–5140. [Google Scholar] [CrossRef]
  40. Fujita, K.-i.; Sato, J.; Inoue, K.; Tsuchimoto, T.; Yasuda, H. Aqueous media carboxylative cyclization of propargylic amines with CO2 catalyzed by amphiphilic dendritic N-heterocyclic carbene-gold(I) complexes. Tetrahedron Lett. 2014, 55, 3013–3016. [Google Scholar] [CrossRef]
  41. Fujita, K.-i.; Inoue, K.; Sato, J.; Tsuchimoto, T.; Yasuda, H. Carboxylative cyclization of propargylic amines with CO2 catalyzed by dendritic N-heterocyclic carbene-gold(I) complexes. Tetrahedron 2016, 72, 1205–1212. [Google Scholar] [CrossRef]
  42. Sadeghzadeh, S.M. Gold (III) phosphorus complex immobilized on fibrous nano-silica as a catalyst for the cyclization of propargylic amines with CO2. J. Mol. Catal. A: Chem. 2016, 423, 216–223. [Google Scholar] [CrossRef]
  43. Sadeghzadeh, S.M. A green approach for the synthesis of 2-oxazolidinones using gold(I) complex immobilized on KCC-1 as nanocatalyst at room temperature. Appl. Organomet. Chem. 2016, 30, 835–842. [Google Scholar] [CrossRef]
  44. Saadati, S.M.; Sadeghzadeh, S.M. KCC-1 supported ruthenium-salen-bridged ionic networks as a reusable catalyst for the cycloaddition of propargylic amines and CO2. Catal. Lett. 2018, 148, 1692–1702. [Google Scholar] [CrossRef]
  45. Inagaki, F.; Maeda, K.; Nakazawa, K.; Mukai, C. Construction of the oxazolidinone framework from propargylamine and CO2 in air at ambient temperature: Catalytic effect of a gold complex featuring an L2/Z-Type Ligand. Eur. J. Org. Chem. 2018, 2018, 2972–2976. [Google Scholar] [CrossRef]
  46. Yoshida, M.; Mizuguchi, T.; Shishido, K. Synthesis of oxazolidinones by efficient fixation of atmospheric CO2 with propargylic amines by using a silver/1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) dual-catalyst system. Chem. Eur. J. 2012, 18, 15578–15581. [Google Scholar] [CrossRef] [PubMed]
  47. Zhao, Y.; Qiu, J.; Tian, L.; Li, Z.; Fan, M.; Wang, J. New copper(I)/DBU catalyst system for the carboxylative cyclization of propargylic amines with atmospheric CO2: An experimental and theoretical study. ACS Sustainable Chem. Eng. 2016, 4, 5553–5560. [Google Scholar] [CrossRef]
  48. Zhao, Y.; Tian, L.; Qiu, J.; Li, Z.; Wang, H.; Cui, G.; Zhang, S.; Wang, J. Remarkable synergistic effect between copper(I) and ionic liquids for promoting chemical fixation of CO2. J. CO2 Util. 2017, 22, 374–381. [Google Scholar] [CrossRef]
  49. Wang, M.-Y.; Song, Q.-W.; Ma, R.; Xie, J.-N.; He, L.-N. Efficient conversion of carbon dioxide at atmospheric pressure to 2-oxazolidinones promoted by bifunctional Cu(ii)-substituted polyoxometalate-based ionic liquids. Green Chem. 2016, 18, 282–287. [Google Scholar] [CrossRef]
  50. Liu, X.; Wang, M.-Y.; Wang, S.-Y.; Wang, Q.; He, L.-N. In situ generated zinc(II) catalyst for incorporation of CO2 into 2-oxazolidinones with propargylic amines at atmospheric pressure. ChemSusChem 2016, 10, 1210–1216. [Google Scholar] [CrossRef]
  51. Brunel, P.; Monot, J.; Kefalidis, C.E.; Maron, L.; Martin-Vaca, B.; Bourissou, D. Valorization of CO2: Preparation of 2-oxazolidinones by metal-ligand cooperative catalysis with SCS indenediide Pd complexes. ACS Catal. 2017, 7, 2652–2660. [Google Scholar] [CrossRef]
  52. Kamata, K.; Kimura, T.; Sunaba, H.; Mizuno, N. Scope of chemical fixation of carbon dioxide catalyzed by a bifunctional monomeric tungstate. Catal. Today 2014, 226, 160–166. [Google Scholar] [CrossRef]
  53. Hu, J.; Ma, J.; Zhang, Z.; Zhu, Q.; Zhou, H.; Lu, W.; Han, B. A route to convert CO2: Synthesis of 3,4,5-trisubstituted oxazolones. Green Chem. 2015, 17, 1219–1225. [Google Scholar] [CrossRef]
  54. Fujita, K.-i.; Fujii, A.; Sato, J.; Onozawa, S.-y.; Yasuda, H. Synthesis of 2-oxazolidinone by N-heterocyclic carbene-catalyzed carboxylative cyclization of propargylic amine with CO2. Tetrahedron Lett. 2016, 57, 1282–1284. [Google Scholar] [CrossRef]
  55. Fujii, A.; Choi, J.-C.; Fujita, K.-i. Quaternary ammonium salt-catalyzed carboxylative cyclization of propargylic amines with CO2. Tetrahedron Lett. 2017, 58, 4483–4486. [Google Scholar] [CrossRef]
  56. Fujii, A.; Matsuo, H.; Choi, J.-C.; Fujitani, T.; Fujita, K.-i. Efficient synthesis of 2-oxazolidinones and quinazoline-2,4(1H,3H)-diones from CO2 catalyzed by tetrabutylammonium fluoride. Tetrahedron 2018, 74, 2914–2920. [Google Scholar] [CrossRef]
  57. Takeda, Y.; Okumura, S.; Tone, S.; Sasaki, I.; Minakata, S. Cyclizative atmospheric CO2 fixation by unsaturated amines with t-BuOI leading to cyclic carbamates. Org. Lett. 2012, 14, 4874–4877. [Google Scholar] [CrossRef] [PubMed]
  58. Yu, B.; Kim, D.; Kim, S.; Hong, S.H. Cyanuric acid-based organocatalyst for utilization of carbon dioxide at atmospheric pressure. ChemSusChem 2017, 10, 1080–1084. [Google Scholar] [CrossRef]
  59. Zhao, Y.; Qiu, J.; Li, Z.; Wang, H.; Fan, M.; Wang, J. An experimental and theoretical study on the unexpected catalytic activity of triethanolamine for the carboxylative cyclization of propargylic amines with CO2. ChemSusChem 2017, 10, 2001–2007. [Google Scholar] [CrossRef]
  60. Zhang, Z.; Ye, J.-H.; Wu, D.-S.; Zhou, Y.-Q.; Yu, D.-G. Synthesis of oxazolidin-2-ones from unsaturated amines with CO2 by using homogeneous catalysis. Chem. Asian J. 2018, 13, 2292–2306. [Google Scholar] [CrossRef]
  61. Farshbaf, S.; Fekri, L.Z.; Nikpassand, M.; Mohammadi, R.; Vessally, E. Dehydrative condensation of β-aminoalcohols with CO2: An environmentally benign access to 2-oxazolidinone derivatives. J. CO2 Util. 2018, 25, 194–204. [Google Scholar] [CrossRef]
  62. Fujita, S.-i.; Kanamaru, H.; Senboku, H.; Arai, M. Preparation of cyclic urethanes from amino alcohols and carbon dioxide using ionic liquid catalysts with alkali metal promoters. Int. J. Mol. Sci. 2006, 7, 438. [Google Scholar] [CrossRef]
  63. Pulla, S.; Felton, C.M.; Gartia, Y.; Ramidi, P.; Ghosh, A. Synthesis of 2-oxazolidinones by direct condensation of 2-aminoalcohols with carbon dioxide using chlorostannoxanes. ACS Sustainable Chem. Eng. 2013, 1, 309–312. [Google Scholar] [CrossRef]
  64. Tamura, M.; Honda, M.; Noro, K.; Nakagawa, Y.; Tomishige, K. Heterogeneous CeO2-catalyzed selective synthesis of cyclic carbamates from CO2 and aminoalcohols in acetonitrile solvent. J. Catal. 2013, 305, 191–203. [Google Scholar] [CrossRef]
  65. Juárez, R.; Concepción, P.; Corma, A.; García, H. Ceria nanoparticles as heterogeneous catalyst for CO2 fixation by ω-aminoalcohols. Chem. Commun. 2010, 46, 4181–4183. [Google Scholar] [CrossRef] [PubMed]
  66. Li, X.-D.; Cao, Y.; Ma, R.; He, L.-N. Thermodynamically favorable protocol for the synthesis of 2-oxazolidinones via Cu(I)-catalyzed three-component reaction of propargylic alcohols, CO2 and 2-aminoethanols. J. CO2 Util. 2018, 25, 338–345. [Google Scholar] [CrossRef]
  67. Niemi, T.; Fernandez, I.; Steadman, B.; Mannisto, J.K.; Repo, T. Carbon dioxide-based facile synthesis of cyclic carbamates from amino alcohols. Chem. Commun. 2018, 54, 3166–3169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Tambe, P.R.; Yadav, G.D. K-La-MgO as heterogeneous catalyst for synthesis of 3-(2-hydroxyethyl)-1,3-oxazolidin-2-one from diethanol amine and carbon dioxide. Clean Technol. Envir. 2018, 20, 1875–1888. [Google Scholar]
  69. Song, Q.-W.; Zhou, Z.-H.; Wang, M.-Y.; Zhang, K.; Liu, P.; Xun, J.-Y.; He, L.-N. Thermodynamically favorable synthesis of 2-oxazolidinones through silver-catalyzed reaction of propargylic alcohols, CO2, and 2-aminoethanols. ChemSusChem 2016, 9, 2054–2058. [Google Scholar] [CrossRef] [PubMed]
  70. Li, X.D.; Song, Q.W.; Lang, X.D.; Chang, Y.; He, L.N. Ag(I) /TMG-promoted cascade reaction of propargyl alcohols, carbon dioxide, and 2-aminoethanols to 2-oxazolidinones. Chemphyschem 2017, 18, 3182–3188. [Google Scholar] [CrossRef] [PubMed]
  71. Xia, S.; Song, Y.; Li, X.; Li, H.; He, L.N. Ionic liquid-promoted three-component domino reaction of propargyl alcohols, carbon dioxide and 2-aminoethanols: A thermodynamically favorable synthesis of 2-oxazolidinones. Molecules 2018, 23, 3033. [Google Scholar] [CrossRef]
  72. Vessally, E.; Soleimani-Amiri, S.; Hosseinian, A.; Edjlali, L.; Babazadeh, M. Chemical fixation of CO2 to 2-aminobenzonitriles: A straightforward route to quinazoline-2,4(1H,3H)-diones with green and sustainable chemistry perspectives. J. CO2 Util. 2017, 21, 342–352. [Google Scholar] [CrossRef]
  73. Zhao, Y.; Yu, B.; Yang, Z.; Zhang, H.; Hao, L.; Gao, X.; Liu, Z. A protic ionic liquid catalyzes CO2 conversion at atmospheric pressure and room temperature: Synthesis of quinazoline-2,4(1H,3H)-diones. Angew. Chem. Int. Ed. 2014, 53, 5922–5925. [Google Scholar] [CrossRef] [PubMed]
  74. Lang, X.-D.; Zhang, S.; Song, Q.-W.; He, L.-N. Tetra-butylphosphonium arginine-based ionic liquid-promoted cyclization of 2-aminobenzonitrile with carbon dioxide. RSC Adv. 2015, 5, 15668–15673. [Google Scholar] [CrossRef]
  75. Li, W.; Yang, N.; Lyu, Y. A mechanistic study on guanidine-catalyzed chemical fixation of CO2 with 2-aminobenzonitrile to quinazoline-2,4(1H,3H)-dione. Org. Chem. Front. 2016, 3, 823–835. [Google Scholar] [CrossRef]
  76. Nale, D.B.; Saigaonkar, S.D.; Bhanage, B.M. An efficient synthesis of quinazoline-2,4(1H,3H)-dione from CO2 and 2-aminobenzonitrile using [Hmim]OH/SiO2 as a base functionalized supported ionic liquid phase catalyst. J. CO2 Util. 2014, 8, 67–73. [Google Scholar] [CrossRef]
  77. Zheng, H.; Cao, X.; Du, K.; Xu, J.; Zhang, P. A highly efficient way to capture CX2 (O, S) mildly in reusable ReILs at atmospheric pressure. Green Chem. 2014, 16, 3142–3148. [Google Scholar] [CrossRef]
  78. Lang, X.-D.; Yu, Y.-C.; Li, Z.-M.; He, L.-N. Protic ionic liquids-promoted efficient synthesis of quinazolines from 2-aminobenzonitriles and CO2 at ambient conditions. J. CO2 Util. 2016, 15, 115–122. [Google Scholar] [CrossRef]
  79. Patil, Y.P.; Tambade, P.J.; Deshmukh, K.M.; Bhanage, B.M. Synthesis of quinazoline-2,4(1H,3H)-diones from carbon dioxide and 2-aminobenzonitriles using [Bmim]OH as a homogeneous recyclable catalyst. Catal. Today 2009, 148, 355–360. [Google Scholar] [CrossRef]
  80. Hulla, M.; Chamam, S.M.A.; Laurenczy, G.; Das, S.; Dyson, P.J. Delineating the mechanism of ionic liquids in the synthesis of quinazoline-2,4(1H,3H)-dione from 2-aminobenzonitrile and CO2. Angew. Chem. Int. Ed. 2017, 56, 10559–10563. [Google Scholar] [CrossRef]
  81. Shi, G.; Chen, K.; Wang, Y.; Li, H.; Wang, C. Highly efficient synthesis of quinazoline-2,4(1H,3H)-diones from CO2 by hydroxyl functionalized aprotic ionic liquids. ACS Sustainable Chem. Eng. 2018, 6, 5760–5765. [Google Scholar] [CrossRef]
  82. Samanta, S.; Srivastava, R. A novel method to introduce acidic and basic bi-functional sites in graphitic carbon nitride for sustainable catalysis: Cycloaddition, esterification, and transesterification reactions. Sustainable Energy Fuels 2017, 1, 1390–1404. [Google Scholar] [CrossRef]
  83. Sarmah, B.; Srivastava, R. Activation and utilization of CO2 using ionic liquid or amine-functionalized basic nanocrystalline zeolites for the synthesis of cyclic carbonates and quinazoline-2,4(1H,3H)-dione. Ind. Eng. Chem. Res. 2017, 56, 8202–8215. [Google Scholar] [CrossRef]
  84. Wang, B.; Sun, S.; Yu, J.-T.; Jiang, Y.; Cheng, J. Palladium-catalyzed multicomponent reactions of O-alkynylanilines, aryl Iodides, and CO2 toward 3,3-diaryl 2,4-quinolinediones. Org. Lett. 2017, 19, 4319–4322. [Google Scholar] [CrossRef] [PubMed]
  85. Zhang, W.-Z.; Li, H.; Zeng, Y.; Tao, X.; Lu, X. Palladium-catalyzed cyclization reaction of O-haloanilines, CO2 and isocyanides: Access to quinazoline-2,4(1H,3H)-diones. Chin. J. Chem. 2018, 36, 112–118. [Google Scholar] [CrossRef]
  86. Adams, P.; Baron, F.A. Esters of carbamic acid. Chem. Rev. 1965, 65, 567–602. [Google Scholar] [CrossRef]
  87. Frezza, M.; Castang, S.; Estephane, J.; Soulère, L.; Deshayes, C.; Chantegrel, B.; Nasser, W.; Queneau, Y.; Reverchon, S.; Doutheau, A. Synthesis and biological evaluation of homoserine lactone derived ureas as antagonists of bacterial quorum sensing. Bioorg. Med. Chem. 2006, 14, 4781–4791. [Google Scholar] [CrossRef]
  88. Gogoi, P.; Konwar, D. An efficient modification of the Hofmann rearrangement: Synthesis of methyl carbamates. Tetrahedron Lett. 2007, 48, 531–533. [Google Scholar] [CrossRef]
  89. Lin, B.; Waymouth, R.M. Urea anions: Simple, fast, and selective catalysts for ring-opening polymerizations. J. Am. Chem. Soc. 2017, 139, 1645–1652. [Google Scholar] [CrossRef]
  90. Park, J.H.; Yoon, J.C.; Chung, Y.K. Cobalt/rhodium heterobimetallic nanoparticle-catalyzed oxidative carbonylation of amines in the presence of carbon monoxide and molecular oxygen to ureas. Adv. Synth. Catal. 2009, 351, 1233–1237. [Google Scholar] [CrossRef]
  91. Marchegiani, M.; Nodari, M.; Tansini, F.; Massera, C.; Mancuso, R.; Gabriele, B.; Costa, M.; Della Ca’, N. Urea derivatives from carbon dioxide and amines by guanidine catalysis: Easy access to imidazolidin-2-ones under solvent-free conditions. J. CO2 Util. 2017, 21, 553–561. [Google Scholar] [CrossRef]
  92. Tamura, M.; Ito, K.; Nakagawa, Y.; Tomishige, K. CeO2-catalyzed direct synthesis of dialkylureas from CO2 and amines. J. Catal. 2016, 343, 75–85. [Google Scholar] [CrossRef]
  93. Sun, D.-L.; Ye, J.-H.; Fang, Y.-X.; Chao, Z.-S. Green synthesis of N,N′-dialkylureas from CO2 and amines using metal salts of oxalates as catalysts. Ind. Eng. Chem. Res. 2016, 55, 64–70. [Google Scholar] [CrossRef]
  94. Dalei, S.; Kaihong, X.; Yanxiong, F.; Xianghua, Y. One-pot synthesis of N,N-dialkylureas via carbonylation of amines with CO2 applying Y0.08Zr0.92O1.96 mixed oxide (YSZ-8) as a heterogeneous catalyst. Catalysts 2018, 8, 188–198. [Google Scholar]
  95. Xu, M.; Jupp, A.R.; Stephan, D.W. Stoichiometric reactions of CO2 and indium-silylamides and catalytic synthesis of ureas. Angew. Chem. Int. Ed. 2017, 56, 14277–14281. [Google Scholar] [CrossRef] [PubMed]
  96. Wang, P.; Ma, X.; Li, Q.; Yang, B.; Shang, J.; Deng, Y. Green synthesis of polyureas from CO2 and diamines with a functional ionic liquid as the catalyst. RSC Adv. 2016, 6, 54013–54019. [Google Scholar] [CrossRef]
  97. Wang, P.; Fei, Y.; Deng, Y. Transformation of CO2 into polyureas with 3-amino-1,2,4-triazole potassium as a solid base catalyst. New J. Chem. 2018, 42, 1202–1207. [Google Scholar] [CrossRef]
  98. Wu, P.-X.; Cheng, H.-Y.; Shi, R.-H.; Jiang, S.; Wu, Q.-F.; Zhang, C.; Arai, M.; Zhao, F.-Y. Synthesis of polyurea via the addition of carbon dioxide to a diamine catalyzed by organic and inorganic bases. Adv. Synth. Catal. 2018. [Google Scholar] [CrossRef]
  99. Xiong, W.; Qi, C.; He, H.; Ouyang, L.; Zhang, M.; Jiang, H. Base-promoted coupling of carbon dioxide, amines, and N-tosylhydrazones: A novel and versatile approach to carbamates. Angew. Chem. Int. Ed. 2015, 54, 3084–3087. [Google Scholar] [CrossRef]
  100. Hong, J.Y.; Seo, U.R.; Chung, Y.K. Synthesis of carbamates from amines and N-tosylhydrazones under atmospheric pressure of carbon dioxide without an external base. Org. Chem. Front. 2016, 3, 764–767. [Google Scholar] [CrossRef]
  101. Sun, S.; Hu, W.-M.; Gu, N.; Cheng, J. Palladium-catalyzed multi-component reactions of N-tosylhydrazones, 2-iodoanilines and CO2 towards 4-aryl-2-quinolinones. Chem. Eur. J. 2016, 22, 18729–18732. [Google Scholar] [CrossRef]
  102. Xiong, W.; Qi, C.; Peng, Y.; Guo, T.; Zhang, M.; Jiang, H. Base-promoted coupling of carbon dioxide, amines, and diaryliodonium salts: A phosgene- and metal-free route to O-aryl carbamates. Chem. Eur. J. 2015, 21, 14314–14318. [Google Scholar] [CrossRef] [PubMed]
  103. Xiong, W.; Qi, C.; Guo, T.; Zhang, M.; Chen, K.; Jiang, H. A copper-catalyzed oxidative coupling reaction of arylboronic acids, amines and carbon dioxide using molecular oxygen as the oxidant. Green Chem. 2017, 19, 1642–1646. [Google Scholar] [CrossRef]
  104. Bernoud, E.; Company, A.; Ribas, X. Direct use of CO2 for O-arylcarbamate synthesis via mild Cu(II)-catalyzed aerobic C-H functionalization in pincer-like macrocyclic systems. J. Organomet. Chem. 2017, 845, 44–48. [Google Scholar] [CrossRef]
  105. Riemer, D.; Hirapara, P.; Das, S. Chemoselective synthesis of carbamates using CO2 as carbon source. ChemSusChem 2016, 9, 1916–1920. [Google Scholar] [CrossRef] [PubMed]
  106. Yuan, H.-Y.; Choi, J.-C.; Onozawa, S.-y.; Fukaya, N.; Choi, S.J.; Yasuda, H.; Sakakura, T. The direct synthesis of N-phenylcarbamates from CO2, aniline, and dibutyltin dialkoxide. J. CO2 Util. 2016, 16, 282–286. [Google Scholar] [CrossRef]
  107. Zhang, Q.; Yuan, H.-Y.; Fukaya, N.; Yasuda, H.; Choi, J.-C. A simple zinc catalyst for carbamate synthesis directly from CO2. ChemSusChem 2017, 10, 1501–1508. [Google Scholar] [CrossRef] [PubMed]
  108. Tamura, M.; Miura, A.; Honda, M.; Gu, Y.; Nakagawa, Y.; Tomishige, K. Direct catalytic synthesis of N-arylcarbamates from CO2, anilines and alcohols. ChemCatChem 2018, 10, 4821–4825. [Google Scholar] [CrossRef]
  109. Cai, J.; Zhang, M.; Zhao, X. Synthesis of allyl carbamates through the regioselective domino reaction of amines, CO2, and unsymmetrical allyl chlorides under Pd catalysis. Eur. J. Org. Chem. 2015, 2015, 5925–5928. [Google Scholar] [CrossRef]
  110. Qi, C.; Yan, D.; Xiong, W.; Jiang, H. Silver-catalyzed regioselective coupling of carbon dioxide, amines and aryloxyallenes leading to O-allyl carbamates. J. CO2 Util. 2018, 24, 120–127. [Google Scholar] [CrossRef]
  111. Tlili, A.; Frogneux, X.; Blondiaux, E.; Cantat, T. Creating added value with a waste: Methylation of amines with CO2 and H2. Angew. Chem. Int. Ed. 2014, 53, 2543–2545. [Google Scholar] [CrossRef]
  112. Tlili, A.; Blondiaux, E.; Frogneux, X.; Cantat, T. Reductive functionalization of CO2 with amines: An entry to formamide, formamidine and methylamine derivatives. Green Chem. 2015, 17, 157–168. [Google Scholar] [CrossRef]
  113. Yan, G.; Borah, A.J.; Wang, L.; Yang, M. Recent advances in transition metal-catalyzed methylation reactions. Adv. Synth. Catal. 2015, 357, 1333–1350. [Google Scholar] [CrossRef]
  114. Li, Y.; Cui, X.; Dong, K.; Junge, K.; Beller, M. Utilization of CO2 as a C1 building block for catalytic methylation reactions. ACS Catal. 2017, 7, 1077–1086. [Google Scholar] [CrossRef]
  115. Liu, X.-F.; Li, X.-Y.; Qiao, C.; He, L.-N. Transition-metal-free catalysis for reductive functionalization of CO2 with amines. Synlett 2018, 29, 548–555. [Google Scholar]
  116. Tappe, N.A.; Reich, R.M.; D’Elia, V.; Kühn, F.E. Current advances in the catalytic conversion of carbon dioxide by molecular catalysts: An update. Dalton Trans. 2018, 47, 13281–13313. [Google Scholar] [CrossRef] [PubMed]
  117. Fernandez-Alvarez, F.J.; Oro, L.A. Homogeneouscatalytic reduction of CO2 with silicon-hydrides, state of the art. ChemCatChem 2018, 10, 4783–4796. [Google Scholar] [CrossRef]
  118. Pramudita, R.A.; Motokura, K. Transformative reduction of carbon dioxide through organocatalysis with silanes. Green Chem. 2018, 20, 4834–4843. [Google Scholar] [CrossRef]
  119. Jacquet, O.; Frogneux, X.; Das Neves Gomes, C.; Cantat, T. CO2 as a C1-building block for the catalytic methylation of amines. Chem. Sci. 2013, 4, 2127–2131. [Google Scholar] [CrossRef]
  120. Frogneux, X.; Jacquet, O.; Cantat, T. Iron-catalyzed hydrosilylation of CO2: CO2 conversion to formamides and methylamines. Catal. Sci. Technol. 2014, 4, 1529–1533. [Google Scholar] [CrossRef]
  121. Gopakumar, A.; Akçok, I.; Lombardo, L.; Formal, F.L.; Magrez, A.; Sivula, K.; Dyson, P.J. Iron-rich natural mineral gibeon meteorite catalyzed N-formylation of amines using CO2 as the C1 source. ChemSelect 2018, 3, 10271–10276. [Google Scholar] [CrossRef]
  122. Das, S.; Bobbink, F.D.; Bulut, S.; Soudani, M.; Dyson, P.J. Thiazolium carbene catalysts for the fixation of CO2 onto amines. Chem. Commun. 2016, 52, 2497–2500. [Google Scholar] [CrossRef]
  123. Liu, X.-F.; Ma, R.; Qiao, C.; Cao, H.; He, L.-N. Fluoride-catalyzed methylation of amines by reductive functionalization of CO2 with hydrosilanes. Chem. Eur. J. 2016, 22, 16489–16493. [Google Scholar] [CrossRef] [PubMed]
  124. Fang, C.; Lu, C.; Liu, M.; Zhu, Y.; Fu, Y.; Lin, B.-L. Selective formylation and methylation of amines using carbon dioxide and hydrosilane catalyzed by alkali-metal carbonates. ACS Catal. 2016, 6, 7876–7881. [Google Scholar] [CrossRef]
  125. Liu, X.-F.; Li, X.-Y.; Qiao, C.; Fu, H.-C.; He, L.-N. Reductive functionalization of CO2 with amine to afford formamide, aminal, and methylamine. Angew. Chem. Int. Ed. 2017, 56, 7425–7429. [Google Scholar] [CrossRef]
  126. Xie, C.; Song, J.; Wu, H.; Zhou, B.; Wu, C.; Han, B. Natural product glycine betaine as an efficient catalyst for transformation of CO2 with amines to synthesize N-substituted compounds. ACS Sustainable Chem. Eng. 2017, 5, 7086–7092. [Google Scholar] [CrossRef]
  127. Hu, Y.; Song, J.; Xie, C.; Wu, H.; Wang, Z.; Jiang, T.; Wu, L.; Wang, Y.; Han, B. Renewable and biocompatible lecithin as an efficient organocatalyst for reductive conversion of CO2 with amines to formamides and methylamines. ACS Sustainable Chem. Eng. 2018, 6, 11228–11234. [Google Scholar] [CrossRef]
  128. Li, G.; Chen, J.; Zhu, D.-Y.; Chen, Y.; Xia, J.-B. DBU-catalyzed selective N-methylation and N-formylation of amines with CO2 and polymethylhydrosiloxane. Adv. Synth. Catal. 2018, 360, 2364–2369. [Google Scholar] [CrossRef]
  129. Wang, M.-Y.; Wang, N.; Liu, X.-F.; Qiao, C.; He, L.-N. Tungstate catalysis: Pressure-switched 2- and 6-electron reductive functionalization of CO2 with amines and phenylsilane. Green Chem. 2018, 20, 1564–1570. [Google Scholar] [CrossRef]
  130. Li, X.-D.; Xia, S.-M.; Chen, K.-H.; Liu, X.-F.; Li, H.-R.; He, L.-N. Copper catalysis: Ligand-controlled selective N-methylation or N-formylation of amines with CO2 and phenylsilane. Green Chem. 2018, 20, 4853–4858. [Google Scholar] [CrossRef]
  131. Nicholls, R.L.; McManus, J.A.; Rayner, C.M.; Morales-Serna, J.A.; White, A.J.P.; Nguyen, B.N. Guanidine-catalyzed reductive amination of carbon dioxide with silanes: Switching between pathways and suppressing catalyst deactivation. ACS Catal. 2018, 8, 3678–3687. [Google Scholar] [CrossRef]
  132. Frogneux, X.; Blondiaux, E.; Thuery, P.; Cantat, T. Bridging amines with CO2: Organocatalyzed reduction of CO2 to aminals. ACS Catal. 2015, 5, 3983–3987. [Google Scholar] [CrossRef]
  133. Lu, Y.; Gao, Z.-H.; Chen, X.-Y.; Guo, J.; Liu, Z.; Dang, Y.; Ye, S.; Wang, Z.-X. Formylation or methylation: What determines the chemoselectivity of the reaction of amine, CO2, and hydrosilane catalyzed by 1,3,2-diazaphospholene. Chem. Sci. 2017, 8, 7637–7650. [Google Scholar] [CrossRef] [PubMed]
  134. Gredig, S.V.; Koeppel, R.A.; Baiker, A. Synthesis of methylamines from carbon dioxide and ammonia. J. Chem. Soc. Chem. Commun. 1995, 0, 73–74. [Google Scholar] [CrossRef]
  135. Gredig, S.V.; Koeppel, R.; Baiker, A. Comparative study of synthesis of methylamines from carbon oxides and ammonia over Cu/A12O3. Catal. Today 1996, 29, 339–342. [Google Scholar] [CrossRef]
  136. Gredig, S.V.; Koeppel, R.; Baiker, A. Synthesis of methylamines from CO2, H2 and NH3. Catalytic behaviour of various metal-alumina catalysts. Appl. Catal. A 1997, 162, 249–260. [Google Scholar] [CrossRef]
  137. Auer, S.M.; Gredig, S.V.; Köppel, R.A.; Baiker, A. Synthesis of methylamines from CO2, H2 and NH3 over Cu–Mg–Al mixed oxides. J. Mol. Catal. A: Chem. 1999, 141, 193–203. [Google Scholar] [CrossRef]
  138. Cui, X.; Dai, X.; Zhang, Y.; Deng, Y.; Shi, F. Methylation of amines, nitrobenzenes and aromatic nitriles with carbon dioxide and molecular hydrogen. Chem. Sci. 2014, 5, 649–655. [Google Scholar] [CrossRef]
  139. Tamura, M.; Miura, A.; Gu, Y.; Nakagawa, Y.; Tomishige, K. Selective N-methylation of aniline to N-methylaniline with CO2 and H2 by CeO2-supported Cu sub-nanoparticle catalyst. Chem. Lett. 2017, 46, 1243–1246. [Google Scholar] [CrossRef]
  140. Cui, X.; Zhang, Y.; Deng, Y.; Shi, F. N-Methylation of amine and nitro compounds with CO2/H2 catalyzed by Pd/CuZrOx under mild reaction conditions. Chem. Commun. 2014, 50, 13521–13524. [Google Scholar] [CrossRef]
  141. Su, X.; Lin, W.; Cheng, H.; Zhang, C.; Li, Y.; Liu, T.; Zhang, B.; Wu, Q.; Yu, X.; Zhao, F. PdGa/TiO2 an efficient heterogeneous catalyst for direct methylation of N-methylaniline with CO2/H2. RSC Adv. 2016, 6, 103650–103656. [Google Scholar] [CrossRef]
  142. Kon, K.; Siddiki, S.M.A.H.; Onodera, W.; Shimizu, K. Sustainable heterogeneous platinum catalyst for direct methylation of secondary amines by carbon dioxide and hydrogen. Chem. Eur. J. 2014, 20, 6264–6267. [Google Scholar] [CrossRef] [PubMed]
  143. Toyao, T.; Siddiki, S.M.A.H.; Ishihara, K.; Kon, K.; Onodera, W.; Shimizu, K.-I. Heterogeneous platinum catalysts for direct synthesis of trimethylamine by N-methylation of ammonia and its surrogates with CO2/H2. Chem. Lett. 2017, 46, 68–70. [Google Scholar] [CrossRef]
  144. Du, X.L.; Tang, G.; Bao, H.L.; Jiang, Z.; Zhong, X.H.; Su, D.S.; Wang, J.Q. Direct methylation of amines with carbon dioxide and molecular hydrogen using supported gold catalysts. ChemSusChem 2015, 8, 3489–3496. [Google Scholar] [CrossRef] [PubMed]
  145. Tang, G.; Bao, H.L.; Jin, C.; Zhong, X.H.; Du, X.L. Direct methylation of N-methylaniline with CO2/H2 catalyzed by gold nanoparticles supported on alumina. RSC Adv. 2015, 5, 99678–99687. [Google Scholar] [CrossRef]
  146. Toyao, T.; Siddiki, S.M.A.H.; Morita, Y.; Kamachi, T.; Touchy, A.S.; Onodera, W.; Kon, K.; Furukawa, S.; Ariga, H.; Asakura, K.; et al. Rhenium-loaded TiO2: A highly versatile and chemoselective catalyst for the hydrogenation of carboxylic acid derivatives and the N-methylation of amines using H2 and CO2. Chem. Eur. J. 2017, 23, 14848–14859. [Google Scholar] [CrossRef] [PubMed]
  147. Zhang, Y.; Wang, H.; Yuan, H.; Shi, F. Hydroxyl group-regulated active nano-Pd/C catalyst generation via in situ reduction of Pd(NH3)xCly/C for N-formylation of amines with CO2/H2. ACS Sustainable Chem. Eng. 2017, 5, 5758–5765. [Google Scholar] [CrossRef]
  148. Ju, P.; Chen, J.; Chen, A.; Chen, L.; Yu, Y. N-formylation of amines with CO2 and H2 using Pd−Au bimetallic catalysts supported on polyaniline-functionalized carbon nanotubes. ACS Sustainable Chem. Eng. 2017, 5, 2516–2528. [Google Scholar] [CrossRef]
  149. Luo, X.; Zhang, H.; Ke, Z.; Wu, C.; Guo, S.; Wu, Y.; Yu, B.; Liu, Z. N-doped carbon supported Pd catalysts for N-formylation of amines with CO2/H2. Sci. China Chem. 2018, 61, 725–731. [Google Scholar] [CrossRef]
Scheme 1. CO2 conversion through C-N bond formation reactions described in this work.
Scheme 1. CO2 conversion through C-N bond formation reactions described in this work.
Molecules 24 00182 sch001
Scheme 2. The mechanism for three-component reaction of propargylic alcohols, amines and CO2.
Scheme 2. The mechanism for three-component reaction of propargylic alcohols, amines and CO2.
Molecules 24 00182 sch002
Scheme 3. Catalytic systems for the reactions of CO2, propargylic alcohols and primary amines.
Scheme 3. Catalytic systems for the reactions of CO2, propargylic alcohols and primary amines.
Molecules 24 00182 sch003
Figure 1. Some 2-oxazolidinone derivatives by Ag2WO4/Ph3P system (isolated yields are given) [21].
Figure 1. Some 2-oxazolidinone derivatives by Ag2WO4/Ph3P system (isolated yields are given) [21].
Molecules 24 00182 g001
Figure 2. 1H-NMR (a,b) and 13C-NMR (cf) investigation. (a,b) 2-Methyl-4-phenylbut-3-yn-2-ol (8.0 mg), Ag2CO3 (2.8 mg) and Ph3P (10.5 mg) ([D6] DMSO 0.6 mL). (c,d) 2-methylbut-3-yn-2-ol (20.2 mg), AgNO3 (40.8 mg) (CDCl3 0.6 mL). (e) [(Ph3P)2Ag]2CO3 (158.6 mg) in 0.6 mL of CDCl3. (f) [(Ph3P)2Ag]2CO3 (158.6 mg) in 0.6 mL of CDCl3 in the presence of 13CO2 (0.1 MPa). Reprinted with permission from [32] (Copyright 2014 John Wiley and Sons).
Figure 2. 1H-NMR (a,b) and 13C-NMR (cf) investigation. (a,b) 2-Methyl-4-phenylbut-3-yn-2-ol (8.0 mg), Ag2CO3 (2.8 mg) and Ph3P (10.5 mg) ([D6] DMSO 0.6 mL). (c,d) 2-methylbut-3-yn-2-ol (20.2 mg), AgNO3 (40.8 mg) (CDCl3 0.6 mL). (e) [(Ph3P)2Ag]2CO3 (158.6 mg) in 0.6 mL of CDCl3. (f) [(Ph3P)2Ag]2CO3 (158.6 mg) in 0.6 mL of CDCl3 in the presence of 13CO2 (0.1 MPa). Reprinted with permission from [32] (Copyright 2014 John Wiley and Sons).
Molecules 24 00182 g002
Figure 3. The scope of the [(PPh3)2Ag]2CO3 reaction system (isolated yields are given). Reprinted with permission from [32] (Copyright 2014 John Wiley and Sons).
Figure 3. The scope of the [(PPh3)2Ag]2CO3 reaction system (isolated yields are given). Reprinted with permission from [32] (Copyright 2014 John Wiley and Sons).
Molecules 24 00182 g003
Scheme 4. A plausible mechanism for propargylic alcohols with ammonium carbamates [33].
Scheme 4. A plausible mechanism for propargylic alcohols with ammonium carbamates [33].
Molecules 24 00182 sch004
Scheme 5. The mechanism for carboxylative cyclization of propargylic amines with CO2.
Scheme 5. The mechanism for carboxylative cyclization of propargylic amines with CO2.
Molecules 24 00182 sch005
Scheme 6. A summary of coinage metals catalysts in carboxylative cyclization of propargylic amines with CO2.
Scheme 6. A summary of coinage metals catalysts in carboxylative cyclization of propargylic amines with CO2.
Molecules 24 00182 sch006
Scheme 7. Synthesis of 2-oxazolidinones via Zn- and Pd-catalyzed reaction.
Scheme 7. Synthesis of 2-oxazolidinones via Zn- and Pd-catalyzed reaction.
Molecules 24 00182 sch007
Scheme 8. Synthesis of 2-oxazolidinones by metal-free catalytic systems.
Scheme 8. Synthesis of 2-oxazolidinones by metal-free catalytic systems.
Molecules 24 00182 sch008
Scheme 9. Synthesis of 2-oxazolidinones through carbonylation of amino alcohols with CO2.
Scheme 9. Synthesis of 2-oxazolidinones through carbonylation of amino alcohols with CO2.
Molecules 24 00182 sch009
Scheme 10. Synthesis of 2-oxazolidinones through thermodynamically favourable approach [68].
Scheme 10. Synthesis of 2-oxazolidinones through thermodynamically favourable approach [68].
Molecules 24 00182 sch010
Scheme 11. Possible reaction pathway of 2-aminobenzonitrile and CO2.
Scheme 11. Possible reaction pathway of 2-aminobenzonitrile and CO2.
Molecules 24 00182 sch011
Scheme 12. IL-catalyzed transformation of 2-aminobenzonitrile and CO2.
Scheme 12. IL-catalyzed transformation of 2-aminobenzonitrile and CO2.
Molecules 24 00182 sch012
Figure 4. 1H-NMR spectra of (a) 2-aminobenzonitrile with (b) [HMIm][Im] (pKa value of cation was 7.1), (c) [HDBU][Im] (pKa value of cation was 11.7) or (d) [HMTBD][Im] (pKa value of cation was 13.0). The H signal of amino in 2-aminobenzonitrile became broader when the basicity of cation was higher. Reprinted with permission from [81] (Copyright 2018 American Chemical Society).
Figure 4. 1H-NMR spectra of (a) 2-aminobenzonitrile with (b) [HMIm][Im] (pKa value of cation was 7.1), (c) [HDBU][Im] (pKa value of cation was 11.7) or (d) [HMTBD][Im] (pKa value of cation was 13.0). The H signal of amino in 2-aminobenzonitrile became broader when the basicity of cation was higher. Reprinted with permission from [81] (Copyright 2018 American Chemical Society).
Molecules 24 00182 g004
Scheme 13. Synthesis of quinazoline-2,4(1H,3H)-dione by heterogeneous catalysts.
Scheme 13. Synthesis of quinazoline-2,4(1H,3H)-dione by heterogeneous catalysts.
Molecules 24 00182 sch013
Scheme 14. Multicomponent reactions including CO2 to synthesize quinazoline-2,4(1H,3H)-diones.
Scheme 14. Multicomponent reactions including CO2 to synthesize quinazoline-2,4(1H,3H)-diones.
Molecules 24 00182 sch014
Scheme 15. Bicyclic guanidine-catalyzed urea formation.
Scheme 15. Bicyclic guanidine-catalyzed urea formation.
Molecules 24 00182 sch015
Figure 5. Various ureas by CeO2/NMP system. Reprinted with permission from [92] (Copyright 2015 Elsevier Inc.).
Figure 5. Various ureas by CeO2/NMP system. Reprinted with permission from [92] (Copyright 2015 Elsevier Inc.).
Molecules 24 00182 g005
Scheme 16. Synthesis of polyureas via the polymerization of CO2 with diamines.
Scheme 16. Synthesis of polyureas via the polymerization of CO2 with diamines.
Molecules 24 00182 sch016
Figure 6. (1) FTIR spectra of (a) TOTDDA, (b) DBU, and (c) mixture of TOTDDA and DBU (TOTDDA:DBU was 10:1). The protonation of DBU by TOTDDA indicated that DBU could activate the -NH2 group of TOTDDA. (2) In situ ATR-FTIR spectra of DBU (C=N) and DBU·CO2 (C=O) collected at different periods of reaction time. The new band at 1644 cm−1 is assignable to C=O and this implies the formation of the complex of DBU·CO2. Reaction conditions: DBU 2 mmol, CO2 8 MPa, 60 °C. Reprinted with permission from [98] (Copyright 2018 John Wiley and Sons).
Figure 6. (1) FTIR spectra of (a) TOTDDA, (b) DBU, and (c) mixture of TOTDDA and DBU (TOTDDA:DBU was 10:1). The protonation of DBU by TOTDDA indicated that DBU could activate the -NH2 group of TOTDDA. (2) In situ ATR-FTIR spectra of DBU (C=N) and DBU·CO2 (C=O) collected at different periods of reaction time. The new band at 1644 cm−1 is assignable to C=O and this implies the formation of the complex of DBU·CO2. Reaction conditions: DBU 2 mmol, CO2 8 MPa, 60 °C. Reprinted with permission from [98] (Copyright 2018 John Wiley and Sons).
Molecules 24 00182 g006
Scheme 17. N-tosylhydrazones as a building block to construct C-N bonds.
Scheme 17. N-tosylhydrazones as a building block to construct C-N bonds.
Molecules 24 00182 sch017
Scheme 18. Synthesis of O-aryl carbamates by three-component reactions containing amines and CO2.
Scheme 18. Synthesis of O-aryl carbamates by three-component reactions containing amines and CO2.
Molecules 24 00182 sch018
Scheme 19. Syntheses of N-arylcarbamates via various strategies.
Scheme 19. Syntheses of N-arylcarbamates via various strategies.
Molecules 24 00182 sch019
Figure 7. (a) 1H- and (b) 15N-NMR spectra of aniline with and without Zn(OAc)2/phen in CD3CN at room temperature. The chemical shifts in both 1H and 15N-NMR spectra illustrated the interaction between aniline and acetate, and the formation of hydrogen bonds. Reprinted with permission from ref. [107] (Copyright 2017 John Wiley and Sons).
Figure 7. (a) 1H- and (b) 15N-NMR spectra of aniline with and without Zn(OAc)2/phen in CD3CN at room temperature. The chemical shifts in both 1H and 15N-NMR spectra illustrated the interaction between aniline and acetate, and the formation of hydrogen bonds. Reprinted with permission from ref. [107] (Copyright 2017 John Wiley and Sons).
Molecules 24 00182 g007
Scheme 20. Synthesis of allyl carbamates from amines and CO2.
Scheme 20. Synthesis of allyl carbamates from amines and CO2.
Molecules 24 00182 sch020
Scheme 21. Zinc complex-catalyzed methylation of amines using hydrosilanes.
Scheme 21. Zinc complex-catalyzed methylation of amines using hydrosilanes.
Molecules 24 00182 sch021
Scheme 22. Fe-based catalysis for temperature controlled N-formylation/N-methylation.
Scheme 22. Fe-based catalysis for temperature controlled N-formylation/N-methylation.
Molecules 24 00182 sch022
Scheme 23. Thiazolium carbene-based catalyst for N-formylation/N-methylation.
Scheme 23. Thiazolium carbene-based catalyst for N-formylation/N-methylation.
Molecules 24 00182 sch023
Scheme 24. TBAF catalysis for selective N-formylation/N-methylation.
Scheme 24. TBAF catalysis for selective N-formylation/N-methylation.
Molecules 24 00182 sch024
Scheme 25. Cs2CO3 catalysis for selective N-formylation/N-methylation and proposed mechanism.
Scheme 25. Cs2CO3 catalysis for selective N-formylation/N-methylation and proposed mechanism.
Molecules 24 00182 sch025
Scheme 26. Glycine betaine-catalyzed selective N-formylation/N-methylation.
Scheme 26. Glycine betaine-catalyzed selective N-formylation/N-methylation.
Molecules 24 00182 sch026
Scheme 27. DBU-catalyzed selective N-methylation/N-formylation using PMHS as hydride reagent [128].
Scheme 27. DBU-catalyzed selective N-methylation/N-formylation using PMHS as hydride reagent [128].
Molecules 24 00182 sch027
Scheme 28. Pressure-switched N-methylation/N-formylation to methylamine/formamide.
Scheme 28. Pressure-switched N-methylation/N-formylation to methylamine/formamide.
Molecules 24 00182 sch028
Scheme 29. DBU-catalyzed selective N-methylation/N-formylation and possible mechanism [130].
Scheme 29. DBU-catalyzed selective N-methylation/N-formylation and possible mechanism [130].
Molecules 24 00182 sch029
Scheme 30. Proposed mechanistic pathways for N-methylation and N-formylation [131].
Scheme 30. Proposed mechanistic pathways for N-methylation and N-formylation [131].
Molecules 24 00182 sch030
Scheme 31. TBD-catalyzed aminal synthesis from amine, CO2 and hydrosilane [132].
Scheme 31. TBD-catalyzed aminal synthesis from amine, CO2 and hydrosilane [132].
Molecules 24 00182 sch031
Scheme 32. Summarized control experiments in the published work.
Scheme 32. Summarized control experiments in the published work.
Molecules 24 00182 sch032
Scheme 33. Proposed mechanism for reduction of aminal [131].
Scheme 33. Proposed mechanism for reduction of aminal [131].
Molecules 24 00182 sch033
Scheme 34. Heterogeneous metal-based catalysis methylation of ammonia.
Scheme 34. Heterogeneous metal-based catalysis methylation of ammonia.
Molecules 24 00182 sch034
Scheme 35. Au-catalyzed methylation of amine with CO2 and H2.
Scheme 35. Au-catalyzed methylation of amine with CO2 and H2.
Molecules 24 00182 sch035
Scheme 36. Reaction route of PdGa-catalyzed methylation of N-methylaniline [141].
Scheme 36. Reaction route of PdGa-catalyzed methylation of N-methylaniline [141].
Molecules 24 00182 sch036
Scheme 37. Heterogeneous palladium catalysis N-formylation of amines [147].
Scheme 37. Heterogeneous palladium catalysis N-formylation of amines [147].
Molecules 24 00182 sch037
Table 1. Silver-containing catalytic systems.
Table 1. Silver-containing catalytic systems.
Catalyst (catalytic amount/mol%)SolventT/°CP/MPat/hYield/% aRef.
Ag2CO3 (1.5), PPh3 (6)CH3CN300.11668–98[32]
Ag2WO4 (1), PPh3 (2)-500.51230–97[21]
Ag2O (2), PPh3 (8)CH3CN60-1837–96 b[33]
Ag2CO3 (0.01), (p-MeOC6H4)3P (0.04)CHCl32529630.5–93.2 c[34]
AgCl (5), Et4NCl (9)CH3CN600.12429–98 b[35]
a Isolated yield; b NMR yield; c gas chromatography (GC) yield.
Table 2. One-pot three-component reactions of propargylic alcohols, CO2 and 2-aminoethanols.
Table 2. One-pot three-component reactions of propargylic alcohols, CO2 and 2-aminoethanols.
Catalyst (Catalytic Amount/mol%)SolventT/°CP/MPat/hYield(2-oxazolidinones)/% a,bRef.
Ag2CO3 (5), Xantphos (10)CHCl360118–4833–92[69]
Ag2O (5), TMG (30)CH3CN8011224–99[70]
CuI (5), 1,10-phen (5), t-BuOK (10)-800.51224–97[66]
[TBDH][TFE] (15)-800.11224–99[71]
a NMR yield.; b Isolated yield.
Table 3. Metal-catalyzed carbonylation of amines and CO2.
Table 3. Metal-catalyzed carbonylation of amines and CO2.
CatalystSolventT/°CP/MPat/hYield/% aRef.
CeO2NMP130524/48<1–86[92]
Y2(C2O4)3 bNMP150CO2(2)/N2(1)2462.43[93]
Y0.08Zr0.92O1.96 bNMP160CO2(2)/N2(1)12–2447.14–80.60[94]
In(N(SiMe3)2)Cl2·(THF)n, pyridine-1100.33–90-[95]
a GC yield; b 4A zeolite as dehydrant.
Table 4. Methylation of selective amines with CO2 and H2.
Table 4. Methylation of selective amines with CO2 and H2.
Molecules 24 00182 i001
EntrySub.Cat.PCO2 /MPaPH2/MPaTemp./°CT/hSol.Prod.Conv./% a,bYield/% a,bSel./% a,bRef.
1Molecules 24 00182 i002CuAlOx3616024hexaneMolecules 24 00182 i019NA86NA[138]
2Molecules 24 00182 i003CuAlOx3716024hexaneMolecules 24 00182 i020NA86NA[138]
3Molecules 24 00182 i004CuAlOx3716024hexaneMolecules 24 00182 i021NA87NA[138]
16Molecules 24 00182 i005Cu/CeO2171604tolueneMolecules 24 00182 i022282795[139]
6Molecules 24 00182 i006Pd/CuZrOx12.515030octaneMolecules 24 00182 i023NA91NA[140]
7Molecules 24 00182 i007Pd/CuZrOx12.515030octaneMolecules 24 00182 i024NA71NA[140]
8Molecules 24 00182 i008Pd/CuZrOx12.515030octaneMolecules 24 00182 i025NA75NA[140]
15Molecules 24 00182 i009PdGa/TiO25518010octaneMolecules 24 00182 i026989294[141]
9Molecules 24 00182 i010Pt-MoOx/TiO20.5520024neatMolecules 24 00182 i0278585100[142]
10Molecules 24 00182 i011Pt-MoOx/TiO20.5520024neatMolecules 24 00182 i0281006363[142]
4NH4HCO3Pt-MoOx/TiO21525024neatMolecules 24 00182 i029NA65NA[142]
5Molecules 24 00182 i012Pt-MoOx/TiO21525024THFMolecules 24 00182 i030NA56NA[143]
12Molecules 24 00182 i013Au/Al2O3-VS261407hexaneMolecules 24 00182 i031NA92aNA[144]
13Molecules 24 00182 i014Au/Al2O3-VS (Au: 0.5 mol%)261407hexaneMolecules 24 00182 i032NA96NA[144]
14Molecules 24 00182 i015Au/Al2O3-VS (Au 0.5 mol%)261407hexaneMolecules 24 00182 i033NA99NA[144]
11Molecules 24 00182 i016Au/Al2O3 (Au Size 2.0 nm, Au 0.73 wt%)131405cyclohexaneMolecules 24 00182 i03476>75>99[145]
17Molecules 24 00182 i017Re/TiO21520024octaneMolecules 24 00182 i035NA98NA[146]
18Molecules 24 00182 i018Re/TiO21520024octaneMolecules 24 00182 i036NA99NA[146]
a GC conversion and yield based on amine. b Not available (NA).
Table 5. N-Formylation of typical amines with CO2/H2.
Table 5. N-Formylation of typical amines with CO2/H2.
Molecules 24 00182 i037
EntrySub.Cat.PCO2 /MPaPH2/MPaTemp./oCT/hSol.Prod.Conv./% a,bYield/% aRef.
1Molecules 24 00182 i038Pd/C1310524MeOHMolecules 24 00182 i044NA93[147]
2nBuNH2Pd/C1310524MeOHnBuNHCHONA77[147]
3Molecules 24 00182 i039Pd−Au/PANI-CNT3.53.5125481,4-dioxaneMolecules 24 00182 i045>99.998.3[148]
4nBuNH2Pd−Au/PANI-CNT3.53.5125481,4-dioxanenBuNHCHO63.162.2[148]
5Molecules 24 00182 i040Pd−Au/PANI-CNT3.53.5125481,4-dioxaneMolecules 24 00182 i0464.13.4[148]
6Molecules 24 00182 i041Pd−Au/PANI-CNT3.53.5125481,4-dioxaneMolecules 24 00182 i0470.20.1[148]
7Molecules 24 00182 i042Pd/NC3413024ethanolMolecules 24 00182 i048NA94[148]
8Molecules 24 00182 i043Pd/NC3413024ethanolMolecules 24 00182 i049NA11[148]
a GC conversion and yield based on amine. b Not available (NA).

Share and Cite

MDPI and ACS Style

Li, J.-Y.; Song, Q.-W.; Zhang, K.; Liu, P. Catalytic Conversion of Carbon Dioxide through C-N Bond Formation. Molecules 2019, 24, 182. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules24010182

AMA Style

Li J-Y, Song Q-W, Zhang K, Liu P. Catalytic Conversion of Carbon Dioxide through C-N Bond Formation. Molecules. 2019; 24(1):182. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules24010182

Chicago/Turabian Style

Li, Jing-Yuan, Qing-Wen Song, Kan Zhang, and Ping Liu. 2019. "Catalytic Conversion of Carbon Dioxide through C-N Bond Formation" Molecules 24, no. 1: 182. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules24010182

Article Metrics

Back to TopTop