Next Article in Journal
Reduced Graphene Oxide-Supported Pt-Based Catalysts for PEM Fuel Cells with Enhanced Activity and Stability
Previous Article in Journal
Characterization of a Novel Thermophilic Mannanase and Synergistic Hydrolysis of Galactomannan Combined with Swollenin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Copper Phyllosilicates-Derived Catalysts in the Production of Alcohols from Hydrogenation of Carboxylates, Carboxylic Acids, Carbonates, Formyls, and CO2: A Review

Department of Chemical Engineering, National Cheng Kung University, 1 University Road, Tainan 70101, Taiwan
*
Author to whom correspondence should be addressed.
Submission received: 22 January 2021 / Revised: 5 February 2021 / Accepted: 8 February 2021 / Published: 14 February 2021

Abstract

:
Copper phyllosilicates-derived catalysts (CuPS-cats) have been intensively explored in the past two decades due to their promising activity in carbonyls hydrogenation. However, CuPS-cats have not been completely reviewed. This paper focuses on the aspects concerning CuPS-cats from synthesis methods, effects of preparation conditions, and dopant to catalytic applications of CuPS-cats. The applications of CuPS-cats include the hydrogenation of carboxylates, carboxylic acids, carbonates, formyls, and CO2 to their respective alcohols. Besides, important factors such as the Cu dispersion, Cu+ and Cu0 surface areas, particles size, interaction between Cu and supports and dopants, morphologies, and spatial effect on catalytic performance of CuPS-cats are discussed. The deactivation and remedial actions to improve the stability of CuPS-cats are summarized. It ends up with the challenges and prospective by using this type of catalyst.

1. Introduction

Phyllosilicates are versatile materials with numerous applications such as being carriers of enzymes in the synthesis of chemical and pharmaceutical compounds [1,2,3,4,5,6,7,8], precursors of organo-phyllosilicates nanocomposites [9,10,11,12,13,14,15], ferromagnetic [16] and anti-ferromagnetism materials [17], lithium ion batteries [18,19], pharmaceutical excipients [20,21,22,23], adsorbents of heavy metals [24,25], toxic compounds [26,27], industrial dyes [28,29,30,31,32], and CO2 scrubbers [33,34,35]. One of its applications is to be the precursor of heterogeneous catalysts. Figure 1 shows the bibliometric analysis of phyllosilicate-related catalysis papers. An exponential trend underlines the emerging interest of phyllosilicate in catalysis applications.
The burgeoning research of phyllosilicates has been stimulated by its unique physicochemical properties. Phyllosilicate is a laminar-structured silicate mineral which is comprised of one tetrahedral layer of silica (SiO44−), sharing 3 oxygen atoms together in a planar with another oxygen links with a layer of octahedra units (see Figure 2) [36,37]. The core of the octahedron unit could be occupied by divalent or trivalent metallic cations such as Al, Mg, Zn, Mn, Fe, Cu, Ni, and Co [38,39]. Various phyllosilicates can be formed by assembling different tetrahedral and octahedral layers. For example, chrysocolla (Cu2Si2O5(OH)2) is a lamellar structure consisted of layers of SiO4 tetrahedra sandwiched between discontinuous layers of CuO6 octahedra [40]; kaolinite is an 1:1 aluminum phyllosilicates while pyrophyllite is 2:1 [41]; chrysotile is made of 1:1 magnesium phyllosilicates (Mg3Si2O5(OH)4) [37]. The difference between the radius of tetrahedrally coordinated Si and octahedrally coordinated cation can bend the layer-structured phyllosilicates [37,42], resulting in distorted morphologies such as rolling of paper sheets into a tube. Compared to conventional catalysts (metal impregnated on SiO2), the high surface area of laminar structured phyllosilicate allows metal atoms to be finely dispersed. At the same weight loading but higher dispersion, phyllosilicate structure might possess more active sites than the conventional catalysts.
Structural diversity, compatibility with various transition metals, high surface area, and finely dispersed metallic cations makes phyllosilicates attractive in numerous reactions, such as Fischer–Tropsch [43], esterification [44], cellulose conversion to glucose [45], oxidation of water [46,47], oxidation of CO [48,49,50,51,52], CO or CO2 reforming [34,53,54,55,56,57,58,59,60,61], hydrogen generation from methanol steam reforming [62,63,64,65,66], and syntheses of valuable alcohols from hydrogenation of CO [67,68,69], CO2 [53,70,71,72], and hydrogenolysis of esters [73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92]. Copper [70,71,73,75,76,77,78,79,80,81,82,83,84,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120] and nickel phyllosilicates [53,54,55,57,58,59,60,61,69,72,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152] derived catalysts (CuPS-cats and NiPS-cats, respectively) are extensively investigated. NiPS-cats are usually used for the hydrogenation of unsaturated hydrocarbons [120,148], CH4 dry reforming [58,60,129,136,137], CO2 methanation [53], water-gas shift [130,134], deoxygenation [151], steam reforming, [135], and hydrodeoxygenation of biomass [131]. In comparison, CuPS-cats are promising in alcohol synthesis through the hydrogenation of carbonyls such as ester, carboxylic acid, carbonate, and CO2 as well. So far, there are three reviews [153,154,155] that discuss about the catalytic hydrogenation of C=O bond to produce alcohols using copper-based catalysts. However, only a specific aspect was focused in each review. Watari et al. [153] targeted hydrogenation reactions mediated by using copper-based catalysts, containing limited information about CuPS-cats. Ye et al. [154] aimed to combat the deactivation problem of Cu-based catalysts. Bian and Kawi [155] published an insightful review on the synthesis of CoPS, NiPS, and CuPS, but emphasized the usages of NiPS. In this contribution, we focus solely on CuPS-cats, but all aspects of CuPS-cats including the preparation methods, control factors of morphology, effects of hetero-element addition, and reduction behaviors of CuPS were systematically summarized and in-depth discussed. This is followed by the applications of CuPS-cats in the hydrogenation of carboxylates, carboxylic acids, carbonates, formyls and carbon dioxide. Lastly, the existing challenges and future prospective of CuPS-cat were discussed based on latest updated researches till 2020.

2. Formation of CuPS-Cats

The development of CuPS-cats over the past 20 years was summarized in Figure 3. There are five major approaches in CuPS synthesis, including examination of synthesis criterion, design of preparation method, selection of supports, addition of promoters, and control of PS morphology, and these topics will be discussed in Section 2.1, Section 2.2, Section 2.3, Section 2.4 and Section 2.5, respectively. Since a reduction step is necessary to transform CuPS into CuPS-cat, the reduction behavior of CuPS will be addressed in Section 2.6.

2.1. Synthesis Methods

2.1.1. Ammonia Evaporation-Related Approaches

CuPS was firstly synthesized by using the cationic exchange method [93]. The cations of copper tetraamine complex [Cu(NH3)4(H2O)2]2+ were dissolved and then deposited on a silica surface at room temperature. The cationic exchange method is usually accompanied by a deposition-precipitation step because the formation of CuPS was proposed to be a sequential process through the dissolution of silica followed by the precipitation of phyllosilicates. Toupance et al. [40] proposed that the formation mechanism of CuPS is through the selective adsorption of neutral Cu complexes (i.e., Cu(OH)2(H2O)2) on a silica surface). Therefore, the cationic exchange was renamed as selective adsorption. Later, Chen et al. [73] reported an ammonia evaporation (AE) method for CuPS synthesis. In this process, silica sol was mixed with a copper ammonia complex (initial pH value at about 11–12), followed by the gradual removal of ammonia through evaporation till the solution is neutralized, yielding CuPS as the precipitates. After washing, drying, and calcination, CuPS-cats can then be obtained. Mostly, octahedrally coordinated Cu2+ needs to be activated by the reduction to lower oxidation states such as Cu+ or Cu0.
Besides the AE method, an ammonia evaporation hydrothermal (AEH) method which integrated an additional hydrothermal treatment step to ammonia evaporation process was developed by Gong et al [96]. The major difference between AE and AEH method is that the latter had a follow-up hydrothermal treatment for the resultant solid after AE. The difference of CuPS-cats synthesized from AE and AEH has not been fully understood. Another related synthesis method was ammonia hydrothermal (AH). The AH method was firstly described by Yue et al. [97] for synthesizing copper phyllosilicate nanotubes (CuPSNTs). Briefly, by excluding the AE step, the mixture of silica, copper nitrate, and ammonia was treated under hydrothermal conditions (200 °C and 10 MPa), resulting in nanotube-shaped CuPS structure. Excess ammonium by adding NH4Cl [107,113] or NH4NO3 [156] in the mixture is required.

2.1.2. Precipitation

Precipitation methods such as co-precipitation (CP) and deposition-precipitation method were developed to synthesize CuPS-cats. Typically, a mixture of copper and silica sol was precipitated by dropwise adding basic agents such as (NH4)2CO3 [79,92], NH4HCO3 [104], Na2CO3 [117,157], or ammonia solution [158]. Unlike the AE method, the pH value of the aqueous solution in precipitation methods was maintained at 6–7. When using tetraethyl orthosilicate (TEOS) as the silica source, CuPS-cats would be produced through a hydrolysis precipitation by using (NH4)2CO3 as a precipitation agent [81,116]. Zhang et al. [78] invented urea hydrolysis deposition-precipitation (UHDP) method by using urea as the precipitant. Briefly, hydrolysis of urea increased the pH value, and thereby induced precipitation of CuPS. The UHDP method was also named as “urea-assisted gelation” [76,85,90,95,159].

2.1.3. Other Methods

The pioneer study of sol-gel method was developed by He et al. [94]. At the outset, ammonia solution was added into the mixture of copper and silica sol until the pH value of the solution increased to 11–12. The gelation was subsequently conducted at 50 °C. Weight loading of copper can be maximized at 45%. Abbas et al. [84] synthesized CuPS-cats by using novel methods such as sonochemical and stirring-enhanced hydrothermal. Systems such as a pulse-mode ultrasonic reactor and agitation-enhanced hydrothermal bomb were applied to improve intermolecular collision. A combinative method of ultrasound-treated and agitation-enhanced hydrothermal method was also reported [160].
In sum, several methods can be used in CuPS-cats synthesis. Among them, the laminar structure of phyllosilicate is usually prepared by using the AE method, as confirmed by various techniques such as TEM, XRD, and FTIR (Figure 4). Besides, AE-made CuPS-cats had several advantages compared to those made by other methods, including high surface area [99,104,158], good structural regularity [158], strong thermal stability [104], and great reproducibility. All these features render AE method the most effective to fabricate CuPS.
One drawback of using the AE method for CuPS preparation is the emission of waste ammonia. This inspires Popa et al. [79] and Ding et al. [92] to invent ammonia emission-free process, i.e., the ammonium carbonate deposition precipitation (AC) method, through which CuCO3 was precipitated on SiO2 followed by a calcination step to decompose CuCO3. AC-made Cu/SiO2 had a lower surface area and a less phyllosilicate crystallinity than those of AE-made counterparts; however, they had comparable dimethyl oxalate (DMO) hydrogenation activities. More Cu+ species were formed (the content of Cu+ over the total Cu species is greater than 60%) by using AC method [79,92].

2.2. Effect of Synthesis Criteria

2.2.1. Temperature for Ammonia Evaporation

The effect of ammonia evaporation temperature was investigated in the range of 60 to 100 °C [73]. Morphologies of CuPS (sphere-like or slit-like structure) depends heavily on the temperature used in AE. An optimal crystallinity was achieved at 90 °C (see Figure 5). A maximal fraction mole of Cu+ (Cu+–(Cu0 + Cu+) = 0.55) and the highest DMO hydrogenation activity was evidenced on the CuPS-cat made by using 90 °C as the evaporation temperature.

2.2.2. Aging Time

The effect of aging time of CuPS precursors was investigated [100]. The mixture of silica sol and [Cu(NH4)]2+ was stirred for different times (4, 16, 24, and 40 h) at room temperature before ammonia evaporation. Aging time does not significantly influence the surface area and porosity of CuPS-cats. By contrast, the Cu+/(Cu0+Cu+) ratio of 20 wt % CuPS-cats were 0.30, 0.48, and 0.39, with respect to the aging time of 4, 24, and 40 h. An explanation for the effect of aging time on the distribution of Cu+ and Cu0 was still absent. Presumably, the crystallization behaviors of CuPS may vary depending on the aging time [92].

2.2.3. Calcination and Reduction Temperature

Ding et al. [102] investigated the effect of calcination temperatures on AE-synthesized CuPS-cats in the hydrogenation of diethyl malonate (DEM) to 1,3-propanediol (PDO). Elevating the calcination temperature above 623 K may partially damage the structure of CuPS. The decreasing order of Cu+ concentration corresponding to the calcination temperature is estimated to be 47.6% (1023 K) > 45.5% (723 K) > 41.3% (623 K) > 36.9% (823 K) > 35.1% (923 K). However, the hydrogenation rate did not follow the same trend as that of the Cu+ concentration. This indicated that the activity of CuPS-cat depends not only on Cu+ concentration, but also on the nature (such as crystallite size and intimacy of Cu0 and Cu+) of derived Cu species.
Sun et al. [87] systematically investigated the structure-activity correlation of CuPS-cats reduced at different temperatures and their DMO hydrogenation activities. The DMO conversion and ethylene glycol (EG) selectivity showed a volcano-like trend, while Cu+–(Cu0 + Cu+) ratio, surface area, pore volume, Cu dispersion, and Cu particles size did not have a strong correlation with the reduction temperature (Figure 6). The interaction of Cu and the support with a proper reduction treatment may improve the catalytic performances of CuPS-cats, as discussed vide infra.
Calcination and reduction temperatures are important in tailoring the ratio of Cu+–Cu0 of CuPS-cats. Different calcination and reduction temperatures were used since a proper Cu+–Cu0 ratio can be varied with respect to the reaction per se. For example, a ratio of Cu+–(Cu0 + Cu+) close to unity is claimed to be effective in CO2 hydrogenation [70] (see Figure 7); while 0.5 is a more proper ratio for hydrogenation of dimethyl malonate [102]. The varying preferred values of Cu+–(Cu0 + Cu+) ratios with respect to different reactions complicate the developing, preparation, and optimization of CuPS-cats.

2.2.4. Cu Weight Loading

Dong et al. [106] found that agglomeration of CuO particles were formed when the percentage of Cu was higher than 43% in CuPS preparation. They claimed that a maximal loading of Cu for phyllosilicate is at 39%, corresponding to the Cu–Si molar ratio of approximately 0.75. The favorable weight loading of Cu to synthesize CuPS-cats is in line with the content of Cu in chrysocolla (Cu2Si2O5(OH)2). However, the relation of Cu loading and activity of CuPS-cat varies in different systems. Li et al. [108] found that 10% of Cu showed the best performance for hydrogenation of ethylene carbonate (EC) to EG. Jiang et al. [111] reported the 20% CuPS-cats could maximize the conversion of adipic acid (AA) and the yield of its hydrogenated product, i.e., 1,6-hexanediol (HDOL). For the conversion of furfural (FFAL) to 2-methyl furan, 25% of Cu was reported to be appropriate [106]. Much remains to be learned about the correlation between a proper Cu content and different hydrogenation reactions.

2.2.5. Ammonium Cation Concentration

Apart from Cu and Si concentrations, the amount of ammonium cation can tune the crystallization of CuPS. Toupance et al. [40] found that the crystallinity of CuPS showed a 4-fold decrease when the pH value increases from 10 to 11 by using ammonia to control the pH value in the deposition-precipitation process. If the [NH4+]–[Cu2+] ratio surpassed 10, CuPS structure cannot be formed due to excess [Cu(NH3)4]2+ complexes. Di et al. [80] also supported this claim: excess ammonia inhibits crystallization of CuPS. CuPS-cats synthesized by using CP with an abundance of ammonia could be reduced easier than that without adding excess ammonia. They also discovered that adding a trace of Zn cations with ammonium can change the morphology of CuPS from 2:1 to 1:1.

2.3. Effect of Supports

2.3.1. Silica Support

Ultra-fine silica produced from in situ hydrolysis of TEOS was proved to enhance the dispersion of copper and to induce the formation of Cu+ [81]. No diffraction of CuO was observed by XRD even at a high copper loading (40 wt % Cu). A volcano trend of the ratio of CuPS phase to amorphous silica was discovered with its apex at a 30% Cu loading. The concentration of Cu+ in the 30 wt % Cu sample was doubled than that of the 30% CuPS-cats synthesized by using the AE method.
Qi et al. [112] disclosed that Cu precipitated on nanosized silica showed a high dispersion (>69%), regardless with silica particle sizes. In contrast, Wang et al. [88] synthesized CuPS-cats at a 20 wt % Cu by using fine silica particles with varying sizes (5 to 25 nm) through the AE method. They found that the Cu dispersion depends on the particle size of silica: the smaller the size of silica, the higher the Cu dispersion can be. Although the correlation between Cu dispersion and silica particle size is still elusive, the aforementioned groups agreed that the interaction between Cu and SiO2 became stronger by reducing SiO2 particle size. The hydroxyl-enriched surface of silica with small particle size was proposed to enhance the concentration of Cu+ and CuPS crystallinity (Figure 8).
In order to increase the contact intimacy of silica and copper, mesoporous silica was adopted. Zhao et al. [82] investigated the effect of different silica sources by using mesoporous silica and silica sol in the synthesis of CuPS-cats by using the AE method. They claimed that the high surface area of mesoporous silica (SBET = 1019 m2/g) produced higher concentration of Cu+ than the silica sol did. The different dissolution rates of silica sources were attributed to the varying concentration of yielded Cu+.
Other mesoporous silica such as SBA-15, KIT-6, and MCM-41, were used to prepare CuPS-cats, and their influences in Cu dispersion was noticeable [105,161]. A high dispersion of Cu was obtained by using SBA-15 (26.2%) as the silica source, much higher than those made by using KIT-6 (8.2%) and MCM-41 (6.8%). [105]. Ciotonea et al. [161] claimed that Cu nanoparticles can be finely anchored on SBA-15 without sintering. After reduction, the sheet-like structure of phyllosilicate is intact with small metallic Cu particles (~5 nm).
In sum, ultra-fine silica or mesoporous silica with the large surface area and high surface hydroxyl groups might cause a stronger interaction between Cu and SiO2. The stronger interaction can increase the dispersion of Cu and Cu+ concentration as well. Moreover, sintering of Cu can be eased due to the strong interaction with silica.

2.3.2. Binary Oxides Support

Besides pure silica, AE-synthesized binary oxides such as SiO2-TiO2 [162], SiO2-MgO [163], or SiO2-CeO2 [164] was developed to enhance Cu dispersion. Introducing TiO2, MgO, or CeO2 in silica leads to a high Cu0 concentration due to exchanged electrons between Cu and dopant [162,163,164]. Mixing MgO and CeO2 with silica sources could enhance Cu+ concentration. Nevertheless, the formation mechanism of binary oxides phyllosilicate phases has not yet been studied.

2.3.3. Interaction between Silica and Copper

Most studies suggest the performances of CuPS-cats depend heavily on the ratio of Cu+–Cu0. A different viewpoint was claimed by Xu et al. [113], indicating that silica can promote molecular hydrogen dissociation for the hydrogenolysis of esters. In comparison to bare copper particles, a dramatic jump in the hydrogenation rate of the carbonyl group was found in copper particles coated by mesoporous silica. DFT calculation showed that higher partially positive charged Cu cations were formed on silica-coated Cu than its uncoated counterpart. Moreover, the negative charged intermediates, such as CH3O(O =)C–CH(OCH3)Oδ−, and Hδ−, can be stabilized due to the presence of SiO–Hδ+ species at the interfaces between copper oxide and silica. Therefore, the Cu–O–SiOx interface is concluded to be active in hydrogen dissociation and to stabilize the intermediates. In comparison to other oxides such as TiO2, Al2O3, and ZrO2, silica showed a much stronger interaction with copper [96]. Cu dispersion and Cu surface area of the SiO2 supported CuPS-cats were 5 to10 times higher than those of TiO2, Al2O3, and ZrO2 supported counterparts. The descending order of the strength of support-copper interaction was SiO2 > Al2O3 > ZrO2 > TiO2 [96].

2.4. Effect of Promoters—Interaction in a Bimetallic System

Introducing a secondary metal such as B [95,165], Ag [159], Ni [119,120,135,136,157,166], In [90,115], Zn [86,109,116], La [114,167], Sn [110], Ce [85], or Mo [168] in CuPS allows the activity and stability of CuPS-cat to be enhanced because of the inhibition of Cu sintering and the presence of electron transfer between Cu and dopant. Smaller Cu crystalline and higher dispersion of Cu can be obtained by doping the promoters than those of pristine CuPS after reduction. The average particles size of Cu promoted by adding 1% cerium can be reduced to be about 3–4 nm [85]. In the case of Zn- [109,116] and La-doped-CuPS-cats [167], Cu particles were at approximately 3 nm. Moreover, the Cu particles size in Ni5-Cu36PS rose slightly from 4.0 to 4.7 nm, much less than that of its counterpart of Cu41PS [119]. Note that the high dispersion of Cu particles can be reached at a very low amount of additives, usually 1–2 wt % [85,110,115,159,167] but the aggregation of Cu will be substantial when doping excess amount of additives (as demonstrated in Figure 9).
For activity enhancement, electron transfer between dopants and neighboring Cu species can tailor the Cu+–Cu0 ratio. For example, boron [95,165] has a higher electrophilicity than that of copper, and the tandem Cu/B site allows electron transfer from Cu to B, elevating the Cu+ concentration. Conversely, Cu+ concentration is suppressed by forming tandem In/Cu or Sn/Cu sites [90,115]. The electronegativity of the elements might give a rough estimation of Cu+–Cu0 proportion based on electron donating and accepting (Figure 10); however, inconsistent trends were reported [85,86,110,114,167]. Beside the electron interaction, the acidity and basicity of CuPS-cats was influenced by heteroelements. The NH3-TPD and CO2-TPD profiles (Figure 11) indicated that acid sites of B-doped CuPS-cats increased following the trend of increased boron content [95,165]. Sometimes, the alloy of copper and promoters may be formed such as InCu [90], NiCu [120] and AgCu [159].

2.5. Non-Planar CuPS-Cats

CuPS nanotubes (CuPSNTs) can be formed from rolling CuPS sheets into a tubular shape under hydrothermal condition [97]. The CuPSNTs of approximately 200–300 nm of length and 15 nm of thickness were successfully synthesized by using hydrothermal treatment. Similarly, the CuPSNTs with hundreds nm of length and 1–2 nm in thickness were synthesized [107,113] by replacing basic ammonia solution with mild acidic ammonia chloride solution. A hybrid type of tubular- and spherical-shaped phyllosilicate – nanotube-assembled hollow spheres (NAHS) (Figure 12) was designed as well [156]. Moreover, synthesize CuPSNTs by the AEH method using a copolymer of ethylene oxide and propylene oxide (P123) as a template was reported [169].
To evaluate the confinement effect of CuPSNTs on DMO conversion to ethanol (EtOH), Yue et al. [97] impregnated Cu nanoparticles in the inner space of CuPSNTs (Cu/CuPSNTs-in), and a higher thermal stability and activity were found than those made by anchoring Cu on the outer surface (Cu/CuPSNTs-out) did. The Cu particles of the reduced Cu/CuPSNTs-in were twice less than those of the reduced Cu/CuPSNTs-out. Moreover, the concentration of Cu+ was much greater of Cu/CuPSNTs-in (90% of the overall Cu content) than that of Cu/CuPSNTs-out. The curvature effect-induced Cu-support interaction of CuPSNTs-in was proposed [70,73,106].
An excellent thermal stability of CuPSNTs after a 24-h thermal treatment at 350 °C was reported [107]. The space time yield (STY) of CuPSNTs in dimethyl 1,4-cyclohexanedicarboxylate (DMCD) conversion to 1,4-cyclohexane dimethanol (CHDM) was unaltered while that of CuPS-cats prepared by the AE method showed a 7.4% decreasing rate. Cu sintering seemed to be hindered by CuPSNTs with ignorable changes of Cu particle size before and after reaction. The great activity and stability were attributed to the strong Cu-O-Si bond and the confinement effect. Xu et al. [113] claimed that the high percentage of Cu+ in CuPSNTs was related to the enriched Cu-O-SiOx interface, which is highly active in DMO hydrogenation. To demonstrate the metal-support interaction and the confinement effect, a thin layer of mesoporous silica was coated on CuPSNTs (CuPSNTs@m-SiO2) (Figure 13). The ratio of Cu+/Cu0 in CuPSNTs@m-SiO2 dramatically increased, which was approximately 2-fold higher than those of CuPSNTs and CuPS-cats.
The relationship between morphologies of CuPS-cats (laminar, nanotubes, and nanotube-assembled hollow spheres (NAHS)) and their activities of DMO hydrogenation was investigated [156,170]. The laminar-, nanotubes- and NAHS-type have similar surface areas, Cu+–Cu0 ratios, and distributions of active sites. However, the NAHS showed the highest turnover frequency in DMO hydrogenation to EG and a high yield of EG (95%) and stability (>300 h) at a H2–DMO molar ratio as low as 20 while using NTs and laminar CuPS needed to be operated at the H2–DMO ratios greater than 50 to obtain the equivalent amount of EG. Noting that the H2–DMO molar ratio of the feed is usually at 80 to 200 (Table 1). They demonstrated that hydrogen can be preferentially adsorbed on the concave surface of nanotubes and hollow spheres, resulting in the high TOF for the NAHS. The design of NAHS opens a new window of low hydrogen consumption in DMO hydrogenation.

2.6. Reduction in CuPS-Cats

To transform CuPS precursors into CuPS-cats, a designed temperature-programmed reduction in H2 atmosphere (H2-TPR) is needed. The H2-TPR profiles of CuPS-cats show a main reduction peak at around 240–250 °C [70,73,96,111,119,171] and the reduction peak ends at around 300 °C. The reduction peak may vary depending on the crystallinity of CuPS, calcination treatment, and interaction with supports or with promoters. From the H2-TPR profiles, different reduction temperatures were selected depending on reduction degrees. Mostly, CuPS-cats are reduced at 300 or 350 °C to completely deconstruct Cu octahedral, to yield Cu+ and Cu0 species for catalysis applications. In some cases, lower reduction temperatures such as 200 [105], 220 [78], 250 [87,172], 260 [157], 270 [104,106], and 280 °C [84] were also selected for lower reduction degrees. The reduction temperatures and time are listed in Table 1.
The exsolution of Cu particles from CuPS-cats was disclosed by van den Berg et al. [174] by using in situ TEM. In this study, kinetic models such as nucleation-and-growth, diffusion- and reaction-limited autocatalytic reduction were established to describe the formation and growth of Cu particles. In accordance with the TPR results, the TEM images observed Cu particles at 250 °C (see Figure 14C). To ensure the TEM images reflect for the whole sample, different regions of a sample were picked up to monitor the particle growth (Figure 14E). Time-resolved TEM images indicated the mechanism of the autocatalytic reduction is controlled by the diffusion of mobile copper species and the Cu particles size has a strong correlation with the time of isothermal treatment (Figure 14E).
The influences by using different reducing agents on derived Cu species were investigated by our group using in situ XANES analysis in H2 and CO environments. Composition of Cu2+ (in CuPS), Cu+ and Cu0 of XANES spectra were analyzed by using the linear combination method (see Figure 15). The results showed that the evolution of Cu+ and Cu0 species in hydrogen and CO environments are different. H2 is a stronger reducing agent than that of CO, facilitating Cu2+ reduction to Cu0 and yielding a limited amount of Cu+ in the whole temperature range. In contrast, using CO as the reducing agent produced a higher composition of Cu+ even at temperature higher than 400 °C. This indicated that CO is a more suitable reducing agent than H2 to control Cu0 and Cu+ distribution of CuPS-cats.

3. Hydrogenation Activity of CuPS-Cats

CuPS-cats are widely used in the synthesis of alcohol from carbonyl groups hydrogenation. In this review, hydrogenation of esters, carbonates, carboxylic acids, aldehydes, and carbon dioxide are included (Figure 16).

3.1. Esters to Alcohols

3.1.1. Hydrogenation of DMO and DEO

CuPS-cats exhibited an extremely high activity for the hydrogenolysis of DMO and diethyl oxalate (DEO) because many studies claimed that both Cu0 and Cu+ mediate the reduction in carbonyl group [116]: Cu+ polarizes the carbonyl, facilitating the subsequent hydride insertion of dissociated hydrogen on Cu0 [78,84,87,92,94,96]. Conversion of DMO and DEO were operated at approximately 200 °C and H2 pressure of 2 to 3 MPa by using CuPS-cats. The distribution of products derived from DMO and DEO hydrogenation, i.e., methyl glycolate (MG), EG, or EtOH (Figure 17), can be controlled by varying synthesis approaches, adding promoter, and varying morphologies of CuPS-cat. EG is the most favorable product because of its essential application as an antifreeze and a coolant [175].
  • EG as the hydrogenated product
Different synthesis methods, such as AE, AC, and sol-gel, were used in CuPS-cats preparation for EG production (EG yields > 90%). Ding and co-workers [92] reported EG production from DEO catalyzed by CuPS-cats prepared by using deposition precipitation method using ammonium carbonate (CuPS-cats-AC) and ammonia evaporation method (CuPS-cats-AE). Although the CuPS-cats-AE and CuPS-cats-AC have the same conversion of DEO (>95%), the EG selectivity of the CuPS-cats-AE is always 5–10% less than that of the CuPS-cats-AC did [92]. Furthermore, the CuPS-cats-AC showed a higher thermal stability than that of CuPS-cats-AE [79]. The durability tests of CuPS-cats-AC lasted for more than 200 h. The higher activity and stability were correlated to the formation of Cu2O, which is more stable than Cu+ of CuPS-cats-AE. The CuPS-cats prepared by sol-gel method also exhibited a relatively high DEO conversion and EG selectivity (96 and 88%, respectively) [94].
A proper Cu+–Cu0 ratio to maximize the EG yield is a popular hypothesis of CuPS-cats in DMO hydrogenation [78,87]. Sun et al. [87] discovered a strong dependence of DMO hydrogenation performance and the Cu+–Cu0 ratio by controlling the reduction temperature of CuPS-cats. At a reduction temperature of 250 °C, the Cu+–(Cu+ + Cu0) ratio was 0.4, implying the amounts of Cu+ and Cu0 were similar. The CuPS-cats reduced at 250 °C showed the highest hydrogenation activity of DMO to EG with a complete conversion and 95.2% selectivity at 180 °C with a 16 h durability. The selectivity of EG from DMO hydrogenation can be tailored by controlled reduction temperatures. However, the effect of reduction temperature on the physicochemical properties of CuPS is unclear. It is likely that different reduction treatments of CuPS can create various degrees of Cu–Si interaction.
Besides, the use of high surface area of silica such as mesoporous silica [82], nanosphere-shaped silica [88], or foam framework of zeolite [76] demonstrated a significant improvement of catalytic performance of CuPS-cats. The high surface area of silica possesses a large amount of surface hydroxyl groups which were supposed to increase the crystallinity of CuPS and surface Cu dispersion. The relationship between surface hydroxyl groups, CuPS phase, and Cu+ surface area were indicated in Section 2.3.1. Silica Support. The higher dispersion, the better catalytic performance because more active sites are produced. A 98% yield of EG was obtained for 250 h reaction at 180 °C when a mesoporous silica was applied to produce the highest surface areas of Cu0 (36.9 m2/g) and Cu+ (41.6 m2/g) [82].
The reactivity of CuPS-cats for the hydrogenation of DMO to EG was investigated by doping boron [95], silver [159], nickel [157], tin [110], zinc [86], or indium [90]. Adding other metals into CuPS-cats was able to improve the activity and stability of the CuPS-cats due to electron donating or accepting between Cu and the additional dopant or formation of Cu-dopant alloy.
Another factor that needs to be combated with is the need of excess hydrogen in DMO hydrogenation. Normally, the molar ratio of H2–DMO in the feed is greater than 50. Yao et al. [156] proposed a novel design—CuPS-cats based nanotube-assembled hollow spheres, which enrich hydrogen concentration inside the spheres so that the molar ratio of H2–DMO of the feed could be reduced to 20 with similar performances as those of conventional CuPS-cats. An improved EG productivity was conducted by Yue et al. [77] when cylindrical cordierite monoliths with micro-tunnels were dip-coated by CuPS-cats. The operating LHSV of their system can be as high as 16 h−1, more than 5 times higher than those used in other studies (2–3 h−1). That is, the space time yield of EG can be substantially magnified.
  • MG as the hydrogenated product
Sometimes, MG is a desired product of DMO hydrogenation. Partial inhibition of DMO hydrogenation to produce high yield of MG using CuPS-cats was proposed [84,89]. Since Cu0 is supposed to be hydrogenation-active, suppressing Cu0 concentration should improve the selectivity of MG. Thus, sodium silicate was impregnated into CuPS-cats powder prepared by AE method, intentionally to increase the formation of Si-O-Cu bonds, to limit surface concentration of Cu0 [89]. As expected, the reducibility of Cu cations was decreased with additional sodium silicate. Coherently, the surface area of Cu0 decreased along with the increase in Cu+ surface area (the ratio of Cu0–Cu+ = 0.50–0.54). The selectivity of MG increased from 0.8 to, approximately, 100% at 84% of DMO conversion by doping 0.5 wt % of sodium silicate. This highlighted the strong correlation between DMO-derived products and surface Cu species.
  • EtOH as the hydrogenated product
In general, the high Cu surface area (Cu+ and Cu0 active sites) and the high dispersion of Cu are the two key factors to increase the yield of EtOH. By introducing Ce to CuPS-cats, the dispersion and electron density of copper could be enhanced [85]. The 1Ce-Cu/SiO2 [85] can produce 92% of EtOH within 200 h constantly, the highest yield and stability compared to other works [96,97]. Besides, a large amount of H2 (H2/DMO = 200) and higher reaction temperature (280 °C) are required (Table 1) to further hydrogenate EG to EtOH. Hydrogen concentration could be enhanced from the feed-in stream or from an increase in local partial pressure. In the case of CuPSNTs, due to the confinement effect of CuPSNTs, hydrogen pressure inside the tubes is higher than that outside [156]. Moreover, longer contacting time of reactants and catalytic sites inside the tubes allows the deep hydrogenation goes further to EtOH [97]. However, too long a reaction time may decrease the EtOH selectivity and increase by-products such as PDO and 1,2-butanediol [97].

3.1.2. DEM to PDO

The reactivity of CuPS-cats for DEM hydrogenolysis to PDO was investigated [102,103]. Because PDO is an intermediate product between 3-hydroxyl ethyl propionate and n-propanol (see Figure 18), the catalytic activity of CuPS-cats should be tuned to moderate in order to increase the productivity of PDO. For example, the PDO productivity of the CuPS-cats can be controlled by calcination temperature. Calcination temperature at 450 °C was found to be the most appropriate to achieve the highest Cu dispersion of CuPS-cat, which showed the highest selectivity of PDO at around 30%, and over 90% of DEM converted for a duration time of 100 h [102]. Increasing calcination temperature of CuPS will lead to deep hydrogenation of DEM to n-propanol due to the increase in Cu crystallite size.
Another way to improve PDO yield is using promoters to tune the rate of Cu+ and Cu0. Depending on the electron interaction between Cu and the additional elements, Cu+ or Cu0 will be promoted (see effect of promoters in Section 2.4). Taking B, Ag, Zn, Al, Ni, or Mo as examples, productivity of hydrogenation of DEM can be manipulated by doping these elements into CuPS-cats [103]. Incorporating Zn, Al, and Mo into CuPS-cats enhanced the selectivity of 3-hydroxyl ethyl propionate. Meanwhile, doping Ag and Ni into CuPS-cats enhanced deep hydrogenation to n-propanol. Introducing B was found to be favorable for PDO production because of the enriched Cu0 surface (mole fraction of Cu0 = 0.67). A close to balance ratio of Cu0 and Cu+ is good enough for production of an intermediate product like PDO.

3.1.3. Methyl Acetate (MA), Ethyl Acetate (EA) to EtOH

CuPS-cats prepared by the AE method exhibited a higher hydrogenation activity for MA conversion than those prepared by ion exchange and deposition precipitation due to the high dispersion of Cu [158]. Nevertheless, EtOH accounted for half of the yield of total products. Additional dopant can tune the product distribution.
The support effect on EtOH yield was investigated [163,164]. The Cu supported on MgO-SiO2 system (30 wt % of Cu and MgO–SiO2 mass ratio of 9) showed extremely high selectivity of EtOH (80.3% conversion and 99% selectivity) due to a strong interaction between MgO and Cu, which retarded the reduction in Cu+ to Cu0 and improved the surface area of Cu+ (about 65% of Cu+–(Cu++Cu0) fraction) [163]. Ye et al. [164] indicated that adding 30 wt % of cerium to CuPS-cats enhanced interaction of Cu and support, improved copper dispersion and enriched Cu+ concentration. More than 90% yield of EtOH was obtained with a full conversion during the 100-h test. Moreover, Cu+ was stabilized in the Cu/30%CeO2–SiO2 after the reaction (Cu+–Cu0 ratio was 0.72 and 0.65 before and after the reaction).
Besides, indium and zinc exhibited as an effective promoter for CuPS-cats in conversion of MA to EtOH reaction. About 98% of MA was converted steadily over the course of 100 h reaction by 1% In2O3 addition [115] while 92% of conversion and approximately 90% of selectivity can be obtained when adding Zn into CuPS-cats (Zn–Cu ratio = 1:9) [116]. However, if silica was substituted by alumina, the synergy of Cu and Zn will vanish.
Instead of MA, ethyl acetate (EA) was also a reactant for EtOH synthesis. Incorporating ammonium and zinc ions during the CP method to transform 2:1 CuPS to 1:1 CuPS can greatly enhanced the reactivity of the catalyst [80]. Most of EA was converted to EtOH, and space-time yield of 1.23 g EtOH/(gcat*h) were achieved using the CuPS-cats prepared by Zn-NH4-CP method.

3.1.4. DMCD to CHDM

By using CuPSNTs, an excellent stability for more than 1000 h was found with unaltered space time yield of CHDM in DMCD hydrogenation (Figure 19) [107]. In contrast, CuPS-cat synthesized by using the AE method showed a 7.4% decreasing rate. The cooperative influence caused by the strong Cu-O-Si bonding and the confinement effect was thereby proposed.

3.1.5. Dimethyl Maleate (DMM) to Tetrahydrofuran (THF)

For the selective hydrogenation of DMM to THF (Figure 20), Han et al. [99] consolidated the synergic role of Cu+ and Cu0. The CuPS-cat prepared by AE method with both Cu+ and Cu0 exhibited an 8-fold yield of THF than that prepared by wet impregnation made counterpart containing mostly Cu0. Although Cu0 species could promote the hydrogenation of DMM with 100% of THF selectivity, conversion of DMM was trivial (approximately 10%). Cu+ was considered as an oxophilic, Lewis acidic site for the adsorption of carboxyl group of DMM.

3.2. EC to MeOH and EG

Hydrogenation of EC is a potential way to synthesize MeOH because of its thermodynamic favorability [105]. Moreover, not only MeOH but also EG could be synthesized concurrently (Figure 21).
Li et al. [108] investigated different weight loadings of Cu (2–30%) in CuPS-cats for the hydrogenation of EC. They found that the conversion of EC was proportional to the increase in Cu content and the surface area of Cu0, while maximized selectivities of MeOH and EG were obtained by using 10 wt % of Cu. An optimal Cu+–(Cu0 + Cu+) ratio was proposed at 0.51.
In another study of Li and coworkers [105], the dispersion of Cu on different types of mesoporous silica such as SBA-15, KIT-6, and MCM-41 were investigated. By using the same weight loading of Cu (10 wt %), they found that the dispersion of Cu on SBA-15 (26.2%), was much higher than those on KIT-6 (8.2%) and MCM-41 (6.8%). Although the ratio of Cu+–(Cu0 + Cu+) in the SBA-15 supported CuPS-cats was far away from the optimal ratio (0.51), the higher surface area of Cu (170 m2/g) likely played a more important role. After a 4 h reaction at 200 °C of temperature and 5 MPa of H2 pressure, EC was completely converted to 94.7% of EG in yield.

3.3. Carboxylic Acid to Alcohol

CuPS-cats were firstly deployed in the hydrogenation of AA to HDOL by our group [111]. CuPS-cats showed better performances in the hydrogenation of dicarboxylic than that of an impregnation-made Cu/SiO2 catalyst. Approximately 90% of HDOL yield together with an improved recyclability were achieved by using CuPS-cats. The follow-up work [119] demonstrated that the addition of nickel can stabilize Cu catalysts derived from hybrid phyllosilicates containing Cu and Ni in their octahedral units (Figure 22). The catalyst derived from Ni-doped CuPS is proposed to induce electron transfer from Cu cations to Ni cations due to a higher electronegativity of Ni than Cu, resulting in an improved activity and stability of hydrogenation. Introducing nickel seems to stabilize copper and to prevent copper leaching in acidic media. Furthermore, Ni-doped CuPS-cat showed an outstanding performance in hydrogenation of several carboxylic acids such as succinic (SUC), glutaric (GLU), hexanoic (HXA), pentanoic acid (PTA), and butyric acid (BTA) as well (Figure 23).
In another work of our group, CuPS-cats demonstrated an application in hydrogenation of levulinic acid (LA) to γ-valerolactone (GVL) [173]. A new strategy was applied to synthesize hydrophobic CuPS-cats: reduction−silylation−reduction. CuPS-cats grafted octyl triethoxysilane (OTS) on the surface showed a high LA conversion (95.7%) and GVL yield (85.2%) at a mild reaction condition (130 °C with a H2 pressure of 12 bar). The hydrophobic surface of OTS-grafted CuPS-cats was supposed to suppress Cu and Si leaching (Figure 24) and stability of OTS-grafted CuPS-cats, thus, was significantly improved (3 cycles with a 7.5% loss of initial activity). The reduction temperature may influence the activity of the catalyst. The optimal condition was to reduce at 350 °C then graft with OTS and repeat the reduction treatment again at 350 °C.

3.4. Aldehyde to Alcohol

Beside the hydrogenation of carboxylates and carboxylic acids, alcohol can be produced from the hydrogenation of aldehyde catalyzed by CuPS-cats. Taking FFAL [104,106] hydrogenation to FFOH for example, with proper catalyst design and operating condition, FFOH can be maximized as the desired product with limited extent of deep hydrogenation to 2-methylfuran (MF) [104,106] (Figure 25). In case of 5-hydroxymethylfurfural (HMF) [172] (Figure 25), the high conversion of HMF and 2,5-dihydroxymethylfuran (DHMF) yield can be obtained when UDP prepared CuPS-cats was applied.

3.5. CO2 Hydrogenation to Methanol (MeOH)

Many catalysts have been developed for conversion of CO2 into valuable chemicals such as CO [176,177], DME [178,179,180,181,182,183], CH4 [184,185,186] or higher hydrocarbons [187], especially, MeOH [72,188,189,190,191,192,193,194,195,196,197]. Copper-based catalysts were highly prospective in MeOH synthesis [188,189,190,191,193,196,197,198,199]. Wang et al. [70,71] proposed a simple and high-performance catalyst derived from CuPS. They claimed that tuning calcination and reduction temperatures can enhance Cu+–(Cu0 + Cu+) ratio of the reduced CuPS-cats usher in an improved hydrogenation activity. They indicated that Cu+ is the active site for the CO2 hydrogenation and the concentration of Cu+ can be maximized at 500 °C of calcination and at 300 °C of reduction (see figure). At Cu+–(Cu0 + Cu+) ratio close to 1, the CO2 conversion and MeOH selectivity reached 25–28% and about 21%, respectively, at 320 °C, 3.0 MPa, weight hour space velocity (WHSV) = 16 L·gcat−1·h−1, and H2:CO2 feed ratio of 4:1.

4. Deactivation

4.1. Causes of Deactivation

Analogous to supported Cu catalysts, CuPS-cats suffered from deactivation caused by sintering of Cu particles, leaching of Cu and Si, and carbonaceous species deposition. These negative impacts hamper the industrial practice of CuPS-cat mainly because of the unacceptable deactivation rate and poor recyclability.

4.1.1. Copper Sintering

Although CuPS-cats had good thermal stability, Cu sintering is frequently reported [76,77,79,91,95,96,101,102,107,167]. About 50% of copper surface area loss during reduction at 350 °C for 4 h [79] or at calcination temperature higher than 450 °C [102], mainly caused by Cu agglomeration. Gong et al. [96] also found the growth of Cu and Cu2O of post-reaction catalysts (Figure 26). In principle, sintering of Cu and Cu2O sintering occurs at temperature higher than their Hüttig temperatures (THüttig of Cu and Cu2O are 137 and 179 °C, respectively) and lower than their Tamman temperatures (TTamman of Cu and Cu2O are 405 and 481 °C, respectively) [77]. In this range of temperature, Cu atoms at edges and corners or defect positions become mobile. Driving force of the sintering was supposed to be caused by the Ostwald ripening [200]. Besides, the dynamic Cu state in the process of reaction was claimed to cause Cu sintering [95]. In DMO hydrogenation using CuPS-cat, He et al. [95] proposed that Cu0 was oxidized to Cu+ by DMO and Cu+ was reduced to Cu0 by H2 concurrently. The transition of each state enhanced Cu mobility and weakened Cu-SiO2 interaction, resulting in Cu agglomeration. The aggregation of Cu nanoparticles and the changes in distribution of surface Cu species were responsible for the deactivation of CuPS-cats [85,95,115,167].

4.1.2. Leaching of Copper and SiO2

In liquid medium, the deactivation caused not only by the sintering but also by the leaching. Wen et al. [201] and Popa et al. [79] indicated the dissolution of silica from the support evidenced by the formation of tetramethoxysilane (TMOS) in the resultant solution of DMO hydrogenation. ICP analysis indicated that more than 55% of Cu was leached out after three consecutive cycling runs of AA hydrogenation [111]. Deactivation of CuPS-cats was expected in hydrogenation of AA [119] and LA [175] accompanying the decline of Cu loading after each cycling run (Table 2).

4.1.3. Coke Deposition

Coking is a continuous carbon deposition process accompanying the evolution of the extent of reaction. Carbonaceous deposition or coking may be induced by many different reactions on acidity catalysts. In case of CuPS-cats, Cu+ incurred from the reduction in CuPS-cats was Lewis acidic, and can be a potential site of carbonization. Some studies indicated coke fouling on CuPS-cats in dry reforming of methane [136] and DME steam reforming [109]. Especially, a serious coking was reported after DEO hydrogenation reaction with approximately 20% of coke deposition on the used catalysts [92]. The decomposition of hydrocarbons into small active carbon species and accumulation of these carbonaceous species on the catalysts surface increases transfer limitations of reactants and products, and encapsulation of active sites by carbon causes CuPS-cats become less reactive. Fortunately, it is possible to rejuvenate catalyst activity by coke burning.

4.2. Remedial Actions

Sintering, Cu leaching, and coke deposition were responsible for CuPS-cats deactivation. Surpassing sintering and leaching are more complicated than coking because accumulation of carbon can be removed and the catalytic performance of the used CuPS-cats can be regenerated when treated in a flow of 30 vol% O2 at 500 °C for 30 min [202]. In order to inhibit the sintering of Cu, many strategies were proposed. Doping another metal such as Ag [159], La [167], Sn [110], Ce [85], or Zn [86] was supposed to stabilize the chemical state of Cu and alleviate Cu sintering. Binary oxides support such as SiO2-TiO2 [162], SiO2-MgO [163], or SiO2-CeO2 [164] was developed to enhance the Cu-support interaction and to maintain Cu dispersion as its fresh state. Moreover, spatial confinement effect of mesoporous silica [91,113] and CuPSNTs [107] allowed hosted Cu to possess high thermal stability. Surface modification by dextrin [83] and by glucose [203] and silylation [204] were found to hurdle the growth of Cu particles as well.
Leaching of Cu and silica might be caused by polarity of solvents and reactants. By amending the polarity, leaching extent of Cu might be eased. Transforming hydrophilic to hydrophobic surface of alkaline-grafted CuPS-cats was supposed to protect the catalyst from acid and water attacks, and ease the amounts of leached Cu [175]. Moreover, the solvent effect in inhibition of dissolution of SiO2 in DMO hydrogenation reaction was indicated [79,201]. When MeOH was substituted by EtOH as a solvent, the dissolution of SiO2 was inhibited in DMO hydrogenation [79]. The dissolution of SiO2 occurs sequentially by nucleophilic attack of methoxy or ethoxy, that are induced from hydrolysis of MeOH and EtOH, to form mono-, di-, tri-, and tetra-alkoxide silane. Although both methoxy and ethoxy are nucleophilic, stearic effect of ethyl group might hinder the attack of later ethoxy molecules while small molecules of MeOH are more accessible to silica surface. Therefore, to overcome the leaching of silica, EtOH is a more appropriate solvent than MeOH in DMO hydrogenation reaction.

5. Challenges and Prospective

The most challenges hindering CuPS-cats from practical application were the low stability and reusability. Despite high activity, Cu sinter and leaching during the reaction caused deactivation. Cu Sintering at high temperature limited application of CuPS-cats for reactions at moderate conditions. After reaction, CuPS-cats were deconstructed and cannot be regenerated. Our study indicated that CuPS-cats could not be regenerated after a reduction-oxidation cycle. So far, prolongation of the lifespan and stability of CuPS-cats has still challenged scientists.
Optimization of CuPS-cats from synthesis methods and promoters has been ploughed extensively. The next generation of CuPS-cats will mainly focus on the innovations of facile and environmentally friendly synthesis methods. For example, organic-inorganic hybrid [83] or surface modification [175,204]. Another possibility to improve CuPS-cat is the innovation of CuPS-cats with novel topologies such as CuPSNTs. Spatial confinement of CuPS-cat usually has a positive effect in hydrogenation reactions. Meanwhile, in-depth studies of CuPS-cats structure, reaction mechanisms, kinetics and their structure-activity correlation are limited. Only few studies were conducted, such as the observation of exsolution behavior of Cu species of CuPS by using TEM with an in situ manner under temperature-programmed reduction condition [173]. It seems the basic studies of Cu exsolution from the octahedral of CuPS is the key to clarify the unsolved questions of CuPS-cat catalysis. Advanced techniques, especially in situ methods used in Cu species characterization like TEM, XAS, and XAES, should be helpful.
Burgeoning research of CuPS-cats in the hydrogenation of other carbonyls such as carbonate [105,108,169,203,205], carboxylic acid [111,119,175], and aldehyde [104,106,160,172] have shown their potentials recently. Alternative applications of CuPS-cats, should be explored such steam reforming of DME [109,114,166], water-gas-shift reaction [206,207], dehydrogenation ethanol [98,202], dry reforming CH4 [136], ethynylation of formaldehyde [117]. Moreover, process design for large scale, continuous processing by using CuPS-cat as the catalyst is needed.

Author Contributions

Writing—original draft preparation, D.-T.T.; writing—review and editing, Y.-C.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology (Projects 109-2628-E-006-011-MY3, 110-2923-E-006-005-MY3, and 110-2927-I-006-506).

Data Availability Statement

Data available on request.

Acknowledgments

The authors appreciate Jyh-Fu Lee and Chi-Liang Chen (National Synchrotron Radiation Research Center, Taiwan) for XAFS spectra collection.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Nomenclature

AAAdipic acid
ACAmmonium carbonate deposition–precipitation method
AEAmmonia evaporation method
AEHAmmonia evaporation hydrothermal method
AHAmmonium chloride (NH4Cl) assisted hydrothermal method
AM30Sodium-stabilized silica
AS30Ammonium-stabilized silica
CHDM1,4-Cyclohexane dimethanol
CPCo-precipitation method
CuPSCopper phyllosilicate
CuPS-catsCopper phyllosilicate-derived catalysts
CuPSNTsCopper phyllosilicate nanotubes
DEMDiethyl malonate
DEODiethyl oxalate
DHMF2,5-dihydroxymethylfuran
DMCDDimethyl 1,4-cyclohexanedicarboxylate
DMMDimethyl maleate
DMODimethyl oxalate
EAEthyl acetate
ECEthylene carbonate
EGEthylene glycol
EtOHEthanol
FFALFurfural
GVLγ-valerolactone
HDOL1,6-Hexanediol
HMF5-Hydroxymethylfurfural
IMImpregnation
IWIIncipient wetness impregnation
LALevulinic acid
MAMethyl acetate
MeOHMethanol
MF2-Methylfuran
MGMethyl glycolate
m-SiO2Mesoporous silica
OMSOrdered mesoporous silica
OTSOctyl triethoxysilane
PDO1,3-propanediol
TEOSTetraethyl orthosilicate or tetraethoxysilane
THFTetrahydrofuran
UHDPUrea hydrolysis deposition-precipitation
UHUrea hydrolysis hydrothermal
USGUrea-assisted sol-gel

References

  1. Hsu, A.-F.; Shen, S.; Wu, E.; Foglia, T.A. Characterization of soybean lipoxygenase immobilized in cross- linked phyllosilicates. Biotechnol. Appl. Biochem. 1998, 28, 55–59. [Google Scholar] [PubMed]
  2. Shen, S.; Hsu, A.-F.; Foglia, T.A.; Tu, S.-I. Effectiveness of cross-linked phyllosilicates for intercalative immobilization of soybean lipoxygenase. Appl. Biochem. Biotechnol. 1998, 69, 79. [Google Scholar] [CrossRef] [PubMed]
  3. Shen, S.; Tu, S.I. Immobilization of horseradish peroxidase in cross-linked phyllosilicates: Conditions and characterizations. Biotechnol. Appl. Biochem. 1999, 29, 185–189. [Google Scholar]
  4. Hsu, A.-F.; Wu, E.; Shen, S.; Foglia, T.A.; Jones, K. Immobilized lipoxygenase in a packed-bed column bioreactor: Continuous oxygenation of linoleic acid. Biotechnol. Appl. Biochem. 1999, 30, 245–250. [Google Scholar] [PubMed]
  5. Hsu, A.-F.; Foglia, T.A.; Shen, S. Immobilization of Pseudomonas cepacia lipase in a phyllosilicate sol-gel matrix: Effectiveness as a biocatalyst. Biotechnol. Appl. Biochem. 2000, 31, 179–183. [Google Scholar] [CrossRef]
  6. Hsu, A.-F.; Jones, K.; Marmer, W.N.; Foglia, T.A. Production of alkyl esters from tallow and grease using lipase immobilized in a phyllosilicate sol-gel. J. Am. Oil Chem. Soc. 2001, 78, 585–588. [Google Scholar] [CrossRef]
  7. Hsu, A.-F.; Jones, K.; Foglia, T.A. Phyllosilicate sol-gel immobilized lipases for the formation of partial acylglycerides. Biotechnol. Lett. 2002, 24, 1161–1165. [Google Scholar] [CrossRef]
  8. Hsu, A.-F.; Jones, K.; Foglia, T.A.; Marmer, W.N. Immobilized lipase-catalysed production of alkyl esters of restaurant grease as biodiesel. Biotechnol. Appl. Biochem. 2002, 36, 181–186. [Google Scholar] [CrossRef] [PubMed]
  9. Whilton, N.T.; Burkett, S.L.; Mann, S. Hybrid lamellar nanocomposites based on organically functionalized magnesium phyllosilicate clays with interlayer reactivity. J. Mater. Chem. 1998, 8, 1927–1932. [Google Scholar] [CrossRef]
  10. Da Fonseca, M.G.; Airoldi, C. New layered inorganic–organic nanocomposites containing n-propylmercapto copper phyllosilicates. J. Mater. Chem. 2000, 10, 1457–1463. [Google Scholar] [CrossRef]
  11. Lomakin, S.M.; Usachev, S.V.; Koverzanova, E.V.; Shilkina, N.G.; Ruban, L.V.; Zaikov, G.E. Polymer nanocomposites and intercalated flame retardants. New approaches to reduce plastics combustibility. Oxid. Commun. 2002, 25, 165–183. [Google Scholar]
  12. Fujii, K.; Hayashi, S.; Kodama, H. Synthesis of an alkylammonium/magnesium phyllosilicate hybrid nanocomposite consisting of a smectite-like layer and organosiloxane layers. Chem. Mater. 2003, 15, 1189–1197. [Google Scholar] [CrossRef]
  13. Patil, A.J.; Muthusamy, E.; Mann, S. Fabrication of functional protein-organoclay lamellar nanocomposites by biomolecule-induced assembly of exfoliated aminopropyl-functionalized magnesium phyllosilicates. J. Mater. Chem. 2005, 15, 3838–3843. [Google Scholar] [CrossRef]
  14. Holmström, S.C.; Patil, A.J.; Butler, M.; Mann, S. Influence of polymer co-intercalation on guest release from aminopropyl-functionalized magnesium phyllosilicate mesolamellar nanocomposites. J. Mater. Chem. 2007, 17, 3894–3900. [Google Scholar] [CrossRef]
  15. Chabrol, K.; Gressier, M.; Pebere, N.; Menu, M.-J.; Martin, F.; Bonino, J.-P.; Marichal, C.; Brendle, J. Functionalization of synthetic talc-like phyllosilicates by alkoxyorganosilane grafting. J. Mater. Chem. 2010, 20, 9695–9706. [Google Scholar] [CrossRef] [Green Version]
  16. Richard-Plouet, M.; Vilminot, S. Ferromagnetism in talc-like cobalt-phyllosilicates. Solid State Sci. 1999, 1, 381–393. [Google Scholar] [CrossRef]
  17. Richard-Plouet, M.; Vilminot, S.; Guillot, M.; Kurmoo, M. Canted antiferromagnetism in an organo-modified layered nickel phyllosilicate. Chem. Mater. 2002, 14, 3829–3836. [Google Scholar] [CrossRef]
  18. Zhang, X.-Q.; Li, W.-C.; He, B.; Yan, N.; Xu, S.; Cao, Y.; Lu, A.-H. Ultrathin phyllosilicate nanosheets as anode materials with superior rate performance for lithium ion batteries. J. Mater. Chem. A 2018, 6, 1397–1402. [Google Scholar] [CrossRef]
  19. Razmjou, A.; Eshaghi, G.; Orooji, Y.; Hosseini, E.; Korayem, A.H.; Mohagheghian, F.; Boroumand, Y.; Noorbakhsh, A.; Asadnia, M.; Chen, V. Lithium ion-selective membrane with 2D subnanometer channels. Water Res. 2019, 159, 313–323. [Google Scholar] [CrossRef] [PubMed]
  20. Viseras, C.; Meeten, G.H.; Lopez-Galindo, A. Pharmaceutical grade phyllosilicate dispersions: The influence of shear history on floc structure. Int. J. Pharm. 1999, 182, 7–20. [Google Scholar] [CrossRef]
  21. VViseras, C.; López-Galindo, A. Characteristics of pharmaceutical grade phyllosilicate powders. Pharm. Dev. Technol. 2000, 5, 47–52. [Google Scholar] [CrossRef] [PubMed]
  22. Viseras, C.; Cerezo, P.; Meeten, G.; Lopez-Galindo, A. One-dimensional filtration of pharmaceutical grade phyllosilicate dispersions. Int. J. Pharm. 2001, 217, 201–213. [Google Scholar] [CrossRef]
  23. Kim, S.Y.; Choi, Y.-S. Preparation of magnesium-based two-dimensional phyllosilicate materials and simultaneous antioxidant drug intercalation. Colloids Surf. A Physicochem. Eng. Asp. 2019, 569, 164–170. [Google Scholar] [CrossRef]
  24. Lagadic, I.L.; Mitchell, M.K.; Payne, B.D. Highly effective adsorption of heavy metal tons by a thiol-functionalized magnesium phyllosilicate clay. Environ. Sci. Technol. 2001, 35, 984–990. [Google Scholar] [CrossRef]
  25. De Melo, M.A.; Pires, C.T.G.V.M.T.; Airoldi, C. The influence of the leaving iodine atom on phyllosilicate syntheses and useful application in toxic metal removal with favorable energetic effects. RSC Adv. 2014, 4, 41028–41038. [Google Scholar] [CrossRef] [Green Version]
  26. Sarr, A.; Mayura, K.; Kubena, L.; Harvey, R.; Phillips, T. Effects of phyllosilicate clay on the metabolic profile of aflatoxin B1 in Fischer-344 rats. Toxicol. Lett. 1995, 75, 145–151. [Google Scholar] [CrossRef]
  27. Han, H.-K.; Lee, Y.-C.; Lee, M.-Y.; Patil, A.J.; Shin, H.-J. Magnesium and calcium organophyllosilicates: Synthesis and in vitro cytotoxicity study. ACS Appl. Mater. Interfaces 2011, 3, 2564–2572. [Google Scholar] [CrossRef]
  28. Lee, Y.-C.; Kim, E.J.; Yang, J.-W.; Shin, H.-J. Removal of malachite green by adsorption and precipitation using aminopropyl functionalized magnesium phyllosilicate. J. Hazard. Mater. 2011, 192, 62–70. [Google Scholar] [CrossRef]
  29. Moscofian, A.S.O.; Pires, C.T.G.V.M.T.; Vieira, A.P.; Airoldi, C. Organofunctionalized magnesium phyllosilicates as mono- or bifunctitonal entities for industrial dyes removal. RSC Adv. 2012, 2, 3502–3511. [Google Scholar] [CrossRef] [Green Version]
  30. Alencar, J.M.; Oliveira, F.J.; Airoldi, C.; Filho, E.C.S. Organophilic nickel phyllosilicate for reactive blue dye removal. Chem. Eng. J. 2014, 236, 332–340. [Google Scholar] [CrossRef]
  31. Kim, Y.-W.; Kim, J.-H.; Moon, D.H.; Shin, H.-J. Adsorption and precipitation of anionic dye Reactive Red 120 from aqueous solution by aminopropyl functionalized magnesium phyllosilicate. Korean J. Chem. Eng. 2018, 36, 101–108. [Google Scholar] [CrossRef]
  32. Das, T.K.; Ganguly, S.; Ghosh, S.; Remanan, S.; Ghosh, S.K.; Das, N.C. In-situ synthesis of magnetic nanoparticle immobilized heterogeneous catalyst through mussel mimetic approach for the efficient removal of water pollutants. Colloid Interface Sci. Commun. 2019, 33, 100218. [Google Scholar] [CrossRef]
  33. Moura, K.O.; Pastore, H.O. Comparative adsorption of CO2 by mono-, di-, and triamino-organofunctionalized magnesium phyllosilicates. Environ. Sci. Technol. 2013, 47, 12201–12210. [Google Scholar] [CrossRef]
  34. Song, D.; Lee, Y.-C.; Bin Park, S.; Han, J.-I. Gaseous carbon dioxide conversion and calcium carbonate preparation by magnesium phyllosilicate. RSC Adv. 2014, 4, 4037–4040. [Google Scholar] [CrossRef]
  35. Song, D.; Lee, Y.-C.; Bin Park, S.; Han, J.-I. CO2 Capturing and Mineralization by (Mg)-Phyllosilicate Coated with Fabrics. J. Nanosci. Nanotechnol. 2016, 16, 5063–5069. [Google Scholar] [CrossRef]
  36. Jaber, M.; Miehé-Brendlé, J. Synthesis, characterization and applications of 2:1 phyllosilicates and organophyllosilicates: Contribution of fluoride to study the octahedral sheet. Microporous Mesoporous Mater. 2008, 107, 121–127. [Google Scholar] [CrossRef]
  37. McDonald, A.; Scott, B.; Villemure, G. Hydrothermal preparation of nanotubular particles of a 1:1 nickel phyllosilicate. Microporous Mesoporous Mater. 2009, 120, 263–266. [Google Scholar] [CrossRef]
  38. Fan, Q.; Li, P.; Pan, D. Chapter 1—Radionuclides sorption on typical clay minerals: Modeling and spectroscopies. In Interface Science and Technology; Chen, C., Ed.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 1–38. [Google Scholar]
  39. Onnainty, R.; Granero, G. Chapter 14—Chitosan-based nanocomposites: Promising materials for drug delivery applications. In Biomedical Applications of Nanoparticles; Grumezescu, A.M., Ed.; William Andrew Publishing: New York, NY, USA, 2019; pp. 375–407. [Google Scholar]
  40. Toupance, T.; Kermarec, M.; Lambert, J.-F.; Louis, C. Conditions of formation of copper phyllosilicates in silica-supported copper catalysts prepared by selective adsorption. J. Phys. Chem. B 2002, 106, 2277–2286. [Google Scholar] [CrossRef]
  41. Pérez-Maqueda, L.A.; Franco, F.; Pérez-Rodríguez, J.L. Comparative study of the sonication effect on the thermal behaviour of 1:1 and 2:1 aluminium phyllosilicate clays. J. Eur. Ceram. Soc. 2005, 25, 1463–1470. [Google Scholar] [CrossRef]
  42. Krasilin, A.A.; Khrapova, E.K.; Nominé, A.; Ghanbaja, J.; Belmonte, T.; Gusarov, V.V. Cation redistribution along the spiral of Ni-doped phyllosilicate nanoscrolls: Energy modelling and STEM/EDS study. ChemPhysChem 2019, 20, 719–726. [Google Scholar] [CrossRef] [PubMed]
  43. Park, J.C.; Kang, S.W.; Kim, J.-C.; Kwon, J.I.; Jang, S.; Rhim, G.B.; Kim, M.; Chun, D.H.; Lee, H.-T.; Jung, H.; et al. Synthesis of Co/SiO2 hybrid nanocatalyst via twisted Co3Si2O5(OH)4 nanosheets for high-temperature Fischer–Tropsch reaction. Nano Res. 2017, 10, 1044–1055. [Google Scholar] [CrossRef]
  44. Nam, B.; Lee, H.U.; Park, S.Y.; Son, B.-C.; Lee, G.-W.; Park, J.-Y.; Lee, Y.-C. Dual-end-functionalized tin (Sn)-phyllosilicates for the esterification of oleic acid. J. Ind. Eng. Chem. 2016, 41, 50–61. [Google Scholar] [CrossRef]
  45. Kang, H.G.; Lee, K.M.; Choi, S.; Nam, B.; Choi, S.A.; Lee, S.C.; Park, J.Y.; Lee, G.W.; Lee, H.U.; Lee, Y.C. Feasibility tests of-SO3H/-SO3--functionalized magnesium phyllosilicate [-SO3H/-SO3-MP] for environmental and bioenergy applications. RSC Adv. 2015, 5, 63271–63277. [Google Scholar] [CrossRef] [Green Version]
  46. Kim, J.S.; Park, I.; Jeong, E.-S.; Jin, K.; Seong, W.M.; Yoon, G.; Kim, H.; Kim, B.; Nam, K.T.; Kang, K. Amorphous cobalt phyllosilicate with layered crystalline motifs as water oxidation catalyst. Adv. Mater. 2017, 29, 1606893. [Google Scholar] [CrossRef]
  47. Kim, B.; Kim, J.S.; Kim, H.; Park, I.; Seong, W.M.; Kang, K.; Sung, W.M. Amorphous multinary phyllosilicate catalysts for electrochemical water oxidation. J. Mater. Chem. A 2019, 7, 18380–18387. [Google Scholar] [CrossRef]
  48. Zheng, X.; Zhang, X.; Wang, X.; Wang, S.; Wu, S. Preparation and characterization of CuO/CeO2 catalysts and their applications in low-temperature CO oxidation. Appl. Catal. A Gen. 2005, 295, 142–149. [Google Scholar] [CrossRef]
  49. Jia, A.-P.; Jiang, S.-Y.; Lu, J.-Q.; Luo, M.-F. Study of catalytic activity at the CuO−CeO2 interface for CO oxidation. J. Phys. Chem. C 2010, 114, 21605–21610. [Google Scholar] [CrossRef]
  50. Sun, S.; Mao, D.; Yu, J.; Yang, Z.; Lu, G.; Ma, Z. Low-temperature CO oxidation on CuO/CeO2 catalysts: The significant effect of copper precursor and calcination temperature. Catal. Sci. Technol. 2015, 5, 3166–3181. [Google Scholar] [CrossRef]
  51. Shang, H.; Zhang, X.; Xu, J.; Han, Y. Effects of preparation methods on the activity of CuO/CeO2 catalysts for CO oxidation. Front. Chem. Sci. Eng. 2017, 11, 603–612. [Google Scholar] [CrossRef]
  52. Wei, B.; Yang, N.; Pang, F.; Ge, J. Cu2O–CuO hollow nanospheres as a heterogeneous catalyst for synergetic oxidation of CO. J. Phys. Chem. C 2018, 122, 19524–19531. [Google Scholar] [CrossRef]
  53. Sivaiah, M.; Petit, S.; Beaufort, M.; Eyidi, D.; Barrault, J.; Batiot-Dupeyrat, C.; Valange, S. Nickel based catalysts derived from hydrothermally synthesized 1:1 and 2:1 phyllosilicates as precursors for carbon dioxide reforming of methane. Microporous Mesoporous Mater. 2011, 140, 69–80. [Google Scholar] [CrossRef]
  54. Li, Z.; Kathiraser, Y.; Ashok, J.; Oemar, U.; Kawi, S. Simultaneous tuning porosity and basicity of nickel@Nickel–magnesium phyllosilicate core–shell catalysts for CO2 reforming of CH4. Langmuir 2014, 30, 14694–14705. [Google Scholar] [CrossRef]
  55. Li, Z.; Kathiraser, Y.; Kawi, S. Facile synthesis of high surface area yolk-shell Ni@Ni embedded SiO2 via Ni phyllosilicate with enhanced performance for CO2 reforming of CH4. ChemCatChem 2014, 7, 160–168. [Google Scholar] [CrossRef]
  56. Yao, L.; Shi, J.; Hu, C. The structure, carbon deposition and stability of a ZrOx/Ni–MnOx/SiO2 catalyst for the CO2 reforming of methane. RSC Adv. 2015, 5, 90168–90177. [Google Scholar] [CrossRef]
  57. Zhao, B.; Chen, Z.; Yan, X.; Ma, X.; Hao, Q.-Q. CO methanation over Ni/SiO2 catalyst prepared by ammonia impregnation and plasma decomposition. Top. Catal. 2017, 60, 879–889. [Google Scholar] [CrossRef]
  58. Li, Z.W.; Sibudjing, K. Facile synthesis of multi-Ni-core@Ni phyllosilicate@CeO2 shell hollow spheres with high oxygen vacancy concentration for dry reforming of CH4. ChemCatChem 2018, 10, 2994–3001. [Google Scholar] [CrossRef]
  59. Li, Z.W.; Kawi, S. Multi-Ni@Ni phyllosilicate hollow sphere for CO2 reforming of CH4: Influence of Ni precursors on structure, sintering, and carbon resistance. Catal. Sci. Technol. 2018, 8, 1915–1922. [Google Scholar] [CrossRef]
  60. Li, Z.; Wang, Z.; Jiang, B.; Kawi, S. Sintering resistant Ni nanoparticles exclusively confined within SiO2 nanotubes for CH4 dry reforming. Catal. Sci. Technol. 2018, 8, 3363–3371. [Google Scholar] [CrossRef]
  61. Ye, R.-P.; Gong, W.; Sun, Z.; Sheng, Q.; Shi, X.; Wang, T.; Yao, Y.; Razink, J.J.; Lin, L.; Zhou, Z.; et al. Enhanced stability of Ni/SiO2 catalyst for CO2 methanation: Derived from nickel phyllosilicate with strong metal-support interactions. Energy 2019, 188, 116059. [Google Scholar] [CrossRef]
  62. Sarafraz, M.; Safaei, M.R.; Goodarzi, M.; Arjomandi, M. Reforming of methanol with steam in a micro-reactor with Cu–SiO2 porous catalyst. Int. J. Hydrogen Energy 2019, 44, 19628–19639. [Google Scholar] [CrossRef]
  63. Kim, W.; Mohaideen, K.K.; Seo, D.J.; Yoon, W.L. Methanol-steam reforming reaction over Cu-Al-based catalysts derived from layered double hydroxides. Int. J. Hydrogen Energy 2017, 42, 2081–2087. [Google Scholar] [CrossRef]
  64. Das, D.K.; Llorca, J.; Dominguez, M.; Colussi, S.; Trovarelli, A.; Gayen, A. Methanol steam reforming behavior of copper impregnated over CeO2–ZrO2 derived from a surfactant assisted coprecipitation route. Int. J. Hydrogen Energy 2015, 40, 10463–10479. [Google Scholar] [CrossRef] [Green Version]
  65. Tonelli, F.; Gorriz, O.; Tarditi, A.; Cornaglia, L.; Arrúa, L.; Abello, M.C. Activity and stability of a CuO/CeO2 catalyst for methanol steam reforming. Int. J. Hydrogen Energy 2015, 40, 13379–13387. [Google Scholar] [CrossRef]
  66. Bichon, P.; Asheim, M.; Jordal, T.S.A.; Fathi, M.; Holmen, A.; Blekkan, A. Hydrogen from methanol steam-reforming over Cu-based catalysts with and without Pd promotion. Int. J. Hydrogen Energy 2007, 32, 1799–1805. [Google Scholar] [CrossRef]
  67. Loaiza-Gil, A.; Arenas, J.; Villarroel, M.; Imbert, F.; Del Castillo, H.; Fontal, B. Heavier alcohols synthesis on cobalt phyllosilicate catalysts. J. Mol. Catal. A Chem. 2005, 228, 339–344. [Google Scholar] [CrossRef]
  68. Feyzi, M.; Nadri, S.; Joshaghani, M. Catalytic performance of Fe-Mn/SiO2 nanocatalysts for CO hydrogenation. J. Chem. 2013. [Google Scholar] [CrossRef]
  69. Zhang, F.; Li, Y.; Gao, S.; Fang, H.; Liang, X.; Yuan, Y. Synthesis of higher alcohols by CO hydrogenation on a K-promoted Ni–Mo catalyst derived from Ni–Mo phyllosilicate. Catal. Sci. Technol. 2018, 8, 4219–4228. [Google Scholar] [CrossRef]
  70. Wang, Z.-Q.; Xu, Z.-N.; Peng, S.-Y.; Zhang, M.-J.; Lu, G.; Chen, Q.-S.; Chen, Y.; Guo, G.-C. High-performance and long-lived Cu/SiO2 nanocatalyst for CO2 hydrogenation. ACS Catal. 2015, 5, 4255–4259. [Google Scholar] [CrossRef]
  71. Wang, Z.-Q.; Xu, Z.-N.; Zhang, M.-J.; Chen, Q.-S.; Chen, Y.; Guo, G.-C. Insight into composition evolution in the synthesis of high-performance Cu/SiO2 catalysts for CO2 hydrogenation. RSC Adv. 2016, 6, 25185–25190. [Google Scholar] [CrossRef]
  72. Richard, A.R.; Fan, M. Low-pressure hydrogenation of CO2 to CH3OH using Ni-In-Al/SiO2 catalyst synthesized via a phyllosilicate precursor. ACS Catal. 2017, 7, 5679–5692. [Google Scholar] [CrossRef]
  73. Chen, L.-F.; Guo, P.-J.; Qiao, M.; Yan, S.-R.; Li, H.-X.; Shen, W.; Xu, H.; Fan, K.-N. Cu/SiO2 catalysts prepared by the ammonia-evaporation method: Texture, structure, and catalytic performance in hydrogenation of dimethyl oxalate to ethylene glycol. J. Catal. 2008, 257, 172–180. [Google Scholar] [CrossRef]
  74. Xue, J.; Cui, F.; Huang, Z.; Zuo, J.; Chen, J.; Xia, C. Liquid phase hydrogenolysis of biomass-derived lactate to 1,2-propanediol over silica supported cobalt nanocatalyst. Chin. J. Chem. 2011, 29, 1319–1325. [Google Scholar] [CrossRef]
  75. Yin, A.; Guo, X.; Dai, W.-L.; Fan, K. Effect of initial precipitation temperature on the structural evolution and catalytic behavior of Cu/SiO2 catalyst in the hydrogenation of dimethyl oxalate. Catal. Commun. 2011, 12, 412–416. [Google Scholar] [CrossRef]
  76. Lin, H.; Zheng, X.; He, Z.; Zheng, J.; Duan, X.; Yuan, Y. Cu/SiO2 hybrid catalysts containing HZSM-5 with enhanced activity and stability for selective hydrogenation of dimethyl oxalate to ethylene glycol. Appl. Catal. A Gen. 2012, 287–296. [Google Scholar] [CrossRef]
  77. Yue, H.; Zhao, Y.; Zhao, L.; Lv, J.; Wang, S.; Gong, J.; Ma, X. Hydrogenation of dimethyl oxalate to ethylene glycol on a Cu/SiO2/cordierite monolithic catalyst: Enhanced internal mass transfer and stability. AIChE J. 2012, 58, 2798–2809. [Google Scholar] [CrossRef]
  78. Zhang, B.; Hui, S.; Zhang, S.; Ji, Y.; Li, W.; Fang, D. Effect of copper loading on texture, structure and catalytic performance of Cu/SiO2 catalyst for hydrogenation of dimethyl oxalate to ethylene glycol. J. Nat. Gas Chem. 2012, 21, 563–570. [Google Scholar] [CrossRef]
  79. Popa, T.; Zhang, Y.; Jin, E.; Fan, M. An environmentally benign and low-cost approach to synthesis of thermally stable industrial catalyst Cu/SiO2 for the hydrogenation of dimethyl oxalate to ethylene glycol. Appl. Catal. A Gen. 2015, 505, 52–61. [Google Scholar] [CrossRef] [Green Version]
  80. Di, W.; Cheng, J.; Tian, S.; Li, J.; Chen, J.; Sun, Q. Synthesis and characterization of supported copper phyllosilicate catalysts for acetic ester hydrogenation to ethanol. Appl. Catal. A Gen. 2016, 510, 244–259. [Google Scholar] [CrossRef]
  81. Zhao, Y.; Li, S.; Wang, Y.; Shan, B.; Zhang, J.; Wang, S.; Ma, X. Efficient tuning of surface copper species of Cu/SiO2 catalyst for hydrogenation of dimethyl oxalate to ethylene glycol. Chem. Eng. J. 2017, 313, 759–768. [Google Scholar] [CrossRef]
  82. Zhao, Y.; Zhang, Y.; Wang, Y.; Zhang, J.; Xu, Y.; Wang, S.; Ma, X. Structure evolution of mesoporous silica supported copper catalyst for dimethyl oxalate hydrogenation. Appl. Catal. A Gen. 2017, 539, 59–69. [Google Scholar] [CrossRef]
  83. Ye, R.-P.; Lin, L.; Yang, J.-X.; Sun, M.-L.; Li, F.; Li, B.; Yao, Y.-G. A new low-cost and effective method for enhancing the catalytic performance of Cu–SiO2 catalysts for the synthesis of ethylene glycol via the vapor-phase hydrogenation of dimethyl oxalate by coating the catalysts with dextrin. J. Catal. 2017, 350, 122–132. [Google Scholar] [CrossRef]
  84. Abbas, M.; Chen, Z.; Zhang, J.; Chen, J. Highly dispersed, ultra-small and noble metal-free Cu nanodots supported on porous SiO2 and their excellent catalytic hydrogenation of dimethyl oxalate to methyl glycolate. N. J. Chem. 2018, 42, 10290–10299. [Google Scholar] [CrossRef]
  85. Ai, P.; Tan, M.; Reubroycharoen, P.; Wang, Y.; Feng, X.; Liu, G.; Yang, G.; Tsubaki, N. Probing the promotional roles of cerium in the structure and performance of Cu/SiO2 catalysts for ethanol production. Catal. Sci. Technol. 2018, 8, 6441–6451. [Google Scholar] [CrossRef]
  86. Qi, W.; Ling, Q.; Ding, D.; Yazhong, C.; Chengwu, S.; Peng, C.; Ye, W.; Qinghong, Z.; Rong, L.; Hao, S. Performance enhancement of Cu/SiO2 catalyst for hydrogenation of dimethyl oxalate to ethylene glycol through zinc incorporation. Catal. Commun. 2018, 108, 68–72. [Google Scholar] [CrossRef]
  87. Sun, Y.; Meng, F.; Ge, Q.; Sun, J. Importance of the initial oxidation state of copper for the catalytic hydrogenation of dimethyl oxalate to ethylene glycol. Chem. Open 2018, 7, 969–976. [Google Scholar] [CrossRef] [PubMed]
  88. Wang, Y.; Yang, W.; Yao, D.; Wang, S.; Xu, Y.; Zhao, Y.; Ma, X. Effect of surface hydroxyl group of ultra-small silica on the chemical states of copper catalyst for dimethyl oxalate hydrogenation. Catal. Today 2020, 350, 127–135. [Google Scholar] [CrossRef]
  89. Huang, H.; Wang, B.; Wang, Y.; Zhao, Y.; Wang, S.; Ma, X. Partial hydrogenation of dimethyl oxalate on Cu/SiO2 catalyst modified by sodium silicate. Catal. Today 2020, 358, 68–73. [Google Scholar] [CrossRef]
  90. Yu, X.; Vest, T.A.; Gleason-Boure, N.; Karakalos, S.G.; Tate, G.L.; Burkholder, M.; Monnier, J.R.; Williams, C.T. Enhanced hydrogenation of dimethyl oxalate to ethylene glycol over indium promoted Cu/SiO2. J. Catal. 2019, 380, 289–296. [Google Scholar] [CrossRef]
  91. Ding, J.; Zhang, J.; Zhang, C.; Liu, K.; Xiao, H.; Kong, F.; Chen, J. Hydrogenation of diethyl oxalate over Cu/SiO2 catalyst with enhanced activity and stability: Contribution of the spatial restriction by varied pores of support. Appl. Catal. A Gen. 2015, 508, 68–79. [Google Scholar] [CrossRef]
  92. Ding, J.; Popa, T.; Tang, J.; Gasem, K.A.; Fan, M.; Zhong, Q. Highly selective and stable Cu/SiO2 catalysts prepared with a green method for hydrogenation of diethyl oxalate into ethylene glycol. Appl. Catal. B Environ. 2017, 209, 530–542. [Google Scholar] [CrossRef] [Green Version]
  93. Toupance, T.†; Kermarec, A.M.; Louis, C. Metal particle size in silica-supported copper catalysts. Influence of the conditions of preparation and of thermal pretreatments. J. Phys. Chem. B 2000, 104, 965–972. [Google Scholar] [CrossRef]
  94. He, L.; Chen, X.; Ma, J.; He, H.; Wang, W. Characterization and catalytic performance of sol–gel derived Cu/SiO2 catalysts for hydrogenolysis of diethyl oxalate to ethylene glycol. J. Sol-Gel Sci. Technol. 2010, 55, 285–292. [Google Scholar] [CrossRef]
  95. He, Z.; Lin, H.; He, P.; Yuan, Y. Effect of boric oxide doping on the stability and activity of a Cu–SiO2 catalyst for vapor-phase hydrogenation of dimethyl oxalate to ethylene glycol. J. Catal. 2011, 277, 54–63. [Google Scholar] [CrossRef]
  96. Gong, J.; Yue, H.; Zhao, Y.; Zhao, S.; Zhao, L.; Lv, J.; Wang, S.; Ma, X. Synthesis of ethanol via syngas on Cu/SiO2 satalysts with balanced Cu0–Cu+ sites. J. Am. Chem. Soc. 2012, 134, 13922–13925. [Google Scholar] [CrossRef] [PubMed]
  97. Yue, H.; Zhao, Y.; Zhao, S.; Wang, B.; Ma, X.; Gong, J. A copper-phyllosilicate core-sheath nanoreactor for carbon–oxygen hydrogenolysis reactions. Nat. Commun. 2013, 4, 2339. [Google Scholar] [CrossRef]
  98. Yu, X.; Zhai, S.; Zhu, W.; Gao, S.; Yan, J.; Yuan, H.; Chen, L.; Luo, J.; Zhang, W.; Wang, Z. The direct transformation of ethanol to ethyl acetate over Cu/SiO2 catalysts that contain copper phyllosilicate. J. Chem. Sci. 2014, 126, 1013–1020. [Google Scholar] [CrossRef]
  99. Han, X.-Q.; Zhang, Q.-F.; Feng, F.; Lu, C.-S.; Ma, L.; Li, X.-N. Selective hydrogenation of dimethyl maleate to tetrahydrofuran over Cu/SiO2 catalyst: Effect of Cu+ on the catalytic performance. Chin. Chem. Lett. 2015, 26, 1150–1154. [Google Scholar] [CrossRef]
  100. Wang, Y.; Shen, Y.; Zhao, Y.; Lv, J.; Wang, S.; Ma, X. Insight into the balancing effect of active Cu species for hydrogenation of carbon–oxygen bonds. ACS Catal. 2015, 5, 6200–6208. [Google Scholar] [CrossRef]
  101. Zhu, S.; Gao, X.; Zhu, Y.; Fan, W.; Wang, J.; Li, Y. A highly efficient and robust Cu/SiO2 catalyst prepared by the ammonia evaporation hydrothermal method for glycerol hydrogenolysis to 1,2-propanediol. Catal. Sci. Technol. 2015, 5, 1169–1180. [Google Scholar] [CrossRef]
  102. Ding, T.; Tian, H.; Liu, J.; Wu, W.; Yu, J. Highly active Cu/SiO2 catalysts for hydrogenation of diethyl malonate to 1,3-propanediol. Chin. J. Catal. 2016, 37, 484–493. [Google Scholar] [CrossRef]
  103. Ding, T.; Tian, H.; Liu, J.; Wu, W.; Zhao, B. Effect of promoters on hydrogenation of diethyl malonate to 1,3-propanediol over nano copper-based catalysts. Catal. Commun. 2016, 74, 10–15. [Google Scholar] [CrossRef]
  104. Dong, F.; Ding, G.; Zheng, H.; Xiang, X.; Chen, L.; Zhu, Y.; Li, Y. Highly dispersed Cu nanoparticles as an efficient catalyst for the synthesis of the biofuel 2-methylfuran. Catal. Sci. Technol. 2016, 6, 767–779. [Google Scholar] [CrossRef]
  105. Li, F.; Wang, L.; Han, X.; He, P.; Cao, Y.; Li, H. Influence of support on the performance of copper catalysts for the effective hydrogenation of ethylene carbonate to synthesize ethylene glycol and methanol. RSC Adv. 2016, 6, 45894–45906. [Google Scholar] [CrossRef]
  106. Dong, F.; Zhu, Y.; Zhao, H.; Tang, Z. Ratio-controlled synthesis of phyllosilicate-like materials as precursors for highly efficient catalysis of the formyl group. Catal. Sci. Technol. 2017, 7, 1880–1891. [Google Scholar] [CrossRef]
  107. Gong, X.; Wang, M.; Fang, H.; Qian, X.; Ye, L.; Duan, X.; Yuan, Y. Copper nanoparticles socketed in situ into copper phyllosilicate nanotubes with enhanced performance for chemoselective hydrogenation of esters. Chem. Commun. 2017, 53, 6933–6936. [Google Scholar] [CrossRef]
  108. Li, F.; Wang, L.; Han, X.; Cao, Y.; He, P.; Li, H. Selective hydrogenation of ethylene carbonate to methanol and ethylene glycol over Cu/SiO2 catalysts prepared by ammonia evaporation method. Int. J. Hydrogen Energy 2017, 42, 2144–2156. [Google Scholar] [CrossRef]
  109. Wang, X.; Ma, K.; Guo, L.; Tian, Y.; Cheng, Q.; Bai, X.; Huang, J.; Ding, T.; Li, X. Cu/ZnO/SiO2 catalyst synthesized by reduction of ZnO-modified copper phyllosilicate for dimethyl ether steam reforming. Appl. Catal. A Gen. 2017, 540, 37–46. [Google Scholar] [CrossRef]
  110. Zhang, C.; Wang, D.; Dai, B. Promotive effect of Sn2+ on Cu0/Cu+ ratio and stability evolution of Cu/SiO2 catalyst in the hydrogenation of dimethyl oxalate. Catalysts 2017, 7, 122. [Google Scholar] [CrossRef] [Green Version]
  111. Jiang, J.W.; Tu, C.C.; Chen, C.H.; Lin, Y.C. Highly selective silica-supported copper catalysts derived from copper phyllosilicates in the hydrogenation of adipic acid to 1,6-hexanediol. ChemCatChem 2018, 10, 5449–5458. [Google Scholar] [CrossRef]
  112. Ye, T.; Zhe, C.H.; Ning, X. Effect of silica particle size on texture, structure, and catalytic performance of Cu/SiO2 catalysts for glycerol hydrogenolysis. Russ. J. Phys. Chem. A 2018, 92, 449–455. [Google Scholar]
  113. Xu, C.; Chen, G.; Zhao, Y.; Liu, P.; Duan, X.; Gu, L.; Fu, G.; Yuan, Y.; Zheng, N. Interfacing with silica boosts the catalysis of copper. Nat. Commun. 2018, 9, 1–10. [Google Scholar] [CrossRef] [Green Version]
  114. Huang, J.; Ding, T.; Ma, K.; Cai, J.; Sun, Z.; Tian, Y.; Jiang, Z.; Zhang, J.; Zheng, L.; Li, X. Modification of Cu/SiO2 catalysts by La2O3 to quantitatively tune Cu+-Cu0 dual sites with improved catalytic activities and stabilities for dimethyl ether steam reforming. ChemCatChem 2018, 10, 3862–3871. [Google Scholar] [CrossRef]
  115. Zhang, Y.; Ye, C.; Guo, C.; Gan, C.; Tong, X. In2O3-modified Cu/SiO2 as an active and stable catalyst for the hydrogenation of methyl acetate to ethanol. Chin. J. Catal. 2018, 39, 99–108. [Google Scholar] [CrossRef]
  116. Zhao, Y.; Shan, B.; Wang, Y.; Zhou, J.; Wang, S.; Ma, X. An effective CuZn–SiO2 bimetallic catalyst prepared by hydrolysis precipitation method for the hydrogenation of methyl acetate to ethanol. Ind. Eng. Chem. Res. 2018, 57, 4526–4534. [Google Scholar] [CrossRef]
  117. Li, H.; Ban, L.; Wang, Z.; Meng, P.; Zhang, Y.; Wu, R.; Zhao, Y. Regulation of Cu species in CuO/SiO2 and its structural evolution in ethynylation reaction. Nanomaterials 2019, 9, 842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Scotti, N.; Finocchio, E.; Evangelisti, C.; Marelli, M.; Psaro, R.; Ravasio, N.; Zaccheria, F. Some insight on the structure/activity relationship of metal nanoparticles in Cu/SiO2 catalysts. Chin. J. Catal. 2019, 40, 1788–1794. [Google Scholar] [CrossRef]
  119. Tu, C.-C.; Tsou, Y.-J.; To, T.D.; Chen, C.-H.; Lee, J.-F.; Huber, G.W.; Lin, Y.-C. Phyllosilicate-derived CuNi/SiO2 catalysts in the selective hydrogenation of adipic acid to 1,6-hexanediol. ACS Sustain. Chem. Eng. 2019, 7, 17872–17881. [Google Scholar] [CrossRef]
  120. Wang, C.; Tian, Y.; Wu, R.; Li, H.; Yao, B.; Zhao, Y.; Xiao, T. Bimetallic synergy effects of phyllosilicate-derived NiCu@SiO2 catalysts for 1,4-butynediol direct hydrogenation to 1,4-butanediol. ChemCatChem 2019, 11, 4777–4787. [Google Scholar] [CrossRef]
  121. Gómez-Reynoso, R.; Ramirez, J.; Nares, R.; Luna, R.; Murrieta, F. Characterization and catalytic activity of Ni/SBA-15, synthesized by deposition-precipitation. Catal. Today 2005, 107–108, 926–932. [Google Scholar] [CrossRef]
  122. Nares, R.; Ramírez, J.; Gutiérrez-Alejandre, A.; Cuevas, R.; Ramı́rez, J.; Cuevas-García, R. Characterization and hydrogenation activity of Ni/Si(Al)- MCM-41 catalysts prepared by deposition−precipitation. Ind. Eng. Chem. Res. 2009, 48, 1154–1162. [Google Scholar] [CrossRef]
  123. Gabrovska, M.; Krstić, J.B.; Tzvetkov, P.; Tenchev, K.; Shopska, M.; Vukelić, N.; Jovanović, D. Effect of the support and the reduction temperature on the formation of metallic nickel phase in Ni/silica gel precursors of vegetable oil hydrogenation catalysts. Russ. J. Phys. Chem. A 2011, 85, 2392–2398. [Google Scholar] [CrossRef]
  124. Kim, M.; Park, J.C.; Kim, A.; Park, K.H.; Song, H. Porosity control of Pd@SiO2 yolk-shell nanocatalysts by the formation of nickel phyllosilicate and its influence on Suzuki coupling reactions. Langmuir 2012, 28, 6441–6447. [Google Scholar] [CrossRef] [PubMed]
  125. Zhang, C.; Yue, H.; Huang, Z.; Li, S.; Wu, G.; Ma, X.; Gong, J. Hydrogen production via steam reforming of ethanol on phyllosilicate-derived Ni/SiO2: Enhanced metal–support interaction and catalytic stability. ACS Sustain. Chem. Eng. 2013, 1, 161–173. [Google Scholar] [CrossRef]
  126. Chen, B.-H.; Liu, W.; Li, A.; Liu, Y.-J.; Chao, Z.-S. A simple and convenient approach for preparing core–shell-like silica@nickel species nanoparticles: Highly efficient and stable catalyst for the dehydrogenation of 1,2-cyclohexanediol to catechol. Dalton Trans. 2014, 44, 1023–1038. [Google Scholar] [CrossRef] [PubMed]
  127. Kong, X.; Zhu, Y.; Zheng, H.; Li, X.; Zhu, Y.; Li, Y.-W. Ni nanoparticles inlaid nickel phyllosilicate as a metal–acid bifunctional catalyst for low-temperature hydrogenolysis reactions. ACS Catal. 2015, 5, 5914–5920. [Google Scholar] [CrossRef]
  128. Yang, M.; Jin, P.; Fan, Y.; Huang, C.; Zhang, N.; Weng, W.; Chen, M.; Wan, H. Ammonia-assisted synthesis towards a phyllosilicate-derived highly-dispersed and long-lived Ni/SiO2 catalyst. Catal. Sci. Technol. 2015, 5, 5095–5099. [Google Scholar] [CrossRef]
  129. Zhang, Q.; Wang, M.; Zhang, T.; Wang, Y.; Tang, X.; Ning, P. A stable Ni/SBA-15 catalyst prepared by the ammonia evaporation method for dry reforming of methane. RSC Adv. 2015, 5, 94016–94024. [Google Scholar] [CrossRef]
  130. Ashok, J.; Ang, M.L.; Terence, P.Z.L.; Kawi, S. Promotion of the water-gas-shift reaction by nickel hydroxyl species in partially reduced nickel-containing phyllosilicate catalysts. ChemCatChem 2016, 8, 1308–1318. [Google Scholar] [CrossRef]
  131. Ma, B.; Cui, H.; Wang, D.; Wu, P.; Zhao, C. Controllable hydrothermal synthesis of Ni/H-BEA with a hierarchical core–shell structure and highly enhanced biomass hydrodeoxygenation performance. Nanoscale 2017, 9, 5986–5995. [Google Scholar] [CrossRef]
  132. Ma, B.; Cui, H.; Zhao, C. A nickel-phyllosilicate core–echinus catalyst via a green and base additive free hydrothermal approach for hydrogenation reactions. Chem. Commun. 2017, 53, 10358–10361. [Google Scholar] [CrossRef]
  133. Rodriguez-Gomez, A.; Caballero, A. Identification of outer and inner nickel particles in a mesoporous support: How the channels modify the reducibility of Ni/SBA-15 catalysts. ChemNanoMat 2017, 3, 94–97. [Google Scholar] [CrossRef]
  134. Bian, Z.; Li, Z.; Ashok, J.; Kawi, S. A highly active and stable Ni–Mg phyllosilicate nanotubular catalyst for ultrahigh temperature water-gas shift reaction. Chem. Commun. 2015, 51, 16324–16326. [Google Scholar] [CrossRef] [PubMed]
  135. Ashok, J.; Kathiraser, Y.; Ang, M.L.; Kawi, S. Ni and/or Ni–Cu alloys supported over SiO2 catalysts synthesized via phyllosilicate structures for steam reforming of biomass tar reaction. Catal. Sci. Technol. 2015, 5, 4398–4409. [Google Scholar] [CrossRef]
  136. Wu, T.; Zhang, Q.; Cai, W.; Zhang, P.; Song, X.; Sun, Z.; Gao, L. Phyllosilicate evolved hierarchical Ni- and Cu–Ni/SiO2 nanocomposites for methane dry reforming catalysis. Appl. Catal. A Gen. 2015, 503, 94–102. [Google Scholar] [CrossRef]
  137. Bian, Z.; Kawi, S. Highly carbon-resistant Ni–Co/SiO2 catalysts derived from phyllosilicates for dry reforming of methane. J. CO2 Util. 2017, 18, 345–352. [Google Scholar] [CrossRef]
  138. Wang, Z.; Ashok, J.; Pu, Z.; Kawi, S. Low temperature partial oxidation of methane via BaBi0.05Co0.8Nb0.15O3−δ-Ni phyllosilicate catalytic hollow fiber membrane reactor. Chem. Eng. J. 2017, 315, 315–323. [Google Scholar] [CrossRef]
  139. Bian, Z.; Kawi, S. Sandwich-like silica@Ni@Silica multicore–shell catalyst for the low-temperature dry reforming of methane: confinement effect against carbon formation. ChemCatChem 2017, 10, 320–328. [Google Scholar] [CrossRef]
  140. Li, Z.; Jiang, B.; Wang, Z.; Kawi, S. High carbon resistant Ni@Ni phyllosilicate@SiO2 core shell hollow sphere catalysts for low temperature CH4 dry reforming. J. CO2 Util. 2018, 27, 238–246. [Google Scholar] [CrossRef]
  141. Soghrati, E.; Ong, T.K.C.; Poh, C.K.; Kawi, S.; Borgna, A. Zeolite–supported nickel phyllosilicate catalyst for CO hydrogenolysis of cyclic ethers and polyols. Appl. Catal. B Environ. 2018, 235, 130–142. [Google Scholar] [CrossRef]
  142. Wang, X.; Zhu, S.; Wang, S.; Wang, J.; Fan, W.; Lv, Y. Ni nanoparticles entrapped in nickel phyllosilicate for selective hydrogenation of guaiacol to 2-methoxycyclohexanol. Appl. Catal. A Gen. 2018, 568, 231–241. [Google Scholar] [CrossRef]
  143. Ashok, J.; Bian, Z.; Wang, Z.; Kawi, S. Ni-phyllosilicate structure derived Ni-SiO2-MgO catalysts for bi-reforming applications: Acidity, basicity and thermal stability. Catal. Sci. Technol. 2018, 8, 1730–1742. [Google Scholar] [CrossRef]
  144. Fujita, T.; Peng, X.; Yamaguchi, A.; Cho, Y.; Zhang, Y.; Higuchi, K.; Yamamoto, Y.; Tokunaga, T.; Arai, S.; Miyauchi, M.; et al. Nanoporous Nickel Composite Catalyst for the Dry Reforming of Methane. ACS Omega 2018, 3, 16651–16657. [Google Scholar] [CrossRef]
  145. Das, S.; Ashok, J.; Bian, Z.; Dewangan, N.; Wai, M.; Du, Y.; Borgna, A.; Hidajat, K.; Kawi, S. Silica–Ceria sandwiched Ni core–shell catalyst for low temperature dry reforming of biogas: Coke resistance and mechanistic insights. Appl. Catal. B Environ. 2018, 230, 220–236. [Google Scholar] [CrossRef]
  146. Ghiat, I.; Boudjemaa, A.; Saadi, A.; Bachari, K.; Coville, N.J. Efficient hydrogen generation over a novel Ni phyllosilicate photocatalyst. J. Photochem. Photobiol. A Chem. 2019, 382, 111952. [Google Scholar] [CrossRef]
  147. Tan, J.J.; Xia, X.; Cui, J.; Yan, W.; Jiang, Z.; Zhao, Y. Efficient tuning of surface nickel species of the Ni-phyllosilicate catalyst for the hydrogenation of maleic anhydride. J. Phys. Chem. C 2019, 123, 9779–9787. [Google Scholar] [CrossRef]
  148. Wang, M.; Qian, X.; Xie, L.; Fang, H.; Ye, L.; Duan, X.; Yuan, Y. Synthesis of a Ni phyllosilicate with controlled morphology for deep hydrogenation of polycyclic aromatic hydrocarbons. ACS Sustain. Chem. Eng. 2019, 7, 1989–1997. [Google Scholar] [CrossRef]
  149. Wang, T.; Liu, C.; Ma, X.; Zhu, W.; Lv, X.; Zhang, H. Synthesis of Ni3Si4O10(OH)2 porous microspheres as support of Pd catalyst for hydrogenation reaction. Nanomaterials 2019, 9, 998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Wang, Z.; Li, Z.; Cui, Y.; Chen, T.; Hu, J.; Kawi, S. Highly efficient NO decomposition via dual-functional catalytic perovskite hollow fiber membrane reactor coupled with partial oxidation of methane at medium-low temperature. Environ. Sci. Technol. 2019, 53, 9937–9946. [Google Scholar] [CrossRef] [PubMed]
  151. Yang, F.; Wang, H.; Han, J.; Ge, Q.; Zhu, X. Enhanced selective deoxygenation of m-cresol to toluene on Ni/SiO2 catalysts derived from nickel phyllosilicate. Catal. Today 2019, 330, 149–156. [Google Scholar] [CrossRef]
  152. Shi, D.C.; Yang, Q.F.; Peterson, C.; Lamic-Humblot, A.F.; Girardon, J.S.; Griboval-Constant, A.; Stievano, L.; Sougrati, M.T.; Briois, V.; Bagot, P.A.J.; et al. Bimetallic Fe-Ni/SiO2 catalysts for furfural hydrogenation: Identification of the interplay between Fe and Ni during deposition-precipitation and thermal treatments. Catal. Today 2019, 334, 162–172. [Google Scholar] [CrossRef]
  153. Watari, R.; Kayaki, Y. Copper Catalysts Unleashing the Potential for Hydrogenation of Carbon−Oxygen Bonds. Asian J. Org. Chem. 2018, 7, 2005–2014. [Google Scholar] [CrossRef]
  154. Ye, R.-P.; Lin, L.; Li, Q.; Zhou, Z.; Wang, T.; Russell, C.K.; Adidharma, H.; Xu, Z.; Yao, Y.-G.; Fan, M. Recent progress in improving the stability of copper-based catalysts for hydrogenation of carbon–oxygen bonds. Catal. Sci. Technol. 2018, 8, 3428–3449. [Google Scholar] [CrossRef]
  155. Bian, Z.; Kawi, S. Preparation, characterization and catalytic application of phyllosilicate: A review. Catal. Today 2020, 339, 3–23. [Google Scholar] [CrossRef]
  156. Yao, D.; Wang, Y.; Li, Y.; Zhao, Y.; Lv, J.; Ma, X. A high-performance nanoreactor for carbon–oxygen bond hydrogenation reactions achieved by the morphology of nanotube-assembled hollow spheres. ACS Catal. 2018, 8, 1218–1226. [Google Scholar] [CrossRef]
  157. ZZhu, J.; Ye, Y.; Tang, Y.; Chen, L.; Tang, K. Efficient hydrogenation of dimethyl oxalate to ethylene glycol via nickel stabilized copper catalysts. RSC Adv. 2016, 6, 111415–111420. [Google Scholar] [CrossRef]
  158. Dong, X.; Ma, X.; Xu, H.; Ge, Q. Comparative study of silica-supported copper catalysts prepared by different methods: Formation and transition of copper phyllosilicate. Catal. Sci. Technol. 2016, 6, 4151–4158. [Google Scholar] [CrossRef]
  159. Huang, Y.; Ariga, H.; Zheng, X.; Duan, X.; Takakusagi, S.; Asakura, K.; Yuan, Y. Silver-modulated SiO2-supported copper catalysts for selective hydrogenation of dimethyl oxalate to ethylene glycol. J. Catal. 2013, 307, 74–83. [Google Scholar] [CrossRef]
  160. Li, B.; Li, L.; Sun, H.; Zhao, C. Selective deoxygenation of aqueous furfural to 2-methylfuran over Cu0/Cu2O·SiO2 sites via a copper phyllosilicate precursor without extraneous gas. ACS Sustain. Chem. Eng. 2018, 6, 12096–12103. [Google Scholar] [CrossRef]
  161. Ciotonea, C.; Dragoi, B.; Ungureanu, A.; Chirieac, A.; Petit, S.; Royer, S.; Dumitriu, E. Nanosized transition metals in controlled environments of phyllosilicate–mesoporous silica composites as highly thermostable and active catalysts. Chem. Commun. 2013, 49, 7665–7667. [Google Scholar] [CrossRef]
  162. Wen, C.; Yin, A.; Cui, Y.; Yang, X.; Dai, W.-L.; Fan, K. Enhanced catalytic performance for SiO2–TiO2 binary oxide supported Cu-based catalyst in the hydrogenation of dimethyloxalate. Appl. Catal. A Gen. 2013, 458, 82–89. [Google Scholar] [CrossRef]
  163. Qin, H.; Guo, C.; Sun, C.; Zhang, J. Influence of the support composition on the hydrogenation of methyl acetate over Cu/MgO-SiO2 catalysts. J. Mol. Catal. A Chem. 2015, 409, 79–84. [Google Scholar] [CrossRef]
  164. Ye, C.-L.; Guo, C.-L.; Zhang, J.-L. Highly active and stable CeO2–SiO2 supported Cu catalysts for the hydrogenation of methyl acetate to ethanol. Fuel Process. Technol. 2016, 143, 219–224. [Google Scholar] [CrossRef]
  165. Zhu, S.; Gao, X.; Zhu, Y.; Zhu, Y.; Zheng, H.; Li, Y. Promoting effect of boron oxide on Cu/SiO2 catalyst for glycerol hydrogenolysis to 1,2-propanediol. J. Catal. 2013, 303, 70–79. [Google Scholar] [CrossRef]
  166. Ma, K.; Cui, Z.; Zhang, Z.; Huang, J.; Sun, Z.; Tian, Y.; Ding, T.; Li, X. Alloy-mediated ultra-low CO selectivity for steam reforming over Cu−Ni bimetallic catalysts. ChemCatChem 2018, 10, 4010–4017. [Google Scholar] [CrossRef]
  167. Zheng, X.; Lin, H.; Zheng, J.; Duan, X.; Yuan, Y. Lanthanum oxide-modified Cu/SiO2 as a high-performance catalyst for chemoselective hydrogenation of dimethyl oxalate to ethylene glycol. ACS Catal. 2013, 3, 2738–2749. [Google Scholar] [CrossRef]
  168. Yang, Y.; Yao, D.; Zhang, M.; Li, A.; Gao, Y.; Fayisa, B.A.; Wang, M.-Y.; Huang, S.; Wang, Y.; Ma, X. Efficient hydrogenation of CO2-derived ethylene carbonate to methanol and ethylene glycol over Mo-doped Cu/SiO2 catalyst. Catal. Today 2020. [Google Scholar] [CrossRef]
  169. Li, H.; Cui, Y.; Liu, Y.; Zhang, L.; Zhang, Q.; Zhang, J.; Dai, W.-L. Highly efficient Ag-modified copper phyllosilicate nanotube: Preparation by co-ammonia evaporation hydrothermal method and application in the selective hydrogenation of carbonate. J. Mater. Sci. Technol. 2020, 47, 29–37. [Google Scholar] [CrossRef]
  170. Yao, D.; Wang, Y.; Hassan-Legault, K.; Li, A.; Zhao, Y.; Lv, J.; Huang, S.; Ma, X. Balancing effect between adsorption and diffusion on catalytic performance inside hollow nanostructured catalyst. ACS Catal. 2019, 9, 2969–2976. [Google Scholar] [CrossRef]
  171. Zhao, S.; Yue, H.; Zhao, Y.; Wang, B.; Geng, Y.; Lv, J.; Wang, S.; Gong, J.; Ma, X. Chemoselective synthesis of ethanol via hydrogenation of dimethyl oxalate on Cu/SiO2: Enhanced stability with boron dopant. J. Catal. 2013, 297, 142–150. [Google Scholar] [CrossRef]
  172. Zhu, Y.; Kong, X.; Yin, J.; You, R.; Zhang, B.; Zheng, H.; Wen, X.; Zhu, Y.; Li, Y.-W. Covalent-bonding to irreducible SiO2 leads to high-loading and atomically dispersed metal catalysts. J. Catal. 2017, 353, 315–324. [Google Scholar] [CrossRef]
  173. Tsou, Y.-J.; To, T.D.; Chiang, Y.-C.; Lee, J.-F.; Kumar, R.; Chung, P.-W.; Lin, Y.-C. Hydrophobic copper catalysts derived from copper phyllosilicates in the hydrogenation of levulinic acid to γ-valerolactone. ACS Appl. Mater. Interfaces 2020, 12, 54851–54861. [Google Scholar] [CrossRef] [Green Version]
  174. van den Berg, R.; Elkjaer, C.F.; Gommes, C.J.; Chorkendorff, I.; Sehested, J.; de Jongh, K.P.; Helveg, S.; Elkjær, C.F. Revealing the formation of copper nanoparticles from a homogeneous solid precursor by electron microscopy. J. Am. Chem. Soc. 2016, 138, 3433–3442. [Google Scholar]
  175. Yue, H.; Zhao, Y.; Ma, X.; Gong, J. Ethylene glycol: Properties, synthesis, and applications. Chem. Soc. Rev. 2012, 41, 4218–4244. [Google Scholar]
  176. Gonçalves, R.V.; Vono, L.L.; Wojcieszak, R.; Dias, C.S.; Wender, H.; Teixeira-Neto, E.; Rossi, L.M. Selective hydrogenation of CO2 into CO on a highly dispersed nickel catalyst obtained by magnetron sputtering deposition: A step towards liquid fuels. Appl. Catal. B Environ. 2017, 209, 240–246. [Google Scholar] [CrossRef]
  177. Zhao, B.; Yan, B.; Jiang, Z.; Yao, S.; Liu, Z.; Wu, Q.; Ran, R.; Senanayake, S.D.; Weng, D.; Chen, J.G. High selectivity of CO2 hydrogenation to CO by controlling the valence state of nickel using perovskite. Chem. Commun. 2018, 54, 7354–7357. [Google Scholar] [CrossRef] [PubMed]
  178. Zhang, Y.; Li, D.; Zhang, S.; Wang, K.; Wu, J. CO2 hydrogenation to dimethyl ether over CuO–ZnO–Al2O3/HZSM-5 prepared by combustion route. RSC Adv. 2014, 4, 16391–16396. [Google Scholar] [CrossRef]
  179. Witoon, T.; Permsirivanich, T.; Kanjanasoontorn, N.; Akkaraphataworn, C.; Seubsai, A.; Chareonpanich, M.; Faungnawakij, K.; Warakulwit, C.; Limtrakul, J. Direct synthesis of dimethyl ether from CO2 hydrogenation over Cu–ZnO–ZrO2/SO42−–ZrO2 hybrid catalysts: Effects of sulfur-to-zirconia ratios. Catal. Sci. Technol. 2015, 5, 2347–2357. [Google Scholar] [CrossRef]
  180. Zhou, X.; Su, T.; Jiang, Y.; Qin, Z.; Ji, H.; Guo, Z. CuO-Fe2O3-CeO2/HZSM-5 bifunctional catalyst hydrogenated CO2 for enhanced dimethyl ether synthesis. Chem. Eng. Sci. 2016, 153, 10–20. [Google Scholar] [CrossRef]
  181. Suwannapichat, Y.; Numpilai, T.; Chanlek, N.; Faungnawakij, K.; Chareonpanich, M.; Limtrakul, J.; Witoon, T. Direct synthesis of dimethyl ether from CO2 hydrogenation over novel hybrid catalysts containing a CuZnOZrO2 catalyst admixed with WOx/Al2O3 catalysts: Effects of pore size of Al2O3 support and W loading content. Energy Convers. Manag. 2018, 159, 20–29. [Google Scholar] [CrossRef]
  182. Michailos, S.; Mccord, S.; Sick, V.; Stokes, G.; Styring, P. Dimethyl ether synthesis via captured CO2 hydrogenation within the power to liquids concept: A techno-economic assessment. Energy Convers. Manag. 2019, 184, 262–276. [Google Scholar] [CrossRef]
  183. Li, W.; Wang, K.; Zhan, G.; Huang, J.; Li, Q. Hydrogenation of CO2 to dimethyl ether over tandem catalysts based on biotemplated hierarchical ZSM-5 and Pd/ZnO. ACS Sustain. Chem. Eng. 2020, 8, 14058–14070. [Google Scholar] [CrossRef]
  184. Falcinelli, S.; Capriccioli, A.; Pirani, F.; Vecchiocattivi, F.; Stranges, S.; Martì, C.; Nicoziani, A.; Topini, E.; Laganà, A. Methane production by CO2 hydrogenation reaction with and without solid phase catalysis. Fuel 2017, 209, 802–811. [Google Scholar] [CrossRef]
  185. Martin, N.M.; Velin, P.; Skoglundh, M.; Bauer, M.; Carlsson, P.-A. Catalytic hydrogenation of CO2 to methane over supported Pd, Rh and Ni catalysts. Catal. Sci. Technol. 2017, 7, 1086–1094. [Google Scholar] [CrossRef] [Green Version]
  186. Liu, H.; Xu, S.; Zhou, G.; Xiong, K.; Jiao, Z.; Wang, S. CO2 hydrogenation to methane over Co/KIT-6 catalysts: Effect of Co content. Fuel 2018, 217, 570–576. [Google Scholar] [CrossRef]
  187. Xie, C.; Chen, C.; Yu, Y.; Su, J.; Li, Y.; Somorjai, G.A.; Yang, P. Tandem catalysis for CO2 hydrogenation to C2–C4 hydrocarbons. Nano Lett. 2017, 17, 3798–3802. [Google Scholar] [CrossRef] [Green Version]
  188. An, X.; Zuo, Y.; Zhang, Q.; Wang, J. Methanol Synthesis from CO2 Hydrogenation with a Cu/Zn/Al/Zr Fibrous Catalyst. Chin. J. Chem. Eng. 2009, 17, 88–94. [Google Scholar] [CrossRef]
  189. Dong, X.; Li, F.; Zhao, N.; Tan, Y.; Wang, J.; Xiao, F. CO2 hydrogenation to methanol over Cu/Zn/Al/Zr catalysts prepared by liquid reduction. Chin. J. Catal. 2017, 38, 717–725. [Google Scholar] [CrossRef]
  190. Phongamwong, T.; Chantaprasertporn, U.; Witoon, T.; Numpilai, T.; Poo-arporn, Y.; Limphirat, W.; Donphai, W.; Dittanet, P.; Chareonpanich, M.; Limtrakul, J. CO2 hydrogenation to methanol over CuO–ZnO–ZrO2–SiO2 catalysts: Effects of SiO2 contents. Chem. Eng. J. 2017, 316, 692–703. [Google Scholar] [CrossRef]
  191. Chen, D.; Mao, D.; Xiao, J.; Guo, X.; Yu, J. CO2 hydrogenation to methanol over CuO–ZnO–TiO2–ZrO2: A comparison of catalysts prepared by sol–gel, solid-state reaction and solution-combustion. J. Sol-Gel Sci. Technol. 2018, 86, 719–730. [Google Scholar] [CrossRef]
  192. Tan, Q.; Shi, Z.; Wu, D. CO2 Hydrogenation to Methanol over a Highly Active Cu–Ni/CeO2–Nanotube Catalyst. Ind. Eng. Chem. Res. 2018, 57, 10148–10158. [Google Scholar] [CrossRef]
  193. Wang, G.; Mao, D.; Guo, X.; Yu, J. Enhanced performance of the CuO-ZnO-ZrO2 catalyst for CO2 hydrogenation to methanol by WO3 modification. Appl. Surf. Sci. 2018, 456, 403–409. [Google Scholar] [CrossRef]
  194. Ting, K.W.; Toyao, T.; Siddiki, S.M.A.H.; Shimizu, K.-I. Low-temperature hydrogenation of CO2 to methanol over heterogeneous TiO2-supported Re catalysts. ACS Catal. 2019, 9, 3685–3693. [Google Scholar] [CrossRef]
  195. Chou, C.-Y.; Lobo, R.F. Direct conversion of CO2 into methanol over promoted indium oxide-based catalysts. Appl. Catal. A Gen. 2019, 583, 117144. [Google Scholar] [CrossRef]
  196. Wang, G.; Mao, D.; Guo, X.; Yu, J. Methanol synthesis from CO2 hydrogenation over CuO-ZnO-ZrO2-MxOy catalysts (M=Cr, Mo and W). Int. J. Hydrogen Energy 2019, 44, 4197–4207. [Google Scholar] [CrossRef]
  197. Xiao, J.; Mao, D.; Wang, G.; Guo, X.; Yu, J. CO2 hydrogenation to methanol over CuO-ZnO-TiO2-ZrO2 catalyst prepared by a facile solid-state route: The significant influence of assistant complexing agents. Int. J. Hydrogen Energy 2019, 44, 14831–14841. [Google Scholar] [CrossRef]
  198. Camposmartin, J.M.; Guerreroruiz, A.; Fierro, J.L.G. Structural and Surface Properties of CuO-ZnO-Cr2O3 Catalysts and Their Relationship with Selectivity to Higher Alcohol Synthesis. J. Catal. 1995, 156, 208–218. [Google Scholar] [CrossRef]
  199. Zegkinoglou, I.; Pielsticker, L.; Han, Z.-K.; Divins, N.J.; Kordus, D.; Chen, Y.-T.; Escudero, C.; Pérez-Dieste, V.; Zhu, B.; Gao, Y.; et al. Surface segregation in CuNi nanoparticle catalysts during CO2 hydrogenation: The role of CO in the reactant mixture. J. Phys. Chem. C 2019, 123, 8421–8428. [Google Scholar] [CrossRef] [Green Version]
  200. Van den Berg, R.; Parmentier, T.E.; Elkjær, C.F.; Gommes, C.J.; Sehested, J.; Helveg, S.; de Jongh, P.E.; de Jong, K.P. Support Functionalization To Retard Ostwald Ripening in Copper Methanol Synthesis Catalysts. ACS Catal. 2015, 5, 4439–4448. [Google Scholar] [CrossRef]
  201. Wen, C.; Cui, Y.; Dai, W.-L.; Xie, S.; Fan, K. Solvent feedstock effect: The insights into the deactivation mechanism of Cu/SiO2 catalysts for hydrogenation of dimethyl oxalate to ethylene glycol. Chem. Commun. 2013, 49, 5195–5197. [Google Scholar] [CrossRef]
  202. Zhang, H.; Tan, H.-R.; Jaenicke, S.; Chuah, G.-K. Highly efficient and robust Cu catalyst for non-oxidative dehydrogenation of ethanol to acetaldehyde and hydrogen. J. Catal. 2020, 389, 19–28. [Google Scholar] [CrossRef]
  203. Chen, W.; Song, T.; Tian, J.; Wu, P.; Li, X. An efficient Cu-based catalyst for the hydrogenation of ethylene carbonate to ethylene glycol and methanol. Catal. Sci. Technol. 2019, 9, 6749–6759. [Google Scholar] [CrossRef]
  204. Wang, M.; Yao, D.; Li, A.; Yang, Y.; Lv, J.; Huang, S.; Wang, Y.; Ma, X. Enhanced selectivity and stability of Cu/SiO2 catalysts for dimethyl oxalate hydrogenation to ethylene glycol by using silane coupling agents for surface modification. Ind. Eng. Chem. Res. 2020, 59, 9414–9422. [Google Scholar] [CrossRef]
  205. Chen, X.; Cui, Y.; Wen, C.; Wang, B.; Dai, W.-L. Continuous synthesis of methanol: Heterogeneous hydrogenation of ethylene carbonate over Cu/HMS catalysts in a fixed bed reactor system. Chem. Commun. 2015, 51, 13776–13778. [Google Scholar] [CrossRef] [PubMed]
  206. Chen, C.S.; Lai, Y.T.; Lai, T.W.; Wu, J.H.; Chen, C.H.; Lee, J.F.; Kao, H.M. Formation of Cu nanoparticles in SBA-15 functionalized with carboxylic acid groups and their application in the water–gas shift reaction. ACS Catal. 2013, 3, 667–677. [Google Scholar] [CrossRef]
  207. Chen, C.; Ren, H.; He, Y.; Zhan, Y.; Au, C.; Luo, Y.; Lin, X.; Liang, S.; Jiang, L. Unraveling the role of Cu0 and Cu+ sites in Cu/SiO2 catalysts for water-gas shift reaction. ChemCatChem 2020, 12, 4672–4679. [Google Scholar] [CrossRef]
Figure 1. Number of publications concerning “phyllosilicate” (dots) and “phyllosilicate + catal” (bars) over years (compiled from statistical data of Scopus on 8 December 2020).
Figure 1. Number of publications concerning “phyllosilicate” (dots) and “phyllosilicate + catal” (bars) over years (compiled from statistical data of Scopus on 8 December 2020).
Catalysts 11 00255 g001
Figure 2. Structural scheme of the 2:1 phyllosilicates and 1:1 phyllosilicates (chrysocolla) (reproduced from [36] with permission from Elsevier and from [40] with permission from American Chemical Society).
Figure 2. Structural scheme of the 2:1 phyllosilicates and 1:1 phyllosilicates (chrysocolla) (reproduced from [36] with permission from Elsevier and from [40] with permission from American Chemical Society).
Catalysts 11 00255 g002
Figure 3. The chronograph of the development of copper phyllosilicates-derived catalysts (CuPS-cats).
Figure 3. The chronograph of the development of copper phyllosilicates-derived catalysts (CuPS-cats).
Catalysts 11 00255 g003
Figure 4. Characterization of Cu/SiO2 catalysts synthesized by impregnation (IM), deposition-precipitation (DP) and ammonia evaporation (AE) (reproduced from [117] with permission from MDPI).
Figure 4. Characterization of Cu/SiO2 catalysts synthesized by impregnation (IM), deposition-precipitation (DP) and ammonia evaporation (AE) (reproduced from [117] with permission from MDPI).
Catalysts 11 00255 g004
Figure 5. Phyllosilicate crystallinity degree of calcined samples at evaporation temperatures, the crystallinity degree was estimated from I663/I800 of FTIR (adapted from [73]).
Figure 5. Phyllosilicate crystallinity degree of calcined samples at evaporation temperatures, the crystallinity degree was estimated from I663/I800 of FTIR (adapted from [73]).
Catalysts 11 00255 g005
Figure 6. The correlation of reduction temperature and physicochemical properties of CuPS-cats (adapted from [87]).
Figure 6. The correlation of reduction temperature and physicochemical properties of CuPS-cats (adapted from [87]).
Catalysts 11 00255 g006
Figure 7. The relationship between Cu+/(Cu0 + Cu+) ratio and (a) calcination and (b) reduction temperatures (adapted from Figure 4 of [71] and Figure S25 of [70]). The Cu+–(Cu0 + Cu+) ratio was estimated from XPS and N2O titration.
Figure 7. The relationship between Cu+/(Cu0 + Cu+) ratio and (a) calcination and (b) reduction temperatures (adapted from Figure 4 of [71] and Figure S25 of [70]). The Cu+–(Cu0 + Cu+) ratio was estimated from XPS and N2O titration.
Catalysts 11 00255 g007
Figure 8. The correlation between surface hydroxyl groups, CuPS phase and Cu+ surface area (adapted from [88]). Surface hydroxyl groups and CuPS phase were, respectively, defined from I960/I800 and I670/I800 relative intensity of FTIR peaks. Cu+ surface area was estimated from XPS and N2O titration.
Figure 8. The correlation between surface hydroxyl groups, CuPS phase and Cu+ surface area (adapted from [88]). Surface hydroxyl groups and CuPS phase were, respectively, defined from I960/I800 and I670/I800 relative intensity of FTIR peaks. Cu+ surface area was estimated from XPS and N2O titration.
Catalysts 11 00255 g008
Figure 9. Sinter of Cu particles when absence of Ni (a), adding trace Ni (b), adding a proper amount of Ni (c), and adding excess Ni (d) into the CuPS (adapted from [157]).
Figure 9. Sinter of Cu particles when absence of Ni (a), adding trace Ni (b), adding a proper amount of Ni (c), and adding excess Ni (d) into the CuPS (adapted from [157]).
Catalysts 11 00255 g009
Figure 10. A prediction of the interaction of Cu and other metals based on electronegativity by Pauling scale.
Figure 10. A prediction of the interaction of Cu and other metals based on electronegativity by Pauling scale.
Catalysts 11 00255 g010
Figure 11. NH3-TPD and CO2 TPD profiles of xB-Cu-SiO2 catalysts. (reproduced from [95] with permission from Elsevier)
Figure 11. NH3-TPD and CO2 TPD profiles of xB-Cu-SiO2 catalysts. (reproduced from [95] with permission from Elsevier)
Catalysts 11 00255 g011
Figure 12. Scheme of nanotube-assembled hollow-sphere nanoreactor formation. (reproduced from [156] with permission from American Chemical Society).
Figure 12. Scheme of nanotube-assembled hollow-sphere nanoreactor formation. (reproduced from [156] with permission from American Chemical Society).
Catalysts 11 00255 g012
Figure 13. Scheme of CuPSNTs coated mesoporous silica (reproduced from [113] with permission from Springer Nature).
Figure 13. Scheme of CuPSNTs coated mesoporous silica (reproduced from [113] with permission from Springer Nature).
Catalysts 11 00255 g013
Figure 14. The in situ mannered reduction in CuPS-cats under hydrogen atmosphere monitored by TEM at different temperatures (AD) and isothermal (E) with 3 subregions. (Reprinted with permission from [174]. Copyright (2016) American Chemical Society).
Figure 14. The in situ mannered reduction in CuPS-cats under hydrogen atmosphere monitored by TEM at different temperatures (AD) and isothermal (E) with 3 subregions. (Reprinted with permission from [174]. Copyright (2016) American Chemical Society).
Catalysts 11 00255 g014aCatalysts 11 00255 g014b
Figure 15. The non-isothermal reduction (X-ray adsorption spectra and temperature programed reduction) of CuPS-cats under H2 (upper) and CO (lower) as reductive agents (our group’s results).
Figure 15. The non-isothermal reduction (X-ray adsorption spectra and temperature programed reduction) of CuPS-cats under H2 (upper) and CO (lower) as reductive agents (our group’s results).
Catalysts 11 00255 g015
Figure 16. Scheme of alcohol production from the hydrogenation of carboxylate, carbonate, carboxylic, aldehyde, and CO2 by using CuPS-cats.
Figure 16. Scheme of alcohol production from the hydrogenation of carboxylate, carbonate, carboxylic, aldehyde, and CO2 by using CuPS-cats.
Catalysts 11 00255 g016
Figure 17. Reaction pathway of hydrogenation of DMO and DEO to MG, EG, and EtOH (adapted from [73,97,159] and [91,92]).
Figure 17. Reaction pathway of hydrogenation of DMO and DEO to MG, EG, and EtOH (adapted from [73,97,159] and [91,92]).
Catalysts 11 00255 g017
Figure 18. Reaction pathway of hydrogenolysis of DEM to 3-hydroxy ethyl propionate, PDO, and n-propanol (adapted from [103]).
Figure 18. Reaction pathway of hydrogenolysis of DEM to 3-hydroxy ethyl propionate, PDO, and n-propanol (adapted from [103]).
Catalysts 11 00255 g018
Figure 19. Scheme of CHDM synthesis from DMCD (adapted from [107]).
Figure 19. Scheme of CHDM synthesis from DMCD (adapted from [107]).
Catalysts 11 00255 g019
Figure 20. Scheme of THF synthesis from DMM (reproduced from [99] with permission from Elsevier).
Figure 20. Scheme of THF synthesis from DMM (reproduced from [99] with permission from Elsevier).
Catalysts 11 00255 g020
Figure 21. EC hydrogenolysis reaction to EG and MeOH (adapted from [105]).
Figure 21. EC hydrogenolysis reaction to EG and MeOH (adapted from [105]).
Catalysts 11 00255 g021
Figure 22. Illustration of conversion AA to HDOL via CuPS-cats and Ni-CuPS-cats (reproduced from [119] by permission of American Chemical Society).
Figure 22. Illustration of conversion AA to HDOL via CuPS-cats and Ni-CuPS-cats (reproduced from [119] by permission of American Chemical Society).
Catalysts 11 00255 g022
Figure 23. Conversion (X%) of different acids and turnover frequency (TOF h−1) of the Cu41PS-cats and Cu36Ni5PS-cats (adapted from Table S2 of [119] by permission of American Chemical Society).
Figure 23. Conversion (X%) of different acids and turnover frequency (TOF h−1) of the Cu41PS-cats and Cu36Ni5PS-cats (adapted from Table S2 of [119] by permission of American Chemical Society).
Catalysts 11 00255 g023
Figure 24. OTS-grafted CuPS-cats with a hydrophobic surface was protected from water incursion (reproduced from [173] by permission of American Chemical Society).
Figure 24. OTS-grafted CuPS-cats with a hydrophobic surface was protected from water incursion (reproduced from [173] by permission of American Chemical Society).
Catalysts 11 00255 g024
Figure 25. Reduction of FFAL to MF [104,106] and HMF to DHMF (adapted from [172]).
Figure 25. Reduction of FFAL to MF [104,106] and HMF to DHMF (adapted from [172]).
Catalysts 11 00255 g025
Figure 26. (a) The XRD and TEM of CuPS-cats (b) before and (c) after reaction (reproduced from [96] with permission from American Chemical Society).
Figure 26. (a) The XRD and TEM of CuPS-cats (b) before and (c) after reaction (reproduced from [96] with permission from American Chemical Society).
Catalysts 11 00255 g026
Table 1. The CuPS-cats and their catalytic performance for hydrogenation reactions.
Table 1. The CuPS-cats and their catalytic performance for hydrogenation reactions.
Reactions Reaction ConditionsPerformanceCatalysts
(prep. Method)
Activation bActivity Ref.
T (°C)P H2 (MPa)H2/Reactant(L/W/G)
HSV (h−1)
Conv. (%)Sel. (%)TOS (h)Tcalc. (°C) Time (h)Tred. (°C) Time (h)XCu+ c
DMO to EG2002.550(L)
0.25
10098-18Cu/SiO2 (AE)450
4
350
4
0.55Cu+Chen 2008 [73]
2002.550(L)
1.2
10095-19Cu/SiO2
(AE)
-300
4
0.14Cu0Yin 2011 [75]
1803.080(L)
0.5
10095-Cu/SiO2 (AE)450
4
250
4
0.40Cu0 and Cu+Sun 2018 [87]
2002.080(L)
1.0
10095-30Cu/SiO2 (UHDP)450
4
220
18
0.62Cu0 and Cu+Zhang 2012 [78]
1902.580(L)
3.25
10091-30Cu/SiO2
(AC)
400
4
350
4
0.52Cu0 and Cu+Zhao 2017 [81]
2102.080(W)
0.8
100~86~20020Cu/SiO2
(AC)
450
4
350
4
0.60Cu dispersionPopa 2015 [79]
1802.580(W)
1.0
1009825020Cu/m-SiO2
(AE)
400
4
300
4
0.53Formation of more CuPSZhao 2017 [82]
2002.580(L)
3.5
1009520020Cu/5 nm SiO2
(AE)
400
4
350
4
0.53Formation of more CuPSWang 2020 [88]
1903.080(W)
2.0
1009530030CuPS/HZSM-5
(USG)
350
2
350
4
0.66Increase in specific surface areaLin 2012 [76]
2003.0100(L)
0.6
1009070Cu/SiO2–TiO2
(AE)
450
4
300
4
0.66Ti-Cu interactionWen 2013 [162]
1902.050(W)
1.0
99.8972000.05Dextrin coated Cu/SiO2
(AE)
450
4
350
5
0.84Cu dispersionYe 2017 [83]
1903.080(W)
3.5
99.7933001B-27Cu/SiO2
(UHDP)
350
2
350
4
0.64B-Cu interactionHe 2011 [95]
1903.080(W)
0.6
100861500.65Ag-8Cu/SiO2
(UHDP)
350
4
350
4
0.53Ag-Cu alloysHuang 2013 [159]
2153.0100(L)
0.55
1009720002Ni-Cu/SiO2
(CP)
400
4
260
-
-Cu dispersionZhu 2016 [157]
2002.590(L
)1.0
100961800.6Sn-Cu/SiO2
(AE)
450
5
300
4
0.49Sn-Cu interactionZhang 2017 [110]
2003.080(L)
0.4
99.592-0.2Zn-0.8Cu/SiO2
(CP)
450
4
300
4
0.53Cu dispersionQi 2018 [86]
1902.595(W)
1.0
>95>701000.5In-10Cu/SiO2
(UHDP)
400
3
350
4
-In-Cu alloysYu 2019 [90]
2003.080(L)
4.2
10098-CuPSNTs
(AH)
-300
4
0.39 Xu 2018 [113]
2003.080(L)
2.0
10096200m-SiO2-coated-CuPSNTs
(AH)
500
2
300
4
0.50Cu-Si Interaction
2002.520(W)
2.0
10095300NAHS CuPS-cats450
4
300
4
0.25Spatial confinementYao 2018 [156]
2002.580(L)
16
10095100monolithic CuPS-cats
(dip coating)
400
4
350
4
-Scale up researchYue 2012 [77]
DMO to MG2202.5200(W)
0.3
8694-50Cu/SiO2 (ultrasonic)400
-
280
4
0.54Cu0 and Cu+Abbas 2018 [84]
2002.580(W)
1.5
8399.8-0.5 Na2SiO3 coated CuPS-cats (AE) 400
4
350
4
0.50Spatial confinementHuang 2020 [89]
DMO to EtOH2802.5200(W)
0.8
100922001Ce-Cu/SiO2
(UHDP)
350
3
300
3
0.48Ce-Cu interactionAi 2018 [85]
2802.5200(L)
2.0
98.792-Cu/CuPSNTs-in (AH)450
4
350
4
0.90Confinement effectYue 2013 [97]
2802.5200(L)
2.0
95.274-Cu/CuPSNTs-out (AH)450
4
350
4
0.67
2802.5200(L)
2.0
100>8020020CuPS-cats
(AEH)
400
4
350
4
0.33Cu0 and Cu+Gong 2012 [96]
DEO to EG2202.070(G)
6500
~96~88<545Cu/SiO2
(sol-gel)
450
5
300
5
-Cu0 and Cu+He 2010 [94]
2202.480(L)
1.0
9995<8020Cu/m-SiO2
(IWI)
550
3
350
6
-smaller Cu particlesDing 2015 [91]
2402.980(L)
1.0
>99~95200Cu/SiO2 (AC)450
4
350
6
0.69Cu0 and Cu+Ding 2017 [92]
2402.980(L)
1.0
9985<100Cu/SiO2 (AE)450
4
350
6
0.57Cu0 and Cu+
DEM to PDO2002.0440(L)
1.8
9730-M-Cu/SiO2 (M=B, Ag, Zn, Al, Ni, Mo)
(AE 90 °C)
450
4
300
4
0.33B increase the formation of Cu0Ding 2016 [103]
2002.0440(L)
1.8
913210022Cu/SiO2 (AE 90 °C)450
4
300
4
0.45Dispersion of CuDing 2016 [102]
MA to EtOH2201.040(W)
1.0
8151-20Cu/SiO2
(AE)
450
4
350
3
-Cu dispersionDong 2016 [158]
3403.010(L)
2.0
~80~9550Cu/9MgO-SiO2
(AE)
450
3
350
4
0.64Metal-support interactionenrichment of Cu+ surfaceQin 2015 [163]
2503.015(L)
2.0
~97~96100Cu/30%CeO2–SiO2
(AE)
450
4
350
4
0.42Metal-support interactionenrichment of Cu+ surfaceYe 2016 [164]
2503.015(L)
2.0
98981001%In2O3-30%Cu/SiO2 (AE)450
4
350
4
0.57In-Cu interaction
enrichment of Cu+ surface
Zhang 2018 [115]
2202.020(L)
2.0
9289-1Zn-9Cu/SiO2 (AC)400
4
300
-
0.41Zn-Cu interaction
enrichment of Cu+ surface
Zhao 2018 [116]
EA to EtOH2503.029(L)
1.24
10099-Zn, NH4 modified Cu/SiO2
(CP)
450
4
300
1
0.952:1 CuPSs reduces Cu surface areaDi 2016 [80]
DMCD to CHDM2205260(L)
1.0
~99~951000CuPSNTs450
4
350
4
0.65Proper Cu+/Cu0 and
Small Cu Particles
Gong 2017 [107]
DMM to THF240560(L)
0.9
84100-Cu/SiO2 (AE)450
4
300
4
0.61Formation of more CuPSs and Cu+ Han 2015 [99]
EC to EG2005--
(Batch reactor)
10095-10Cu/SBA-15 (AE)400
4
200
4
0.65Cu0 and Cu+Li 2016 [105]
1805--
(Batch reactor)
10098-10Cu/SiO2 (AE)400
4
350
4
0.51Cu0 and Cu+Li 2017 [108]
AA to HDOL2406--
(Batch reactor)
10097-20Cu/SiO2
(AEH)
400
4
350
3
0.19Cu0 and Cu+Jiang 2018 [111]
2406--
(Batch reactor)
10095-5Ni-36Cu/SiO2 (AEH)400
4
350
3
-Ni-Cu interaction Tu 2019 [119]
LA to GVL13012--
(Batch reactor)
95.785.2-OTS-grafted Cu/SiO2 (AEH)400
4
350
3
-Cu0 and Cu+Tsou 2020 [173]
FFAL to FFOH1400.117(W)
2.0
9588-24Cu/SiO2 (AE) 450
5
270
-
-Phyllosilicates formationDong 2016 [104]
1400.117(W)
7.0
3999.5-24Cu/SiO2 (AE)450
5
270
16
-Optimal Cu/Si ratio at 0.75 for Phyllosilicates formation Dong 2017 [106]
HMF to DHMF805--
(Batch reactor)
80100-15Cu/SiO2
(UH)
450
5
250
2
-SiO2-Cu interactionZhu 2017 [172]
CO2 to MeOH32034(W)
16
~28~2112011Cu/SiO2 (AE)500
5
300
2
1Cu+Wang 2015, 2016 [70,71]
a Catalysts and preparation methods. The prefix represents the weight loading of copper and the characters between parenthesis represents the preparation method. For example: 20CuPS-cats (AE) means the copper phyllosilicate-derived catalysts with 20 wt % of Cu dispersing on SiO2 was prepared by ammonia evaporation method. b Activation treatment includes calcination and reduction at different temperatures and time (Tcalc: calcination temperature, Tred: reduction temperature). c After activation, the active sites are supposed to the ratio of Cu+ and Cu0. The ratio of Cu+ over total surface area of copper was calculated from XAES data by the following formula X Cu+ = Cu+/(Cu0 + Cu+).
Table 2. Residual and leached amount of Cu, Ni in Cu41PS-cats and Cu36Ni5PS-cats after each round of activity test (reproduced from [119] with permission from American Chemical Society). The figures were determined by ICP-AES analysis.
Table 2. Residual and leached amount of Cu, Ni in Cu41PS-cats and Cu36Ni5PS-cats after each round of activity test (reproduced from [119] with permission from American Chemical Society). The figures were determined by ICP-AES analysis.
Cu41PS-RedCu36Ni5PS-Red
CuLeached CuCuLeached CuNiLeached Ni
(wt%)(ppm)(wt%)(ppm)(wt%)(ppm)
Fresh40.4-35.6 3.9
1st cycle37.23.933.13.13.8ND
2nd cycle34.83.531.82.93.9ND
3rd cycle32.93.231.12.93.8ND
ND = undetectable.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

To, D.-T.; Lin, Y.-C. Copper Phyllosilicates-Derived Catalysts in the Production of Alcohols from Hydrogenation of Carboxylates, Carboxylic Acids, Carbonates, Formyls, and CO2: A Review. Catalysts 2021, 11, 255. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020255

AMA Style

To D-T, Lin Y-C. Copper Phyllosilicates-Derived Catalysts in the Production of Alcohols from Hydrogenation of Carboxylates, Carboxylic Acids, Carbonates, Formyls, and CO2: A Review. Catalysts. 2021; 11(2):255. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020255

Chicago/Turabian Style

To, Dien-Thien, and Yu-Chuan Lin. 2021. "Copper Phyllosilicates-Derived Catalysts in the Production of Alcohols from Hydrogenation of Carboxylates, Carboxylic Acids, Carbonates, Formyls, and CO2: A Review" Catalysts 11, no. 2: 255. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020255

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

Article Metrics

Back to TopTop