首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A poly(methyl methacrylate)-supported Pd0 nanocatalyst was successfully prepared from solution reaction of Pd(CH3COO)2 with a copolymer acid, poly(methyl methacrylate-ran-methacrylic acid) (MMA–MAA). The reaction was carried out in a benzene/methanol mixed solvent in the dark at room temperature (∼25 °C) in the absence of a typical chemical reductant. There was coordination between the Pd0 nanoclusters and MMA–MAA, resulting in Pd0 nanoclusters being stably and uniformly dispersed in the MMA–MAA matrix, with an average particle size of ∼2.5 ± 0.5 nm. Mechanistically, it can tentatively be proposed that PMMA-ionomerization of the Pd2+ ions produces intramolecular –2COO–Pd2+ aggregate cross-links in the solution. On swelling of the chain-segments that are covalently bound via multiple C–C bonds, the resultant elastic forces cause instantaneous dissociation at the O–Pd coordination bonds to give transient bare (i.e., uncoordinated), highly-oxidative Pd2+ ions and H+-associative carboxylate groups, both of which rapidly scavenge electrons and protons, respectively, of the active α-H atoms abstracted from the methanol molecules of the solvent to make Pd0 nanoclusters supported by the re-formed MMA–MAA. The MMA–MAA acid copolymer, without itself undergoing any permanent chemical change, serves as a mechanical activator or catalyst for the mechanochemical reduction of Pd(CH3COO)2 under mild conditions. Compared with traditional Pd/C catalysts, this Pd0 nanocatalyst exhibited more excellent catalytic efficiency and reusability in the Heck reaction between iodobenzene and styrene, and it could be easily separated. The supported Pd0 nanocatalyst prepared using this novel and simple preparation method may display high-efficiency catalytic properties for other cross coupling reactions.

A polymer-supported Pd0 nanocatalyst is prepared by using mechanochemical reduction as the driving force for the reaction.  相似文献   

2.
A highly efficient heterogeneous catalyst was synthesized by delicate engineering of NH2-functionalized and N-doped hollow mesoporous carbon spheres (NH2–N-HMCS), which was used for supporting AuPd alloy nanoparticles with ultrafine size and good dispersion (denoted as AuPd/NH2–N-HMCS). Without using any additives, the prepared AuPd/NH2–N-HMCS catalytic formic acid dehydrogenation possesses superior catalytic activity with an initial turnover frequency value of 7747 mol H2 per mol catalyst per h at 298 K. The excellent performance of AuPd/NH2–N-HMCS derives from the unique hollow mesoporous structure, the small particle sizes and high dispersion of AuPd nanoparticles and the modified Pd electronic structure in the AuPd/NH2–N-HMCS composite, as well as the synergistic effect of the modified support.

Anchoring ultrafine AuPd on NH2-functionalized and N-doped hollow mesoporous carbon spheres for formic acid dehydrogenation.  相似文献   

3.
Monometallic (Pd, Ru or Rh) and bimetallic (Pd0.5–Ru0.5) alloy NPs catalysts were examined for the hydrogenation of quinoline. Pd–Ru alloy catalyst showed superior catalytic activity to the traditional Rh catalyst. The characterization of Pd0.5–Ru0.5 catalysts, HAADF-EDX mapping and XPS analysis suggested that the alloy state of PdRu catalysts remained unchanged in the recovered catalyst. Furthermore, the catalyst was highly selective for the hydrogenation of different arenes.

Recyclable Pd0.5Ru0.5–PVP catalyst showed higher activity than monometallic Pd or Ru catalyst for the hydrogenation of quinoline. Furthermore, Pd0.5Ru0.5–PVP was able to hydrogenate different arenes.

1,2,3,4-Tetrahydroquinolines are important building blocks for the synthesis of fine chemicals, pharmaceuticals and petrochemicals.1–3 Traditionally, 1,2,3,4-tetrahydroquinolines have been synthesized by catalytic cyclization or the Beckman rearrangement method.3–6 Direct hydrogenation of readily available quinoline by molecular hydrogen is a simple and atom-economical route for the hydrogenated quinoline compounds. Nevertheless, the hydrogenation of quinolines reaction suffers from catalyst deactivation caused by the strong interaction between the nitrogen atom of quinoline and active sites of the catalyst. Recently, several heterogeneous (Co, Au, Ru, Pt, Pd and Rh) catalysts have been developed to overcome the catalyst deactivation problem (Table S1).7–24 Specifically, Rh catalysts exhibits remarkably excellent activity in comparison with other noble metal catalysts but some of them require high temperature (>80 °C)13,21 and/or high pressure (>10 bar).14–21 Later, Rh/AlOOH and Rh nanoparticles (NPs) catalysts have been reported for efficient hydrogenation of quinolines under ambient conditions (1 bar H2 and 25 °C).22–24 Although high selectivity and activity of Rh catalysts, the high cost of Rh make it unfavourable for the industrial purposes. From the economical and environmental point of view, the cost-effective catalyst is necessary for the hydrogenation of quinolines under mild condition.Solid-solution alloy method is a promising way for the synthesis of alloy NPs which offers an opportunity to control the electronic state of alloy by changing the composition ratio and/or different combinations of alloy in this method. In search of a cheap alternative for Rh catalyst, Pd and Ru are important metals because both metals are relatively cheaper than Rh metal and their well-known activity for hydrogenation reactions. We believe that Pd0.5–Ru0.5 solid solution alloy NPs could be alternative for the traditional Rh catalyst. However, a solid-solution alloy of Pd and Ru is a challenging task due to the immiscibility of Pd and Ru in bulk state.25,26 Recently, our research group led by H. Kitagawa has reported the first successful example of Pd0.5–Ru0.5 NPs via a chemical reduction method using a nano-size effect which showed almost similar catalytic activity than those of Rh for CO oxidation and automobile exhaust purification.27,28 As part of our continuing interest in the applications of Pd–Ru as pseudo Rh catalyst, we herein report that chemoselective hydrogenation of heteroarenes and arenes under mild condition.In preliminary experiments, we carried out the hydrogenation of quinoline to 1,2,3,4-tetrahydroquinoline as model reaction (27,28 We found the similar trend using Pd0.5Ru0.5 catalyst and it showed full conversion with 95% yield of 1,2,3,4-tetrahydroquinolines (entry 3). The effect of the solvent showed that the reactivity of Pd0.5Ru0.5 was improved using methanol as solvent. Probably, methanol provides the better dispersion of Pd0.5–Ru0.5 NPs in the reactions and/or high solubility of hydrogen gas in methanol. Next, we examined the activity of previously reported Rh catalyst. Although, the particle size of Rh and Pd0.5–Ru0.5 was similar, the moderate yield of tetrahydroquinoline was obtained in the presence of Rh–PVP (entry 7). Due to the high amount of PVP, the catalytic activity of Rh was slightly declined via less interaction between quinoline and metal sites.29Optimization of reaction conditions for hydrogenation of quinolinea
EntryCatalystParticle sizeb (nm)Conv. (%)GC yield (%)
1Ru–PVP8.4 ± 2.41313
2Pd–PVP4.8 ± 0.84132
3Pd0.5Ru0.5–PVP5.6 ± 1.69995
4cPd0.5Ru0.5–PVP5.6 ± 1.66360
5dPd0.5Ru0.5–PVP5.6 ± 1.64540
6ePd0.5Ru0.5–PVP5.6 ± 1.62113
7Rh–PVP5.4 ± 1.0f6666
Open in a separate windowaReaction conditions: quinoline (1 mmol), catalyst (2 mol%), CH3OH (1 mL), 5 bar H2, 25 °C, 6 h, CH3OH.bDetermined by STEM analysis.cTHF.dToluene.e1,4-Dioxane.fRef. 38.The substrate scope of Pd0.5–Ru0.5PVP catalyst was explored under mild reaction conditions (5 bar H2 and 25 °C). As summarized in
EntryHeteroareneProductConv. (%)GC yield (%)
1 9996
2 9890
3 9992
4 6460
5 9976
6 9999
7 9445b
8 9980
Open in a separate windowaReaction conditions: substrate (1 mmol), catalyst (2 mol%), 5 bar H2, 25 °C, 24 h.b3,4-Dihydroquinazoline was obtained.Hydrogenation of arenes to saturated compounds is an important transformation for petrochemical and pharmaceutical industries. For instance, hydrogenation of benzoic acid to cyclohexanecarboxylic acid, it is used for the synthesis of pharmaceutical drugs such as ansatrienin.30 Hydrogenation of benzoic acid has been reported by using supported and polymer stabilized metal nanoparticles.31–38 Among them, Rh catalyst demonstrated the excellent activity for the hydrogenation of benzoic acid. The scope of Pd0.5–R0.5 catalyst was examined for different arene hydrogenation for the hydrogenation of benzoic acid. The scope of Pd0.5–R0.5 catalyst was examined for different arene hydrogenation under solvent-free condition (EntryAreneProductConv. (%)GC yield (%)1 99942 99923 99994 95955 99926 8060Open in a separate windowaReaction conditions: substrate (1 mmol), catalyst (2 mol%), 10 bar H2, 150 °C, 24 h.The reusability of Pd0.5–Ru0.5 catalyst was investigated using optimized condition (5 bar H2 and 25 °C) for the hydrogenation of quinoline. The catalyst was washed by acetone multiple times after completion of reaction and separated by centrifugation. The recovered catalyst was dried under vacuum at 40 °C for 12 h and reused without further purification. The catalyst was reused three times without loss in activity (Fig. 1).Open in a separate windowFig. 1Recycle experiment for the hydrogenation of quinoline over Pd0.5Ru0.5–PVP catalyst.To study morphological changes, the fresh and recovered catalyst after the 1st cycle was analysed by high-angle annular dark field. The reconstructed overlaying image of fresh Pd0.5–Ru0.5 catalyst revealed that Pd and Ru atom was distributed equally in the particle which indicated the formation of Pd0.5Ru0.5 alloy via homogeneous mixing at atomic-level. Dealloying and/or aggregation were not observed in Pd0.5–Ru0.5 (Fig. 2). Next, the electronic state of fresh and used Pd0.5–Ru0.5 catalyst was examined by XPS (Fig. 3). In our previous reports, the binding energy of Pd0.5–Ru0.5 was shifted positively in 3d5/2 (334.5 eV) from 3d5/2 (334.30 eV) of monontellic Pd and negatively in 3P3/2 (460.6 eV) from 3P3/2 (461.4 eV) of monometallic Ru.39 Such shifting was also observed in fresh and used Pd0.5–Ru0.5 catalyst. From these results, we concluded that the electronic state of used catalysts was not changed significantly after 1st cycle.Open in a separate windowFig. 2HAADF-STEM images Pd-L and Ru-L STEM-EDX maps and reconstructed overlay images of fresh (A) and recovered (B) Pd0.5–Ru0.5–PVP catalyst.Open in a separate windowFig. 3XPS spectra for Pd0.5–Ru0.5̲fresh (blue) and Pd0.5–Ru0.5̲used (red) with Pd 3d (A) and Ru 3p (B).The superiority of Pd0.5–Ru0.5 alloy over Pd and Ru can be explained by unique electronic state. Our group has used first-principles methods to study the electronic states of Pd, Ru, Rh, and the Pd0.5–Ru0.5 alloy. Interestingly, we have found that the electronic state of the Pd–Ru alloy differs from those of the parental Pd or Ru metal. This result has indicated that a new electronic state is produce by alloying Pd and Ru. The density of state of the Pd0.5–Ru0.5 alloy is similar to that of Rh.Furthermore, Pd0.5–Ru0.5 alloy shows the electron transfer from Pd to Ru and produced the partial positive (δ+) charge at Pd and partial negative charge (δ−) at Ru. This unique electronic state with bifunctional sites of Pd0.5–Ru0.5 alloy made it favorable for chemoselective for hydrogenation of quinoline. In case of monometallic Pd or Ru catalyst, the strong adsorption of quinoline might be responsible for low conversion of quinoline.It is well-known that the hydrogenation of organic functional group such as C Created by potrace 1.16, written by Peter Selinger 2001-2019 C and C Created by potrace 1.16, written by Peter Selinger 2001-2019 N proceeded via heterolytic cleavage of H2. Such cleavage of H2 yielded to H+ and H using metal and support/ligand.8,37 In colloidal Pd0.5Ru0.5 catalyst, the heterolytic cleavage of hydrogen proceeded using bifunctional sites of Pd0.5–Ru0.5 catalysis (Scheme 1). The better H2 cleavage over bifunctional sites of Pd0.5Ru0.5 catalyst may have been responsible for the higher catalytic activity of the Pd0.5–Ru0.5 catalyst for the hydrogenation of quinoline.Open in a separate windowScheme 1Plausible transition state for the hydrogenation of quinoline by Pd0.5–Ru0.5–PVP.In summary, we have developed a successful example of Pd0.5–Ru0.5 NPs for the hydrogenation of N-,O-heteroarenes and arenes. This catalyst exhibited the higher activity than Rh catalyst which was extensively studied for the hydrogenation of quinoline. A parametric study demonstrated that methanol was the best solvent for the hydrogenation of heteroarenes, whereas solvent-free conditions facilitated for the hydrogenation of arenes. The catalyst was recycled three times recycled without loss in activity. It is likely that Pd0.5–Ru0.5 NPs could be efficient catalysts for the hydrogenation of nitroarenes and synthesis of N-heterocycles. Currently, we are exploring this possibility.  相似文献   

4.
Peptide-directed Pd-decorated Au and PdAu nanocatalysts for degradation of nitrite in water     
Imann Mosleh  Alireza Abbaspourrad 《RSC advances》2021,11(52):32615
In this work, a palladium binding peptide, Pd4, has been used for the synthesis of catalytically active palladium-decorated gold (Pd-on-Au) nanoparticles (NPs) and palladium–gold (PdxAu100−x) alloy NPs exhibiting high nitrite degradation efficiency. Pd-on-Au NPs with 20% to 300% surface coverage (sc%) of Au showed catalytic activity commensurate with sc%. Additionally, the catalytic activity of PdxAu100−x alloy NPs varied based on palladium composition (x = 6–59). The maximum nitrite removal efficiency of Pd-on-Au and PdxAu100−x alloy NPs was obtained at sc 100% and x = 59, respectively. The synthesized peptide-directed Pd-on-Au catalysts showed an increase in nitrite reduction three and a half times better than monometallic Pd and two and a half times better than PdxAu100−x NPs under comparable conditions. Furthermore, peptide-directed NPs showed high activity after five reuse cycles. Pd-on-Au NPs with more available activated palladium atoms showed high selectivity (98%) toward nitrogen gas production over ammonia.

In this work, a palladium binding peptide, Pd4, has been used for the synthesis of catalytically active palladium-decorated gold (Pd-on-Au) nanoparticles (NPs) and palladium–gold (PdxAu100−x) alloy NPs exhibiting high nitrite degradation efficiency.  相似文献   

5.
Reusable magnetic PdxCoy nanoalloys confined in mesoporous carbons for green Suzuki–Miyaura reactions     
Mohamed Enneiymy  Claude Le Drian  Camlia Matei Ghimbeu  Jean-Michel Becht 《RSC advances》2018,8(31):17176
We report herein PdxCoy nanoalloys confined in mesoporous carbons (Pdx–Coy@MC) prepared by an eco-friendly one-pot approach consisting in the co-assembly of readily available and non-toxic carbon precursors (phloroglucinol, glyoxal) with a porogen template (pluronic F-127) and metallic salts (H2PdCl4 and Co(NO3)2·6H2O) followed by thermal annealing. Three PdxCoy@MC materials with different alloy compositions were prepared (C1: x/y = 90/10; C2: x/y = 75/25; C3 and C4: x/y = 50/50). The nanoalloys were uniformly distributed in the carbon framework and the particle sizes depended on the alloy composition. These composites were then used for Suzuki–Miyaura reactions using either H2O or a 1 : 1 H2O/EtOH mixture as solvent. The Pd50Co50@MC catalyst C3 proved to be the most efficient catalyst (in terms of efficiency and magnetic recovery) affording the coupling products in good to excellent yields. After reaction, C3 was recovered quantitatively by simple magnetic separation and reused up to six times without loss of efficiency. The amount of palladium lost in the reaction mixture after magnetic separation was very low (ca. 0.1 % wt of the amount initially used).

(Pdx–Coy)@MC were prepared in one-pot via an eco-friendly route and used many times for Suzuki reactions in H2O or H2O/EtOH mixture.  相似文献   

6.
Palladium decorated on a new dendritic complex with nitrogen ligation grafted to graphene oxide: fabrication,characterization, and catalytic application     
Mohsen Golestanzadeh  Hossein Naeimi 《RSC advances》2019,9(47):27560
Immobilized Pd nanoparticles on a new ligand, namely, tris(pentaethylene-pentamine)triazine supported on graphene oxide (Pdnp-TPEPTA(L)-GO) was introduced as a novel and robust heterogeneous catalyst for use in C–C bond formation reaction. The Pdnp-TPEPTA(L)-GO catalyst was synthesized by complexation of Pd with TPEPTA as a ligand with high N-ligation sites that were supported on graphene oxide through 3-chloropropyltrimethoxysilane. The prepared catalyst was characterized using some microscopic and spectroscopic techniques. The TPEPTA(L)-GO substrate is a 2D heterogeneous catalyst with a high specific surface area and a large amount of N-ligation sites. The Pdnp-TPEPTA(L)-GO catalyst used in the C–C bond formation reaction between aryl or heteroaryl and phenylboronic acid derivatives was applied towards the synthesis of biaryl units in high isolated yields. Notably, a series of competing experiments were performed to establish the selectivity trends of the presented method. Also, this catalyst system was reusable at least six times without a significant decrease in its catalytic activity.

Immobilized Pd nanoparticles on a new ligand, namely, tris(pentaethylene-pentamine)triazine supported on graphene oxide (Pdnp-TPEPTA(L)-GO) was introduced as a novel and robust heterogeneous catalyst for use in C–C bond formation reaction.  相似文献   

7.
Comparison of electrocatalytic activity of Pt1−xPdx/C catalysts for ethanol electro-oxidation in acidic and alkaline media     
Qiang Zhang  Ting Chen  Rongyan Jiang  Fengxing Jiang 《RSC advances》2020,10(17):10134
In this paper, a comparision of Pt1−xPdx/C catalysts for ethanol-oxidation in acidic and alkaline media has been investigated. We prepared Pt1−xPdx/C catalysts with different ratios of Pt/Pd (x at% = 0, 27, 53, 77 and 100) by the formic acid reduction method. The obtained Pt1−xPdx/C catalysts were characterized by X-ray diffraction (XRD), energy dispersive X-ray spectroscopy (EDX), induced coupled plasma-atomic emission spectroscopy (ICP-AES), X-ray photoelectron spectroscopy (XPS) and transmission electron microscopy (TEM). Structural and morphological investigations of the as-prepared catalysts revealed that the metallic particle size increases with increasing Pd content in the catalyst. The electrocatalytic performances and stabilities of Pt1−xPdx/C catalysts were tested by cyclic voltammetry (CV), linear sweep voltammetry (LSV) and chronoamperometry (CA) measurements for ethanol oxidation in acidic and alkaline media. The electrochemical measurements demonstrate that Pt1−xPdx/C catalysts exhibit much higher electrocatalytic activity for alcohol oxidation in alkaline media than that in acidic media. The composition of Pt/Pd has a significant impact on the ethanol-oxidation in both acidic and alkaline media. The Pt23Pd77/C catalyst shows the highest electrocatalytic performance with a mass specific peak current of 2453.7 mA mgPtPd−1 in alkaline media, which is higher than the Pt77Pd23/C with the maximum of peak current of 339.7 mA mgPtPd−1 in acidic media. Meanwhile, the effect of electrolyte, CH3CH2OH concentrations and scan rates was also studied for ethanol-oxidation in acidic and alkaline media.

The Pt1−xPdx/C catalysts exhibit much higher electrocatalytic activity in alkaline media than in acid media.  相似文献   

8.
Reaction induced robust PdxBiy/SiC catalyst for the gas phase oxidation of monopolistic alcohols     
Pengwei Wang  Lijun Xu  Jianming Zhu  Kunqi Gao  Yan Zhang  Jifen Wang 《RSC advances》2020,10(69):42564
Reaction induced PdxBiy/SiC catalysts exhibit excellent catalytic activity (92% conversion of benzyl alcohol and 98% selectivity of benzyl aldehyde) and stability (time on stream of 200 h) in the gas phase oxidation of alcohols at a low temperature of 240 °C due to the formation of Pd0–Bi2O3 species. TEM indicates that the agglomeration of the 5.8 nm nanoparticles is inhibited under the reaction conditions. The transformation from inactive PdO–Bi2O3 to active Pd0–Bi2O3 under the reaction conditions is confirmed elaborately by XRD and XPS.

Reaction induced PdxBiy/SiC catalysts exhibit excellent catalytic activity and stability in the gas phase oxidation of monopolistic alcohols at a low temperature of 240 °C due to the formation of Pd0–Bi2O3 species.  相似文献   

9.
Enhanced oxygen reduction activity of Pt shells on PdCu truncated octahedra with different compositions     
Xingqiao Wu  Qingfeng Xu  Yucong Yan  Jingbo Huang  Xiao Li  Yi Jiang  Hui Zhang  Deren Yang 《RSC advances》2018,8(61):34853
Pd@Pt core–shell nanocrystals with ultrathin Pt layers have received great attention as active and low Pt loading catalysts for oxygen reduction reaction (ORR). However, the reduction of Pd loading without compromising the catalytic performance is also highly desired since Pd is an expensive and scarce noble-metal. Here we report the epitaxial growth of ultrathin Pt shells on PdxCu truncated octahedra by a seed-mediated approach. The Pd/Cu atomic ratio (x) of the truncated octahedral seeds was tuned from 2, 1 to 0.5 by varying the feeding molar ratio of Pd to Cu precursors. When used as catalysts for ORR, these three PdxCu@Pt core–shell truncated octahedra exhibited substantially enhanced catalytic activities compared to commercial Pt/C. Specifically, Pd2Cu@Pt catalysts achieved the highest area-specific activity (0.46 mA cm−2) and mass activity (0.59 mA μgPt−1) at 0.9 V, which were 2.7 and 4.5 times higher than those of the commercial Pt/C. In addition, these PdxCu@Pt core–shell catalysts showed a similar durability with the commercial Pt/C after 10 000 cycles due to the dissolution of active Cu and Pd in the cores.

PdxCu@Pt core–shell truncated octahedra were synthesized and exhibited substantially enhanced catalytic properties for oxygen reduction reaction relative to Pt/C.  相似文献   

10.
Enhanced catalytic activity over palladium supported on ZrO2@C with NaOH-assisted reduction for decomposition of formic acid     
Tong Wang  Fang Li  Hualiang An  Wei Xue  Yanji Wang 《RSC advances》2019,9(6):3359
A ZrO2@C support based on t-ZrO2 embedded in amorphous carbon was obtained via the pyrolysis of a UiO-66 precursor. Highly dispersed Pd nanoparticles (NPs) were subsequently deposited onto this support, using NaOH-assisted reduction, to obtain a formic acid (FA) decomposition catalyst. This material showed a turnover frequency (TOF) for the heterogeneously-catalyzed decomposition of FA of 8588 h−1 at 60 °C, with 100% H2 selectivity. This performance is ascribed to the uniform dispersion of smaller palladium nanoparticles and a synergistic effect between the metal NPs and support. Even at 30 °C, the complete decomposition of FA was achievable in FA/SF (SF, sodium formate) solution, with a TOF as high as 1857 h−1.

Pd/ZrO2@C was prepared employing UiO-66-derived ZrO2@C as the support and showed high catalytic activity for formic acid decomposition.  相似文献   

11.
Confinement of a Au–N-heterocyclic carbene in a Pd6L12 metal–organic cage     
Lihua Zeng  Shujian Sun  Zhang-Wen Wei  Yu Xin  Liping Liu  Jianyong Zhang 《RSC advances》2020,10(64):39323
A Au(i)–N-heterocyclic-carbene (NHC)-edged Pd6L12 molecular metal–organic cage is assembled from a Au(i)–NHC-based bipyridyl bent ligand and Pd2+. The octahedral cage structure is unambiguously established by NMR, electrospray ionization-mass spectrometry and single crystal X-ray crystallography. The electrochemical behaviour was analyzed by cyclic voltammetry. The octahedral cage has a central cavity for guest binding, and is capable of encapsulating PF6 and BF4 anions within the cavity.

A Au(i)–NHC-edged Pd6L12 molecular cage is assembled from a Au(i)–NHC-based bipyridyl bent ligand and Pd2+.  相似文献   

12.
A comparative study on atomically precise Au nanoclusters as catalysts for the aldehyde–alkyne–amine (A3) coupling reaction: ligand effects on the nature of the catalysis and efficiency     
Ying-Zhou Li  Weng Kee Leong 《RSC advances》2019,9(10):5475
Atomically precise Au13 nanoclusters stabilized by stibines catalyze the aldehyde–alkyne–amine coupling reaction more efficiently than those stabilized by thiols or phosphines. The nature of the catalytic activity is also different, and may be attributed to the weaker coordinating ability of the stibine ligands.

A stibine-stabilised gold nanocluster which acts as a homogeneous catalyst in the A3 coupling reaction.

Atomically precise Au nanoclusters have recently emerged as a new class of catalysts for a range of reactions, including the A3 coupling reaction, which involves a highly efficient assembly, in a single step, of an aldehyde, an amine and an acetylene to form a propargylamine. Propargylamines are valuable precursors to a number of organic heterocyclic substrates, natural products and therapeutic drugs, and have thus found broad applications in synthetic and pharmaceutical chemistry.1 The A3 coupling reaction is generally catalyzed by transition metal species, for example, those of Cu, Zn, Rh, Ru, Ir, Ni, Fe, Ag and Au.2 Among them, noble-metal salts, complexes or nanoparticles, especially those of Au, are found to be more efficient as they can more readily activate the acetylene as the corresponding acetylide.3 Recently, recyclable Au(0) nanoparticles have been shown to be more efficient than the widely used Au(i) and Au(iii) complexes as catalysts for this reaction.4A challenge with the Au nanoparticle catalysts is the establishment of correlations between the particle structure and its catalytic performance, due to the fluidity of the surface structure (coordination pattern between surface metal and ligands).5 Compared to larger gold nanoparticles, therefore, atomically precise Au nanoclusters with their well-defined molecular structures, make it possible to correlate their surface structures with catalytic performance. The higher surface area to volume associated with their smaller metal core size (<2 nm) also translates to a reduced amount of the catalyst needed. For example, Jin, et al., have reported that the thiolate-protected nanocluster [Au38(SC2H4Ph)24] (Au38) could catalyze the A3 coupling reaction, and the high catalytic efficiency was ascribed to the synergistic effect between the electron-rich Au core and the electron-deficient surface of the nanoclusters.6 The group of Obora reported the use of the thiolate-protected nanocluster [Au25(SC2H4Ph)18][TOA] (Au25, TOA = tetraoctylammonium) as an efficient catalyst for the reaction and they believed that the surface Au atoms, sterically unblocked by the thiolate ligands, served as active sites in the catalytic process,7 and Li, et al., reported that the reaction could also be catalyzed by supported Au nanoparticles derived from [Au25(PPh3)10(C Created by potrace 1.16, written by Peter Selinger 2001-2019 CPh)5]X2 (Au′25, X = Cl or Br), which has phosphine and alkyne as protecting ligands. It was demonstrated in the latter that removal of one phosphine ligand from the supported Au′25 cluster was necessary to enable catalysis.8 Finally, Wu, et al., reported an Au–Cd nanocluster, [Au13Cd2(PPh3)6(SC2H4Ph)6(NO3)2]2Cd(NO3)4 (Au26Cd4), with high catalytic activity which was ascribed to the cooperative effect between the peripheral Cd and adjacent Au atoms on the surface of the Au13 icosahedral core.9It is clear from the foregoing that the surface structure of the Au nanocluster, including the ligand binding strength, may have a significant influence on the catalytic performance. In this regard, Au nanoclusters protected by more labile ligands may be expected to be more efficient catalysts. We have recently reported the preparation of an Au13 nanocluster stabilized by stibine ligands, viz., [Au13{Sb(p-tolyl)3}8Cl4][Cl] (Au13).10 We anticipated that it may catalyze the A3 coupling reaction more readily than the previously reported phosphine- and thiolate-protected clusters [Au11(PPh3)8Cl2][Cl] (Au11),11 and Au25;12 the weaker bond between the Au13 core and stibine ligands should lead to more ready ligand dissociation to expose catalytically active Au sites.The thermogravimetric analysis (TGA) curves for crystalline Au25, Au11 and Au13 show that weight loss at 300 °C is ca. 37, 50 and 56%, respectively (Fig. 1), matching the theoretical values (37, 50 and 57%, respectively) corresponding to loss of all coordinating ligands. While the loss for Au13 begins at ca. 145 °C, that for Au11 and Au25 begin at ca. 165 °C and 195 °C, respectively, clearly demonstrating that the stibine-protected nanocluster is thermally less stable than the phosphine- and thiolate-protected ones.Open in a separate windowFig. 1TGA curves for Au25, Au13 and Au11.A comparative study was carried out for the A3-coupling reaction of phenylacetylene, piperidine and benzaldehyde with Au38, Au25, Au′25 and Au26Cd4 as catalysts (i) salts as catalysts.13 (2) It has been suggested that elevated reaction temperatures were required as it favoured formation of the cationic imine intermediate, and facilitated catalyst activation via removal of the protecting ligands (entry 4). (3) An inert atmosphere was needed in order to avoid rapid oxidation of the substrates exacerbated by the elevated temperature. From the economic and environmental perspectives, therefore, an efficient catalyst which can catalyze the A3 coupling reaction under ambient conditions, without an organic solvent (neat), and in air, would be desirable.A3 coupling reaction of benzaldehyde, piperidine, and phenylacetylene with Au nanoclusters as catalystsa
EntryCat.bCat. loadingc (mol%)Time (h)Sol. T (°C)Conv.d (%)Ref.
1eAu250.1024Tol8089f 7
2Au250.0135None80∼95%g 6
3Au38/CeO20.0105None8099
0.0105H2O8094
0.0105Tol8080
4Au′25/TiO20.01218H2O8066 8
0.01218H2O10090
5Au26Cd40.505DCMr.t.80f 9
Entries below are from this work
6Au250.5012DCMh210
7Au110.5012DCMh210
8Au130.5024DCM2147
9Au130.5019CHCl32125
10Au130.5012None2179
11Au250.5012None214
12Au110.5012None21Trace
13Au130.0315None2132
14Au250.0315None210
15Au110.0315None210
16Au130.035None5081
12None5092
17Au250.035None507
12None5069
18Au110.0312None50Trace
Open in a separate windowaBenzaldehyde (1.0 mmol), piperidine (1.2 mmol), phenylacetylene (1.3 mmol), solvent (1.0 ml, if present).bSupported catalysts are at 1 wt%, i.e., 1 mg of Au nanocluster on 100 mg of support.cBased on Au nanoclusters, with respect to amount of benzaldehyde.dConversion of benzaldehyde, determined by 1H NMR.eDifferent reactant ratios used: benzaldehyde (0.5 mmol), piperidine (1.0 mmol), phenylacetylene (1.5 mmol).fIsolated yields.gConversion not given in original report; estimated from 1H NMR plot provided in ESI.hThree other solvents were also tested: DCE, CDCl3 and toluene.In this work, we have found that Au25 and Au11 (0.5 mol%) failed to afford any product at room temperature (21 °C), under aerobic conditions, in any of the organic solvents tested (dichloromethane (DCM), 1,2-dichloroethane (DCE), chloroform (CHCl3) or toluene (Tol)) (entries 6, 7) even after 12 h. Another previously reported icosahedral cluster Au13(PPh3)4(SC2H4Ph)4 also showed no catalytic activity under similar reaction conditions.9 In contrast, Au13 gave the desired product under the same reaction conditions, albeit at a relatively low conversion of benzaldehyde: 47% in DCM after 24 h, and 25% in CHCl3 after 19 h (entries 8, 9). These figures were obtained by monitoring the reactions in CD2Cl2 and CDCl3 by 1H NMR spectroscopy (Fig. S1–S4). The lack of activity in the thiolate and phosphine-stabilised nanoclusters is probably related to the lack of catalytically active sites at room temperature since these ligands are more strongly bound.The neat reactions (all the three reactants are liquids at room temperature) showed obvious improvement for Au13, with the conversion rate reaching 79% after 12 h (entry 10), and ∼90% with prolonged reaction time (36 h). This can be ascribed to good solubility of the catalyst, and the higher substrate concentrations. The reaction kinetics monitored through 1H NMR spectroscopy (Fig. S5), showed that conversion to propargylamine rapidly increased to 57% in the first 5 h and further to >90% over a prolonged reaction time (Fig. 2). Fitting the data for the first 10 h to first-order reaction kinetics gave a rate constant k = 0.14 h−1 (Fig. S6). The turnover frequency (TOFs) at 6 min (2.9% conversion), was estimated to be 58 h−1 per Au13 cluster or 4.5 h−1 per Au atom. Both Au25 and Au11 showed very low conversion at 12 h of reaction under the same conditions (0.5 mol% catalyst loading, neat) (Fig. 2). As may be expected, lowering the catalyst loading to 0.03 mol% led to a significant drop in conversion rates (entries 13–15), and increasing the temperature to 50 °C had a dramatic effect on the conversion for Au13 and Au25 but not Au11 (entries 16–18). At 50 °C, the reaction with Au25 was still within the induction period for the first 5 h. A comparison of the performance of Au13 with Au25 at the 5 h mark showed that the former (81% conversion, TON = 208 per Au atom) outperformed the latter (7% conversion, TON = 9.3 per Au atom) (Fig. S20 and S21).Open in a separate windowFig. 2Conversion (%) of benzaldehyde as a function of reaction time (h) catalysed by the various gold nanoclusters. Timescale at the top is for Au11 and that at the bottom is for Au13 and Au25.It is known that the active catalyst in many reactions catalyzed by Au salts or complexes are actually Au(0) nanoparticles, or other Au clusters formed during the reaction; these generally exhibit an induction period during which the catalytically active species is formed, at a slower rate than for the product-formation reaction.14 In the case of Au13 here, no induction period was observed (Fig. 2). Since the induction period is related to the concentration of the precursor to the active catalyst, it is also possible that there may have been a short induction period. To rule this out, the reaction was repeated with a lower catalyst loading of 0.03 mol%; although the conversion rate was decreased significantly (32% conversion after 15 h, entry 13), an induction period was still not observed (Fig. S8). In all these reactions, there were no obvious signs of nanoparticle formation or precipitation. The catalyst could be recovered from the room temperature reaction by precipitation with hexane (after 60 h, at 93% conversion), and the UV-Vis spectrum of the recovered catalyst showed the characteristic absorption peaks of Au13 and some decomposition, although we have not been able to determine the identity of the decomposition products (Fig. S7). Monitoring of the reaction by 1H NMR spectroscopy also showed that Au13, or a structural analogue (a “less intact” Au13), was present throughout the reaction; there were no obvious change in intensity or position of the resonances (Fig. S9). In addition, we have also found that although (p-tolyl)3SbAuCl, the most likely dissociated fragment from Au13, could catalyze the reaction it had a much lower efficiency (Fig. S10). Taken together, these results suggest that the catalytically active species is most likely Au13, or a “less intact” Au13 nanocluster in which one or more ligands have dissociated or been replaced. Thus Au13 behaves as a homogeneous, or what has been termed as quasi-homogeneous, catalyst,15 and its higher catalytic activity may be attributed to the weaker coordination of the stibine ligands to the Au13 core.For Au25, monitoring the reaction by UV-Vis spectroscopy showed that it gradually decomposed during reaction, as reflected by the replacement of its characteristic peak at ca. 670 nm by another at ca. 840 nm (Fig. 3). The absence of a surface plasmonic resonance (SPR) peak suggests that little or no larger Au(0) nanoparticles (>4 nm) were formed.16 Monitoring the same reaction by 1H NMR spectroscopy showed that the Au25 gradually disappeared after the induction period, accompanied by an obvious increase in the conversion (Fig. S11). The oxidative decomposition of Au25 to generate some Au(i) species during the catalysis of styrene oxidation has already been noted previously.17 Together with the results here, we propose that Au25 is not the active catalyst for the A3 coupling reaction, at least under the given reaction conditions here. Instead, in the presence of air, there is decomposition into the active catalyst which is probably a mixture of smaller Au clusters.Open in a separate windowFig. 3Electronic spectra of the crudes catalyzed by Au25 (0.5 mol%) in neat conditions at 21 °C. The percentages in parentheses are the corresponding conversion of benzaldehyde.Similarly, the much lower catalytic efficiency for Au11 may be partly attributed to its conversion to neutral Au11(PPh3)7Cl3 (Au′11) which is poorly soluble in the reaction mixture and precipitated out, leading to a much lower precursor catalyst concentration; the orange Au′11 could be collected by washing with hexane and was identified by its 1H NMR spectrum (Fig. S12 and S13).11 Just as in the case of Au25, no SPR peak was observed in the UV-Vis spectrum throughout the reaction (Fig. S14, S15 and S18). That the catalytic activity of Ph3PAuCl, albeit at a lower catalyst loading based on Au atoms, was observed to be better in comparison initially, may be attributed to the absence of an induction period (Fig. S16), and suggests that the catalytically active species are smaller Au clusters formed through decomposition of Au11 and/or Au′11. Monitoring the A3 coupling reaction with Au11 showed that while the nanocluster was more soluble at 50 °C, it also decomposed to other unknown species as conversion increased (Fig. S17); the UV-Vis spectrum of the crude at 80% conversion showed no distinct absorption peak, indicating almost complete decomposition of Au11 and Au′11 (Fig. S18). Consistent with all these is the observation that initial heating of the reaction mixture at 50 °C for 12 h significantly diminished the induction period (Fig. S19). That Au11 nanoclusters can act as catalyst precursors has also been reported previously.18In conclusion, we have found that the stibine-protected nanocluster Au13 was a more efficient catalyst than the thiolate- and phosphine-protected nanoclusters for the aldehyde–acetylene–amine (A3) coupling reaction. The reaction proceeded under mild reaction conditions and in air. In contrast to the thiolate- and phosphine-protected nanoclusters, the stibine-protected nanocluster behaved as a homogeneous or quasi-homogeneous catalyst. The effect of the ligand on catalytic performance for Au nanoclusters has implications for the design and preparation of more catalytically active Au nanoclusters. Studies into this and with a wider substrate scope are currently underway.  相似文献   

13.
One pot synthesis of aryl nitriles from aromatic aldehydes in a water environment     
Hongyan Lin  Ziling Zhou  Xiaopeng Ma  Qingqing Chen  Hongwei Han  Xiaoming Wang  Jinliang Qi  Yonghua Yang 《RSC advances》2021,11(39):24232
In this study, we found a green method to obtain aryl nitriles from aromatic aldehyde in water. This simple process was modified from a conventional method. Compared with those approaches, we used water as the solvent instead of harmful chemical reagents. In this one-pot conversion, we got twenty-five aryl nitriles conveniently with pollution to the environment being minimized. Furthermore, we confirmed the reaction mechanism by capturing the intermediates, aldoximes.

In a formic acid–H2O solution (60% : 40%), most aromatic aldehydes react efficiently with hydroxylamine hydrochloride and sodium acetate to form nitriles, where formic acid acts as both catalyst and solvent.  相似文献   

14.
Microwave-assisted pyrolysis of Pachira aquatica leaves as a catalyst for the oxygen reduction reaction     
Sun-Tang Chang  Huan-Ping Jhong  Yu-Chung Chang  Chia-Chi Liu  Tai-Chin Chiang  Hsin-Chih Huang  Chen-Hao Wang 《RSC advances》2020,10(20):11543
In this study, biomimetic Mg–Nx–Cy from Pachira aquatica leaves were mixed with carbon black (L/C catalyst), in which the mixture was treated by a conventional microwave oven at 700 W and 2 min, exhibiting high catalytic activity for the oxygen reduction reaction (ORR). By using a microwave-assisted process, it not only offers a cheaper and faster way to synthesize the catalyst compared to the conventional furnace process but also avoids the decomposition of the N4-structure. Using the optimized conditions, the L/C catalyst exhibits an electron transfer number of 3.90 and an HO2 yield of only 5% at 0.25 V vs. RHE, which is close to the perfect four electron-transfer pathway. Besides, the L/C catalyst offers superior performance and long-term stability up to 20 000 s. The L/C catalyst contains a high proportion of quaternary-type nitrogen, Mg–Nx–Cy, and –C–S–C– which can be the active sites for the ORR.

In this study, biomimetic Mg–Nx–Cy from Pachira aquatica leaves were mixed with carbon black, in which the mixture was treated by a conventional microwave oven at 700 W and 2 min, exhibiting high catalytic activity for the oxygen reduction reaction.  相似文献   

15.
Preparation and investigation of highly selective solid acid catalysts with sodium lignosulfonate for hydrolysis of hemicellulose in corncob     
Xun Li  Fengyao Shu  Chao He  Shuna Liu  Noppol Leksawasdi  Qiong Wang  Wei Qi  Md. Asraful Alam  Zhenhong Yuan  Yi Gao 《RSC advances》2018,8(20):10922
Saccharification of lignocellulose is a necessary procedure for deconstructing the complex structure for building a sugar platform that can be used for producing biofuel and high-value chemicals. In this study, a carbon-based solid acid catalyst derived from sodium lignosulfonate, a waste by-product from the paper industry, was successfully prepared and used for the hydrolysis of hemicellulose in corncob. The optimum preparation conditions for the catalyst were determined to be carbonization at 250 °C for 6 h, followed by sulfonation with concentrated H2SO4 (98%) and oxidation with 10% H2O2 (solid–liquid ratio of 1 : 75 g mL−1) at 50 °C for 90 min. SEM, XRD, FT-IR, elemental analysis and acid–base titration were used for the characterization of the catalysts. It was found that 0.68 mmol g−1 SO3H and 4.78 mmol g−1 total acid were loaded onto the catalyst. When corncob was hydrolyzed by this catalyst at 130 °C for 12 h, the catalyst exhibited high selectivity and produced a relatively high xylose yield of up to 84.2% (w/w) with a few by-products. Under these conditions, the retention rate of cellulose was 82.5%, and the selectivity reached 86.75%. After 5 cycles of reuse, the catalyst still showed high catalytic activity, with slightly decreased yields of xylose from 84.2% to 70.7%.

A novel carbon-based catalyst with high catalytic ability and xylose selectivity was prepared from sodium lignosulfonate.  相似文献   

16.
The influence of the polymerization approach on the catalytic performance of novel porous poly (ionic liquid)s for green synthesis of pharmaceutical spiro-4-thiazolidinones     
Zahra Elyasi  Javad Safaei Ghomi  Gholam Reza Najafi  Mohammad Reza Zand Monfared 《RSC advances》2020,10(72):44159
Although poly (ionic liquids) (PILs) have attracted great research interest owing to their various applications, the performance of nanoporous PILs has been rarely developed in the catalysis field. To this end, a micro–mesoporous PIL with acid–base bifunctional active sites was designed and fabricated by two different polymerization protocols including hydrothermal and classical precipitation polymerization in this paper. Based on our observations, hydrothermal conditions (high temperature and pressure) enabled the proposed sonocatalyst to possess a great porous structure with a high specific surface area (SBET: 315 m2 g−1) and thermal stability (around 450 °C for 45% weight loss) through strengthening cross-linking. In a comparative study, the preferred nanoporous PIL was selected and utilized as the sonocatalyst in a multicomponent reaction of isatins, primary amines, and thioglycolic acid. In the following, a variety of new and known pharmaceutical spiro-4-thiazolidinone derivatives were synthesized at room temperature and obtained excellent yields (>90%) within short reaction times (4–12 min) owing to the substantial synergistic effect between ultrasound irradiation and magnetically separable catalyst.

Sustainable synthesize of a new mesoporous poly (ionic liquid) as acid–base bifunctional catalyst for environmental being preparation of monospiro derivatives has been developed.  相似文献   

17.
Copper-catalyzed aerobic decarboxylative coupling between cyclic α-amino acids and diverse C–H nucleophiles with low catalyst loading     
Jing Guo  Ying Xie  Qiao-Lei Wu  Wen-Tian Zeng  Albert S. C. Chan  Jiang Weng  Gui Lu 《RSC advances》2018,8(29):16202
An aerobic decarboxylative cross-coupling of α-amino acids with diverse C–H nucleophiles has been realized using Cu2(OH)2CO3 (1 mol%) as the catalyst under air. This protocol enables highly efficient formation of various C(sp3)–C(sp3), C(sp3)–C(sp2) and C(sp3)–C(sp) bonds under simple conditions without the use of any ligand or extra oxidant, providing a practical approach to numerous nitrogen-containing compounds in good to excellent yields. The efficiency and practicability were also demonstrated by the gram-scale experiment and three-step synthesis of a Rad51 inhibitor.

An aerobic decarboxylative cross-coupling of α-amino acids was realized using 1 mol% Cu2(OH)2CO3 catalyst under ligand free conditions.  相似文献   

18.
Structural exploration of AuxM− (M = Si,Ge, Sn; x = 9–12) clusters with a revised genetic algorithm     
Ping Huang  Yan Jiang  Tianquan Liang  Enhui Wu  Jun Li  Jing Hou 《RSC advances》2019,9(13):7432
We used a revised genetic algorithm (GA) to explore the potential energy surface (PES) of AuxM (x = 9–12; M = Si, Ge, Sn) clusters. The most interesting finding in the structural study of AuxSi (x = 9–12) is the 3D (Au9Si and Au10Si) → quasi-planar 2D (Au11Si and Au12Si) structural evolution of the Si-doped clusters, which reflects the competition of Au–Au interactions (forming a 2D structure) and Au–Si interactions (forming a 3D structure). The AuxM (x = 9–12; M = Ge, Sn) clusters have quasi-planar structures, which suggests a lower tendency of sp3 hybridization and a similarity of electronic structure for the Ge or Sn atom. Au9Si and Au10Si have a 3D structure, which can be viewed as being built from Au8Si and Au9Si with an extra Au atom bonded to a terminal gold atom, respectively. In contrast, the quasi-planar structures of AuxM (x = 9–12; M = Ge, Sn) reflect the domination of the Au–Au interactions. Including the spin–orbit (SO) effects is very important to calculate the simulated spectrum (structural fingerprint information) in order to obtain quantitative agreement between theoretical and future experimental PES spectra.

We used a revised genetic algorithm (GA) to explore the potential energy surface (PES) of AuxM (x = 9–12; M = Si, Ge, Sn) clusters.  相似文献   

19.
Acid–base sites synergistic catalysis over Mg–Zr–Al mixed metal oxide toward synthesis of diethyl carbonate     
Tingting Yan  Weihan Bing  Ming Xu  Yinwen Li  Yusen Yang  Guoqing Cui  Lan Yang  Min Wei 《RSC advances》2018,8(9):4695
In heterogeneous catalysis processes, development of high-performance acid–base sites synergistic catalysis has drawn increasing attention. In this work, we prepared Mg/Zr/Al mixed metal oxides (denoted as Mg2ZrxAl1−x–MMO) derived from Mg–Zr–Al layered double hydroxides (LDHs) precursors. Their catalytic performance toward the synthesis of diethyl carbonate (DEC) from urea and ethanol was studied in detail, and the highest catalytic activity was obtained over the Mg2Zr0.53Al0.47MMO catalyst (DEC yield: 37.6%). By establishing correlation between the catalytic performance and Lewis acid–base sites measured by NH3-TPD and CO2-TPD, it is found that both weak acid site and medium strength base site contribute to the overall yield of DEC, which demonstrates an acid–base synergistic catalysis in this reaction. In addition, in situ Fourier transform infrared spectroscopy (in situ FTIR) measurements reveal that the Lewis base site activates ethanol to give ethoxide species; while Lewis acid site facilitates the activated adsorption of urea and the intermediate ethyl carbamate (EC). Therefore, this work provides an effective method for the preparation of tunable acid–base catalysts based on LDHs precursor approach, which can be potentially used in cooperative acid–base catalysis reaction.

Mg/Zr/Al mixed metal oxides were prepared via a facile phase transformation process of hydrotalcite precursors, which showed acid–base sites synergistic catalytic performance toward the synthesis of diethyl carbonate from ethanol and urea.  相似文献   

20.
Au-based bimetallic catalysts: how the synergy between two metals affects their catalytic activity     
Jin Sha  Sbastien Paul  Franck Dumeignil  Robert Wojcieszak 《RSC advances》2019,9(51):29888
Supported bimetallic nanoparticles are particularly attractive catalysts due to increased activity and stability compared to their monometallic counterparts. In this work, gold-based catalysts have been studied as catalysts for the selective base-free oxidation of glucose. TiO2-supported Au–Pd and Au–Cu series prepared by the sol-immobilization and precipitation-reduction methods, respectively, showed a significant synergistic effect, particularly when the theoretical weight ratio of the two metals was close to 1 : 1 (with an actual experimental bulk Au/Pd molar ratio of ca. 0.8 and ca. 0.4 for Au/Cu) in both cases. XPS analysis showed that the presence of Auδ+, Pd2+ and CuOH species played an important role in the base-free glucose oxidation.

Supported bimetallic nanoparticles are particularly attractive catalysts due to increased activity and stability compared to their monometallic counterparts.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号