Sulfur-Containing Polymer as a Platform for ... - ACS Publications

Sulfur-Containing Polymer as a Platform for Synthesis of Size-. Controlled Pd Nanoparticles for Selective Semihydrogenation of. Alkynes. Yuangong Zhan...
0 downloads 0 Views 4MB Size
Article Cite This: Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

pubs.acs.org/IECR

Sulfur-Containing Polymer As a Platform for Synthesis of SizeControlled Pd Nanoparticles for Selective Semihydrogenation of Alkynes Yuangong Zhang,† Xin Wen,*,† Yongqing Shi,† Ru Yue,† Libin Bai,† Qingtao Liu,‡ and Xinwu Ba*,† †

Downloaded via UNIV OF LIVERPOOL on January 10, 2019 at 11:30:59 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Key Laboratory of Chemical Biology of Hebei Province, College of Chemistry and Environmental Science, Hebei University, Baoding 071002, People’s Republic of China ‡ Hebei Chemical and Pharmaceutical College, Shijiazhuang 050026, People’s Republic of China S Supporting Information *

ABSTRACT: Sulfur-containing polymer supported palladium nanoparticles (Pd/SPMB) were prepared by using the crosslinked poly(N,N′-methylene bis(acrylamide)) (SPMB) as the platform. The SPMB was synthesized via the reversible addition−fragmentation chain transfer (RAFT) polymerization, from which the sulfur inherently exists in the RAFT polymer due to the use of thiocarbonate as the chain transfer agent, and could be further utilized as a modifier of Pd nanoparticles. The Pd/SPMB showed size-dependent activity and selectivity in the semihydrogenation of alkynes to alkenes. The catalyst with an average Pd size of 5.3 nm was active in the semihydrogenation of both internal and terminal alkynes. Most of the surface of Pd NPs was covered (or poisoned) by sulfide and thiolate due to the moderate complexation strength between Pd NPs and SPMB, thus leading to the selectivity enhancement toward alkene. Furthermore, the catalyst was highly stable and can be reused 10 times without an appreciable loss in either activity or selectivity.

1. INTRODUCTION Semihydrogenation of alkynes to alkenes is a highly important reaction in fine chemical synthesis, such as for bioactive molecules, pharmaceutical drugs, and natural products.1−7 Selectivity control is one of the key issues in this transformation. Palladium has proved to be the most active catalyst. In fact, the surface of the unmodified Pd catalyst is highly unselective and readily leads to the overhydrogenation of alkene. Toward a good improvement in the selectivity for alkene, two common approaches are always adopted in the literature. One is the use of alloy or an intermetallic compound with a second metal8 (Ga,9 Bi,10 Zn,11 Sn,12 In,13 and Pb14,15) and core−shell nanostructures (Pd@Ag,16 (Pd@Ag)@CeO2,17 and Pd@sulfoxide-SiO218). Another is to modify the surface of the Pd catalyst with heteroatom-containing (N,19−21 P,22−25 and B26) organic ligands or macromolecular modifiers,26−30 of which weakly bound ligands are preferred to suppress overpoisoning.31 With this strategy, the selectivity can be precisely tuned by the topological structure of ligands (steric effect) as well as the preferable poisoning of the active catalytic sites (electronic effect).32 Generally, the preparation of the catalyst usually requires a multistep procedure and sophisticated organic ligands. Meanwhile, overhydrogenation as well as harsh reaction conditions limits their practical applications. Reversible addition−fragmentation chain transfer (RAFT) polymerization is one of the most versatile reversibledeactivation radical polymerizations and emerges as a powerful © XXXX American Chemical Society

approach in the synthesis of polymers as well as nanomaterials, such as micelles, vesicles, and hybrid nanoparticles.33−36 It exhibits good tolerance to a diverse range of functional groups in monomers, solvents, and initiators.37−40 Specifically, to achieve the deactivation−activation equilibrium of the radical, it utilizes a thiocarbonylthio-containing chain transfer agent (CTA) to tune the radically induced degenerative chain transfer reaction between a thiocarbonylthio moiety and a propagating radical, thus RAFT polymer features a thiocarbonylthio moiety in the main chain or terminal. In this context, the sulfur that inherently exists in the RAFT polymer due to the use of thiocarbonate as the CTA could be further used as a selectivity modifier or activity promoter in the hydrogenation catalyst.41 The catalytic performance of the RAFT polymer modified Pd catalyst may be markedly enhanced under the promotional effect of a metal−sulfur complex.42−47 In a continuing attempt to achieve a higher selectivity for alkene,48 we herein demonstrate the preparation of sizecontrolled Pd nanoparticles (NPs) supported on a RAFT polymer gel and their size-dependent activity and selectivity in the semihydrogenation of alkynes to alkenes. It is worth noting that the thiocarbonylthio moiety in RAFT polymer gel plays Received: Revised: Accepted: Published: A

October 6, 2018 December 19, 2018 December 24, 2018 December 24, 2018 DOI: 10.1021/acs.iecr.8b04913 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research Scheme 1. Schematic Illustration for the Preparation of Pd/SPMB

Figure 1. TEM images and particle size distributions of (a) Pd/SPMB-0.9, (b) Pd/SPMB-1.8, (c) Pd/SPMB-2.7, and (d) Pd/SPMB-3.6.

selectivity for the semihydrogenation of both internal and terminal alkynes, due to the formation of a Pdx-Sy/thiolate complex on the surface of Pd NPs. To the best of our

dual functionalities: the chain transfer agent in the RAFT polymerization and the precursor of the Pd modifier. The optimum catalyst Pd/SPMB-2.7 exhibits excellent activity and B

DOI: 10.1021/acs.iecr.8b04913 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research knowledge, this is the first report on the RAFT polymer stabilized Pd catalyst and its application in the semihydrogenation of alkynes. The results show the RAFT polymer simultaneously acts as a well-sourced catalyst support and an excellent selectivity modifier.

TWIN instrument at a voltage of 200 kV. Scanning electron microscope (SEM) images were acquired on a JEOL JSM7500 electron microscope. Brunauer−Emmet−Teller (BET) surface area was assessed using a Micromeritics Tristar II 3020 surface area and pore analyzer. CO and H2 chemisorption were obtained with a Micromeritics AutoChem II 2920 equipped with a thermal conductivity detector. CO and H2 chemisorption were performed at 20 and 55 °C, respectively. Elemental analysis was determined on a Carlo-Erba 1106 elemental analyzer. Thermogravimetric analysis (TGA) was carried out on a PerkinElmer Pyris 6 type thermogravimetric analyzer under a nitrogen atmosphere (10 mL min−1) at a heating rate of 20 °C min−1. X-ray diffraction (XRD) patterns were acquired with a Bruker D8 diffractometer using a Cu Kα radiation source at 40 kV and 100 mA with a step size of 0.02° over the range 10−80°. Inductively coupled plasma optical emission spectrometry (ICP-OES) measurement was conducted using an Agilent 5110 instrument. X-ray photoelectron spectroscopy (XPS) analysis was performed on a Thermo Scientific ESCALab 250Xi using 200 W monochromated Al Kα radiation with a 500 μm X-ray spot, and the hydrocarbon C 1s line at 284.8 eV from adventitious carbon was used for energy referencing.

2. EXPERIMENTAL SECTION 2.1. Materials. N,N′-methylenebis(acrylamide) (99%) and palladium acetate (99%) were purchased from J&K Scientific Ltd., and sodium borohydride (99%), phenylacetylene (98%), and the reactants (unless otherwise noted) were purchased from Aladdin Reagent Co. Ltd. and used without further purification. 1-(Phenylethynyl)-4-(trifluoromethyl) benzene, 1fluoro-4-(phenylethynyl) benzene,49 and the chain transfer agent, S,S′-bis(α,α′-dimethyl-α″-acetic acid)-trithiocarbonate,50 were synthesized according to a reported method. 2.2. Synthesis of RAFT Polymer Gels. The sulfurcontaining polymer gel (SPMB) was synthesized via a RAFTmediated cross-linking polymerization. In brief, N,N′methylenebis(acrylamide) (1.0 g, 6.5 mmol) and chain transfer agent S,S′-bis(α,α′-dimethyl-α″-acetic acid)-trithiocarbonate (CTA; 183.1 mg, 0.65 mmol) were dissolved in 20 mL of ethanol, and oxygen was removed using three repeated vacuum-nitrogen cycles. After the AIBN initiator (53.3 mg, 0.33 mmol) was added, the polymerization was conducted at 70 °C in an oil bath for 2 h. Afterward, the gel was washed with water three times and lyophilized to get a yellow product. The sulfur-free polymer gel PMB was prepared using the same method as that of SPMB but without using the chain transfer agent. 2.3. Synthesis of RAFT Polymer Gels Supported Pd NPs. In a representative experiment, to an acetonitrile solution (5.0 μM, 2 mL) of palladium acetate, a certain amount of SPMB [the mole ratio of S and Pd (S/Pd) was tuned to be 0.9, 1.8, 2.7 and 3.6] was added and stirred for 1 h at room temperature. After sodium borohydride (1 mg, 0.03 mmol) was added, the color of the solution immediately turned to dark brown, and the reaction mixture was stirred for another 1 h. Finally, the Pd/SPMB catalysts were denoted as Pd/SPMB0.9, Pd/SPMB-1.8, Pd/SPMB-2.7, and Pd/SPMB-3.6 and dispersed in 2 mL of acetonitrile. The Pd/PMB catalyst was prepared using the same method as for Pd/SPMB-2.7. 2.4. Semihydrogenation of Alkynes. To a 2 mL volume of the above Pd/SPMB catalyst, another 2 mL of acetonitrile solution containing alkyne (1 mmol) was added. Then, the reaction system was thoroughly degassed by evacuation and filled with 1 atm of H2. The solution was then stirred vigorously using a magnetic stirrer at room temperature. During the course of the reaction, 20 μL of samples were drawn with a syringe inserted on one arm of the reactor at 5− 10 min intervals. The reaction was monitored by GC. The crude products were purified by column chromatography on silica gel to withhold the catalyst. Reaction samples were analyzed by Agilent 7820A gas chromatography equipped with an FID detector using mesitylene as an internal standard. The carbon balance for the calculation method was 100 ± 5%. The products were identified using gas chromatography−mass spectrometry (GC-MS) on an Agilent 7890−5977a spectrometer. The Z- and E-isomers were determined with a Bruker Avance III 600 spectrometer (NMR) using tetramethylsilane as the reference. 2.5. Characterizations. Transmission electron microscope (TEM) images were obtained with an FEI Tecnai G2 F20 S-

3. RESULTS AND DISCUSSION Pd/SPMB catalysts were facilely synthesized as presented in Scheme 1. First, the cross-linked poly(N,N′-methylene bisTable 1. Semihydrogenation of Phenylacetylene (1a) to Styrene (1b) Using Pd/SPMB Catalystsa entry

catalyst

1

Pd/ SPMB0.9 Pd/ SPMB1.8 Pd/ SPMB2.7 Pd/ SPMB3.6 Pd/PMB

2 3 4 5c

mole ratio (S/Pd)

size of Pd (nm)

T (min)

conv. (%)b

sel. for 1b (%)b

0.9

9.1

70

94

86

1.8

7.2

110

96

86

2.7

5.3

140

>99

96

3.6

2.7

180

96

93

90

96

84

a

Reaction conditions: phenylacetylene 1a (1.0 mmol), Pd catalyst, CH3CN (4 mL), room temperature, and the carbon balance is 100 ± 5%. bThe conversion and selectivity were determined by GC using mesitylene as an internal standard. cThe sulfur-free catalyst support PMB was used, and the weight of the catalyst support was the same as that of Pd/SPMB-2.7.

Scheme 2. Semihydrogenation of Phenylacetylene (1a) to Styrene (1b) over Pd/SPMB

(acrylamide)) (SPMB) was prepared via a RAFT-mediated cross-linking polymerization initiated by 2,2-azobis(isobutyronitrile) (AIBN), where S,S′-bis(α,α′-dimethyl-α″acetic acid)-trithiocarbonate was used as the CTA. The resultant cross-linked polymer SPMB showed a porous C

DOI: 10.1021/acs.iecr.8b04913 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

Figure 4. Recyclability of Pd/SPMB-2.7 in the semihydrogenation of phenylacetylene (1a).

Figure 2. Dependence of the catalytic performance of Pd/SPMB toward the semihydrogenation of phenylacetylene (1a) on the size of Pd NPs.

can be considered as a qualified support for the Pd-based catalysts in the semihydrogenation process under ambient conditions. Subsequently, different amounts of SPMB (S/Pd = 0.9, 1.8, 2.7 and 3.6) were added to the palladium acetate solution, and the Pd(II) was reduced by sodium borohydride to form Pd NPs on the surface of SPMB. TEM images of Pd/ SPMB catalysts were presented in Figure 1 along with the particle size distribution, and the average sizes of Pd NPs were summarized in Table 1. As presented in Figure 1a, Pd/SPMB0.9 showed a maximum average size (9.1 nm) of Pd NPs, and a broad size distribution was found to be 4.2−35.0 nm. Upon the increase of the S/Pd molar ratio (Figure 1b, c, and d), the average size of Pd NPs decreased to 7.2, 5.3, and 2.7 nm (Table 1), while the size distribution narrowed to 2.2−20.2, 2.1−10.5, and 0.5−8.7 nm. Therefore, these results suggested that the average size of Pd NPs and the size distribution can be simply tuned by varying the S/Pd molar ratio. With the size-controlled Pd/SPMB catalysts in hands, we investigated the catalytic performance in the semihydrogenation of phenylacetylene (1a) to styrene (1b; Scheme 2). Screening of Pd/SPMB catalysts with different metal sizes toward a unique combination of high activity and good selectivity for 1b provided the results shown in Table 1 and Figure 2. Notably, Pd/SPMB-2.7 displayed the highest activity (>99% conversion) and excellent selectivity (96%) for 1b and hence resulted in the highest yield of 1b (Table 1, entry 3; Figure 2). Using Pd/SPMB catalysts with a larger size of Pd NPs (9.1 and 7.2 nm), the activity increased but the selectivity for 1b lowered to 86% (Figure 2). In contrast, when the size of Pd NPs decreased to 2.7 nm (Figure 2), lower activity and selectivity were obtained. These results indicated that the medium size of Pd NPs (5.3 nm) favored the enhancement of selectivity for alkene. In a control experiment, Pd/PMB afforded a low selectivity (84%) for 1b (Table 1, entry 5), demonstrating the crucial role of the thiocarbonylthio moiety in selectivity enhancement. During the solvent-screening (Table S1), reactions proceeded slowly in the tested solvents, such as 1,4-dioxane, CH2Cl2, cyclohexane, and n-hexane. Although the conversion in acetone increased to >99% within 50 min, the selectivity for 1b lowered to 84%. Therefore, we can conclude that the catalytic system involving Pd/SPMB-2.7 and acetonitrile was an efficient and selective protocol for 1b synthesis. The most effective catalyst Pd/SPMB-2.7 was selected for further characterization to better understand the behavior. The SEM image of Pd/SPMB-2.7 showed the porous surface morphology (Figure S5) was similar to that of SPMB. The

Figure 3. HRTEM images of Pd/SPMB-2.7 (a) and XPS spectra of Pd 3d (b) and S 2p (c) in Pd/SPMB-2.7.

morphology as presented in the SEM image (Figure S1). According to the nitrogen adsorption−desorption analysis (Figure S2), the SPMB support displayed a mesoporous structure (pore volume: 0.14 cm3 g−1) and a BET surface area of 65 m2 g−1. Elemental analysis revealed that SPMB contained 3.89 wt % of sulfur, indicating that numerous thiocarbonylthio moieties existed in the bulk SPMB. TGA results showed that SPMB had a good thermal stability (Figure S3). Thus, SPMB D

DOI: 10.1021/acs.iecr.8b04913 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research Table 2. Semihydrogenation of Various Alkynes Catalyzed by Pd/SPMB-2.7a

a Reaction conditions: substrate (1.0 mmol), Pd/SPMB-2.7 (0.5 mol % of Pd, the Pd loading in Pd/SPMB-2.7 was 3.37 wt %), CH3CN (4 mL), room temperature. bThe conversion and selectivity were determined by GC. cRelative ratio of Z-alkene and E-alkene.

CS (about 160.8 eV) was detected after deconvolution,54 indicating the reduction of trithiocarbonate to thiol.55 Instead, two main sulfur components at 162.4 and 163.7 eV were observed, which can be assigned to S2−47 and the C−S bond in thiolate,47,52 respectively. On the other hand, the formation of S2− might be ascribed to the cleavage of the C−S bond in the thiolate group, and the generated S2− then entered into the lattice of Pd NPs to form the Pdx−Sy compounds.47 Thus, we can deduce that the surface of Pd NPs could be ligated (or poisoned) by S2− and thiolate (Scheme 1), which played an important role in selectivity enhancement in the semihydrogenation of 1a. In order to gain more insight into the high selectivity for alkene, CO and H2 chemisorption experiments were carried out to investigate the surface accessibility of the supported Pd catalysts. It was found that Pd/SPMB-2.7 displayed a low chemisorption value (CO/Pd = 16.6 cm3 g−1; H/Pd = 0.1 cm3 g−1), whereas the sulfur-free Pd/PMB catalyst showed appreciable chemisorption amounts (CO/Pd = 378.3 cm3

formed Pd NPs that can be clearly observed in the TEM image (Figure 1c) were well dispersed on the surface of SPMB with an average size of 5.3 nm. The high-resolution TEM image (Figure 3a) exhibited a lattice spacing of 0.23 nm, which corresponded to the (111) plane distance of the pure Pd.51 Typical peaks for Pd NPs were not strong enough to be seen in the XRD results (Figure S6), which could be attributed to the low Pd loading of 3.37 wt %, as determined by ICP-OES.46 The BET surface area of Pd/SPMB-2.7 increased to 74 m2 g−1 (Figure S7) compared with that of SPMB (65 m2 g−1). The XPS spectra of Pd/SPMB-2.7 revealed the interaction between Pd NPs and SPMB support. In the Pd 3d spectrum of Pd/ SPMB-2.7 (Figure 3b), the Pd 3d signal was fitted to two pairs of doublets with a spin−orbit separation of 5.3 eV. One pair of peaks at the binding energies of 340.6 eV (Pd 3d3/2) and 335.3 eV (Pd 3d5/2) was assigned to the Pd metal, and the other pair of peaks at 342.6 eV (Pd 3d3/2) and 337.3 eV (Pd 3d5/2) has previously been assigned to Pdx−Sy compounds.52,53 In the S 2p spectrum of Pd/SPMB-2.7 (Figure 3c), no peak assigned to E

DOI: 10.1021/acs.iecr.8b04913 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research g−1, H/Pd = 1.0 cm3 g−1). The results suggested that a significant portion of active sites on the surface of Pd NPs was covered (or poisoned) with S2− and thiolate, thereby reducing the H−H dissociation ability of Pd/SPMB-2.7 and resulting in the low surface accessibility to H2. Although the Pd NPs were supported on the surface of gel-like SPMB, we believe that there were still some flexible chain segments on the surface of SPMB which may have covered the nearby Pd NPs to poison the metal surface. Moreover, the crystalline Pd was isolated by sulfur atoms, resulting in a high barrier for the segregation of Pd atoms, affecting the behavior of the hydride phase. Taking the comparative experiment (Table 1, entry 5) into consideration, one does believe that the copresence of S2− and thiolate on the Pd/SPMB-2.7 was responsible for the enhancement of selectivity for 1b in the semihydrogenation of 1a. As features of the supported catalyst, the recyclability and durability of Pd/SPMB-2.7 were also investigated in the model reaction. After the completion of each cycle, Pd/SPMB-2.7 was separated by centrifugation and washed with acetonitrile, and a new portion of acetonitrile and 1a was added for the next batch of reaction. As presented in Figure 4, Pd/SPMB-2.7 can be used in 10 runs without a loss of activity and selectivity for 1b. Interestingly, the reaction time was markedly reduced from 140 to 40 min during the cycle from the first to seventh, yet increased to 110 min in the 10th cycle. It has been, in fact, reported that Pdx−Sy compounds, such as Pd16S7, Pd4S, and Pd3S, were active and durable in the hydrogenation, whereas the PdS phase would be deactivated completely, following the order of Pd16S7 ≥ Pd4S > Pd3S ≫ PdS.56 Taking this into consideration, we propose that the increasing of activity during the cycle from first to seventh may be attributed to the reconstruction of Pdx−Sy to form more active phases,57,58 which can also be proved by the shift of Pdx−Sy in the XPS spectra of recycled catalysts (Figure S8). However, the activity of Pd/SPMB-2.7 may also be decreased by the aggregation of Pd NPs. As demonstrated by the TEM results, the average size of Pd NPs increased to 5.9 nm after the seventh cycle (Figure S9a) and further increased to 13 nm after the 10th cycle (Figure S9b). Since the Pd/SPMB-2.7 catalyst displayed the best activity for the production of 1b, its general applicability was investigated with various alkynes under optimized conditions (Table 2). The tested substrates including terminal and internal alkynes were converted smoothly and selectively to the corresponding alkenes in moderate to excellent yields. A broad range of functional groups, such as trifluoromethyl, fluoro, chloro, bromo, cyano, alkyl, ester, ether, thienyl, pyridyl, and tertiary amine, was all well-tolerated in the alkyne hydrogenation process. In detail, the Pd/SPMB-2.7 catalyst showed satisfactory catalytic performance in the semihydrogenation of terminal alkynes to terminal alkenes (Table 2, entries 1−8). 2Pyridylacetylene afforded lower selectivity for alkene products probably due to the coordination with Pd NPs (Table 2, entry 9), resulting in the more difficult desorption of alkene from the catalyst surface.59 The internal alkynes with either aromatic or alkyl groups all provided the corresponding alkene products in high yields (Table 2, entries 10−15). Notably, the Z/E ratios of obtained alkenes were higher than 96:4, revealing that isomerization of Z-alkene was hard to take place in such a reaction system. Additionally, the reaction of dipropylacetylene also readily proceeded in 99% conversion with a >99% Z/E ratio. The above results basically demonstrated that such a

protocol can be successfully extended to the semihydrogenation of typical alkynes.

4. CONCLUSIONS In summary, we have presented the use of RAFT polymer as the support and modifier to prepare size-controlled Pd NPs for the semihydrogenation of alkynes to their corresponding alkenes. The average size of Pd NPs can be simply tuned by varying the S/Pd mole ratio in the range of 2.7−9.1 nm. The Pd/SPMB-2.7 with a medium Pd particle size of 5.3 nm exhibited excellent activity, stability, and selectivity for alkenes in the semihydrogenation of both internal and terminal alkynes. Notably, the selectivity enhancement toward alkene was attributed to the coverage of S2− and thiolate on the Pd surface with the appropriate degree of poisoning. This strategy for the selectivity control using the RAFT polymer would provide useful insights for the development of selective nanocatalysts.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.iecr.8b04913. Characterization results of SPMB and Pd/SPMB-2.7 (SEM, N2 adsorption−desorption isotherm, XRD, and TGA); effect of solvents on the catalytic performance of Pd/SPMB-2.7; time course of the semihydrogenation of 1a; XPS spectra and TEM images of recycled catalyst after the 7th and 10th cycle; NMR spectra data of products; and 1H and 13C NMR spectra of products (PDF)



AUTHOR INFORMATION

Corresponding Authors

*Phone: +86-312-5079359. Fax: +86-312-5937102. E-mail: [email protected]. *Phone: +86-312-5079359. Fax: +86-312-5937102. E-mail: [email protected]. ORCID

Xin Wen: 0000-0001-9043-8724 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Financial support by the National Natural Science Foundation of China (21706049 and 21274037), the Natural Science Foundation of Hebei Province (B2016418005, B2015201193), and the Hebei University construction project for comprehensive strength promotion of Midwest colleges and universities is gratefully acknowledged.



REFERENCES

(1) Chinchilla, R.; Nájera, C. Chemicals from alkynes with palladium catalysts. Chem. Rev. 2014, 114, 1783−1826. (2) Nikoshvili, L. Z.; Bykov, A. V.; Khudyakova, T. E.; LaGrange, T.; Héroguel, F.; Luterbacher, J. S.; Matveeva, V. G.; Sulman, E. M.; Dyson, P. J.; Kiwi-Minsker, L. Promotion effect of alkali metal hydroxides on polymer-stabilized Pd nanoparticles for selective hydrogenation of C-C triple bonds in alkynols. Ind. Eng. Chem. Res. 2017, 56, 13219−13227. (3) García-Mota, M.; Bridier, B.; Pérez-Ramírez, J.; López, N. Interplay between carbon monoxide, hydrides, and carbides in F

DOI: 10.1021/acs.iecr.8b04913 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research selective alkyne hydrogenation on palladium. J. Catal. 2010, 273, 92− 102. (4) Ellis, I. T.; Wolf, E. H.; Jones, G.; Lo, B.; Meng-Jung Li, M; York, A. P.; Edman Tsang, S. C. Lithium and boron as interstitial palladium dopants for catalytic partial hydrogenation of acetylene. Chem. Commun. 2017, 53, 601−604. (5) Delgado, J. A.; Benkirane, O.; Claver, C.; Curulla-Ferre, D.; Godard, C. Advances in the preparation of highly selective nanocatalysts for the semi-hydrogenation of alkynes using colloidal approaches. Dalton Trans 2017, 46, 12381−12403. (6) Crespo-Quesada, M.; Cárdenas-Lizana, F.; Dessimoz, A. L.; Kiwi-Minsker, L. Modern trends in catalyst and process design for alkyne hydrogenations. ACS Catal. 2012, 2, 1773−1786. (7) Tejeda-Serrano, M.; Cabrero-Antonino, J. R.; Mainar-Ruiz, V.; López-Haro, M.; Hernández-Garrido, J. C.; Calvino, J. J.; Leyva-Pérez, A.; Corma, A. Synthesis of supported planar iron oxide nanoparticles and their chemo-and stereoselectivity for hydrogenation of alkynes. ACS Catal. 2017, 7, 3721−3729. (8) Hafner, J.; Krajčí, M. Surfaces of complex intermetallic compounds: insights from density functional calculations. Acc. Chem. Res. 2014, 47, 3378−3384. (9) Armbrüster, M.; Kovnir, K.; Behrens, M.; Teschner, D.; Grin, Y.; Schlögl, R. Pd-Ga intermetallic compounds as highly selective semihydrogenation catalysts. J. Am. Chem. Soc. 2010, 132, 14745− 14747. (10) Cherkasov, N.; Ibhadon, A. O.; Rebrov, E. V. Solvent-free semihydrogenation of acetylene alcohols in a capillary reactor coated with a Pd−Bi/TiO2 catalyst. Appl. Catal., A 2016, 515, 108−115. (11) Zhou, H.; Yang, X.; Li, L.; Liu, X.; Huang, Y.; Pan, X.; Wang, A.; Li, J.; Zhang, T. PdZn intermetallic nanostructure with Pd-Zn-Pd ensembles for highly active and chemoselective semi-hydrogenation of acetylene. ACS Catal. 2016, 6, 1054−1061. (12) Johnston, S. K.; Cherkasov, N.; Pérez-Barrado, E.; Aho, A.; Murzin, D. Y.; Ibhadon, A. O.; Francesconi, M. G. Pd3Sn nanoparticles on TiO2 and ZnO supports as catalysts for semihydrogenation: Synthesis and catalytic performance. Appl. Catal., A 2017, 544, 40−45. (13) Cao, Y.; Sui, Z.; Zhu, Y.; Zhou, X.; Chen, D. Selective hydrogenation of acetylene over Pd-In/Al2O3 catalyst: promotional effect of indium and composition-sependent performance. ACS Catal. 2017, 7, 7835−7846. (14) Niu, W.; Gao, Y.; Zhang, W.; Yan, N.; Lu, X. Pd-Pb alloy nanocrystals with tailored composition for semihydrogenation: taking advantage of catalyst poisoning. Angew. Chem. 2015, 127, 8389−8392. (15) Anderson, J. A.; Mellor, J.; Wells, R. P. Pd catalysed hexyne hydrogenation modified by Bi and by Pb. J. Catal. 2009, 261, 208− 216. (16) Mitsudome, T.; Urayama, T.; Yamazaki, K.; Maehara, Y.; Yamasaki, J.; Gohara, K.; Maeno, Z.; Mizugaki, T.; Jitsukawa, K.; Kaneda, K. Design of core-Pd/shell-Ag nanocomposite catalyst for selective semihydrogenation of alkynes. ACS Catal. 2016, 6, 666−670. (17) Song, S.; Li, K.; Pan, J.; Wang, F.; Li, J.; Feng, J.; Yao, S.; Ge, X.; Wang, X.; Zhang, H. Achieving the trade-off between selectivity and activity in semihydrogenation of alkynes by fabrication of (asymmetrical Pd@ Ag core)@(CeO2 shell) nanocatalysts via autoredox reaction. Adv. Mater. 2017, 29, 1605332. (18) Mitsudome, T.; Takahashi, Y.; Ichikawa, S.; Mizugaki, T.; Jitsukawa, K.; Kaneda, K. Metal-ligand core-shell nanocomposite catalysts for the selective semihydrogenation of alkynes. Angew. Chem., Int. Ed. 2013, 52, 1481−1485. (19) Chen, F.; Kreyenschulte, C.; Radnik, J.; Lund, H.; Surkus, A. E.; Junge, K.; Beller, M. Selective semihydrogenation of alkynes with Ngraphitic-modified cobalt nanoparticles supported on silica. ACS Catal. 2017, 7, 1526−1532. (20) Karakhanov, E.; Maximov, A.; Kardasheva, Y.; Semernina, V.; Zolotukhina, A.; Ivanov, A.; Abbott, G.; Rosenberg, E.; Vinokurov, V. Pd nanoparticles in dendrimers immobilized on silica-polyamine composites as catalysts for selective hydrogenation. ACS Appl. Mater. Interfaces 2014, 6, 8807−8816.

(21) Hauwert, P.; Boerleider, R.; Warsink, S.; Weigand, J. J.; Elsevier, C. J. Mechanism of Pd(NHC)-catalyzed transfer hydrogenation of alkynes. J. Am. Chem. Soc. 2010, 132, 16900−16910. (22) Carenco, S.; Leyva-Pérez, A.; Concepción, P.; Boissière, C.; Mézailles, N.; Sanchez, C.; Corma, A. Nickel phosphide nanocatalysts for the chemoselective hydrogenation of alkynes. Nano Today 2012, 7, 21−28. (23) Ren, M.; Li, C.; Chen, J.; Wei, M.; Shi, S. Preparation of a ternary Pd-Rh-P amorphous alloy and its catalytic performance in selective hydrogenation of alkynes. Catal. Sci. Technol. 2014, 4, 1920− 1924. (24) Zhao, M. Fabrication of ultrafine palladium phosphide nanoparticles as highly active catalyst for chemoselective hydrogenation of alkynes. Chem. - Asian J. 2016, 11, 461−464. (25) Liu, Y.; McCue, A. J.; Miao, C.; Feng, J.; Li, D.; Anderson, J. A. Palladium phosphide nanoparticles as highly selective catalysts for the selective hydrogenation of acetylene. J. Catal. 2018, 364, 406−414. (26) Chan, C. W. A.; Mahadi, A. H.; Li, M. M. J.; Corbos, E. C.; Tang, C.; Jones, G.; Kuo, W. C. H.; Cookson, J.; Brown, C. M.; Bishop, P. T.; Tsang, S. C. E. Interstitial modification of palladium nanoparticles with boron atoms as a green catalyst for selective hydrogenation. Nat. Commun. 2014, 5, 5787. (27) Karakhanov, E. A.; Maximov, A. L.; Zolotukhina, A. V.; Yatmanova, N.; Rosenberg, E. Alkyne hydrogenation using Pd-Ag hybrid nanocatalysts in surface-immobilized dendrimers. Appl. Organomet. Chem. 2015, 29, 777−784. (28) Sajiki, H.; Mori, S.; Ohkubo, T.; Ikawa, T.; Kume, A.; Maegawa, T.; Monguchi, Y. Partial hydrogenation of alkynes to cisolefins by using a novel Pd0-polyethyleneimine catalyst. Chem. - Eur. J. 2008, 14, 5109−5111. (29) Long, W.; Brunelli, N. A.; Didas, S. A.; Ping, E. W.; Jones, C. W. Aminopolymer-silica composite-supported Pd catalysts for selective hydrogenation of alkynes. ACS Catal. 2013, 3, 1700−1708. (30) Nishio, R.; Sugiura, M.; Kobayashi, S. Semi-hydrogenation of alkynes using phosphinated polymer incarcerated (PI) palladium catalysts. Org. Biomol. Chem. 2006, 4, 992−995. (31) Quintanilla, A.; Butselaar-Orthlieb, V. C. L.; Kwakernaak, C.; Sloof, W. G.; Kreutzer, M. T.; Kapteijn, F. Weakly bound capping agents on gold nanoparticles in catalysis: Surface poison? J. Catal. 2010, 271, 104−114. (32) da Silva, F. P.; Fiorio, J. L.; Rossi, L. M. Tuning the catalytic activity and selectivity of Pd nanoparticles using ligand-modified supports and surfaces. ACS Omega 2017, 2, 6014−6022. (33) Hill, M. R.; Carmean, R. N.; Sumerlin, B. S. Expanding the scope of RAFT polymerization: recent advances and new horizons. Macromolecules 2015, 48, 5459−5469. (34) Huang, J.; Xu, J.; Chen, K.; Wang, T.; Cui, C.; Wei, X.; Zhang, R.; Li, L.; Guo, X. Synthesis of triblock copolymers via RAFT polymerization and their application as surfactants for crude oil-inwater emulsion. Ind. Eng. Chem. Res. 2015, 54, 1564−1575. (35) Moad, G. RAFT polymerization to form stimuli-responsive polymers. Polym. Chem. 2017, 8, 177−219. (36) Kurek, P. N.; Kloster, A. J.; Weaver, K. A.; Manahan, R.; Allegrezza, M. L.; De Alwis Watuthanthrige, N.; Boyer, C.; Reeves, J. A.; Konkolewicz, D. How do reaction and reactor conditions affect photoinduced electron/energy transfer reversible addition-fragmentation transfer polymerization? Ind. Eng. Chem. Res. 2018, 57, 4203− 4213. (37) Wang, J.; Zhu, H.; Chen, G.; Hu, Z.; Weng, Y.; Wang, X.; Zhang, W. Controlled synthesis and self-assembly of dopaminecontaining copolymer for honeycomb-like porous hybrid particles. Macromol. Rapid Commun. 2014, 35, 1061−1067. (38) Boyer, C.; Stenzel, M. H.; Davis, T. P. Building nanostructures using RAFT polymerization. J. Polym. Sci., Part A: Polym. Chem. 2011, 49, 551−595. (39) Ma, M.; Zheng, S.; Chen, H.; Yao, M.; Zhang, K.; Jia, X.; Mou, J.; Xu, H.; Wu, R.; Shi, J. A combined “RAFT” and “Graft From” polymerization strategy for surface modification of mesoporous silica G

DOI: 10.1021/acs.iecr.8b04913 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research nanoparticles: towards enhanced tumor accumulation and cancer therapy efficacy. J. Mater. Chem. B 2014, 2, 5828−5836. (40) Zhang, Z.; Schepens, B.; Nuhn, L.; Saelens, X.; Schotsaert, M.; Callewaert, N.; De Rycke, R.; Zhang, Q.; Moins, S.; Benali, S.; Mespouille, L.; Hoogenboom, R.; De Geest, B. G. Influenza-binding sialylated polymer coated gold nanoparticles prepared via RAFT polymerization and reductive amination. Chem. Commun. 2016, 52, 3352−3355. (41) McCue, A. J.; Anderson, J. A. Sulfur as a catalyst promoter or selectivity modifier in heterogeneous catalysis. Catal. Sci. Technol. 2014, 4, 272−294. (42) Kahsar, K. R.; Schwartz, D. K.; Medlin, J. W. Control of metal catalyst selectivity through specific noncovalent molecular interactions. J. Am. Chem. Soc. 2014, 136, 520−526. (43) Pang, S. H.; Schoenbaum, C. A.; Schwartz, D. K.; Medlin, J. W. Directing reaction pathways by catalyst active-site selection using selfassembled monolayers. Nat. Commun. 2013, 4, 2448. (44) Schoenbaum, C. A.; Schwartz, D. K.; Medlin, J. W. Controlling the surface environment of heterogeneous catalysts using selfassembled monolayers. Acc. Chem. Res. 2014, 47, 1438−1445. (45) McCue, A. J.; Guerrero-Ruiz, A.; Rodríguez-Ramos, I.; Anderson, J. A. Palladium sulphide-A highly selective catalyst for the gas phase hydrogenation of alkynes to alkenes. J. Catal. 2016, 340, 10−16. (46) Yun, S.; Lee, S.; Yook, S.; Patel, H. A.; Yavuz, C. T.; Choi, M. Cross-linked “poisonous” polymer: thermochemically stable catalyst support for tuning chemoselectivity. ACS Catal. 2016, 6, 2435−2442. (47) Zhao, X.; Zhou, L.; Zhang, W.; Hu, C.; Dai, L.; Ren, L.; Wu, B.; Fu, G.; Zheng, N. Thiol treatment creates selective palladium catalysts for semihydrogenation of internal alkynes. Chem. 2018, 4, 1080− 1091. (48) Wen, X.; Shi, X.; Qiao, X.; Wu, Z.; Bai, G. Ligand-free nickelcatalyzed semihydrogenation of alkynes with sodium borohydride: a highly efficient and selective process for cis-alkenes under ambient conditions. Chem. Commun. 2017, 53, 5372−5375. (49) Gieshoff, T. N.; Welther, A.; Kessler, M. T.; Prechtl, M. H.; Jacobi von Wangelin, A. Stereoselective iron-catalyzed alkyne hydrogenation in ionic liquids. Chem. Commun. 2014, 50, 2261−2264. (50) Willcock, H.; O’Reilly, R. K. End group removal and modification of RAFT polymers. Polym. Chem. 2010, 1, 149−157. (51) Zhang, J.; Cheng, Y.; Lu, S.; Jia, L.; Shen, P. K.; Jiang, S. P. Significant promotion effect of carbon nanotubes on the electrocatalytic activity of supported Pd NPs for ethanol oxidation reaction of fuel cells: the role of inner tubes. Chem. Commun. 2014, 50, 13732−13734. (52) Mori, A.; Mizusaki, T.; Kawase, M.; Maegawa, T.; Monguchi, Y.; Takao, S.; Takagi, Y.; Sajiki, H. Novel palladium-on-carbon/ diphenyl sulfide complex for chemoselective hydrogenation: preparation, characterization, and application. Adv. Synth. Catal. 2008, 350, 406−410. (53) Zeng, Q.; Li, Y.; Wu, K. H.; Huang, N.; Dalapati, S.; Su, B. J.; Jang, L. Y.; Gentle, I. R.; Jiang, D.; Wang, D. W. Long-chain solid organic polysulfide cathode for high-capacity secondary lithium batteries. Energy Storage Mater. 2018, 12, 30−36. (54) Markevich, E.; Salitra, G.; Rosenman, A.; Talyosef, Y.; Chesneau, F.; Aurbach, D. The effect of a solid electrolyte interphase on the mechanism of operation of lithium-sulfur batteries. J. Mater. Chem. A 2015, 3, 19873−19883. (55) Yang, L.; Xu, J.; Han, J.; Shen, Y.; Luo, Y. A novel method for preparing click-ready latex and latex with stability against high electrolyte concentrations. Ind. Eng. Chem. Res. 2015, 54, 5536−5542. (56) Zhang, Q.; Xu, W.; Li, X.; Jiang, D.; Xiang, Y.; Wang, J.; Cen, J.; Romano, S.; Ni, J. Catalytic hydrogenation of sulfur-containing nitrobenzene over Pd/C catalysts: In situ sulfidation of Pd/C for the preparation of PdxSy catalysts. Appl. Catal., A 2015, 497, 17−21. (57) McCue, A. J.; Guerrero-Ruiz, A.; Ramirez-Barria, C.; Rodríguez-Ramos, I.; Anderson, J. A. Selective hydrogenation of mixed alkyne/alkene streams at elevated pressure over a palladium sulfide catalyst. J. Catal. 2017, 355, 40−52.

(58) Albani, D.; Shahrokhi, M.; Chen, Z.; Mitchell, S.; Hauert, R.; López, N.; Pérez-Ramírez, J. Selective ensembles in supported palladium sulfide nanoparticles for alkyne semi-hydrogenation. Nat. Commun. 2018, 9, 2634. (59) Fu, S.; Chen, N. Y.; Liu, X.; Shao, Z.; Luo, S. P.; Liu, Q. Ligandcontrolled cobalt-catalyzed transfer hydrogenation of alkynes: Stereodivergent synthesis of Z- and E-alkenes. J. Am. Chem. Soc. 2016, 138, 8588−8594.

H

DOI: 10.1021/acs.iecr.8b04913 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX