Next Article in Journal
Modification of Alkali Lignin with Poly(Ethylene Glycol) Diglycidyl Ether to Be Used as a Thickener in Bio-Lubricant Formulations
Next Article in Special Issue
PDDA-Montmorillonite Composites Loaded with Ru Nanoparticles: Synthesis, Characterization, and Catalytic Properties in Hydrogenation of 2-Butanone
Previous Article in Journal
Visco-Hyperelastic Model with Damage for Simulating Cyclic Thermoplastic Elastomers Behavior Applied to an Industrial Component
Previous Article in Special Issue
Thermal Degradation Kinetics and Viscoelastic Behavior of Poly(Methyl Methacrylate)/Organomodified Montmorillonite Nanocomposites Prepared via In Situ Bulk Radical Polymerization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Heterogeneous Catalytic Composites from Palladium Nanoparticles in Montmorillonite Intercalated with Poly (Vinyl Pyrrolidone) Chains

Zhejiang Key Laboratory of Alternative Technologies for Fine Chemicals Process, College of Chemistry and Chemical Engineering, Shaoxing University, Shaoxing 312000, China
*
Author to whom correspondence should be addressed.
Polymers 2018, 10(6), 669; https://doi.org/10.3390/polym10060669
Submission received: 26 May 2018 / Revised: 11 June 2018 / Accepted: 13 June 2018 / Published: 15 June 2018
(This article belongs to the Special Issue Polymer-Clay (Nano)Composites)

Abstract

:
In this study, poly (vinyl pyrrolidone) (PVP) chains intercalated montmorillonite (MMT) matrices has been demonstrated as an excellent scaffolding material for the immobilization of palladium (Pd) nanoparticles to prepare efficient heterogeneous catalysts for Heck reactions. Multiple layers (up to four) of PVP chains can intercalate the interlayer space of the MMT, resulting in an increase therein from 1.25 to 3.22 nm. MMT/PVP with PVP loading (20%) was selected as the platform for the immobilization of Pd. The in-situ reduction of the chelated Pd2+ into Pd0 in the interlayer space of MMT/PVP composite could be easily achieved. For the prepared Pd@MMT/PVP catalytic composite, a unique maze-like microstructure of Pd nanoparticles tightly encaged by PVP chains and by lamellae of layered silica has been detected by high resolution transmission electron microscopy (HR-TEM) and X-ray diffraction (XRD). Furthermore, the microstructure is well elucidated in molecular level by positron annihilation lifetime analysis of the Pd@MMT/PVP catalytic composite. The prepared Pd@MMT/PVP catalysts were highly active for the Heck coupling reactions between aromatic halides and alkenes, and could be recycled 9 times without significant decreases in coupling yields. The excellent comprehensive catalytic performances of the Pd@MMT/PVP catalytic composites are mainly attributed to their unique maze-like microstructure.

Graphical Abstract

1. Introduction

Palladium (Pd) catalyzed Heck cross-coupling reactions are among of the most versatile methods for the construction of C–C bonds linked to unsaturated double bonds in organic synthesis [1,2,3]. Nowadays, the Heck reaction is widely used for the synthesis of many natural products, biologically active compounds, fine chemicals, and their intermediates with alkene structure units [4,5,6,7,8,9]. Great attention has been given to the development of novel and advantageous catalytic systems for Pd-catalyzed Heck reactions [10,11,12]. Among them, homogeneous Pd-based catalysts are preferred due to their excellent catalytic activity and selectivity. However, the homogeneous catalysts for Heck reactions often have present the problems of difficult separation, recovery, easy deactivation, and high costs, which greatly limit their wide applications in industry. Such drawbacks can be effectively reduced in heterogeneous catalysis processes. On the one hand, the immobilization of Pd nanoparticles on a support can significantly reduce the aggregation of the nanoparticles to prevent the loss of the activity. On the other hand, the heterogeneously supported Pd catalysts can easily be separated from the reaction system and recycled to the next reaction. Therefore, preparations of heterogeneously supported catalysts by immobilization of Pd species on various supports with excellent recovery performance have become the focus of the catalytic system for Heck reactions [13,14,15].
Many solid supporting materials, such as carbon materials [16,17], silica [18,19], TiO2 [20,21], Fe3O4 [22,23], clay [24,25], polymers [26,27,28], and so on, have been reported as palladium supports applied in Heck reactions. Among them, a type of natural layered silicate clay of montmorillonite (MMT) which contains negatively charged layers and interlayers with metal cations, like Na+, Ca2+, etc., are often used as transition metal catalyst supports. The transition metal cations can be easily introduced into the interlayers of MMT by an ion-exchange process for high cation exchange capacity. In addition, MMT has remarkable stability against mechanical agitations, good swelling capacities, strong adsorption properties, and high thermal stabilities. Kaneda et al. [29] synthesized sub-nano ordered Pd clusters within the interlayer space of MMT, and the prepared Pd-MMT heterogeneous catalyst showed high catalytic activity for allylic substitution reactions. Dutta et al. [30] also prepared a Pd-MMT heterogeneous catalyst by the introduction of Pd0-nanoparticles into the nanopores of sulfuric acid modified MMT, which showed excellent activity when applied in a Heck reaction and a Sonogashira reaction. However, as an inorganic support, although there are a number of hydroxyl groups located on the surface of the silicate layer, the chelation capability of MMT to transition metals which encaged in the interlayer space is generally not strong enough, and the recyclable times of such a Pd-MMT heterogeneous catalyst is limited (3 times [30]). This indicates that MMT needs further modification to enlarge the interlayer space and improve its interactions with the encaged transition metal species. Cationic surfactants, such as cetyl trimthyl ammonium bromide (CTAB), are often used as organic modifiers for MMT. For example, Zhou et al. [31] reported that Ru nanoparticles supported on CTAB modified MMT supports showed high activity and selectivity in the hydrogenation of quinolone reactions. Recently, we have succeeded in preparing a high stable Pd heterogeneous catalyst for coupling reactions by encaging Pd0 nanoparticles within a cationic polymer of chitosan (CS) modified MMT supports [32]. On the one hand, the intercalation of CS is beneficial for expanding the interlayer space of MMT from 1.25 nm (MMT) to 1.97 nm (Pd@MMT/CS), which is advantageous for the reaction substrates to easily access the entrapped Pd species. On the other hand, CS chains have strong chelation capabilities with Pd species, which is good for the tight encaging of the Pd species within the interlayer space of MMT. It revealed that the construction compatible inorganic MMT/polar polymer composite supports to immobilize metal catalysts is a good solution for achieving satisfactory comprehensive properties.
Besides CS, other nonionic polymers, such as poly (vinyl pyrrolidone) (PVP) can also effectively intercalate into the MMT interlayer using a similar solution intercalation method [33]. PVP is one of the most used polymer stabilizers for transition metal nanoparticles to prevent aggregation into big-sized particles [34,35,36]. The amide groups within the macromolecular backbone of PVP have good chelation capabilities with transition metals. Furthermore, PVP itself is demonstrated as an excellent polymer support for transition metals in preparing heterogeneous catalysts [37,38]. For example, Bai et al. reported [37] a novel catalyst containing Pd nanoparticles supported on PVP nanofiber films, showing fairly good efficiency in hydrogenation reactions of nitro compounds and Heck reactions. However, as a polymer support like CS, PVP support has limited stability against mechanical agitations, heat, and solvents, etc. The Pd@PVP heterogeneous catalyst prepared by Bai et al. can be only recycled 3 times. Drawing from the above works, we believe that heterogeneous catalysts with excellent comprehensive properties could be also prepared by the immobilization of Pd nanoparticles in the interlayer space of a PVP intercalated MMT composite. Previously, Dekany et al. [39,40,41] carried out some preliminary examinations of the influences of the preparation process conditions on the size of metal particles formed on the layer surface and interlayer space of MMT intercalated with PVP. The results showed that the average size of transition metal nanoparticles can be kept within 1–6 nm. However, application of the novel Pd nanoparticles supported on PVP intercalated MMT composites in catalysis is still lacking, which motivated us to systemically explore the microstructure and catalytic performance of this novel, heterogeneous catalytic composite.
In this study, the preparation process of the novel Pd0@MMT/PVP catalytic composites can be illustrated in Scheme 1, which includes three steps: (i) intercalation of PVP into Na+-MMT by interactions with hydroxyl groups on the surface of MMT and exchanges with the solvents molecules previously diffused in MMT, (ii) immobilization of Pd2+ on the MMT/PVP composite by cations exchanges of Pd2+ with Na+ and chelation with PVP chains, (iii) in situ reduction of Pd2+ into Pd0 encaged in the interlayer space of PVP intercalated MMT. The unique microstructure of the prepared Pd0@MMT/PVP catalysts has been characterized by several methods, including X-ray diffraction (XRD), X-ray photo electron spectroscopy (XPS), high resolution transmission electron microscopy (HR-TEM), thermal gravity analysis (TGA), and inductively coupled plasma-atomic emission spectroscopy (ICP). Moreover, positron annihilation lifetime spectroscopy (PALS) is further used to elucidate the sub-nano level microstructure of Pd@MMT/PVP catalytic composites. The catalytic performances of the Pd@MMT/PVP catalytic composite catalyzed Heck reaction in DMSO/ethylene glycol mixed solution were investigated.

2. Materials and Methods

2.1. Materials

G-105 type of Na+−MMT was obtained from Nanocor Co., Arlington Hts, IL, USA (cationic exchange capacity: 145 mmol/100 g). K-30 type of PVP was obtained from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China (Mn = 4 × 104). PdCl2 was obtained from Zhejiang Metallurgical Research Institute Co., Ltd., Zhejiang, China. The Heck reaction substrates of aromatic halides, acrylates and alkenes were obtained from Energy Chemical, Sun Chemical Technology (Shanghai) Co., Ltd., Shanghai, China. Other chemical reagents and solvents in analytical purity grade were obtained from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China.

2.2. Preparation of MMT/PVP Supports and Pd@MMT/PVP Catalytic Composite

A homogeneous suspension solution was prepared by dispersing 2 g of MMT into 100 mL of deionized water under magnetic stirring. A specific amount of PVP was dissolved in 100 mL acetic acid solution (2 wt %). The PVP solution was then added into the MMT suspension solution and stirred magnetically at 55 °C for 12 h. The mass ratios of MMT/PVP were 100/0, 90/10, 80/20, 70/30, 60/40, 50/50. Then, 2 mL of Pd2+ solution (0.3 g of PdCl2 and 2 g of NaCl co-dissolved in 100 mL deionized water) was added drop-wise into the MMT/PVP (80/20) mixture and stirred for another 6 h. The Pd2+@MMT/PVP composite was centrifugally separated and washed repeatedly with deionized water to neutralize (pH = 7). The resultant Pd2+@MMT/PVP composite was dried at 60 °C. Before catalysis use, Pd2+@MMT/PVP composite was reduced to Pd0@MMT/PVP catalytic composite by ethylene glycol at 80 °C for 30 min. The effect of reduction with alcohols is similar to that with hydrogen gas [42]. ICP-AES determination of Pd content in the Pd@MMT/PVP catalyst showed the content of the metal was about 0.2 wt %.

2.3. Characterizations of the MMT/PVP Supports and Pd@MMT/PVP Catalytic Composite

The crystalline structure of the MMT/PVP supports and Pd@MMT/PVP catalytic composite were measured with an Empyrean X-ray diffraction system (PANalytical, Eindhoven, the Netherlands). The recorded 2θ range was 2–70° and the scanning rate was 2°/min. The thermal stabilities of MMT/PVP supports and Pd@MMT/PVP catalytic composites were measured with a Mettler Toledo TGA/DSC 2 STAR system. The temperature range was 30–700 °C and the heating rate was 20 °C/min at air atmosphere. The XPS analysis of the binding energy of Pd elements within Pd@MMT/PVP catalyst was measured with a Thermos Scientific ESCALAB 250 Xi X-ray photoelectron spectrometer. The morphology of Pd@MMT/PVP catalytic composite was measured with a JEM-2100F HR-TEM (JEOL Ltd., Tokyo, Japan). The contents of Pd in the Pd@MMT/PVP catalytic composite were determined by Leemann ICP-AES Prodigy XP inductively coupled plasma atomic emission spectrometry. The PALS analysis of MMT/PVP supports and Pd@MMT/PVP catalysts was performed with an EG & G ORTEC fast-slow system. The time resolution of the PALS analysis was about 198 ps. The samples of MMT/PVP supports and Pd@MMT/PVP catalysts were pressured into thin disks firstly (diameter × thickness: 1 cm × 2 mm). The positron source (22 NaCl, 16 μCi) was deposited between two Kapton foils. This was then sandwiched between two pre-pressured samples disks. To obtain good statistics, the total counts of the lifetime spectra of each sample are at least 2 × 106. The positron annihilation spectra were analyzed by Lifetime9.0 and MELT-4 (Maximum Entropy for Lifetime Analysis-4) program.

2.4. Catalytic Performance of the Pd@MMT/PVP Catalytic Composite

In a 50 mL round bottom flask, a mixture of aromatic halide (1 mmol), acrylates or alkenes (2 mmol), Pd0@MMT/PVP catalyst (3 μmol of Pd), CH3COOK (3 mmol), and solvent (5 mL DMSO + 0.2 mL ethylene glycol) was stirred magnetically at 110 °C (heated in oil bath) for 5 h. The reaction progress was detected with the layer chromatography (TLC) method, and a gas chromatography-mass spectrometry (GC/MS) analysis was performed. The Heck coupling yield was obtained from a GC/MS quantitative analysis of the reaction mixture. The chemical structure of all products was confirmed by H1 NMR and GC/MS analysis, which is consistent with our recent work [32]. In the recycling experiment, the Pd@MMT/PVP catalytic composite was filtrated out from the reaction system after each reaction run, repeatedly washed with ethanol 3–5 times, and then reused for the next reaction run.

3. Results and Discussion

Figure 1 compares the X-ray diffraction patterns from pure MMT, composites of MMT/PVP with different PVP loading, and Pd@MMT/PVP catalytic composites with reduction treatment. The characteristic reflection peak related to the basal spacing of d001 of pure MMT is found to be 2θ of 7.03° (Figure 1), indicating the typical layer-to-layer distance of pure MMT is 1.25 nm from Bragg equation [33]. This reflection peak shifts to lower angles as the PVP polymer loading increases, i.e., 5.96° (10% loading), 4.22° (20% loading), 3.51° (30% loading), 2.85° (40% loading), 2.74° (50% loading), respectively. The corresponding layer-to-layer distances were calculated to be 1.48 nm (10% loading), 2.09 nm (20% loading), 2.51 nm (30% loading), 3.10 nm (40% loading), 3.22 nm (50% loading). This reveals that the intercalation of PVP polymers in the interlayer space of MMT leads to an efficient expansion of the interlayer distance.
As a kind of highly polar polymer, PVP macromolecules can form strong interactions with –OH groups on the surface of MMT layers, which is the main driving force for PVP macromolecules intercalation into the interlayer space of MMT in solution. During the intercalation process, the PVP chains preferentially take extended conformation in order to maximize their interactions with the MMT layers. Taking into account the thickness of the silica layer (about 0.96 nm) [33,43] and the extended PVP chain (about 0.52 nm) [44,45], the increased interlayer spacing of MMT/PVP (90/10) (d001 of 1.48 nm, 0.52 nm larger than MMT) is attributed to the formation of intercalation of mono-layer of PVP chain into the interlayer space of MMT. Increasing PVP loading results in the formation of intercalation of double layers of PVP chains into MMT, as shown in the diffraction result on MMT/PVP composite with 20 wt % PVP loading (d001 of 2.09 nm, 1.03 nm larger than MMT). It was found that the number of the intercalated PVP chains layers could be further increased with an increase of PVP loading. In the case of MMT/PVP (70/30), the d001 spacing is 2.51 nm (1.55 nm larger than MMT), indicating three layers of PVP chains have intercalated into the MMT. Similarly, four layers of PVP chains have intercalated into the MMT in the case of the MMT/PVP composite with 40 wt % PVP loading (d001 of 3.10 nm, 2.14 nm larger than MMT). This is difficult in the intercalation of more than four layers of PVP chains in MMT, as the interlayer spacing increased limitedly when PVP loading is 50% (d001 of 3.22 nm, 2.26 nm larger than MMT).
Too much PVP content in the MMT/PVP composite could be disadvantageous for high physical properties (such as thermal stability, mechanical properties, and resistance to solvents etc.) because of the lower stability of PVP than MMT. In addition, too much PVP content might also lead to an exfoliation structure of MMT layers; the stabilization effects of both MMT and PVP on the Pd nanoparticles would then decrease significantly. Therefore, we chose MMT/PVP with a moderate PVP loading (20%) as the platform for immobilization of Pd. As shown in Figure 1, after chelation of Pd2+ with the amide groups of PVP that intercalated in MMT, the interlayer spacing increases from 2.09 nm (MMT/PVP (80/20)) to 2.40 nm (Pd2+@MMT/PVP (80/20)). The immobilized Pd2+ can be reduced by ethylene glycol. According to the XPS analysis results (Figure 2), it is evidenced that both Pd2+ (characteristic Pd 3d5/2 electron binding energy at 337.6 eV and 336.8 eV) and Pd0 (characteristic Pd 3d5/2 electron binding energy at 335.5 eV) species co-exist in the Pd0@MMT/PVP (80/20) composite [46]. The interlayer spacing of Pd0@MMT/PVP (80/20) composite further increases to 2.48 nm, which should be caused by local distortions of the MMT layered structures by the in-situ generation of Pd0 nanoparticles.
Figure 3 shows the morphology of pure MMT, MMT/PVP, and Pd0@MMT/PVP catalytic composite, as observed by HR-TEM. MMT shows a two-dimensional porous layer structure, but the layer spacing is small and the interlayer contract is not clear, as visualized by HR-TEM. After the intercalation of PVP chains, the interlayer spacing of MMT increases significantly. The interlayer contract then becomes much clearer than that of MMT. After immobilization of Pd2+ and reduction with ethylene glycol, many Pd0 nanoparticles can be visualized in the MMT/PVP matrices. The dispersed Pd0 nanoparticles are mainly in three forms: (i) Pd0 nanoparticles sized 3–5 nm clipped in 2–3 layers of MMT, (ii) Pd0 nanoparticles sized 1–2 nm dispersed in the interlayer space of MMT, which show up as dark bands between neighboring galleries of the image, (iii) Pd0 nanoparticles sized 3–5 nm dispersed on the surface of MMT layer. The microstructure in the interlayer space of the Pd0@MMT/PVP catalytic composite is much like a “maze”, in which the intercalated PVP chains and encaged Pd species acts as the “hedges”. Such a maze-like microstructure might be responsible for good catalytic performance. On the one hand, the much expanded interlayer space of MMT by multi layers of PVP intercalation makes the diffusion of reactant substrates into the “maze” much easier. On the other hand, the encaged Pd species is difficult to leach out to the reaction solution from the “maze”.
To obtain a good understanding of the microstructure of MMT/PVP matrices and Pd@MMT/PVP catalytic composites in molecular level, PALS was used as a testing method to provide information on the micro defects of the composite [47,48,49]. In molecular solids, a part of the implanted positrons (e+) will combine with electrons (e) of the surrounding molecules to form positroniums. According to the different electron and positron spin states, positroniums are classified into two types: para-positronium (p-Ps, singlet spin state), and ortho-positronium (o-Ps, triplet spin). Generally, three positron annihilation processes are possible, i.e., p-Ps annihilation (shortest lifetime component, τ1), free positron annihilation (intermediate lifetime component, τ2), and o-Ps annihilation (longest lifetime component, τ3). The mean diameter of micro-defects within composite D can be calculated from the lifetime τ3 through Equations (1) and (2), according to the Tao-Eldrup free volume model [50,51]. Where ΔR = 0.1656 nm is the fitted empirical electron layer thickness.
1 τ 3 = 2 { 1 R R + Δ R + 1 2 π sin ( 2 π R R + Δ R ) }
D = 2R
As shown in Table 1, the PALS spectra was fitted well in three-component lifetimes, p-Ps annihilation (τ1, and its intensity, I1), free positron annihilation (τ2, and its intensity, I2), o-Ps annihilation in the micro-defects of MMT/PVP and/or Pd@MMT/PVP composites (τ3, and its intensity, I3). For pure MMT, the o-Ps annihilation lifetime is 2.806 ns, attributing to the o-Ps annihilation in the interlayer space of MMT. The D value can be calculated to be 0.6988 nm from Equations (1) and (2). After intercalation with mono layer of PVP chains, o-Ps annihilation mainly occurs in the micro defects between the intercalated PVP mono layer and MMT layer, and the D value of MMT/PVP (90/10) is 0.6072 nm. This reveals that the size of the micro defects obtained by PALS decreases though the interlayer space between MMT interlayers has been expanded by the intercalated mono layer of PVP chain. After intercalation of 2–4 layers of PVP chains, o-Ps annihilation will occur in the micro defects not only in the space between the intercalated PVP chains and MMT layers, but also in the space between intercalated PVP chains themselves. Therefore, a similar trend of a slight decrease in D value is observed as PVP loading increases. The changes in the microstructure after immobilization of Pd species are sensitively detected by PALS. After immobilization of Pd2+, the D value increases from 0.6062 nm (MMT/PVP (80/20)) to 0.6158 nm (Pd2+@MMT/PVP (80/20)). For the chelation of Pd2+ cations with amide groups of PVP chains, they are preferred to be entrapped between the PVP chain layers. Over all, the interlayer spacing is further expanded (also detected in XRD measurement), which leads to an increase of the size of all the micro defects. The effects of the reduction of Pd2+ to Pd0 on the changes of the microstructure are sensitively detected by PALS, too. As shown in Table 1, the D value increases from 0.6158 nm (Pd2+@MMT/PVP (80/20)) to 0.6208 nm (Pd0@MMT/PVP (80/20)). Usually, the Pd0 species has poorer chelation capability with PVP chains than Pd2+ species. Therefore, the compactness of the PVP chains will undergo a decrease after the reduction of Pd2+ to Pd0, leading to an increase in the size of micro defects. In addition, it is likely that the interlayer spacing enlargement caused by the formation of large-sized Pd0 nanoparticles with distorted microstructures is another reason for this increase in the size of micro defects.
The PALS spectra of the Pd@MMT/PVP composites are further analyzed by MELT-4 program to obtain the distribution of the lifetimes. As shown in Figure 4A, there are three lifetime peaks for each sample and the longest lifetime peak is attributed to the o-Ps annihilation of the composites. For MMT/PVP (80/20), the o-Ps lifetime ranges from 1.55 ns to 1.98 ns. After immobilization of Pd2+, the o-Ps lifetime shifts to longer lifetime range (1.55–2.05 ns). Reduction of Pd2+ to Pd0 leads a further longer shift of the o-Ps lifetime (1.65–2.15 ns). The size distribution of the micro defects as calculated from Equation1 and Equation2 of MMT/PVP (0.4786–0.5644 nm), Pd2+@MMT/PVP (0.4780–0.5770 nm), and Pd0@MMT/PVP (0.5148–0.6026 nm) is shown in Figure 4B. Clearly, the variances trend of size distribution by MELT analysis corresponds with that of mean size by LT analysis. In summary, the microenvironment results based on PALS analysis provides strong evidence that both the extensive PVP chain layers and most of the Pd species are well incorporated into the expanded interlayer space of MMT.
The thermal stabilities of pure MMT, PVP, MMT/PVP (80/20), and Pd@MMT/CS (80/20) catalyst were measured with TGA (Figure 5). MMT shows excellent thermal stability, except during a weight loss stage before 125 °C, which was attributed to the evaporation of absorbed and/or bonded H2O. The first weight loss stage of PVP is observed from 330 to 410 °C. Then, PVP show serious decomposition of PVP chains at higher temperature than 410 °C. The thermal degradation curves of MMT/PVP, Pd2+@MMT/PVP, and Pd0@MMT/PVP are almost overlapped, indicating similar thermal stabilities with each other. In these three cases, the starting decomposition temperature of PVP component in the composites is improved to 430 °C, exhibiting much higher thermal stability than pure PVP after intercalation into the MMT layers. The high thermal stability of the Pd@MMT/PVP catalytic composite should be advantageous for good recyclability.
The catalytic activities of the Pd0@MMT/PVP were explored in the Heck coupling reactions of aromatic halides with acrylates or alkenes in DMSO solution. As shown in Table 2, the novel Pd0@MMT/PVP catalyst shows high catalytic activity for the reaction iodo benzene with n-butyl acrylate (entry 1) with a yield of 98%. It also shows high catalytic efficiencies for the substituted aryl iodides having an electron-donating group like p-CH3 (entry 2, 93% yield), o-CH3 (entry 3, 89% yield), m-OCH3 (entry 4, 87% yield), or having an electron-withdrawing group such as p-F (entry 5, 90% yield), m-F (entry 6, 89% yield), and o-Cl (entry 7, 87% yield). The coupling reactions of substituted iodo benzene with styrene can be also well catalyzed by the Pd0@MMT/PVP catalyst (entry 8–10). These catalysis application results indicate that all the reactant substrates can easily diffuse into the interlayer space of MMT to get access to the entrapped Pd species. For a higher bond strength of C-Br than C-I bonding, the Pd0@MMT/PVP catalyst has low catalytic activity for the coupling reaction of bromo benzene with n-butyl acrylate and/or styrene (entries 11, 12, trace yield). Nevertheless, C-Br bond can be effectively activated with an electron-withdrawing group such as the m-COCH3 group (entries 13, 14, moderate yields close to 50%). Such Pd0@MMT/PVP catalysts have similar high catalytic activities with most of other recently reported Pd heterogeneous catalysts, such as Pd@PVP nanofiber membrane catalyst [37], Pd@MMT catalyst [30,31], and Pd@MMT/CS catalyst [32], etc.
As shown in Figure 6, the stability and reusability of Pd0@MMT/PVP catalyst were assessed with a model reaction of iodo benzene coupling with n-butyl acrylate. During the recycling experiments, the Pd0@MMT/PVP catalyst could easily be filtrated out from the reaction system for use in the next reaction run. An ICP analysis of the reaction solution after filtration confirmed that the Pd content was below the detection limit, indicating that Pd nanoparticles are tightly entrapped within the interlayer space of MMT/PVP supports. Moreover, 67% yield for the recycled Pd0MMT/PVP catalyst catalyzed Heck reaction after 9 cycles was observed. Clearly, the recyclable times of Pd0@MMT/PVP catalyst are fewer than that of recent reported Pd0@MMT/CS catalyst (30 times) [32], which is mainly due to extremely high chelation capacities of CS with Pd species. Nevertheless, with similar Pd loading amounts, the Pd0@MMT/PVP catalyst has much higher recyclability than many of other Pd heterogeneous catalysts, such as Pd@PVP nanofiber films (3 times) [37], and/or Pd@MMT catalysts (3 times) [30,31], etc. This indicates that the unique maze-like microstructure of the Pd0@MMT/PVP catalyst is advantageous for the slow leaching of the Pd species.

4. Conclusions

In this study, it was demonstrated that the intercalation of multi layers of PVP chains into Na+-MMT provides novel MMT/PVP matrices for immobilization of Pd species. The prepared Pd0@MMT/PVP catalysts show fairly good catalytic activities for Heck reactions, and can be recycled 9 times with high yields. The excellent comprehensive catalytic performances of the Pd@MMT/PVP catalysts are mainly attributed to their unique maze-like microstructure, which is well elucidated by positron annihilation lifetime spectrum and other methods. On the one hand, the reactant substrates can easily diffuse into the enlarged interlayer space of MMT to get access to the tightly entrapped Pd species. On the other, a perfect combination of the excellent stabilizing of PVP for Pd nanoparticles and remarkable stability of MMT was achieved in the well-designed Pd0@MMT/PVP catalytic composite. Such a novel catalyst overcomes the shortages of Pd leaching and difficult Pd recovery in conventional Pd homogeneous catalysis, and shows a broad foreground in both experimental and industrial applications.

Author Contributions

M.Z. conceived and designed the experiments; M.X., and J.Z. performed the materials preparation and characterization experiments; G.S. and Q.L., performed the organic synthesis experiments.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 11475114) and the Zhejiang Provincial Fundamental Public Welfare Research Project (Grant No. LGG18E030004).

Acknowledgments

We are grateful to Professor Yao Lin (Institute of Materials Science and Department of Chemistry, University of Connecticut) for his valuable suggestions and discussions. And we are also grateful to Professor XingZhong Cao and Baoyi Wang (Institute of High Energy Physics, The Chinese Academy of Science, China) for the positron annihilation characterization supports.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Heck, R.F.; Nolley, J.P. Palladium-catalyzed vinylic hydrogen substitution reactions with aryl, benzyl, and styryl halides. J. Org. Chem. 1972, 37, 2320–2322. [Google Scholar] [CrossRef]
  2. Mizoroki, T.; Mori, K.; Ozaki, A. Arylation of olefin with aryl iodide catalyzed by palladium. Bull. Chem. Soc. Jpn. 1971, 44, 581. [Google Scholar] [CrossRef]
  3. Yin, L.; Liebscher, L.J. Carbon-carbon coupling reactions catalyzed by heterogeneous palladium catalysts. Chem. Rev. 2007, 107, 133–173. [Google Scholar] [CrossRef] [PubMed]
  4. Ohno, H.; Inuki, S. Recent progress in palladium-catalyzed cascade cyclizations for natural product synthesis. Synth. Stuttg. 2018, 50, 700–710. [Google Scholar] [CrossRef]
  5. Heravi, M.M.; Moradi, R.; Malmir, M. Recent advances in the application of the heck reaction in the synthesis of heterocyclic compounds: An update. Curr. Org. Chem. 2018, 22, 165–198. [Google Scholar] [CrossRef]
  6. Budarin, V.L.; Shuttleworth, P.S.; Clark, J.H.; Luque, R. Industrial applications of C-C coupling reactions. Curr. Org. Chem. 2010, 7, 614–627. [Google Scholar] [CrossRef]
  7. Qureshi, Z.; Weinstabl, H.; Suhartono, M.; Liu, H.Q.; Thesmar, P.; Lautens, M. Application of the palladium-catalysed norbornene-assisted catellani reaction towards the total synthesis of (+)-linoxepin and isolinoxepin. Eur. J. Org. Chem. 2014, 19, 4053–4069. [Google Scholar] [CrossRef]
  8. Vulovic, B.; Watson, D.A. Heck-like reactions involving heteroatomic electrophiles. Eur. J. Org. Chem. 2017, 34, 4996–5009. [Google Scholar] [CrossRef] [PubMed]
  9. Torborg, C.; Beller, M. Recent Applications of palladium-catalyzed coupling reactions in the pharmaceutical, agrochemical, and fine chemical industries. Adv. Synth. Catal. 2009, 351, 3027–3043. [Google Scholar] [CrossRef]
  10. Biffis, A.; Centomo, P.; Del Zotto, A.; Zeccal, M. Pd metal catalysts for cross-couplings and related reactions in the 21st century: A critical review. Chem. Rev. 2018, 118, 2249–2295. [Google Scholar] [CrossRef] [PubMed]
  11. Nasrollahzadeh, M.; Issaabadi, Z.; Tohidi, M.M.; Sajadi, S.M. Recent progress in application of graphene supported metal nanoparticles in C-C and C-X coupling reactions. Chem. Rec. 2018, 18, 165–229. [Google Scholar] [CrossRef] [PubMed]
  12. Roy, D.; Uozumi, Y. Recent advances in palladium-catalyzed cross-coupling reactions at ppm to ppb molar catalyst loadings. Adv. Synth. Catal. 2018, 360, 602–625. [Google Scholar] [CrossRef]
  13. Balanta, A.; Godard, C.; Claver, C. Pd nanoparticles for C-C coupling reactions. Chem. Soc. Rev. 2011, 40, 4973–4985. [Google Scholar] [CrossRef] [PubMed]
  14. Lamblin, M.; Nassar-Hardy, L.; Hierso, J.C.; Fouquet, E.; Felpin, F.X. Recyclable heterogeneous palladium catalysts in pure water: Sustainable developments in Suzuki, Heck, Sonogashira and Tsuji–Trost reactions. Adv. Synth. Catal. 2010, 352, 33–79. [Google Scholar] [CrossRef]
  15. Nasrollahzadeh, M.; Atarod, M.; Alizadeh, M.; Hatamifard, A.; Sajadi, S.M. Recent advances in the application of heterogeneous nanocatalysts for sonogashira coupling reactions. Curr. Org. Chem. 2017, 21, 708–749. [Google Scholar] [CrossRef]
  16. Tang, Z.; Shen, S.; Zhuang, J.; Wang, X. Noble-metal-promoted three-dimensional macro assembly of single-layered graphene oxide. Angew. Chem. Int. Ed. 2010, 49, 4603–4607. [Google Scholar] [CrossRef] [PubMed]
  17. Movassagh, B.; Parvis, F.S.; Navidi, M. Pd(II) salen complex covalently anchored to multi-walled carbon nanotubes as a heterogeneous and reusable precatalyst for Mizoroki–Heck and Hiyama cross-coupling reactions. Appl. Organomet. Chem. 2015, 29, 40–44. [Google Scholar] [CrossRef]
  18. Polshettiwar, V.; Len, C.; Fihri, A. Silica-supported palladium: Sustainable catalysts for cross-coupling reactions. Coord. Chem. Rev. 2009, 253, 2599–2626. [Google Scholar] [CrossRef]
  19. Veerakumar, P.; Thanasekaran, P.; Lu, K.-L.; Liu, S.B.; Rajagopal, S. Computational studies of versatile heterogeneous palladium-catalyzed Suzuki, Heck, and Sonogashira coupling reactions. ACS Sustain. Chem. Eng. 2017, 5, 6357–6376. [Google Scholar] [CrossRef]
  20. Nasrollahzadeh, M.; Azarian, A.; Ehsani, A.; Khalaj, M. Preparation, optical properties and catalytic activity of TiO2@ Pd nanoparticles as heterogeneous and reusable catalysts for ligand-free Heck coupling reactions. J. Mol. Catal. A Chem. 2014, 394, 205–210. [Google Scholar] [CrossRef]
  21. Martinez-Klimov, M.E.; Hernandez-Hipolito, P.; Klimova, T.E.; Solis-Casados, D.A.; Martinez-Garcia, M. Development of reusable palladium catalysts supported on hydrogen titanate nanotubes for the Heck reaction. J. Catal. 2016, 342, 138–150. [Google Scholar] [CrossRef]
  22. Veisi, H.; Sedrpoushan, A.; Hemmati, S. Palladium supported on diaminoglyoxime-functionalized Fe3O4 nanoparticles as a magnetically separable nanocatalyst in Heck coupling reaction. Appl. Organomet. Chem. 2015, 29, 825–828. [Google Scholar] [CrossRef]
  23. Firouzabadi, H.; Iranpoor, N.; Gholinejad, M.; Akbari, S.; Jeddi, N. Palladium nanoparticles supported on agarose-functionalized magnetic nanoparticles of Fe3O4 as a recyclable catalyst for C–C bond formation via Suzuki–Miyaura, Heck–Mizoroki and Sonogashira–Hagihara coupling reactions. RSC Adv. 2014, 4, 17060–17070. [Google Scholar] [CrossRef]
  24. Martinez, A.V.; Leal-Duaso, A.; Garcia, J.I.; Mayoral, J.A. An extremely highly recoverable clay-supported Pd nanoparticle catalyst for solvent-free Heck–Mizoroki reactions. RSC Adv. 2015, 5, 59983–59990. [Google Scholar] [CrossRef] [Green Version]
  25. Singh, V.; Ratti, R.; Kaur, S. Synthesis and characterization of recyclable and recoverable MMT-clay exchanged ammonium tagged carbapalladacycle catalyst for Mizoroki–Heck and Sonogashira reactions in ionic liquid media. J. Mol. Catal. A Chem. 2011, 334, 13–19. [Google Scholar] [CrossRef]
  26. Guibal, E. Heterogeneous catalysis on chitosan-based materials: A review. Prog. Polym. Sci. 2005, 30, 71–109. [Google Scholar] [CrossRef]
  27. El Kadib, A. Chitosan as a sustainable organocatalyst: A concise overview. ChemSusChem 2015, 8, 217–244. [Google Scholar] [CrossRef] [PubMed]
  28. Mohammadi, E.; Movassagh, B. Synthesis of polystyrene-supported Pd(II)-NHC complex derived from theophylline as an efficient and reusable heterogeneous catalyst for the Heck-Matsuda cross-coupling reaction. J. Mol. Catal. A Chem. 2016, 418, 158–167. [Google Scholar] [CrossRef]
  29. Mitsudome, T.; Nose, K.; Mori, K.; Mizugaki, T.; Ebitani, K.; Jitsukawa, K.; Kaneda, K. Montmorillonite-entrapped sub-nanoordered Pd clusters as a heterogeneous catalyst for allylic substitution reactions. Angew. Chem. Int. Ed. 2007, 46, 3288–3290. [Google Scholar] [CrossRef] [PubMed]
  30. Borah, B.J.; Dutta, D.K. In situ stabilization of Pd0-nanoparticles into the nanopores of modified montmorillonite: Efficient heterogeneous catalysts for Heck and Sonogashira coupling. J. Mol. Catal. A Chem. 2013, 366, 202–209. [Google Scholar] [CrossRef]
  31. Zhou, L.M.; Qi, X.L.; Jiang, X.H.; Zhou, Y.F.; Fu, H.Y.; Chen, H. Organophilic worm-like ruthenium nanoparticles catalysts by the modification of CTAB on montmorillonite supports. J. Colloid Interface Sci. 2013, 392, 201–205. [Google Scholar] [CrossRef] [PubMed]
  32. Zeng, M.F.; Wang, Y.D.; Liu, Q.; Yuan, X.; Zuo, S.F.; Feng, R.K.; Yang, J.; Wang, B.Y.; Qi, C.Z.; Lin, Y. Encaging palladium nanoparticles in chitosan modified montmorillonite for efficient, recyclable catalysts. ACS Appl. Mater. Interface 2016, 8, 33157–33164. [Google Scholar] [CrossRef] [PubMed]
  33. Sinha Ray, S.; Okamoto, M. Polymer/layered silicate nanocomposites: A review from preparation to processing. Prog. Polym. Sci. 2003, 28, 1539–1641. [Google Scholar] [CrossRef]
  34. Ott, L.S.; Hornstein, B.J.; Finke, R.G. A test of the transition-metal nanocluster formation and stabilization ability of the most common polymeric stabilizer, poly (vinyl pyrrolidone), as well as four other polymeric protectants. Langmuir 2006, 22, 9357–9367. [Google Scholar] [CrossRef] [PubMed]
  35. Narayanan, R.; El-Sayed, M.A. Effect of catalysis on the stability of metallic nanoparticles: Suzuki reaction catalyzed by PVP-palladium nanoparticles. J. Am. Chem. Soc. 2003, 125, 8340–8347. [Google Scholar] [CrossRef] [PubMed]
  36. Song, H.; Rioux, R.M.; Hoefelmeyer, J.D.; Komor, R.; Niesz, K.; Grass, M.; Yang, P.; Somorjai, G.A. Hydrothermal growth of mesoporous SBA-15 silica in the presence of PVP-stabilized Pt nanoparticles: Synthesis, characterization, and catalytic properties. J. Am. Chem. Soc. 2006, 128, 3027–3037. [Google Scholar] [CrossRef] [PubMed]
  37. Guo, L.P.; Bai, J.; Li, C.P.; Meng, Q.R.; Liang, H.O.; Sun, W.Y.; Li, H.Q.; Liu, H. A novel catalyst containing palladium nanoparticles supported on PVP composite nanofiber films: Synthesis, characterization and efficient catalysis. Appl. Surf. Sci. 2013, 283, 107–114. [Google Scholar] [CrossRef]
  38. Jiang, C.L.; Nie, J.; Ma, G.P. A polymer/metal core–shell nanofiber membrane by electrospinning with an electric field, and its application for catalyst support. RSC Adv. 2016, 6, 22996–23007. [Google Scholar] [CrossRef]
  39. Papp, S.; Patakfalvi, R.; Dékány, I. Metal nanoparticle formation on layer silicate lamellae. Colloid Polym. Sci. 2008, 286, 3–14. [Google Scholar] [CrossRef]
  40. Papp, S.; Szél, J.; Oszkó, A.; Dékány, I. Synthesis of polymer-stabilized nanosized rhodium particles in the interlayer space of layered silicates. Chem. Mater. 2004, 16, 1674–1685. [Google Scholar] [CrossRef]
  41. Papp, S.; Dékány, I. Stabilization of palladium nanoparticles by polymers and layer silicates. Colloid Polym. Sci. 2003, 281, 727–737. [Google Scholar] [CrossRef]
  42. Zeng, M.F.; Du, Y.J.; Qi, C.Z.; Zuo, S.F.; Li, X.D.; Shao, L.J.; Zhang, X.-M. An efficient and recyclable heterogeneous palladium catalyst utilizing naturally abundant pearl shell waste. Green Chem. 2011, 13, 350–356. [Google Scholar] [CrossRef]
  43. Darder, M.; Colilla, M.; Ruiz-Hitzky, E. Biopolymer−clay nanocomposites based on chitosan intercalated in montmorillonite. Chem. Mater. 2003, 15, 3774–3780. [Google Scholar] [CrossRef]
  44. Komori, Y.; Sugahara, Y.; Kuroda, K. Direct intercalation of poly (vinyl pyrrolidone) into kaolinite by a refined guest displacement method. Chem. Mater. 1999, 11, 3–18. [Google Scholar] [CrossRef]
  45. Kerr, T.A.; Wu, H.; Nazar, L.F. Concurrent polymerization and insertion of aniline in molybdenum trioxide: Formation and properties of a [poly(aniline)]0.24 MoO3 nanocomposite. Chem. Mater. 1996, 8, 2005–2015. [Google Scholar] [CrossRef]
  46. Wagner, C.D.; Riggs, W.M.; Davis, L.E.; Mullenberg, G.E. Handbook of X-ray Photoelectron Spectroscopy; Perkin-Elmer: Waltham, MA, USA, 1979. [Google Scholar]
  47. Pethrick, R.A. Positron annihilation—A probe for nanoscale voids and free volume? Prog. Polym. Sci. 1997, 22, 1–47. [Google Scholar] [CrossRef]
  48. Jean, Y.C. Positron annihilation spectroscopy for chemical analysis: A novel probe for microstructural analysis of polymers. Microchem. J. 1990, 42, 72–102. [Google Scholar] [CrossRef]
  49. Jean, Y.C.; Van Horn, J.D.; Hung, W.S.; Lee, K.R. Perspective of positron annihilation spectroscopy in polymers. Macromolecules 2013, 46, 7133–7145. [Google Scholar] [CrossRef]
  50. Tao, S.J. Positronium annihilation in molecular substances. J. Chem. Phys. 1972, 56, 5499–5510. [Google Scholar] [CrossRef]
  51. Eldrup, M.; Lightbody, D.; Sherwood, J.N. The temperature-dependence of positron lifetimes in solid pivalic acid. Chem. Phys. 1981, 63, 51–58. [Google Scholar] [CrossRef]
Scheme 1. Process of the preparation of the Pd@MMT/PVP catalytic composite.
Scheme 1. Process of the preparation of the Pd@MMT/PVP catalytic composite.
Polymers 10 00669 sch001
Figure 1. XRD patterns of MMT, MMT/PVP support and Pd@MMT/PVP catalytic composite.
Figure 1. XRD patterns of MMT, MMT/PVP support and Pd@MMT/PVP catalytic composite.
Polymers 10 00669 g001
Figure 2. XPS spectra of the Pd0@MMT/PVP catalytic composite.
Figure 2. XPS spectra of the Pd0@MMT/PVP catalytic composite.
Polymers 10 00669 g002
Figure 3. HR-TEM images of MMT (A), MMT/PVP support (B) and Pd@MMT/PVP catalytic composite (C).
Figure 3. HR-TEM images of MMT (A), MMT/PVP support (B) and Pd@MMT/PVP catalytic composite (C).
Polymers 10 00669 g003
Figure 4. The distribution of positron annihilation lifetime (A) and calculated micro defects size (B) of MMT/PVP support, Pd2+@MMT/PVP and Pd0@MMT/PVP catalytic composite.
Figure 4. The distribution of positron annihilation lifetime (A) and calculated micro defects size (B) of MMT/PVP support, Pd2+@MMT/PVP and Pd0@MMT/PVP catalytic composite.
Polymers 10 00669 g004
Figure 5. TGA curves of MMT, PVP and Pd@MMT/PVP catalytic composite.
Figure 5. TGA curves of MMT, PVP and Pd@MMT/PVP catalytic composite.
Polymers 10 00669 g005
Figure 6. Dependence of the Heck reaction yield on the recycle times of the Pd0@MMT/PVP catalytic composite.
Figure 6. Dependence of the Heck reaction yield on the recycle times of the Pd0@MMT/PVP catalytic composite.
Polymers 10 00669 g006
Table 1. Variation of positron annihilation lifetimes, intensities, and mean diameter of micro-defects of the MMT, MMT/PVP support and Pd@MMT/PVP catalytic composite analyzed by LT-9 program.
Table 1. Variation of positron annihilation lifetimes, intensities, and mean diameter of micro-defects of the MMT, MMT/PVP support and Pd@MMT/PVP catalytic composite analyzed by LT-9 program.
Sampleτ1 (ns)I1 (%)τ2 (ns)I2 (%)τ3 (ns)I3 (%)D (nm)
MMT0.235663.70.512533.22.8063.10.6988
MMT/PVP (90/10)0.197751.60.44943.32.2135.10.6072
MMT/PVP (80/20)0.202557.30.475536.32.2086.40.6062
MMT/PVP (70/30)0.188156.60.459135.92.1147.50.5900
MMT/PVP (60/40)0.190852.40.463337.82.0069.80.5708
MMT/PVP (50/50)0.179650.70.427838.51.92810.80.5562
Pd2+@MMT/PVP (80/20)0.186163.20.436430.82.2656.00.6158
Pd0@MMT/PVP (80/20)0.200755.60.43140.42.2954.00.6208
Table 2. Heck reactions between aryl halides and acrylates or alkenes catalyzed with Pd@MMT/PVP catalysts. a
Table 2. Heck reactions between aryl halides and acrylates or alkenes catalyzed with Pd@MMT/PVP catalysts. a
Polymers 10 00669 i001
EntryAryl HalidesAcrylates or AlkenesProductsYield b
1Polymers 10 00669 i002Polymers 10 00669 i003Polymers 10 00669 i00494%
2Polymers 10 00669 i005Polymers 10 00669 i006Polymers 10 00669 i00793%
3Polymers 10 00669 i008Polymers 10 00669 i009Polymers 10 00669 i01089%
4Polymers 10 00669 i011Polymers 10 00669 i012Polymers 10 00669 i01387%
5Polymers 10 00669 i014Polymers 10 00669 i015Polymers 10 00669 i01690%
6Polymers 10 00669 i017Polymers 10 00669 i018Polymers 10 00669 i01989%
7Polymers 10 00669 i020Polymers 10 00669 i021Polymers 10 00669 i02287%
8Polymers 10 00669 i023Polymers 10 00669 i024Polymers 10 00669 i02593%
9Polymers 10 00669 i026Polymers 10 00669 i027Polymers 10 00669 i02890%
10Polymers 10 00669 i029Polymers 10 00669 i030Polymers 10 00669 i03189%
11Polymers 10 00669 i032Polymers 10 00669 i033Polymers 10 00669 i034trace
12Polymers 10 00669 i035Polymers 10 00669 i036Polymers 10 00669 i037trace
13Polymers 10 00669 i038Polymers 10 00669 i039Polymers 10 00669 i04044%
14Polymers 10 00669 i041Polymers 10 00669 i042Polymers 10 00669 i04347%
a Reaction conditions: 1 mmol aryl halides, 2 mmol acrylates or alkenes, 3 μmol Pd@MMT/PVP catalyst, 3 mmol CH3COOK, in 5 mL DMSO + 0.2 mL ethylene glycol solution, 110 °C, 5 h. b GC/MS yield.

Share and Cite

MDPI and ACS Style

Xu, M.; Zhao, J.; Shu, G.; Liu, Q.; Zeng, M. Heterogeneous Catalytic Composites from Palladium Nanoparticles in Montmorillonite Intercalated with Poly (Vinyl Pyrrolidone) Chains. Polymers 2018, 10, 669. https://doi.org/10.3390/polym10060669

AMA Style

Xu M, Zhao J, Shu G, Liu Q, Zeng M. Heterogeneous Catalytic Composites from Palladium Nanoparticles in Montmorillonite Intercalated with Poly (Vinyl Pyrrolidone) Chains. Polymers. 2018; 10(6):669. https://doi.org/10.3390/polym10060669

Chicago/Turabian Style

Xu, Mengdie, Jing Zhao, Guiqing Shu, Qi Liu, and Minfeng Zeng. 2018. "Heterogeneous Catalytic Composites from Palladium Nanoparticles in Montmorillonite Intercalated with Poly (Vinyl Pyrrolidone) Chains" Polymers 10, no. 6: 669. https://doi.org/10.3390/polym10060669

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