G Model
ARTICLE IN PRESS
CATTOD-10612; No. of Pages 8
Catalysis Today xxx (2017) xxx–xxx
Contents lists available at ScienceDirect
Catalysis Today journal homepage: www.elsevier.com/locate/cattod
Magnetite nanoparticles coated with tannic acid as a green reductant and stabilizer sorbent for palladium ions: Synthetic application of Fe3 O4 @TA-Pd NPs as magnetically separable and reusable nanocatalyst for reduction of 4-nitrophenol and Suzuki reactions Hojat Veisi a,∗ , Mozhgan Pirhayati b , Pourya Mohammadi c , Mohammad Reza Abdi a , Javad Gholami d a
Department of Chemistry, Payame Noor University, Tehran, Iran Young Researchers and Elite Club, Malayer Branch, Islamic Azad University, Malayer, Iran c Department of Chemistry, Shahid Bahonar University of Kerman, Kerman 76169, Iran d Department of Applied Chemistry, Faculty of Science, Malayer University, Malayer, Iran b
a r t i c l e
i n f o
Article history: Received 2 December 2016 Received in revised form 29 January 2017 Accepted 18 February 2017 Available online xxx Keywords: Magnetite nanoparticles Tannic acid Palladium 4-Nitrophenol Suzuki
a b s t r a c t Tannic acid as a humic-like substance having a lot of phenolic hydroxyl and carbonyl functional groups on the framework can be used in the modification of Fe3 O4 NPs. Moreover, the feasibility of complexation of polyphenols with palladium ions in aqueous solution can improve the surface properties and capacity of the Fe3 O4 @TA NPs for sorbent and reduction of palladium ions. Therefore, the novel nano-adsorbent (Fe3 O4 @TA NPs) has potential ability as both the reducing and stabilizing agent for immobilization of palladium nano particles to make a novel magnetic palladium nanocatalyst. Inductively coupled plasma (ICP), transmission electron microscopy (TEM), scanning electron microscopy (SEM), X-ray photoelectron spectroscopy (XPS), energy-dispersive X-ray spectrum (EDS), vibrating sample magnetometer (VSM) and Fourier transform infrared (FTIR) studies have been used to characterize the catalyst. Fe3 O4 @TA/Pd NPs show high catalytic activity as recyclable nanocatalyst toward Suzuki-Miyaura cross-coupling reactions and reduction of 4-nitrophenol (4-NP) at room temperature. © 2017 Elsevier B.V. All rights reserved.
1. Introduction The Suzuki-Miyaura cross-coupling reaction [1–3] has become arguably one of the most powerful synthetic methods for preparing biaryl compounds, which constitute a wide range of natural products [4,5], pharmaceuticals [6,7], and polymers [8]. Homogeneous palladium complexes have been widely explored as catalysts for this reaction [9–12]. Despite the high activity and selectivity of these systems, they have limited utility in industrial processes because of challenges associated with their removal, recovery, and recycling. Complicating matters, many ligands used in catalyst design (e.g., phosphines) are air sensitive, challenging to prepare and/or costly, limiting suitability of such materials for large-scale application [13]. To overcome these issues, many heterogeneous catalysts have been developed, including inorganic
∗ Corresponding author. E-mail address:
[email protected] (H. Veisi).
solid (e.g., charcoal, zeolites and metal oxides) supported, and polymer-stabilized systems [14,15]. Palladium nanoparticles prepared in situ also show promise [16,17]. As a result of the toxicity and high price of Pd, in recent years, interest has grown in the organization and use of recoverable heterogeneous Pd catalysts [18–23]. For this purpose, different solid materials such as modified silicas [24,25], polymers [26,27], mesoporous materials [28–31], ionic liquids [32–35], and natural supports [36,37] have been designated as supports for the heterogenization of Pd catalysts. Even with significant achievements in this area, the separation and efficient recovery of the heterogeneous catalyst from the reaction medium by conventional methods such as filtration or centrifugation is not always an easy assignment. One of the best methods to solve this problem is the use of magnetic nanoparticles (NPs) of Fe3 O4 as a very convenient support for heterogenization of Pd catalysts [38–43]. Recently, magnetic nanoparticles have been extensively employed as alternative catalyst supports, in view of their high surface area resulting in high catalyst loading capacity, high dispersion, outstanding stability, and convenient catalyst
http://dx.doi.org/10.1016/j.cattod.2017.02.023 0920-5861/© 2017 Elsevier B.V. All rights reserved.
Please cite this article in press as: H. Veisi, et al., Magnetite nanoparticles coated with tannic acid as a green reductant and stabilizer sorbent for palladium ions: Synthetic application of Fe3 O4 @TA-Pd NPs as magnetically separable and reusable nanocatalyst for reduction of 4-nitrophenol and Suzuki reactions, Catal. Today (2017), http://dx.doi.org/10.1016/j.cattod.2017.02.023
G Model CATTOD-10612; No. of Pages 8 2
ARTICLE IN PRESS H. Veisi et al. / Catalysis Today xxx (2017) xxx–xxx
recycling [44,45]. Magnetic separation renders the recovery of catalysts from liquid phase reactions much easier than by cross-flow filtration and centrifugation. In addition, the magnetic properties of the particles are stable enough to tolerate most chemical environments, with the exception of those that are extremely acidic or corrosive. The focus today is to make the experimental conditions more environmentally friendly, e.g., in water and in the presence of ligand-free catalysts in very small amounts like nanoparticles (NPs). Although various methodologies have been developed for the synthesis of palladium nanoparticles, including chemical and electrochemical reduction [46,47], ion exchange [48], vapor deposition, thermal decomposition [49,50] and the polyol method [51], preparation of metal nanoparticles is usually based on the reduction of a metal salt in the presence of a reducing agent, such as sodium borohydride, hydrazine, dimethylformamide or hydrogen, and a stabilizer, such as polymeric materials, dendrimers, surfactants, organic ligands or polyoxometalates [52–59]. Bioinspired, eco-friendly greener methods for the synthesis of metal nanoparticles is one of the most attractive aspects of today’s nanoscience and nanotechnology [60,61]. In addition, the other advantages of this environmentally benign and safe protocol include a simple reaction setup, very mild reaction conditions, use of nontoxic solvents such as water, elimination of toxic and dangerous materials and cost effectiveness as well as compatibility for biomedical and pharmaceutical applications [62]. Also, in this method there is no need to use high pressure, energy, temperature and toxic chemicals. Despite the availability of methods for the green synthesis of Pd NPs by various plants or gums the potential of plants as biological materials for the synthesis of nanoparticles is yet to be fully explored [63,64]. Tannic acid is a plant polyphenol which is found, along with other condensed tannins, in several beverages including red wine, beer, coffee, black tea, green tea, and many foodstuffs such as grapes, pears, bananas, sorghum, black-eyed peas, lentils and chocolate [65,66]. Similar to many polyphenols, tannic acid has been shown to possess antioxidant [67–70], antimutagenic [68,71,72] and anticarcinogenic properties [73–75]. Nitrophenols are among the most common organic pollutants in industrial and agricultural wastewaters. It can damage the central nervous system, liver, kidney and blood of humans and animals Degradation of the compound to nondangerous product is difficult because of its high stability and low solubility in water [76–78]. Its reduction product, 4-aminophenol (4-AP), is very useful and important in many applications that include analgesic and antipyretic drugs, photographic developer, corrosion inhibitor, anticorrosion lubricant, and so on [79]. Recently, developing new nanocatalysts with high catalytic activity for 4-NP reduction has attracted much attention. For example, Pd graphene nanohybrids [80], AuPd nanoparticles/graphene nanosheets [81], silver supported nanoporous iron oxide microbox hybrids [82], Au-CeO2 @ZrO2 yolk–shell nanoreactors [83], Ni@Pd core-shell nanoparticles modified fibrous silica nanospheres [84], have been synthesized for highly efficient catalytic reduction of 4-NP by NaBH4 . The nm-sized metal oxides are not target selective and are unsuitable for samples with complex matrices [85]. Therefore, a suitable coating is essential to overcome such limitations. Also, surface modification stabilizes the NPs and also prevents their oxidation. However, MNPs can efficiently be functionalized based on the formation of relatively stable linker between hydroxyl groups on the NPs surface and suitable anchoring agents such as phosphonic acid and dopamine derivatives [86][86a]. So, TA as a humic-like substance having a lot of phenolic hydroxyl and carbonyl functional groups on the framework can be used in the modification of MNP (Scheme 1). Moreover, the feasibility of complexation of polyphenols with polyvalent cations in simple aqueous solutions can improve the surface properties and capacity of the Fe3 O4 @TA NPs for to adsorption and reduction of metal ions [86][86b]. Therefore,
the novel nano-adsorbent (Fe3 O4 @TA NPs) has potential ability as both the reducing and stabilizing agent for immobilization of palladium nano particles to make a novel magnetically separable and reusable catalyst (Scheme 1). 2. Materials and methods 2.1. Materials High-purity chemical reagents were purchased from the Merck and Aldrich chemical companies. All materials were of commercial reagent grade. FT-IR spectra were recorded on a FT-IR-8400 infrared spectrometer (Simadzu) using pressed KBr pellets. X-ray diffraction (XRD, Rigaku Corporation, Tokyo, Japan) patterns were obtained at room temperature on a Riga kuD/Max-2550 powder diffractometer with a scanning rate of 5◦ min−1 , and recorded in the 2 range of 10–70 ◦ C. SEM, (AB912 LEO) was used to obtain information on the morphology of nanocomposite. Transmission electron microscopy, TEM, (Zeiss, EM10C, 120 kV) was used to obtain information on the particle size and morphology of the nanocomposite. 2.2. Preparation of the Fe3 O4 and Fe3 O4 @Tannic acid (Fe3 O4 @TA) In our previous works [39a,b,c,d,e] the magnetic Fe3 O4 nanoparticles were prepared by the chemical coprecipitation method, and the detailed procedure is described below. FeCl2 ·4H2 O (4 g) and FeCl3 ·6H2 O (10.4 g) were dissolved into 50 mL deoxygenated water followed by adding 1.7 mL of concentrated hydrochloric acid. The resulting solution was dropped into 500 mL of 1.5 M NaOH solution under vigorous stirring and N2 protection at 80 ◦ C. The obtained magnetic nanoparticles were separated from solution by a powerful external magnet and rinsed with 200 mL deionized water three times. Finally the products were dried at 40 ◦ C to give Fe3 O4 nanoparticles. In the next step for the synthesis of Fe3 O4 @Tannic acid, in the first step magnetite nanoparticle (1 g) was dispersed in 50 cc water and sonicated for 20 min. Next, tannic acid (1 g) was dissolved in water (500 cc) and added to the mixture. Afterwards, the solution was stirred for 12 h at 40 ◦ C, and in the following the Fe3 O4 @TA NPs precipitate obtained was separated by magnetic decantation and washed several times with deionized water. Finally, the Fe3 O4 @TA NPs obtained were dried in a vacuum oven at 40 ◦ C for 12 h. 2.3. Preparation of the Fe3 O4 @TA/Pd The Fe3 O4 @TA (500 mg) were dispersed in CH3 CN (30 mL) by ultrasonic bath for 30 min. Subsequently, a yellow solution of PdCl2 (30 mg) in 30 mL acetonitrile was added to dispersion of Fe3 O4 @TA and the mixture was stirred for 24 h at room temperature in order to ensure complete reduction of Pd(II) ions in the precursor solution [Tannic acid reduces of Pd(II) ions to Pd(0)]. Then, the Fe3 O4 @TA/Pd was separated by magnetic decantation and washed by CH3 CN, H2 O and acetone respectively to remove the unattached substrates. Scheme 1 depicted the synthetic procedure of Fe3 O4 @TA/Pd. The final nanocatalyst was dried in vacuum at 40 ◦ C. The concentration of palladium was 3.75 wt% (0.353 mmol/g), which was determined by ICP-AES. 2.4. General procedure for the reduction of 4-NP For catalytic testing of Fe3 O4 @TA/Pd nanocomposite, the reactant solutions of 4-NP and NaBH4 were freshly prepared in molar concentration of 2.5 mM and 0.2 M, respectively. Followed by, 1 mL of both solutions were added and mixed by magnetic stirring and transferred to a quartz cuvette. Subsequently, 0.2 mg of catalyst was loaded into the cuvette to start the reaction. The intensity of
Please cite this article in press as: H. Veisi, et al., Magnetite nanoparticles coated with tannic acid as a green reductant and stabilizer sorbent for palladium ions: Synthetic application of Fe3 O4 @TA-Pd NPs as magnetically separable and reusable nanocatalyst for reduction of 4-nitrophenol and Suzuki reactions, Catal. Today (2017), http://dx.doi.org/10.1016/j.cattod.2017.02.023
G Model CATTOD-10612; No. of Pages 8
ARTICLE IN PRESS H. Veisi et al. / Catalysis Today xxx (2017) xxx–xxx
3
Scheme 1. Preparation of Fe3 O4 @TA/Pd nanocatalyst.
the absorption peak at 400 nm in ultra violet visible (UV–vis) spectroscopy was used to monitor the process of the conversion of 4-NP to 4-AP. 2.5. Suzuki coupling reactions Catalytic activity of the synthesized Fe3 O4 @TA/Pd nanocomposite was tested through the Suzuki reaction of 4-methylhalobenzene (Cl, Br, I) with phenylboronic acid. In the typical experiment, a mixture of aryl halide (1.0 mmol), phenylboronic acid (1.1 mmol), K2 CO3 (2 mmol), and 6.0 mg Fe3 O4 @TA/Pd nanocomposite (6 mg ≈ 0.021 mol% Pd) were added into 4 mL water/ethanol (1:1) solution in a 25 mL balloon. The mixture was then stirred for the desired time at room temperature. The reaction was monitored by thin layer chromatography (TLC, n-hexane/acetone; 4:1). After completion of the reaction, 5 mL ethanol was added, and the catalyst was removed by external magnet. Further purification was achieved by column chromatography. 3. Results and discussion 3.1. Characterization FT-IR spectral analysis disclosed the surface chemistry of the Fe3 O4 NPs, Fe3 O4 @TA NPs and Fe3 O4 @TA/Pd NPs. In the spectra of Fe3 O4 NPs, the broad band at 1627 and 3446 cm−1 are corresponding to the surface-adsorbed water and hydroxyl groups of Fe3 O4 NPs, while the peaks at 459 and 598 cm-1 are respectively related to the octahedral bending and tetrahedral stretching vibration of the Fe-O function group and peak at 630 cm−1 confirm the existence of Fe3 O4 structure and also confirm that we do not have any other structure of Iron oxide like hematite or goethite. In the spectra of Fe3 O4 @TA NPs, there are several additional peaks in the spectrum belonging to the TA coating. Absorption bands at 1718 and 1080 cm−1 belong to stretching vibrations of C O and C O, respectively. Peaks appearing at 2928 and 2889 cm−1 are characteristic of C H stretching vibrations. The weak absorbtion band at 923 cm−1 is assigned to O H out of plane bending mode of the acid group. The band at 759 cm−1 can be related to the C H out plane bend of phenyl group [87]. According to these results, FT-IR analyses suggest that TA is successfully coated on the surface of Fe3 O4 NPs. The spectrum of Fe3 O4 @TA/Pd NPs was same with the spectra of Fe3 O4 @TA NPs spectra. This is also indicating that the structure was not changed after reduction–absorbtion of palladium nanoparticles. During the reduction process carboxylic acid group (COOH) present in the tannic acid becomes COO− . This COO− along with the rest part of the polymer can act as surfactant to attach on the surface of Pd NPs and it stabilizes Pd NPs through electrosteric stabilization [88]. This may be confirmed by disappearance of the bond at 1718 cm−1 (C O) in the curve c (Fig. 1).
Scanning electron microscopy (SEM) is a primary tool for determining the size distribution, particle shape, surface morphology, and fundamental physical properties. The SEM image of (a) Fe3 O4 ; and (b) Fe3 O4 @TA/Pd was shown that the catalyst was formed in nanometer-sized particles as quasi-spherical with an average diameter of 10–25 nm (Fig. 2). Also, in compared with Fe3 O4 , a continuous layer of biopolymer can be observed on the surface of Fe3 O4 @TA/Pd catalyst (Fig. 2), and the energy-dispersive X-ray spectrum (EDS) confirmed the presence of Pd, Fe, O and C species in the structure of the material (Fig. 3). Furthermore, transmission Electron Microscope (TEM) has been used to study the morphology of Pd nanoparticles. The particles were found to be very small, spherical in shape and dispersed (Fig. 4). The histogram clearly shows that the average hydrodynamic size of the Fe3 O4 @TA/Pd is in the range of 5–25 nm. Fig. 5 shows a representative SEM image and corresponding elemental maps for the synthesized catalyst. It can be seen that Pd metal particles are well dispersed in the composite. The selected area elemental analysis figure reveals the presence of C, Fe and Pd throughout the sample in a homogeneous manner, which confirms the regular uniformity of the prepared sample. X-ray photoelectron spectroscopy (XPS) was used to characterize the electronic properties and chemical state information of Fe3 O4 @TA/Pd NPs (Fig. 6). The binding energies of Fe 2p3/2 and Fe 2p1/2 are 723.8 eV and 711.5 eV, respectively, which correspond well with those of bulk Fe3 O4 . The binding energies (BEs) of 533.2 and 553 eV were assigned to incarcerated Pd(0) species for 3p3/2 and 3p1/2 , respectively. Moreover, the presence of carbon in the structure of Fe3 O4 @TA/Pd NPs confirmed by XPS analysis by the appearance of peaks related to C1 s at BE = 285.4 eV for aliphatic carbon, BE = 286.6 eV for oxygenated carbon, and BE = 288.8 eV for COO. The magnetic property of Fe3 O4 @TA/Pd nanocatalyst containing a magnetite component were studied using a vibrating sample magnetometer (VSM) at room temperature. The saturation magnetization of Fe3 O4 @TA/Pd nanocatalyst was determined as 33.5 emu/g (Fig. 7). 3.2. Catalytic activities and recyclability The reduction of 4-NP was carried out using NaBH4 as a hydrogen producing agent and Fe3 O4 @TA/Pd NPs as a nanocatalyst. As the reaction proceeded, the color of the solution changed gradually from yellow to colorless at room temperature. UV–vis spectra were recorded at short intervals to monitor the progress of the reaction. The intensity of the absorption peak at 400 nm for 4-NP was monitored by UV–vis spectroscopy along with time, at the same time, a new absorption peak for 4-AP appeared at 304 nm (Fig. 8a). The rate of reduction reaction catalyzed by Fe3 O4 @TA/Pd was assumed to be independent of the concentration of NaBH4
Please cite this article in press as: H. Veisi, et al., Magnetite nanoparticles coated with tannic acid as a green reductant and stabilizer sorbent for palladium ions: Synthetic application of Fe3 O4 @TA-Pd NPs as magnetically separable and reusable nanocatalyst for reduction of 4-nitrophenol and Suzuki reactions, Catal. Today (2017), http://dx.doi.org/10.1016/j.cattod.2017.02.023
G Model CATTOD-10612; No. of Pages 8 4
ARTICLE IN PRESS H. Veisi et al. / Catalysis Today xxx (2017) xxx–xxx
Fig. 1. FTIR spectrum of (a) Fe3 O4 , (b) Fe3 O4 @TA, and (c) Fe3 O4 @TA/Pd.
Fig. 2. SEM image of (a) Fe3 O4 , and (b) Fe3 O4 @TA/Pd.
because this reagent was used in large excess compared to 4-NP. Therefore, the kinetic data were fitted by a first-order rate law. Linear relationship between ln(A/A0 ) and reaction time is obtained in the reduction catalyzed by Fe3 O4 @TA/Pd nano composite (Fig. 8b), and the rate constant k is calculated to be 0.018 s−1 (1.1 min−1 ) and the catalyst activity parameter (Ka) of the composite materials have also been calculated from the ratio of rate constants for the catalysts to the amount of given catalyst added, where Ka = k/m, that Ka is calculated to be 90 s−1 g−1 . The activity parameters used for the comparison of the catalytic activity of Fe3 O4 @TA/Pd with the reported other based catalysts employed for the reduction of 4-NP are listed in Table 1. The catalytic activity of Fe3 O4 @TA/Pd, obtained in this study, is 90 s−1 g−1 which is significantly higher than the catalytic activities of the mentioned nano catalysts (Table 1). So, the prepared nanocatalyst displayed superior catalytic activity for the reduction reactions.
Table 1 The comparison of catalytic activities for the reduction of 4-NP with catalysts reported in literatures. Catalyst TAC-Ag-10 P(AMPS)-Ni P(AMPS)-Co P(AMPS)-Cu Fe3 O4 @Tannic/Pd(0)
K (s−1 ) −3
5.19 × 10 9.38 × 10−4 2 × 10−3 1.72 × 10−3 1.8 × 10−2
Ka (s−1 g−1 )
Ref.
1.03 0.019 0.04 0.172 90
[89] [90] [91] [92] This work
The catalyst was separated by magnet external after monitoring the whole reduction process, washed several times with water and ethanol, respectively, and then reused. For practical applications of heterogeneous systems, the level of reusability and the activity of the catalyst is an important factor. The catalyst exhib-
Please cite this article in press as: H. Veisi, et al., Magnetite nanoparticles coated with tannic acid as a green reductant and stabilizer sorbent for palladium ions: Synthetic application of Fe3 O4 @TA-Pd NPs as magnetically separable and reusable nanocatalyst for reduction of 4-nitrophenol and Suzuki reactions, Catal. Today (2017), http://dx.doi.org/10.1016/j.cattod.2017.02.023
G Model CATTOD-10612; No. of Pages 8
ARTICLE IN PRESS H. Veisi et al. / Catalysis Today xxx (2017) xxx–xxx
Fig. 3. EDS spectrum of Fe3 O4 @TA/Pd.
Fig. 4. TEM image of Fe3 O4 @TA/Pd and its particle size distribution.
Fig. 5. SEM image of Fe3 O4 @TA/Pd elemental maps of C, Fe, Pd and Fe/C, C/Pd atoms in the catalyst.
Please cite this article in press as: H. Veisi, et al., Magnetite nanoparticles coated with tannic acid as a green reductant and stabilizer sorbent for palladium ions: Synthetic application of Fe3 O4 @TA-Pd NPs as magnetically separable and reusable nanocatalyst for reduction of 4-nitrophenol and Suzuki reactions, Catal. Today (2017), http://dx.doi.org/10.1016/j.cattod.2017.02.023
5
G Model
ARTICLE IN PRESS
CATTOD-10612; No. of Pages 8
H. Veisi et al. / Catalysis Today xxx (2017) xxx–xxx
6
Fig. 6. XPS spectrum of Fe3 O4 @TA/Pd.
40 32
Magnetization(emu/g)
24 16 8 0 -8 -16 -24 -32 -40 -10000 -8000 -6000 -4000
-2000
0
2000
4000
6000
8000
Applied Field(Oe) Fig. 7. The magnetic hysteresis loops of Fe3 O4 @TA/Pd.
10000
ited similar catalytic performance without significant reduction in the conversion even after running 10 cycles (Fig. 9). To investigate another activity of this prepared nanocatalyst, we examined the Suzuki reaction of phenylboronic acid with a variety of different 4-methyl aryl halides (I, Br, Cl) in EtOH/H2 O (1:1) at room temperature (Table 2). All the reactions afford the corresponding products in high yields. Aryl chlorides are well known to be least reactive of the aryl halides during the Suzuki reaction (relative reactivity: I > OTf > Br ≫ Cl) because of the strength of the Ar–Cl bond (Ph–X: Cl (96) > Br (81) > I (65 kcal/mol)). The results of this work were therefore in accordance with those reported previously in the literature by Wu et al. [93]. The recyclability of Fe3 O4 @TA/Pd was further studied because the recycling of the heterogeneous catalyst was an important issue for practical applications. The recovered catalyst was added to reaction mixture under the same conditions (Table 2, Entry 2) for eight cycles without a significant loss of yield and catalytic activity (Fig. 10). This result also demonstrated that the palladium leach-
Fig. 8. (a) UV–vis spectra for catalytic reduction of 4-NP to 4-AP, (b) Plot of the Ln(At /A0 ) against the reaction different times.
Please cite this article in press as: H. Veisi, et al., Magnetite nanoparticles coated with tannic acid as a green reductant and stabilizer sorbent for palladium ions: Synthetic application of Fe3 O4 @TA-Pd NPs as magnetically separable and reusable nanocatalyst for reduction of 4-nitrophenol and Suzuki reactions, Catal. Today (2017), http://dx.doi.org/10.1016/j.cattod.2017.02.023
G Model
ARTICLE IN PRESS
CATTOD-10612; No. of Pages 8
H. Veisi et al. / Catalysis Today xxx (2017) xxx–xxx
7
Table 2 Suzuki cross-coupling reaction between phenylboronic acid and 4-methyl-halobenzene.
B(OH)2
Fe3O4@Tannic/Pd(0)
Me
X
Me
K2CO3, EtOH:H2O (1:1), rt X= Cl, Br, I
.
Entry
R
X
Time
Yield (%)
1 2 3
Me Me Me
I Br Cl
15 min 1h 6h
98 96 86
leaching is good characteristics. Catalyst was reused for many times with any loosing of catalytic activity. The result showed that the green approach to synthesizing Pd NPs is useful for removing toxic pollutants and dyes such as nitroaromatics from the environment. Moreover, the feasibility of complexation of polyphenols with metal ions in aqueous solution can improve the surface properties and capacity of the Fe3 O4 @TA NPs for adsorption and reduction of other ions [65,85,86]. References
Fig. 9. Recyclability of Fe3 O4 @TA/Pd.
Fig. 10. Reusability of catalyst for Suzuki–Miyaura coupling reaction.
ing of the catalyst was low. In order to regenerate the catalyst, after each cycle, it was separated by a magnet and reused. In continuation of our works, a blank test by Fe3 O4 @Pd was performed under same condition for this reaction coupling. Interestingly, we have found that the catalyst was not recoverable after two reaction cycles. So, these results shown that the tannic acids are responsible for stabilization of palladium nanoparticles. 4. Conclusions Magnetite (Fe3 O4 ) nanoparticles were coated with tannic acid (TA) to give a novel adsorbent as reducing and capping/stabilizing magnetic agent to deposition of palladium nanoparticles. Inductively coupled plasma (ICP), transmission electron microscopy (TEM), scanning electron microscopy (SEM), X-ray photoelectron spectroscopy (XPS), energy-dispersive X-ray spectrum (EDS), vibrating sample magnetometer (VSM) and Fourier transform infrared (FTIR) studies have been used to characterize the prepared nanocatalyst. We have established that the ligand-free heterogeneous catalyst Fe3 O4 @TA/Pd is a highly efficient and stable catalyst for the synthesis of biaryls by Suzuki–Miyaura cross-coupling reaction at mild conditions. Also, the catalyst was used for the reduction of 4-NP to 4-AP in water at room temperature. Notably, the recyclability of the catalyst by an external magnet, with no palladium
[1] N. Miyaura, K. Yamada, A. Suzuki, Tetrahedron Lett. 36 (1979) 3437–3440. [2] N. Miyaura, A. Suzuki, Chem. Rev. 95 (1995) 2457–2483. [3] A. de Meijere, F. Diederich, Metal-Catalyzed Cross-Coupling Reaction, Wiley-VCH, New York, 2004. [4] R.H. Thomson, The Chemistry of Natural Products, Blackie and Son, Glasgow, 1985. [5] O. Baudoin, M. Cesario, D. Guenard, F. Gueritte, J. Org. Chem. 67 (2002) 1199–1207. [6] S. Kotha, K. Lahiri, D. Kashinath, Tetrahedron 58 (2002) 9633–9695. [7] J. Albaneze-Walker, J.A. Murry, A. Soheili, S. Ceglia, S.A. Springfield, C. Bazaral, P.G. Dormer, D.L. Hughes, Tetrahedron 61 (2005) 6330–6336. [8] M. Kertesz, C.H. Choi, S. Yang, Chem. Rev. 105 (2005) 3448–3481. [9] A.D. Littke, C. Dai, G.C. Fu, J. Am. Chem. Soc. 122 (2000) 4020–4028. [10] J. Yin, M.P. Rainka, X.-X. Zhang, S.L. Buchwald, J. Am. Chem. Soc. 124 (2002) 1162–1163. [11] R. Martin, S.L. Buchwald, Acc. Chem. Res. 41 (2008) 1461–1473. [12] O. Navarro, R.A. Kelly III, S.P. Nolan, J. Am. Chem. Soc. 125 (2003) 16194–16195. [13] J. Hagen, Industrial Catalysis: A Practical Approach, 2nd ed., Wiley-VCH, Weinheim, 2006. [14] L. Yin, J. Liebscher, Chem. Rev. 133 (2007) 133–173. [15] D. Astruc, F. Lu, J.R. Aranzaes, Angew. Chem. Int. Ed. 44 (2005) 7852–7872. [16] W. Han, C. Liu, Z.-L. Jin, Org. Lett. 9 (2007) 4005–4007. [17] T.M. Razler, Y. Hsiao, F. Qian, R. Fu, R.K. Khan, W. Doubleday, J. Org. Chem. 74 (2009) 1381–1384. [18] M. Pagliaro, V. Pandarus, F. Bland, R. Ciriminna, G. Palmisano, P. Demma Cara, Catal. Sci. Technol. 1 (2011) 736–739. [19] C.J. Pink, H.-T. Wong, F.C. Ferreira, A.G. Livingston, Org. Process Res. Dev. 12 (2008) 589–595. [20] Y. Yabe, Y. Sawama, Y. Monguch, H. Sajiki, Catal. Sci. Technol. 4 (2014) 260–271. [21] M. Mora, C. Jimnez-Sanchidrin, J.R. Ruiz, Curr. Org. Chem. 16 (2012) 1128–1150. [22] L. Yin, J. Liebscher, Chem. Rev. 107 (2007) 133–173. [23] M. Pagliaro, V. Pandarus, R. Ciriminna, F. Bland, P. Demma Car, Chem. Cat. Chem. 4 (2012) 432–445. [24] V. Polshettiwar, A. Molnr, Tetrahedron 63 (2007) 6949–6976. [25] A. Fihri, D. Cha, M. Bouhrara, N. Almana, V. Polshettiwar, Chem. Sus. Chem. 5 (2012) 85–89. [26] S. Navaln, M. lvaro, H. Garca, Chem. Cat. Chem. 5 (2013) 3460–3480. [27] Y. Gong, M. Li, H. Li, Y. Wang, Green Chem. 17 (2015) 715–736. [28] B. Karimi, H. Behzadnia, H. Vali, Chem. Cat. Chem. 6 (2014) 745–748. [29] R. Ali Molla, M.A. Iqubal, K. Ghosh, A.S. Roy Kamaluddin, S.M. Islam, RSC Adv. 4 (2014) 48177–48190. [30] P. Puthiaraj, K. Pitchumani, Green Chem. 16 (2014) 4223–4233. [31] H. Veisi, A. Amini Manesh, N. Eivazi, A.R. Faraji, RSC Adv. 5 (2015) 20098–20107. [32] F. Bellina, C. Chiappe, Molecules 15 (2010) 2211–2245. [33] S.B. Park, H. Alper, Chem. Commun. (2004) 1306–1307. [34] N. Iranpoor, H. Firouzabadi, Y. Ahmadi, Eur. J. Org. Chem. (2012) 305–311. [35] P. Nehra, B. Khungar, K. Pericherla, S.C. Sivasubramaniana, A. Kumar, Green Chem. 16 (2014) 4266–4271. [36] A. Ghaderi, M. Gholinejad, H. Firouzabadi, Curr. Org. Chem. 8 (2016) 441–447. [37] A. Hassine, M. Bouhrara, S. Sebti, A. Solhy, R. Mahfouz, D. Luart, C. Len, A. Fihri, Curr. Org. Chem. 18 (2014) 3141–3148. [38] B. Karimi, F. Mansouri, H. Mirzaei, Chem. Cat. Chem. 7 (2015) 1736–1789.
Please cite this article in press as: H. Veisi, et al., Magnetite nanoparticles coated with tannic acid as a green reductant and stabilizer sorbent for palladium ions: Synthetic application of Fe3 O4 @TA-Pd NPs as magnetically separable and reusable nanocatalyst for reduction of 4-nitrophenol and Suzuki reactions, Catal. Today (2017), http://dx.doi.org/10.1016/j.cattod.2017.02.023
G Model CATTOD-10612; No. of Pages 8 8
ARTICLE IN PRESS H. Veisi et al. / Catalysis Today xxx (2017) xxx–xxx
[39] (a) H. Veisi, J. Gholami, H. Ueda, P. Mohammadi, M. Noroozi, J. Mol. Catal. A: Chem. 396 (2015) 216; (b) H. Veisi, A. Sedrpoushan, B. Maleki, M. Hekmati, M. Heidariand, S. Hemmati, Apply. Organomet. Chem. 29 (2015) 834; (c) H. Veisi, P. Mohammadi, J. Gholami, Appl. Organomet. Chem. 28 (2014) 868; (d) M. Pirhayati, H. Veisi, A. Kakanejadifard, RSC Adv. 6 (2016) 27252; (e) B. Abbas Khakiani, K. Pourshamsian, H. Veisi, Appl. Organometal. Chem. 29 (2015) 259. [40] S. Verma, D. Verma, A.K. Sinha, S.L. Jain, Appl. Catal. A 489 (2015) 17–23. [41] Z. Yinghuai, S.C. Peng, A. Emi, S.Z. Monalisa, R.A. Kemp, Adv. Synth. Catal. 349 (2007) 1917–1922. [42] C.G. Tan, R.N. Grass, Chem. Commun. (2008) 4297–4299. [43] S. Shylesh, L. Wang, W.R. Thiel, Adv. Synth. Catal. 352 (2010) 425–432. [44] S. Ko, J. Jang, Angew. Chem. 45 (2006) 7564–7567. [45] V. Polshettiwar, R. Luque, A. Fihri, H. Zhu, M. Bouhrara, J.-M. Basset, Chem. Rev. 111 (2011) 3036–3075. [46] M. Tristany, J. Courmarcel, P. Dieudonne, M. Moreno-Manas, R. Pleixats, M. Rimola, S. Sodupe, S. Villarroya, Chem. Mater. 18 (2006) 716–722. [47] Y. Xiong, J. Chen, B. Wiley, Y. Xia, Y. Yin, Z.-Y. Li, Nano Lett. 5 (2005) 1237–1242. [48] A. Sartre, M. Phaner, L. Porte, G.N. Sauvion, Appl. Surf. Sci. 70 (1993) 402–406. [49] S.U. Son, Y. Jang, K.Y. Yoon, E. Kang, T. Hyeon, Nano Lett. 4 (2004) 1147–1151. [50] S.W. Kim, J. Park, Y. Jang, Y. Chung, S. Hwang, T. Hyeon, Y.W. Kim, Nano Lett. 3 (2003) 1289–1291. [51] Y. Xiong, J. Chen, B. Wiley, Y. Xia, J. Am. Chem. Soc. 127 (2005) 7332–7333. [52] N. Karousis, G.E. Tsotsou, F. Evangelista, P. Rudolf, N. Ragoussis, N. Tagmatarchis, J. Phys. Chem. C 112 (2008) 13463–13469. [53] K. Patel, S. Kapoor, D. Purshottam, T. Mukherjee, J. Chem. Sci. 117 (2005) 311–316. [54] P. Pallavicini, G. Chirico, M. Collini, G. Dacarro, A. Dona, L. D’Alfonso, A. Falqui, Y. Diaz-Fernandez, S. Freddi, B. Garofalo, A. Genovese, L. Sironi, A. Taglietti, Chem. Commun. 47 (2011) 1315–1317. [55] D. de Luna Martins, H.M. Alvarez, L.C.S. Aguiar, Tetrahedron Lett. 51 (2010) 6814–6817. [56] M.A. Albiter, R. Morales, F. Zaera, Appl. Catal. A 391 (2011) 386–393. [57] B. Keita, I.-M. Mbomekalle, L. Nadjo, C. Haut, Electrochem. Commun. 6 (2004) 978–983. [58] G. Maayan, R. Neumann, Catal. Lett. 123 (2008) 41–45. [59] M.N. Nadagouda, R.S. Varma, Green Chem. 10 (2008) 859–862. [60] P. Mohanpuria, N.K. Rana, S.K. Yadav, J. Nanopart. Res. 10 (2008) 507–517. [61] M. Gericke, A. Pinches, Hydrometallurgy 83 (2006) 132–140. [62] E.S. Abdel-Halim, M.H. El-Rafie, S.S. Al-Deyab, Carbohydr. Polym. 85 (2011) 692–697. [63] L. Jia, Q. Zhang, Nanotechnology 20 (2009) 385601. [64] X. Yang, Q. Li, H. Wang, J. Huang, L. Lin, W. Wang, et al., J. Nanopart. Res. 12 (2009) 589. [65] K.T. Chung, T.Y. Wong, C.I. Wei, Y.W. Huang, Y. Lin, Crit. Rev. Food Sci. Nutr. 38 (1998) 421–464.
[66] A. King, G. Young, J. Am. Diet. Assoc. 99 (1999) 213–218. [67] G.K.B. Lopes, H.M. Schulman, M. Hermes-Lima, Biochim. Biophys. Acta 1472 (1999) 142–152. [68] L.R. Ferguson, Mutat. Res. 75 (2001) 89–111. [69] L.T. Wu, C.C. Chu, J.G. Chung, C.H. Chen, L.S. Hsu, J.K. Liu, S.C. Chen, Mutat. Res. 556 (2004) 75–82. [70] R.G. Andrade, L.T. Dalvi, J.M.C. Silva, G.K.B. Lopes, A. Alonso, M. Hermes-Lima, Arch. Biochem. Biophys. 437 (2005) 1–9. [71] K. Horikawa, T. Mohri, Y. Tanaka, H. Tokiwa, Mutagenesis 9 (1994) 523–526. [72] S.C. Chen, K.T. Chung, Food Chem. Toxicol. 38 (2000) 1–5. [73] M. Athar, W.A. Khan, H. Mukhtar, Cancer Res. 49 (1989) 5784–5788. [74] H.U. Gali, E.M. Perchellet, D.S. Klish, J.M. Johnson, J.P. Perchellet, Int. J. Cancer 51 (1992) 425–432. [75] C. Nepka, E. Sivridis, O. Antonoglou, A. Kortsaris, A. Georgellis, I. Taitzoglou, P. Hytiroglou, C. Papadimitrou, I. Zintzaras, D. Kouretas, Cancer Lett. 141 (1999) 57–62. [76] http://www.epa.gov/ttn/atw/hlthef/nitrophe.html. [77] W. Zhang, X. Xiao, T. An, Z. Song, J. Fu, G. Sheng, M. Cui, Kinetics, J. Chem. Technol. Biotechnol. 78 (2003) 788–794. [78] N. Mand, D.G. Crosby, J. Agric. Food Chem. 22 (1974) 849–853. [79] Y.-T. Woo, D.Y. Lai, Aromatic Amino and Nitro-Amino Compounds and Their Halogenated Derivatives. Patty’s Toxicology, Wiley, New York, 2001. [80] Z.M. Wang, C.L. Xu, G.Q. Gao, X. Li, RSC Adv. 4 (2014) 13644–13651. [81] X.M. Chen, Z.X. Cai, X. Chen, M. Oyama, J. Mater. Chem. A 2 (2014) 5668–5674. [82] Z.F. Jiang, D.L. Jiang, A.M.S. Hossain, K. Qian, J.M. Xie, Z.F. Jiang, D.L. Jiang, A.M.S. Hossain, K. Qian, J.M. Xie, Phys. Chem. Chem. Phys. 17 (2015) 2550–2559. [83] V. Evangelista, B. Acosta, S. Miridonov, E. Smolentseva, S. Fuentes, A. Simakov, Appl. Catal. B 166–167 (2015) 518–528. [84] Z.P. Dong, X. Le, C.X. Dong, W. Zhang, X.L. Li, J.T. Ma, Appl. Catal. B 162 (2015) 372–380. [85] K. Aguilar-Arteaga, J.A. Rodriguez, E. Barrado, Anal. Chim. Acta 674 (2010) 157. [86] (a) MazurM, A. Barras, V. Kuncser, A. Galatanu, V. Zaitzev, K.V. Turcheniuk, P. Woisel, J. Lyskawa, W. Laure, A. Siriwardena, R. Boukherroub, S. Szunerits, Nanoscale 5 (2013) 2692; (b) M. Bagtash, Y. Yamini, E. Tahmasebi, J. Zolgharnein, Z. Dalirnasab, Microchim Acta 183 (1) (2016) 449–456. [87] I. Kostova, I.H. Joe, S.C. Pinzaru, J. Optoelectron. Biomed. Mater. 1 (2009) 188–199. [88] S.A. Aromal, D. Philip, Physica E 44 (2012) 1692–1696. [89] M.H. Rashid, T.K. Mandal, J. Phys. Chem. C 111 (2007) 16750–16760. [90] N. Sahiner, H. Ozay, O. Ozay, N. Aktas, Appl. Catal. A: Gen. 385 (2010) 201–207. [91] N. Sahiner, H. Ozay, O. Ozay, N. Aktas, Appl. Catal. B: Environ. 101 (2010) 137–143. [92] N. Sahiner, O. Ozay, Curr. Nanosci. 8 (2012) 367–374. [93] X.F. Wu, P. Anbarasan, H. Neumann, M. Beller, Angew. Chem. Int. Ed. 49 (2010) 9047–9050.
Please cite this article in press as: H. Veisi, et al., Magnetite nanoparticles coated with tannic acid as a green reductant and stabilizer sorbent for palladium ions: Synthetic application of Fe3 O4 @TA-Pd NPs as magnetically separable and reusable nanocatalyst for reduction of 4-nitrophenol and Suzuki reactions, Catal. Today (2017), http://dx.doi.org/10.1016/j.cattod.2017.02.023