Well-dispersed rhenium nanoparticles on three ...

17 downloads 0 Views 2MB Size Report
Jun 22, 2017 - Vlasenko, I.V. Pirog, O.N. Shishilov, O.V. Popkov, Synthesis and properties of rhenium– polyethylene nanocomposite, Composites Part B 43 ...
Accepted Manuscript Well-dispersed rhenium nanoparticles on three-dimensional carbon nanostructures: Efficient catalysts for the reduction of aromatic nitro compounds Pitchaimani Veerakumar, Pounraj Thanasekaran, King-Chuen Lin, Shang-Bin Liu PII: DOI: Reference:

S0021-9797(17)30833-0 http://dx.doi.org/10.1016/j.jcis.2017.07.065 YJCIS 22594

To appear in:

Journal of Colloid and Interface Science

Received Date: Revised Date: Accepted Date:

4 May 2017 22 June 2017 17 July 2017

Please cite this article as: P. Veerakumar, P. Thanasekaran, K-C. Lin, S-B. Liu, Well-dispersed rhenium nanoparticles on three-dimensional carbon nanostructures: Efficient catalysts for the reduction of aromatic nitro compounds, Journal of Colloid and Interface Science (2017), doi: http://dx.doi.org/10.1016/j.jcis.2017.07.065

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Well-dispersed

rhenium

nanoparticles

on

three-dimensional

carbon

nanostructures: Efficient catalysts for the reduction of aromatic nitro compounds Pitchaimani Veerakumar,a,b,* Pounraj Thanasekaran,c King-Chuen Lina,b,* and Shang-Bin Liua a

Institute of Atomic and Molecular Sciences, Academia Sinica, Taipei 10617, Taiwan

b

Department of Chemistry, National Taiwan University, Taipei 10617, Taiwan

c

Institute of Chemistry, Academia Sinica, Taipei, 11529 Taiwan

Abstract: Rhenium nanoparticles (ReNPs) supported on ordered mesoporous carbon (OMC) as a catalyst (Re/OMC) through a solvent-evaporation induced self-assembly (ELSA) method were prepared. The synthesized heterogonous catalyst was fully characterized using X-ray diffraction, field

emission

thermogravimetric

transmission analysis,

electron Raman,

microscopy,

N2

Fourier-transform

sorption, infrared,

metal and

dispersion,

X-ray

photon

spectroscopies. In addition, the catalyst was applied to reduce the aromatic nitro compounds (ANCs) for the first time in aqueous media and the reactions were monitored by following the intensity changes in the UV-Vis absorption spectra with respect to time. This method provides the advantages of obtaining a high rate constant (k), green reaction conditions, simple methodology, easy separation and easy workup procedures. Moreover, the catalyst can be easily recovered by centrifugation, recycled several times and reused without any loss of activity. The higher activity of this catalyst was attributed to higher dispersion and smaller particle size of ReNPs as observed from FE-TEM and XRD results. Keywords: Catalytic reduction, Rhenium nanoparticles, Sodium borohydride, Rate constant, Toxic pollutants

*Corresponding author. E-mail addresses: [email protected] (P.Veerakumar); [email protected] (K.C. Lin).

1

1.

Introduction

Over the past decade, transition metal nanoparticles (MNPs) supported catalysts have received wide attention as heterogeneous catalysts due to their environmental benignity, low cost, structural integrity and high catalytic activities in various reactions [1-4]. By taking advantages of extraordinary properties of these nanomaterials including large specific surface area, high electrical conductivity and superior strength, the assembled 3D porous structures offer excellent opportunities for building high performance supercapacitors and electrochemical sensors [5]. Nevertheless, the catalytic activity of rhenium (Re) based nanomaterials used in the catalyst is limited compared to other metal catalysts [6-10]. Extensive effort has thus been devoted herein to synthesizing the carbon supported rhenium nanoparticles (ReNPs) and studied their catalytic activities. Several new techniques including pulsed-laser decomposition [11], electrodeposition [12], gamma irradiation [13], chemical vapour deposition (CVD) [14], wet chemical reduction [15,16], solvothermal [17], colloidal and microemulsion and other methods [18-20] were adopted to synthesize ReNPs materials. Apart from rhenium nanoparticles, other catalytic active MNPs, such as RuNPs [21], OsNPs [22,23], IrNPs [24], AuNPs [25] and, AgNPs [26] were also used as catalysts for the reduction of aromatic nitro compounds (ANCs). Rhenium metals are the most promising catalytic materials that could be used as versatile catalysts. Apart from bare metals, ReNPs may serve as catalysts while they usually supported on polymer [27], DNA scaffold [28], carbon [29-31], and γ- Al2O3 [32], etc. Ordered mesoporous carbon materials (OMCs) as a support for transition metal catalysts [33,34] offer several advantages including easy modification, elimination of nanoparticle agglomeration, relatively low cost, high mechanical strength, chemical stability, and a pore structure along with an attractive surface chemistry. In addition, these nanocatalysts can make the products easily removable and the catalysts recyclable [35]. Over the last several years, numerous review articles have specifically addressed the toxicity and mutagenicity of ANCs. For example, the remediation of ANCs is of interest because it is a mutagen, anaemia and skin irritation [36]. Moreover, the reduced intermediates and products (such as amino-, and hydroxylamino-ANCs) often have greater toxicity and/or aqueous solubility than their parent compounds. Its ingestion can cause various health problems including eye, liver 2

or kidney damages, aplastic anaemia, cyanosis, respiratory organs, as well as damage to neurological system [37]. Some ANCs are high-energy explosives. For example, 2,4,6trinitrotoluene (TNT), 2,4,6-trinitrophenol (TNP) or picric acid are the primary constituents that were adopted in many unexploded landmines worldwide [38]. Therefore, it is necessary to remove such highly toxic pollutants. As the reduced products of ANCs, aromatic amino compounds (AACs) have attracted enormous attention in the chemical industry. They can also be used to synthesize drugs, pesticides, and explosives. In addition, these are key intermediates in many important fine chemicals, agrochemicals, pharmaceuticals, and polymer building blocks [39]. In the course of our studies on the development of heterogeneous catalysts, we report the potentiality of the Re/OMC catalyst that has been tested for the first time as a catalyst for the reduction of ANCs at room temperature, as illustrated in Scheme 1. In addition to the good catalytic properties, the synthesized catalyst could be easily separated from the reaction mixture and reused six times without loss of activity, indicative of its potential application in chemical industries.

Scheme 1. Illustrated synthesis of Re/OMC nanocomposite from phloroglucinol-formaldehyde resin and its morphology and catalytic application.

2. Results and Discussion 2.1. Structural properties 3

A highly mesoporous 3D structure of OMC was prepared via pyrolysis of phloroglucinolformaldehyde resol by evaporation induced self-assembly (EISA) strategy (see Supporting Information, hereafter denoted as SI). Decomposition of dirhenium decacarbonyl (Re2(CO)10, 217 mg, 0.33 mmol) used as a metal precursor may produce small ReNPs, which are embedded on the interior pore walls of the 3D carbon matrix (Scheme 1). Unlike previous methods adopted to synthesize on mesoporous carbon [8] and MWCNTs [29], the rhenium precursor Re2(CO)10 decomposes to the ReNPs upon microwave irradiation as shown in equation (1).

Structural characterization based on X-ray diffraction pattern reveals that OMC showed two strong peaks appeared at 2θ = 24.6° and 44.6°, which can be attributed to the amorphous graphitic carbon (002) and (100), respectively (Fig.1a) [40]. The related XRD pattern of the Re/OMC catalyst showed the peaks at 2θ = 37.7°, 40.6°, 43.0°, 56.6°, 68.4°, 75.3, and 82.5°, confirming the formation of hexagonal rhenium. These results indicate that Re/OMC catalyst possessed a very small size  2–5 nm of Re particle with a hexagonal closely-packed (hcp) structure of the bulk Re, which is in good agreement with previously published reports [11,40].

4

Fig. 1. (a) XRD patterns, (b) Raman spectra, (c) N2 sorption isotherms and their pore size distributions (inset), and (d) TGA and DTG curves of OMC and Re/OMC catalyst.

This fact verifies that ReNPs have been successfully loaded on to the three dimensional OMC material with rather good dispersion of the metallic phase. No other peaks have been observed; it is indicated that the rhenium dispersions have a highly textured, nanocrystalline structure with preferential growth orientation [20]. The crystallite size of Re in Re/OMC sample was determined according to the Scherrer formula (2):

where

is the wavelength,

is the full width at half maximum (FWHM) of the Bragg’s peak

corrected using the corresponding peak in sample, and

is the Bragg’s angle.

Besides, pristine OMC and Re/OMC samples were also characterized by Raman spectroscopy (Fig. 1b). Both samples show the D and G bands at 1351 and 1598 cm1, respectively, which are in agreement with those reported [41]. Moreover, we observed an additional peak at  2742 cm−1 for the 2D and 2902 cm−1 for the D+G band, implying that the

Table 1. Physical properties of the as-prepared OMC and Re/OMC materials. Surface areaa (m2 g1) Sample

Pore volumea (cm3 g1)

Dpc

Dmd

ID/IG

STotal

SMicrob

SMesob

VTotal

VMicrob

(nm)

(%)

OMC

1468

344.1

1123.9

0.1499

1.0207

6.19



0.99

Re/OMC

1386

275.3

1110.7

0.1103

0.8883

4.95

0.22

0.99

a

SBET surface area from BET method, and total pore volume (VTot) calculated at P/P0 = 0.99. Microporous surface areas (SMicro) and pore volumes (VMicro) obtained from t-plot analysis. c Average pore size (Dp) derived by the BJH method using the adsorption branch of the isotherm. dMetal dispersion measured by H2 chemisorption at 323 K. b

5

ReNPs are successfully imbedded inside the OMC matrix. Given the N2 adsorption/desorption isotherm measurements, both OMC and Re/OMC samples exhibited type-IV isotherm (Fig. 1c) with type-H1 hysteresis loop, indicating the existence of mesoporosity [42]. This result indicated notable decreases in the surface area and pore volume of the Re/OMC sample upon loading of ReNPs onto the 3D OMC matrix. Note that a substantial fraction of small mesopores in OMC sample was blocked by ReNPs as the pore diameter of OMC decreased. The textural properties of as-prepared catalyst are summarized in Table 1. Thermogravimetric analysis (TGA; Fig. 1d) of as prepared samples showed the initial weight loss of water around ca.11.6 (OMC) and 13.1 wt% (Re/OMC) at about 110 °C. Then, the nanocomposite became stable until it reached to 510 °C, after which a weight loss ca. 75.7 wt% (Re/OMC) at 640 °C and 88.4 wt% (OMC) at ca. 680 °C was observed. These weight losses are probably caused by the decomposition of the OMC framework and the presence of rhenium oxides [33]. The morphology of these samples was observed by field-emission transmission electron microscopy (FE-TEM) analysis, which showed that the highly well-ordered 3D hexagonal carbon frameworks (Fig. 2a–c) were fully loaded with well-dispersed ReNPs with an average size range of 5  0.2 nm (Fig. 2d–h). This result was also confirmed by H2 chemisorption study. A single ReNP with a d-spacing of 0.377 nm corresponding to the (101) plane was also shown in Fig. 2h. In the electron diffraction (SAED) pattern, the diffraction spots of Re observed from Re/OMC catalyst (Fig. 2i) revealed a mean particle size of 5  0.2 nm.

6

Fig. 2. FE-TEM images of (a-c) bare OMC, (d-g) Re/OMC nanocatalysts, (h) single Re NP, and (i) their SAED pattern. This result is consistent with the view of FE-TEM, showing that the highly dispersed ReNPs are successfully fixed into the interior part of the OMC walls. Energy dispersive X-ray spectroscopy (EDX) analysis of the Re/OMC catalyst verified the existence of rhenium, carbon, oxygen and copper elements (See Fig. S1, SI). The peaks of elemental Cu at about 1 and 8 keV corresponding to the copper grid were excluded during elemental analysis. The weight % of Re, C, and O was confirmed in the range of 32.94, 62.76, 4.30, respectively, as expected with a previously published report [11]. The surface properties of as-prepared materials were further examined by X-ray photoelectron spectroscopy (XPS), as shown in Fig. 3a. The intense Re 4f peaks from the Re loading sample were displayed in Fig. 3b. 7

Fig. 3. XPS survey spectra of (a) OMC and Re/OMC, (b) expanded XPS profiles of Re 4f, (c) C 1s, and (d) O 1s regions of the Re/OMC catalyst. In general, rhenium exists in five positive oxidation states, Re0, Re3+, Re4+, Re6+ and Re7+[43]. The XPS survey spectrum of Re/OMC showed that it exhibits two well-resolved spectral lines at 40.7 and 42.9 eV, which are assigned to the Re 4f7/2 and Re 4f5/2 spin-orbit components, respectively. These results imply that the metallic Re0 was thermally reduced from the Re2(CO)10 under inert conditions. The high resolution C1s spectrum (Fig. 3c) can be deconvoluted into three peaks located at 284.7, 285.8, and 288.7 eV, which are ascribed to C–C, C–O, and –CC=O, respectively. The oxygen O 1s spin−orbits yield the characteristic peaks with binding energy (BE) at 531.2 (C–O), 532.8 (C=O) and 533.2 eV (C-O-C/C-OH), as shown in Fig. 3d. Moreover, the Re 4f7/2 BE (40.7 eV) corresponding to the metallic rhenium in the Re/OMC sample is similar to that determined previously for rhenium nanopowder (40.5 eV) [44].

8

2.2. H2-TPR analysis The TPR profiles of as-prepared bare OMC and Re/OMC catalysts are shown in Fig. 4. It is an ideal tool for examining the reducibility, and the interaction between the metal species and support, which provides the information regarding the active state of the solid catalysts [8,45]. Mitra et al. reported that consumption of H2 over the bare carbon catalyst showed a broad maximum peak at the range of 500–800 °C, accompanied by methane emission and the hydrogenation of surface functional groups on the carbon support [46].

Fig. 4. H2-TPR profiles of (a) pristine OMC, and (b) Re/OMC catalysts.

A similar phenomenon was observed for OMC, whereas the reduction profile of Re/OMC consisted of two peaks. Obviously, at a low temperature, the peaks at approximately 320 °C was caused by the reduction of small rhenium oxide species, while gasification/methane (CH4) formation on the carbon support generated a broader peak centred at ~600 °C [47]. 2.3.

FT-IR analysis The FT-IR spectra of Re2(CO)10, polymeric resin containing Re2(CO)10, and OMC supported

rhenium catalysts (Re/OMC) are shown in (Fig. S2, SI) before and after carbonization. The pristine Re2(CO)10 exhibited C=O horizontal and C=O vertical stretching vibrations at 2071, and 1978 cm−1, and 2011 cm−1, respectively. The Re–C stretching vibration band in Re precursor of 9

Re2(CO)10 was observed at 590 cm−1 [48]. The intensities of these characteristic peaks mostly disappeared, when carbonized at high temperature, indicating that the Re 2(CO)10 was completely decomposed into metallic Re. This result is consistent with the previous reports [49], and also indicates the strong interaction as well as active involvement during incorporation of ReNPs onto the 3D carbon matrix. 2.4. Catalytic studies The catalytic behavior of these samples was studied in the reduction reaction among the following compounds: nitrobenzene (NB), 4-nitrophenol (4-NP), 2-nitroaniline (2-NA), 4nitroaniline (4-NA), 2,4-di-nitrophenol (2,4-DNP) and 2,4,6-tri-nitrophenol (2,4,6-TNP)) in aqueous medium using sodium borohydride (NaBH4) as a hydrogen generator. Currently, general conversion of a nitro group into an amino group is mostly achieved by treatment with various reducing reagents, such as NaBH4 [33,50,51], H2 and CO gas [52], N2H4 [53], ammonia-borane (NH3BH3) [54], 1,1,3,3-tetramethyl disiloxane (TMDS) [55], and triethylsilicon hydride (Et3SiH) [56], under acidic/basic/neutral conditions. Among the various reducing agents, NaBH4 is the most commonly used as a mild, selective reducing agent, which is adopted to be the best economical option for the reduction reactions under normal pressure and room temperature in the presence of transition metal catalysts, although preparation is an intricate procedure (high temperature and pressure were always required) in the case of silicon-based reagents. In some studies, a relatively high H2 pressure was required, but might lead to explosion if careless during experiments. It is well known that the final products (AACs) are very useful and important industrial intermediate, which are applicable in many fields including analgesic and antipyretic drugs, photographic developer, corrosion inhibitor, anticorrosion-lubricant, and manufacture of various dyes and polymers, etc., [57]. To optimize the reaction conditions, the reduction of 4-NP was first carried out as a model reaction in the presence of different source of catalysts under various conditions with the aim of optimizing the yield, and the results are summarized in Table 2. Recently, various functional groups containing bare carbon materials are employed as metal free catalyst in catalytic reduction reactions [58]. In order for optimization, the catalytic performances during reduction of 4-NP over OMC without ReNPs involved were also investigated. The bare OMC showed less catalytic activity that resulted in trace yield over a 10

reaction time of 20 s (Table 2, entry 1). By comparison, the same reaction was carried out using the rhenium precursor Re2(CO)10 as the catalyst, but no reaction was observed. Table 2. Optimization of the reaction conditions for the reduction of 4-NPa Entry

Catalyst

Amount (mg)

Time (s)

k (s1)

Conversion (%)b

1

OMC

0.02

20

-

trace

2

Re2(CO)10

0.02

20

-

NRc

3

ReO3

0.02

20

0.1148

68

4

Re/OMC

0.01

20

0.1375

95

5

Re/OMC

0.02

20

0.1498

99

6

Re/OMC

0.03

20

0.1500

99

7

Re/OMC

0.05

20

0.1502

99

8

Re/OMC

0.02

30

0.1499

99

9

Re/OMC

0.05

30

0.1508

99

a

Reaction conditions: 4-NP (1.8 mL, 0.1 mM), NaBH4 (0.5 mL, 0.1 M) and water (3 mL). bGC conversion on the basis of 4-nitrophenol. cNo reaction.

Generally, metal carbonyls are soluble in nonpolar and polar organic solvents but insoluble in water [59], and hence, they do not cause the reduction reaction in aqueous medium (Table 2, entry 2). The commercially available Re2O3 catalyst gave a low yield of reduction product, (typically ca. 68%) as shown in (Table 2, entry 3). In contrast, the current Re/OMC catalyst shows a much better reactivity, and gives an excellent product yield as shown in (Table 2, entry 4-9). While increasing the amount of Re/OMC catalyst from 0.01 to 0.05 mg, a little effect on the reduction efficiency was found (Table 2, entries 4-7). Upon increasing the reaction time (30 s) under the same condition, there are no appreciable changes in the product yields as compared to 20 s (Table 2, entries 7 and 8). The catalytic reduction of NB could be apparently considered as a pseudo-first order reaction with respect to the concentration of NaBH4. Thus, the kinetic equation of this catalytic reaction is readily solved and the reaction rate constant (k) was obtained from the slope of the plots [10]. As the concentration of NaBH4 is higher than that of 4-NP, the reduction rate is found to be independent of NaBH4 concentration, and the reaction could be considered as a pseudo-first order with respect to the concentration of 4-NP. It is worth noting 11

that our protocol catalytic conditions are significantly milder (0.1M of NaBH4 and room temperature) than those used in other studies, in which similar reduction reactions were examined; however, either large excess of NaBH4 (100-fold excess) [60] or hast conditions (>100 C) are required [61]. Since ReNPs are more sensitive to oxygen, the ReNPs are covered with a thin layer of ReO3. Although partial oxidation of the ReNPs on the surface of ReO3 particle may have a strong influence on the catalytic efficiency to some extent, the catalyst can be reused for six cycles still with a high activity. 2.5.Reduction of nitrobenzene Recently, considerable interest has been focused on MNPs supported materials as heterogeneous catalysts for the reduction of NB system and for evaluation of the catalytic activity [51]. The development of an efficient and reliable technique for the catalytic reduction of aromatic nitro compounds is a demanding task. In this system, the nitro compounds act as an electronic relay system, in which the electron transfer takes place from donor BH4− to acceptor nitro (-NO2) groups. We explored the catalytic activity of our synthesized Re/OMC catalyst using NaBH4 as a reducing agent. Consequently, the progress of the reduction reaction can be readily followed by means of UV−vis absorption spectrophotometry (Fig. 5a). Upon the reduction of NB using Re/OMC catalyst in the presence of NaBH4, the absorption spectra of NB showed that the main peak at

max

= 275 nm disappeared gradually while the peak at

max

= 232

nm of aniline (AN) was slowly increased (indicated by arrows) within a period 25 s, indicating a successive reduction of NB to AN [62]. As shown in Fig. 5a, two isobestic points at

max

= 216

and 241 nm are clearly visible. Obviously, the appearance of isobestic point at 241 nm indicates the formation of aniline peak without any side reactions [63]. A good linear correlation (R2 > 0.9995) of ln (At/A0) vs. reduction time was observed, confirming the pseudo first-order kinetics. The rate constant (k) is calculated as 0.1482 s1, and the corresponding kinetic plot is given in Fig. 5a (inset). The reduction reaction did not proceed in the absence of catalyst as well as in the absence of reductant. We selected a few important studies from the literature [51,62-64] for the purpose of comparison with our results. These findings are summarized together with our data (Table S1, SI). In particular, Guan et al [64] have reported multifunctional core–shell catalysts

12

(Au@C yolk–shell) for the catalytic reduction of NB, yielding k = 2.8 × 10−3 (12 min), which is significantly lower than our catalytic system. 2.6. Reduction of 4-nitrophenol The reduction of phenolic aromatic nitro compounds with noble metal catalysts in the presence of NaBH4 has been extensively investigated for the efficient production of aromatic amine compounds [65]. Inspired by excellent activity of Re/OMC for the catalytic reduction of NB, we have extended the catalytic activity measurement for another five model ANCs, namely 4-NP, 2-NA, 4-NA, 2,4-DNP and 2,4,6-TNP under identical experimental conditions. The catalytic activity of Re/OMC catalyst was tested for the reduction of 4-NP. In this case, the 4-NP solution exhibits a distinct absorption maximum at max

max

= 380 nm; however, the peak shifted to

= 400 nm, immediately after the addition of NaBH4 solution as shown in Fig. 5b. The

resulting colour change from light to bright yellow was due to the formation of 4-nitrophenolate ion in alkaline medium (Fig. S3, SI). The maximum absorption peak at

max

= 400 nm did not

change over time, even after NaBH4 was added enough to generate small bubbles from NaBH4 upon reducing water to H2 gas according to the reaction (3):

The reduction of 4-NP (E0 = −0.76 V) by NaBH4 (E0 = −1.33 V) is thermodynamically feasible, but kinetically restricted in the absence of a catalyst in aqueous NaBH4 (Fig. S3, SI) [66]. The result confirms that the reduction reaction did not proceed in aqueous NaBH4 solution alone. Furthermore, the addition of catalyst to reaction mixture led to the decrease in the peak intensity of 4-nitrophenolate ion at

max

= 400 nm with a concomitant appearance of a new peak at

max

=

300 nm (Fig. 5b), revealing that the new peak was ascribed to 4-aminophenol (4-AP), and the corresponding kinetic plot is given in Fig. 5b (inset). Notably, the spectra have well-defined isobestic points, indicative of the sole product formation without any by-product generation. The complete conversion of 4-NA can also be confirmed by the colour change of the solution from originally bright yellow to colourless (inset of Fig. S3, SI). Indeed, the 4-nitrophenolate molecules have an active –OH group (activating character), which remained unchanged, but the NO2 group was totally reduced to amine (4-AP). Given the rate constant estimate [33], the Re/OMC gives the largest value, ca. k = 0.1498 s−1 as compared 13

with other carbon matrix-supported AgNPs, which yields k = 5.33 × 103 s1 (see Table S2, SI). With increasing the catalyst amount (from 0.01mg to 0.05 mg), conversion of 4-NP increased from 95 to 99% as listed in Table 2. It is noteworthy that the synthesized Re/OMCs with an optimal k value of 0.1498 s−1 for the catalytic reduction of 4-NP, indeed, show a better catalytic performances surpassing most of the metal nanoparticles, such as Ru0 nanochains (k = 2.8 × 10−1 min−1) [21], Re nanoclusters (k = 0.06 min−1) [28], and bimetallic Pt–Ni NPs (k = 11.6 × 10−2 min−1) [50] catalysts. In addition, the reduction rate constant using Re/OMC catalyst is also much higher than previously reported carbon based Au@CMK-3-O catalyst (k = 0.132 min1), and Ag doped carbon spheres (k = 2.02  10−2 s−1) [47,67].These results indicate that Re/OMC is more active than the other mono/bimetallic nanomaterials.

14

Fig. 5. UV-vis absorption spectra for the successive reduction of (a) NB, (b) 4-NP, (c) 2-NA, (d) 4-NA, (e) 2,4-DNP and (f) 2,4,6-TNP using NaBH4 in the presence of Re/OMC at different time intervals and their corresponding insets show the plot of ln At/A0 vs. time.

2.7. Reduction of 2-nitroaniline and 4-nitroaniline Apart from the above reactions, the catalytic activity of Re/OMC was also tested for the reduction of 2-NA and 4-NA under the same conditions, and their results are displayed in Figs. 5c and 5d, respectively. In the presence of catalyst, the absorption peaks of 2-NA at and 400 nm [68] gradually decreases but the peak at max

max

max

= 280

= 280 nm is found to be red-shifted to

= 290 nm with reaction time, which indicates that the reduction of 2-NA to ortho-

phenylenediamine (o-PDA) appeared with time. The yellow coloured solution became colorless after the completion of the reaction (Fig. 5c). When 4-NA is reduced by NaBH4 alone, no appreciable change in the absorption spectrum was observed even after standing for 30 s. It indicates that the use of BH4- ions alone is incapable of reducing 4-NA to p-PDA (see Fig. S4, SI). After the addition of a small amount (0.02 mg) of the Re/OMC catalyst, the absorption peak of 4-NP at

= 380 nm significantly decreases while the peak at

max

= 235 nm gradually

increases as the reaction proceeds. Meanwhile a new peak appears at

max

= 308 nm, revealing

max

that the reduction of 4-NA leads to form p-PDA (Fig. 5d) [69]. The time-dependent UV spectra for the successive reduction of 2-NA and 4-NA and their corresponding kinetic plots are given in Figs 5c and 5d (Insets), respectively. The UV-vis spectra of 4-NA show the presence of an isobestic point at

max

= 263 nm, suggesting that the catalytic reduction of 4-NA gives the

product p-PDA only. This reduction reaction proceeds via nitrophenolate ion intermediate in NPs or via hydroxylamine in 4-NA [70]. A linear relationship between ln(At/A0) and reaction time is obtained in the 2-NA and 4-NA reduction by Re/OMC catalyst (Fig. 5c and Fig. 5d). The plots for 2-NA and 4-NA well matched with the pseudo first-order reaction, and their rate constant k were calculated to be 0.1454 and 0.1472 s−1, respectively. In contrast, the rate constant of RuNPs [21] catalysed 2-NA reduction was calculated to be 3.4 × 10−2 min1, which is smaller than that of present catalyst. In another study, Cu/Co3O4 catalyst was reported to require a longer time (21 min) for the achievement of complete reduction of 4-NA (k = 17.1 × 10−2 min1) [69]. Likewise, the reduction rate constant of 4-NA with other monometallic or bimetallic NPs catalysts was 15

compared. For instance, the rate constants k was evaluated to be 34.33 (10 min) for RuNPs [21], 3.87 (14 min) for OsNPs [23], 2.083 (86 min) for AuNPs [25], 7.59 × 10−2 (42 min) for AgNPs [26], 37.4 (8 min) for ReNPs [28], and 6.4 × 10−2 min−1 (8 min) for bimetallic Pt-Ni nanoparticles [50] catalysts. The diversity of such noble metal catalysts showed a potentially great contribution in the reduction reactions. Despite, it is inconvenient to recover the noble metals from the reaction mixture and low kinetic stability, whereas, the use of OMC supports makes it easy to recover for recycling process. However, these noble metallic NPs showed a lower catalytic activity compared to our present catalyst owing to their lower surface area. Additionally, the high dispersion of noble metal catalysts on carbon supports with large surface areas, a benefit for catalytic applications [33]. Hence, we compiled the catalytic activity of Re/OMC with the other reported metal NPs based catalysts employed for the reduction of 2-NA and 4-NA (see Tables S3 and S4, SI). These results clearly suggest that the mesoporous Re catalyst possess a high catalytic activity compared to others reported. A wide variety of ANCs can be reduced to their corresponding AACs in near quantitative yields, and their results are

Table 3. Catalytic performances for reduction of various ANCs using Re/OMC catalysta Entry Reactant Product Time (s) k (s1) Conversion (%)b 1

Nitrobenzene

Aniline

25

0.1482

98

2

4-Nitrophenol

4-aminophenol

20

0.1498

100

3

2-Nitroaniline

o-phenylenediamine

25

0.1454

99

4

4-Nitroaniline

p-phenylenediamine

30

0.1472

99

5

2,4-dinitrophenol

2,4-diaminophenol

20

0.1469

98

6

2,4,6-trinitrophenol

2,4,6-triaminophenol

20

0.1442

97

a

Reaction conditions: 4-NP (1.8 mL, 0.1 mM), NaBH4 (0.5 mL, 0.1 M) and water (3 mL). bGC

conversion on the basis of ANCs. summarized in Table 3. 2.8. Reduction of 2,4-dinitrophenol and 2,4,6-trinitrophenol The reduction of toxic nitro compounds into their amino analogues are one of the most important derivatives, which are receiving commercial attention in a wide range of areas such as

16

photographic, pharmaceutical, and dye industries [71]. Various effective methods have been developed for the treatment of such toxic compounds [72,73]. However, their disposal into nearby water sources from industries cause serious environmental pollution and harm the ecosystem. In this work, we investigated the catalytic reduction of 2,4-DNP [69,70] and 2,4,6TNP [74] using the same manner, and their corresponding linear plots of the products 2,4diaminophenol (2,4-DAP) and 2,4,6-tri-aminophenol (2,4,6-TAP) were obtained by using equation (2), as shown in Fig. 5e and Fig. 5f, respectively. When 0.02 mg of the Re/OMC catalyst was added to the reaction mixture, the peak intensity of 2,4-DNP at

max

= 358 nm

decreased with a concomitant increase in the peaks corresponding to 2,4-diaminophenol (2,4DAP) at

max

= 298 nm. However, upon the addition of aqueous solution of NaBH4, the peak

maxima immediately shifted towards 445 nm (due to generation of nitrophenolate ion). The peak at

max

= 298 nm gradually increased which was attributed to the generation of DAP. The

corresponding pseudo-first-order rate constant was calculated to be k = 0.1469 s1. As for the reduction of 2,4,6-TNP with the catalyst, the yellow colour of the reaction mixture started fading initially, and the absorption intensity shifted towards at

max

intermediate) followed by the appearance of new peak at

= 390 nm (nitrophenolate ion as max

= 305 nm, representing the

formation of 2,4,6-tri-aminophenol (2,4,6-TAP) as shown in Fig. 5e. The reaction was completed within 20 s, giving a rate constant (k) ca. 0.1442 s1 (Fig. 5e), while comparing with MnFe2O4 (k = 0.01134 s−1) [73], and ZnO nanorods (k = 0.141 min1) [74]. The reduction reaction became quite slow without using this catalyst (Figs. S5 and S6, SI). In summary, the reduction rates of the six nitro-aromatics performed using Re/OMC catalyst followed the order: 2,4,6-NP (k = 0.1442 s1) > 2-NA (k = 0.1454 s1) > 2,4 -DNP (k = 0.1469 s1) > 4-NA (k = 0.1472 s1) NB (k = 0.1482 s1) > 4-NP (k = 0.1498 s1). The order of reactivity of ANCs can be realized by the following two factors: (i) the formation of intermediates, and (ii) their reactivity. For the formation of nitrophenolate ion, the acidic strength of nitro phenols plays an important role during the reaction. Here, NaBH4 acts as a Lewis base. The order of acidic strength of the ANCs are: 2,4,6-TNP > 2,4-DNP > 4-NA > 2-NA > NB > 4-NP, and the stability of nitrophenolate ions follow the same order. This is due to the fact that the negative charge of oxygen in nitrophenolate ion is delocalized through the aromatic ring that is stabilized via resonance. In addition, inductive effect of nitro groups further results in the facile liberation of

17

H+ ion. However, the fact that the reduction rate constant of 2,4-DNP exceeds other ANCs, which could be due to the electron donating ability of NH2 as well as decrease of the electronegativity of nitrogen in aniline moieties, as compared to oxygen of hydroxyl group in nitrophenol compounds, and thus, nitrophenolate ions. Also, 2,4-DNP and 2,4,6-TNP (picric acid) suffer from more steric hindrance compared with 2-NA and 4-NA upon orientation on the catalyst surface. Among these nitroaniline molecules, the lone pair of electron on the amino nitrogen atom is attracted toward the strong electron withdrawing NO2 groups, resulting in a resonating structure. Hence, the amino nitrogen becomes positively charged in 4-NA, by which an electrostatic interaction with our catalyst system occurs [75].

2.9. Simultaneous reduction of ANCs We, furthermore, investigated the simultaneous catalytic reactions of mixture of ANCs (4NP, 4-NA, 2,4-DNP and 2,4,6-TNP), using the same reaction conditions as seen in Fig. 6. Upon the reduction of ANCs using this catalyst, the absorption peak of ANCs decreases at nm within 25 s, indicating the complete reduction of ANCs, while the peak at increases due to the formation of amine products with the isobestic points at

max

max

max

= 400

= 300 nm

= 276 and 340

nm. Interestingly, the Re/OMC catalyst has ability to reduce many nitroarenes into their respective amines in aqueous medium simultaneously. The time evolution of UV/Vis absorption spectra for the simultaneous reduction of nitro compounds into their corresponding products (AACs) are shown in Figs 6a and Fig. 6b. This may be due to simultaneous adsorption/diffusion of all ANCs on the surface of Re/OMC catalyst. It was also observed that ANCs cannot be reduced by NaBH4 without Re/OMC catalyst involvement (see Fig. S7, SI). The rate constants, k, were obtained directly from the slopes of the straight lines, and the k values are found to be 0.1278 s−1 (R2 = 0.994; Fig. 6c) and 0.1317 s−1 (R2 = 0.9983; Fig. 6d) for the 0.02 and 0.05 mg of Re/OMC catalyst, respectively.

18

Fig. 6. UV-visible spectra representing simultaneous reduction of mixture of all ANCs in the presence of Re/OMC catalyst (a) 0.02 mg, and (b) 0.05 mg under same experimental conditions.

The reduction rate constant (k) for each nitro compound was found to be higher than that for simultaneous reduction of ANCs. The sole product was formed without any by-product generation during the reduction process. This catalytic system is proved to be useful for reduction of various nitro compounds that were carried out individually as well as simultaneously. Our catalytic system may be an effective catalyst for the rapid reduction of these highly toxic ANCs into less toxic AACs, which are useful for other practical applications.

2.10. Mechanism of reduction reaction The as-prepared Re/OMC catalyst with enhanced catalytic activity can be attributed to the active surface area, and a large adsorption of reactants onto the highly porous 3D carbon matrix, as shown in Scheme 2. Initially, the active nitro groups on the surfaces of the catalyst were converted into nitrophenolate ion in the presence of NaBH4 followed by hydrogen transfer from BH4 (which is adsorbed on an carbon matrix), leading to a rapid reaction between the reactive hydrogen (H+) and 4-(hydroxylamino)phenolate ion [63]. The rate of electron transfer at the 19

catalyst surface can be influenced by two-steps: (i) adsorption of 4-nitrophenolate ion onto the catalyst surface, and (ii) interfacial electron transfer and desorption of 4-aminophenolate ion away from the surfaces. Finally, the AACs were formed, and the catalyst was recovered by centrifugation and reused for subsequent runs. It is noteworthy that the formation of 4-AP from 4-NP proceeds via the generation of 4-nitrophenolate ion and 4-(hydroxylamino)phenolate ion as the intermediates, which involve six-electron (6e), and six-proton (6H+) transfer, as shown in Scheme 2.

Scheme 2. Mechanism for the reduction of 4-NP using Re/OMC catalyst.

2.11. Reusability and stability To investigate the reusability, the Re/OMCs catalysts were separated from the catalytic reaction solution using an ultracentrifugation. In order to keep the amount of catalyst, we added reactants to the resulting mixture for the next six cycles to investigate the reusability of the catalyst under the same experimental conditions. The conversion and activity of the catalyst were found to be slightly decreased for the same reaction time (20 s) even after running for 6 cycles. It was found

20

that the rate constant (k) remained almost the same. Meanwhile, the conversion percentage for complete reaction gradually decreased with increasing the running cycles.

Fig. 7. (a) Recycling test of the Re/OMC catalyst towards the 4-NP reduction, (b) XRD pattern, and (c,d) FE-TEM images of spent Re/OMC catalyst.

For instance, the rate constant was ca. 0.1188 s1 during the 20 s time at sixth cycle (Fig. 7a). The slight decrease in the activity may be caused by the loss of some catalysts at the steps of centrifugation. Besides, the X-ray diffraction pattern of the spent Re catalyst remains the same (Fig. 7b). FE-TEM measurements also confirmed that the 3D micro/mesoporous structures of this catalyst are well maintained after six catalytic processes, further suggesting its excellent stability and long life (Figs. 7c and 7d). Recyclability is the main advantage of heterogeneous 21

catalysis compared with homogeneous catalysis for practical applications. In our catalytic system, the solid catalyst was easily recovered from the reaction mixture by brief centrifugation or sedimentation process. It can be seen that the recyclable reduction of 4-NP in the presence of Re/OMCs catalyst exhibited similar catalytic activity as in the first three cycles of the reaction. The 4-NP conversion decreased from the fourth to sixth cycles, probably due to loss of catalyst during the recycling process. Thus, the Re/OMCs catalyst was recycled and reused for six successive reactions with a conversion of more than 90%, and the product yield was >90%. These facts confirmed that the catalyst is stable and active (Fig. 6b). Moreover, metal dispersion characterization data, before and after the catalytic reactions, are shown in the supporting information (Table S5, SI). The recycled catalyst so prepared from the hot filtration procedure was examined by inductively coupled plasma atomic emission spectroscopy (ICP-AES). Only about 15 ppm loss of Re species was obtained while comparing to that of fresh Re/OMC catalyst (ca. 3.12 wt %; Fig. S1, SI), and a spent catalyst (ca. 2.95 wt %; Fig. S8, SI) obtained after six consecutive runs. These results clearly demonstrate that the Re/OMC catalyst is truly robust and heterogeneous in nature.

3.

Conclusions

In this work, we have reported a new and highly efficient methodology for the reduction of ANCs to AACs using Re/OMC catalyst in the presence of NaBH4 as a hydrogen resource in aqueous medium. We have demonstrated a solvent-evaporation induced self-assembly (ELSA) method to synthesize ReNPs catalysts supported on OMC materials, which possess an accessible ordered mesopore channels. This catalyst was characterized by FE-TEM, XRD and N2 adsorption/desorption isotherm measurements. FE-TEM studies showed the isolated ReNPs with an average diameter of 5 nm were formed and immobilized onto the OMC supports. The ReNPs/OMC catalysts were highly active for the reduction of ANCs to AACs and could be recycled for six consecutive runs without significant loss of catalytic activity. The catalytic activity was found to be higher while compared to some of the monometallic or bimetallic NPs catalysts reported [21-26,50]. The simplicity of the system, easy separation of catalyst, simple workup procedures and excellent yields make this method an attractive, environmentally acceptable synthetic tool for the reduction of various aromatic nitro compounds to their corresponding amines. All these features of OMC support with noble metal nanoparticle 22

encourage researchers to continue to develop new heterogeneous catalysts. Therefore, our results may provide new ways to fabricate 3D-based carbon materials, which can be extended to the fabrication of other sustainable applications such as energy and environmental, and electrochemical/biosensor, etc. Acknowledgments The authors are grateful for the financial support (NSC 102-2113-M-002-009-MY3 to KCL; NSC 104-2113-M-001-020-MY3 to SBL) from the Ministry of Science and Technology (MOST), Taiwan. Appendix A. Supplementary data Supplementary data associated with this article can be found, in the online version, at. References [1]

Y. Wang, W. He, L. Wang, J. Yang, X. Xiang, B. Zhang, F. Li, Highly active supported Pt nanocatalysts synthesized by alcohol reduction towards hydrogenation of cinnamaldehyde: Synergy of metal valence and hydroxyl groups, Chem. Asian J. 10 (2015) 1561–1570.

[2]

Z. Tian, X. Xiang, L. Xie, F. Li, Liquid-Phase hydrogenation of cinnamaldehyde: enhancing selectivity of supported gold catalysts by incorporation of cerium into the support, Ind. Eng. Chem. Res. 52 (2013) 288−296.

[3]

H. Shang, K. Pan, L. Zhang, B. Zhang, X. Xiang, Enhanced activity of supported Ni catalysts promoted by Pt for rapid reduction of aromatic nitro compounds, Nanomaterials 6 (2016) 103, doi:10.3390/nano6060103.

[4]

X. Xiang, W. He, L. Xie, F. Li, A mild solution chemistry method to synthesize hydrotalcite-supported

platinum

nanocrystals

for

selective

hydrogenation

of

cinnamaldehyde in neat water, Catal. Sci. Technol. 3 (2013) 2819–2827. [5]

Y.L. Kim, H.-A. Choi, N.-S. Lee, B. Son, H.J. Kim, J.M. Baik, Y. Lee, C. Lee, M.H. Kim, RuO2–ReO3 composite nanofibers for efficient electrocatalytic responses, Phys. Chem. Chem. Phys. 17 (2015) 7435−7442.

[6]

Y. Yuan, T. Shido, Y. Iwasawa, The new catalytic property of supported rhenium oxides for selective oxidation of methanol to methylol, Chem. Comm. (2000) 1421–1422.

23

[7]

V.G. Kessler, G.A. Seisenbaeva, Rhenium nanochemistry for catalyst preparation, Minerals 2 (2012) 244–257.

[8]

U.G. Hong, H.W. Park, J. Lee, S. Hwang, J. Yi, I.K. Song, Hydrogenation of succinic acid to tetrahydrofuran (THF) over rhenium catalyst supported on H2SO4-treated mesoporous carbon, Appl. Catal. A Gen. 415 (2012) 141–148.

[9]

K. Kon, W. Onodera, T. Toyao, K. Shimizu, Supported rhenium nanoparticle catalysts for acceptorless dehydrogenation of alcohols: structure–activity relationship and mechanistic studies, Catal. Sci. Technol. 6 (2016) 5864–5870.

[10] K. Sakthikumar, S. Anantharaj, S.R. Ede, K. Karthick, S. Kundu, A highly stable rhenium organosol on a DNA scaffold for catalytic and SERS applications, J. Mater. Chem. C 4 (2016) 6309–6320. [11] Y.Y. Chong, W.Y. Chow, W.Y. Fan, Preparation of rhenium nanoparticles via pulsed-laser decomposition and catalytic studies, J. Colloid Interface. Sci. 369 (2012) 164−169. [12] A. Vargas-Uscategui, E. Mosquera, L. Cifuentes, Analysis of the electrodeposition process of rhenium and rhenium oxides in alkaline aqueous electrolyte, Electrochimica Acta 109 (2013) 283−290. [13] J.V. Rojas, C.H. Castano, Synthesis of rhenium oxide nanoparticles (Re xOy) by gamma irradiation, Radiat. Phys. Chem. 99 (2014) 1−5. [14] Y. Tong, S. Bai, H. Zhang, Y. Ye, Rhenium coating prepared on carbon substrate by chemical vapor deposition, Appl. Surf. Sci. 261 (2012) 390–395. [15] K.M. Babu, M.R. Mucalo, XPS studies of freshly prepared rhenium nanoparticle dispersions from hydrazinium hydrate and borohydride reduction of hexachlororhenate solutions, J. Mater. Sci. Lett. 22 (2003) 1755−1757. [16] N. Yang, G.E. Mickelson, N. Greenlay, S. D. Kelly, F. D. Vila, J. Kas, J. J. Rehr, S. R. Bare, Size and shape of rhenium nanoparticles, AIP Conf. Proc. 882 (2007) 591. [17] K. Biswas, C.N.R. Rao, Metallic ReO3 nanoparticles, J. Phys. Chem. B 110 (2006) 842– 845. [18] J. Bedia, L. Calvo, J. Lemus, A. Quintanilla, J.A. Casas, A.F. Mohedano, J.A. Zazo, J.J. Rodriguez, M.A. Gilarranz, Colloidal and microemulsion synthesis of rhenium nanoparticles in aqueous medium, Colloids Surf. A. Physicochem. Eng. Asp. 469 (2015) 202–210. 24

[19] A.A. Revina, M.A. Kuznetsov, A.M. Chekmarev, Physicochemical properties of rhenium nanoparticles obtained in reverse micelles, Dokl. Akad. Nauk 450 (2013) 47–49. [20] T. Ayvalı, P. Lecante, P.-F. Fazzini, A. Gillet, K. Philippot, B. Chaudre, Facile synthesis of ultra-small rhenium nanoparticles, Chem. Commun. 50 (2014) 10809–10811. [21] S. Anantharaj, M. Jayachandran, S. Kundu, Unprotected and interconnected Ru0 nanochain networks: advantages of unprotected surfaces in catalysis and electrocatalysis, Chem. Sci. 7 (2016) 3188–3205. [22] S. Anantharaj, U. Nithiyanantham, S.R. Ede, S. Kundu, Osmium organosol on DNA: Application in catalytic hydrogenation reaction and in SERS studies, Ind. Eng. Chem. Res. 53 (2014) 19228−19238. [23] U. Nithiyanantham, S.R. Ede, S. Kundu, Self-assembled wire-like and honeycomb-like osmium nanoclusters (NCs) in DNA with pronounced catalytic and SERS activities, J. Mater. Chem. C 2 (2014) 3782–3794. [24] S. Kundu, H. Liang, Shape-selective formation and characterization of catalytically active iridium nanoparticles, J. Colloid Interface Sci. 354 (2011) 597–606. [25] S. Kundu, S. Lau, H. Liang, Shape-controlled catalysis by cetyltrimethylammonium bromide terminated gold nanospheres, nanorods, and nanoprisms, J. Phys. Chem. C 113 (2009) 113, 5150–5156. [26] S. Kundu, Formation of self-assembled Ag nanoparticles on DNA chains with enhanced catalytic activity, Phys. Chem. Chem. Phys. 15 (2013) 14107–14119. [27] G.Y. Yurkov, A.V. Kozinkin, Y.A. Koksharov, A.S. Fionov, N.A. Taratanov, V.G. Vlasenko, I.V. Pirog, O.N. Shishilov, O.V. Popkov, Synthesis and properties of rhenium– polyethylene nanocomposite, Composites Part B 43 (2012) 3192−3197. [28] S. Anantharaj, K. Sakthikumar, A. Elangovan, G. Ravi, T. Karthik, S. Kundu, Ultra-small rhenium nanoparticles immobilized on DNA scaffolds: An excellent material for surface enhanced Raman scattering and catalysis studies, J. Colloid Interface Sci. 483 (2016) 360– 373. [29] C.D. Valenzuela, M.L. Valenzuela, S. Caceres, C. O’Dwyer, Solution and surfactant-free growth of supported high index facet SERS active nanoparticles of rhenium by phase demixing, J. Mater. Chem. A 1 (2013) 1566–1572.

25

[30] A.D. Dobrzanska-Danikiewicz, W. Wolany, G. Benke, Z. Rdzawsk, The new MWCNTs– rhenium nanocomposite, Phys. Status Solidi B 251 (2014) 2485–2490. [31] I.T. Ghampson, C. Sepúlveda, R. García, J.L.G. Fierro, N. Escalon, Carbon nanofibersupported ReOx catalysts for the hydrodeoxygenation of lignin-derived compounds, Catal. Sci. Technol. 6 (2016) 4356–4369. [32] A.K. Aboul-Gheit, S.M. Abdel-Hamid, S.M. Aboul-Fotouh, N.A.K. Aboul-Gheit, Cyclohexene hydroconversion using monometallic and bimetallic catalysts supported on Alumina, J. Chin. Chem. Soc. 53 (2006) 793–802. [33] P. Veerakumar, N. Dhenadhayalan, K.C. Lin, S.B. Liu, Highly stable ruthenium nanoparticles on 3D mesoporous carbon: an excellent opportunity for reduction reactions, J. Mater. Chem. A 3 (2015) 23448−23457. [34] P. Tomkins, E. Gebauer-Henke, W. Leitner, T.E. Muller, Concurrent hydrogenation of aromatic and nitro groups over carbon-supported ruthenium catalysts, ACS Catal. 5 (2015) 203−209. [35] T. Ayvali, Rhenium based mono-and bi-metallic nanoparticles: Synthesis, characterization and application in catalysis (Doctoral dissertation, Université de Toulouse, Université Toulouse III-Paul Sabatier), 2015. [36] V. Purohit, A.K. Basu, Mutagenicity of nitroaromatic compounds, Chem. Res. Toxicol. 13 (2000) 673–692. [37] US Department of Health and Human Services, Public Health Service, Agency for Toxic Substances and Disease Registry (1995). [38] J.C. Oxley, The chemistry of explosives, In J. A. Zukas W. Walters, (ed.), Explosive effects and applications. Springer, Berlin, Germany (2002) 137–172. [39] A.S. Travis, Manufacture and uses of the anilines: A vast array of processes and products, In Z. Rappoport (ed.), The chemistry of anilines, part 1. John Wiley and Sons, New York, New York (2007) 715–782. [40] G.H. Lee, S.H. Huh, S.H. Kim, B.J. Choi, B.S. Kim, J.H. Park, Structure and size distribution of Os, Re, and Ru nanoparticles produced by thermal decomposition of Os3(CO)12, Re2(CO)10, and Ru3(CO)12, J. Korean Phys. Soc. 42 (2003) 835−837. [41] Y. Song, Z. Li, K. Guo, T. Shao, Hierarchically ordered mesoporous carbon/graphene composites as supercapacitor electrode materials, Nanoscale 8 (2016) 15671–15680. 26

[42] P. Veerakumar, R. Madhu, S.-M. Chen, C.-T. Hung, P.-H. Tang, C.-B. Wang and S.-B. Liu, Porous carbon-modified electrodes as highly selective and sensitive sensors for detection of dopamine, Analyst 139 (2014) 4994–5000. [43] D.D. Falcone, J.H. Hack, A.Y. Klyushin, A. Knop-Gericke, R. Schl gl, R. J. Davis, Evidence for the bifunctional nature of Pt–Re catalysts for selective glycerol hydrogenolysis, ACS Catal. 5 (2015) 5679−5695. [44] M.T. Greiner, T.C.R. Rocha, B. Johnson, A. Klyushin, A.

nop- ericke, R. Schl gl, The

Oxidation of rhenium and identification of rhenium oxides during catalytic partial oxidation of ethylene: An In-Situ XPS study, Z. Phys. Chem. 228 (2014) 521−541. [45] A. Borodzinski, M. Bonarowska, Relation between crystallite size and dispersion on supported metal catalysts, Langmuir 13 (1997) 5613–5620. [46] B. Mitra, X.T. Gao, I.E. Wachs, A.M. Hirt, G. Deo, Characterization of supported rhenium oxide catalysts: Effect of loading, support and additives, Phys. Chem. Chem. Phys. 3 (2001) 1144–1152. [47] B. Lin, Y. Qi, Y. Guo, J. Lin, J. Ni, Effect of potassium precursors on the thermal stability of K-promoted Ru/carbon catalysts for ammonia synthesis, Catal. Sci. Technol. 5 (2015) 2829–2838. [48] X. Di, Z. Shao, C. Li, W. Li, C. Liang, Hydrogenation of succinic acid over supported rhenium catalysts prepared by the microwave assisted thermolytic method, Catal. Sci. Technol. 5 (2015) 2441–2448. [49] Y. Zhang, H. Jiang, Y. Wang, M. Zhang, Hydrogenation of succinic acid over supported rhenium catalysts prepared by the microwave assisted thermolytic method, Ind. Eng. Chem. Res. 53 (2014) 6380−6387. [50] S.K. Ghosh, M. Mandal, S. Kundu, S. Nath, T. Pal, Bimetallic Pt–Ni nanoparticles can catalyze reduction of aromatic nitro compounds by sodium borohydride in aqueous solution, Appl. Catal. A Gen. 268 (2004) 61–66. [51] H.-S. Shin, S. Huh, Au/Au@Polythiophene core/shell nanospheres for heterogeneous catalysis of nitroarenes, ACS Appl. Mater. Interfaces 4 (2012) 6324–6331. [52] L. He, L.-C. Wang, H. Sun, J. Ni, Y. Cao, H.-Y. He, K.-N. Fan, Efficient and selective room-temperature gold-catalyzed reduction of nitro compounds with CO and H 2O as the hydrogen source, Angew. Chem. Int. Ed. 48 (2009) 9538–9541. 27

[53] N.M. Patil, T. Sasaki, B.M. Bhanage, Immobilized iron metal-containing ionic liquidcatalyzed chemoselective transfer hydrogenation of nitroarenes into anilines, ACS Sustainable Chem. Eng. 4 (2016) 429−436. [54] H.

ksu, S.F. Ho,

. Metin, K. Korkmaz, A.M. Garcia, M.S.

ltekin, S. Sun, Tandem

dehydrogenation of ammonia borane and hydrogenation of nitro/nitrile compounds catalyzed by graphene-supported NiPd alloy nanoparticles, ACS Catal. 4 (2014) 1777−1782. [55] J. Pesti and G. L. Larson, Tetramethyldisiloxane: A practical organosilane reducing agent, Org. Process Res. Dev. 20 (2016) 1164−1181. [56] N. Sakai, K. Fujii, S. Nabeshima, R. Ikeda, T. Konakahara, Highly selective conversion of nitrobenzenes using a simple reducing system combined with a trivalent indium salt and a hydrosilane, Chem. Commun. 46 (2010) 3173–3175. [57] A.M. Tafesh, J. Weiguny, A review of the selective catalytic reduction of aromatic nitro compounds into aromatic amines, isocyanates, carbamates, and ureas using CO, Chem. Rev. 96 (1996) 2035−2052. [58] T.S. Wu, G. Wen, B. Zhong, B. Zhang, X. Gu, N. Wang, D. Su, Reduction of nitrobenzene catalyzed by carbon materials, Chin. J. Catal. 35 (2014) 914–921. [59] K.S. Suslick, P.F. Schubert, Sonochemistry of dimanganese decacarbonyl (Mn2(CO)10) and dirhenium decacarbonyl (Re2(CO)10), J. Am. Chem. Soc. 105 (1983) 6042–6044. [60] S. Saha, A. Pal, S. Kundu, S. Basu, T. Pal, Photochemical green synthesis of calciumalginate-stabilized Ag and Au nanoparticles and their catalytic application to 4-nitrophenol reduction, Langmuir 26 (2010) 2885–2893. [61] I. Nakamula, Y. Yamanoi, T. Imaoka, K. Yamamoto, H. Nishihara, A uniform bimetallic rhodium/iron nanoparticle catalyst for the hydrogenation of olefins and nitroarenes, Angew. Chem. Int. Ed. 50 (2011) 5830–5833. [62] R. Begum, K. Naseem, E. Ahmed, A. Sharif, Z.H. Farooqi, Simultaneous catalytic reduction

of

nitroarenes

using

silver

nanoparticles

fabricated

in

poly(N-

isopropylacrylamide-acrylic acid-acrylamide) microgels, Colloids Surf. A. Physicochem. Eng. Asp. 511 (2016) 17–26.

28

[63] S. Jayabal, R. Ramaraj, Bimetallic Au/Ag nanorods embedded in functionalized silicate sol–gel matrix as an efficient catalyst for nitrobenzene reduction, Appl. Catal. A 470 (2014) 369–375. [64] B. Guan, X. Wang, Y. Xiao, Y. Liu, Q. Huo, A versatile cooperative template-directed coating method to construct uniform microporous carbon shells for multifunctional core– shell nanocomposites, Nanoscale 5 (2013) 2469–2475. [65] H. Wu, X. Huang, M.M. Gao, X.P. Liao, B. Shi, Polyphenol-grafted collagen fiber as reductant and stabilizer for one-step synthesis of size-controlled gold nanoparticles and their catalytic application to 4-nitrophenol reduction, Green Chem. 13 (2011) 651–658. [66] S. Kundu, K. Wang, H. Liang, Size-selective synthesis and catalytic application of polyelectrolyte encapsulated gold nanoparticles using microwave irradiation, J. Phys. Chem. C 113 (2009) 5157–5163. [67] P. Veerakumar, M. Velayudham, K.-L. Lu, S. Rajagopal, Polyelectrolyte encapsulated gold nanoparticles as efficient active catalyst for reduction of nitro compounds by kinetic method, Appl. Catal. A Gen. 439–440 (2012) 197–205. [68] Y. Zhang, X. Yuan, Y. Wang, Y. Chen, One-pot photochemical synthesis of graphene composites uniformly deposited with silver nanoparticles and their high catalytic activity towards the reduction of 2-nitroaniline, J. Mater. Chem. 2 (2012) 7245–7251. [69] P. Deka, R. Choudhury, R.C. Deka, P. Bharali, Influence of Ni on enhanced catalytic activity of Cu/Co3O4 towards reduction of nitroaromatic compounds: Studies on the reduction kinetics, RSC Adv. 6 (2016) 71517–71528. [70] S. Mondal, U. Rana, R.R. Bhattacharjee, S. Malik, One pot green synthesis of polyaniline coated gold nanorods and its applications, RSC Adv. 4 (2014) 57282–57289. [71] X. -Q. Wu, X. -W. Wu, J.-S. Shen, H.-W. Zhang, In situ formed metal nanoparticle systems for catalytic reduction of nitroaromatic compounds, RSC Adv. 4 (2014) 49287–49294. [72] M.T. Qamar, M. Aslam, I.M.I. Ismail, N. Salah, A. Hameed, Synthesis, characterization, and sunlight mediated photocatalytic activity of CuO coated ZnO for the removal of nitrophenols, ACS Appl. Mater. Interfaces 7 (2015) 8757−8769. [73] I. Ibrahim, I.O. Ali, T.M. Salama, A.A. Bahgat, M.M. Mohamed, Synthesis of magnetically recyclable spinel ferrite (MFe2O4, M = Zn, Co, Mn) nanocrystals engineered by sol gel-

29

hydrothermal technology: High catalytic performances for nitroarenes reduction, Appl. Catal. B Environ. 181 (2016) 389–402. [74] A. Bhattacharjee, M. Ahmaruzzaman, A facile and green strategy for the synthesis of 1dimensional luminescent ZnO nanorods and their reduction behavior for aromatic nitrocompounds, RSC Adv. 6 (2016) 527–533. [75] X.-Q. Wu, X.-W. Wu, Q. Huang, J.-S. Shen, H.-W. Zhang, In situ synthesized gold nanoparticles in hydrogels for catalytic reduction of nitroaromatic compounds, Appl. Surf. Sci. 331 (2015) 210–218.

30

Graphical Abstract

Well-dispersed

rhenium

nanoparticles

on

three-dimensional

carbon

nanostructures: Efficient catalysts for the reduction of aromatic nitro compounds Pitchaimani Veerakumar,a,b,* Pounraj Thanasekaran,c King-Chuen Lina,b,* and Shang-Bin Liua

Ordered mesoporous carbon supported ReNPs (Re/OMC) unsupported colloidal metal nanoparticles.

31

exhibited superior performances compared to their