One-Pot Synthesis of Pd-Based Ternary Pd@CdS@TiO2

Dec 28, 2018 - In this work, we report the synthesis of Pd-based ternary Pd@CdS@TiO2 nanocomposites using molecular precursors. This method is facile,...
20 downloads 0 Views 7MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2018, 3, 18663−18672

http://pubs.acs.org/journal/acsodf

One-Pot Synthesis of Pd-Based Ternary Pd@CdS@TiO2 Nanoclusters via a Solvothermal Route and Their Catalytic Reduction Efficiency toward Toxic Hexavalent Chromium Rashmi A. Badhe, Aleem Ansari, and Shivram S. Garje* Department of Chemistry, University of Mumbai, Vidyanagari, Santacruz (East), Mumbai 400 098, India

ACS Omega 2018.3:18663-18672. Downloaded from pubs.acs.org by 95.181.217.163 on 01/01/19. For personal use only.

S Supporting Information *

ABSTRACT: In this work, we report the synthesis of Pdbased ternary Pd@CdS@TiO2 nanocomposites using molecular precursors. This method is facile, less time-consuming, and cost-effective. This catalyst is prepared within 2 h by a solvothermal route using molecular precursors. Information about the phase, morphologies, elemental mapping, and composition of the nanocomposites was obtained using various characterization techniques. The catalytic activity of the as-prepared Pd-based ternary Pd@CdS@TiO2 nanocomposites exhibits effective reduction efficiency for the conversion of toxic Cr(VI) to Cr(III) using formic acid as a reducing agent within 5−7 min. To the best of our knowledge, this is the first report on Pd-based ternary Pd@CdS@TiO2 nanocomposites prepared by a solvothermal route and used as catalysts toward the reduction of hexavalent chromium at room temperature. process to generate CO2 and H2 (HCOOH → CO2 + H2). This H2 adsorbs on the surface of nanocomposites, which is mainly responsible for the reduction of Cr(VI) to Cr(III) (Cr2O72− + 8H+ + 3H2 → 2Cr3+ + 7H2O).2,5,19−25 Generally, noble and transition metals Pd, Pt, Ru, Ag, and so forth are active catalysts for many catalytic hydrogenolyses and catalytic reduction of heavy metal pollutants. Although Pd is a high-cost metal, it has many more uses in electrocatalysis, C−C coupling reactions, and other catalytic reductions. Many researchers have focused on the fabrication of Pd-based nanocomposites with low-cost transition metals, which provide high efficiency in catalytic sites. According to World Health Organization (WHO), hexavalent chromium Cr(VI) is the most hazardous pollutant, hence there is a need to develop practical and costeffective methodologies for reductive transformation of Cr(VI) to Cr(III). The Pd-based catalyst has high efficiency for the reduction of Cr(VI).2,15−18,26,27 So far, various methods such as chemical reduction, ion exchange, photocatalysis, and adsorption have been used for the reduction of Cr(VI) to Cr(III). For example, modified TiO2-mediated photocatalytic reduction of Cr(VI) to Cr(III) has been reported by Acharya et al.28 However, the advantage of a chemical reduction method that has been used in the present study is that there is no need of external source of light

1. INTRODUCTION Nowadays, the synthesis of nanocomposites and their applications in selective organic transformations from wastewater have been investigated extensively. Hexavalent chromium Cr(VI) is the most hazardous pollutant found in soil, wastewater, and ground water.1 Various pigment manufacturing, leather tanning, metal finishing, and electroplating industries make use of hexavalent chromium, Cr(VI). The effluents of such industries contain a large amount of hexavalent chromium Cr(VI) which is nonbiodegradable and exists for a longer period in the environment. Chromium is a heavy metal having variable oxidation states, with two valence states, hexavalent chromium Cr(VI) and trivalent chromium Cr(III). Cr(III) is less harmful as compared to Cr(VI).2,3 Because Cr(VI) has strong oxidizing properties, it is more toxic and contains carcinogenic species as compared to Cr(III) species. Because of high toxicity of Cr(VI) at very low concentration, it has the ability to cause hazardous threats to human and other living organisms such as it can increase the risk of DNA mutation and lung cancer by chronic inhalation.1,4,5 Various semiconductor catalysts such as TiO2,6−8 NiO@ TiO2,9 ZnO@TiO2,10 ZnO,11 Fe3O4@graphene,12 Pd@SiO2− NH2,2 ZnO−TiO2−OCNT,13 and CdS/RGO14 have been reported for the reduction of Cr(VI) to Cr(III). Recently, the use of formic acid (HCOOH) as a reducing agent for the reduction of Cr(VI) to Cr(III) has been reported because of its simplicity and efficiency.5,15−18 In general mechanism, HCOOH undergoes a dehydrogenation decomposition © 2018 American Chemical Society

Received: October 24, 2018 Accepted: December 10, 2018 Published: December 28, 2018 18663

DOI: 10.1021/acsomega.8b02924 ACS Omega 2018, 3, 18663−18672

ACS Omega

Article

Figure 1. XRD patterns of as-prepared nanocomposites obtained from TNC-I (I) and TNC-II (II).

formation of Pd-based ternary Pd@CdS@TiO2 nanocomposites and binary CdS@TiO2 nanocomposites (Figure S4). They match with anatase TiO2 (JCPDS file no. 00-021-01272) and hexagonal CdS (JCPDS file no. 00-041-1049). The additional peak observed at 2θ = 39.4° for TNC-I and 2θ = 39.8° for TNC-II confirmed the successful loading of Pd on the surface of CdS@TiO2 nanocomposites,31,32 and it was observed that the crystallinity is retained after the loading of Pd as shown in Figure 1. The average particle size was calculated using the Scherrer formula,33 which was found to be 10.26 nm (TNC-I), 7.31 nm (TNC-II), 30.5 nm (BNC-I), and 13.3 nm (BNC-II) for as-prepared nanocomposites. In order to examine the morphology and elemental mapping of as-prepared Pd-based ternary nanocomposites, we have carried out high-resolution transmission electron microscopy (HRTEM) (Figure S5), TEM, scanning electron microscopy (SEM), and SEM−energy-dispersive X-ray spectroscopy (EDS) elemental mapping studies. TEM images are shown in Figure 2. From Figure 2Ia,IIa, spherical morphologies of the

and that reduction can be carried out at room temperature (RT) using a simple reducing agent. Herein, we report a facile and efficient route for the synthesis of binary CdS@TiO2 nanocomposites and Pd-based ternary Pd@CdS@TiO2 nanocomposites by a solvothermal route using CdCl2(4-MebenztsczH)2 (precursor I), CdI2(4-MebenztsczH)2 (precursor II) (where 4-MebenztsczH = 4methylbenzaldehyde thiosemicarbazone), and titanium isopropoxide (Ti(OC3H7)4) as molecular precursors. As these molecular precursors contain both metallic and nonmetallic parts, they act as good precursors for the synthesis of semiconductor nanoparticles (NPs). The use of such molecular precursors provides several advantages such as minimization of expensive chemicals, low toxicity, limited or no prereactions, and so forth.29,30 The catalytic reductive efficiency has been further studied using as-prepared Pd-based ternary Pd@CdS@TiO2 nanocomposites (TNC-I obtained from precursor I and TNC-II obtained from precursor II) for the reduction of toxic hexavalent Cr(VI) to Cr(III). For comparison, the catalytic reduction of Cr(VI) has also been studied using binary CdS@ TiO2 nanocomposites (BNC-I obtained from precursor I and BNC-II obtained from precursor II). Two different precursors, that is, (CdCl 2 (4-MebenztsczH) 2 and CdI 2 (4-MebenztsczH)2), were used to prepare the composites. This was carried out in order to find out whether two different precursors result into different composites, which may result into different catalytic activities.

2. RESULTS AND DISCUSSION The characterization of the precursors was done by elemental analysis and 1H and 13C {1H} nuclear magnetic resonance (NMR) (Figures S1 and S2). The IR spectra of precursor I, precursor II, and ligand are shown in Figure S3. The peaks observed at 3433 and 3259 cm−1 (for precursor I) and 3395 and 3287 cm−1 (for precursor II) are attributed to νNH 2 asymmetric and symmetric vibrations, respectively. The bands at 3162 and 3188 cm−1 (for precursor I and precursor II, respectively) are observed because of νNH. The bands at 1536 and 1526 cm−1 (for precursor I and precursor II, respectively) are assigned to νCN. The bands at 946 and 955 cm−1 are attributed to νCS for precursor I and precursor II, respectively. The bands observed for νCN in the case of precursor I and precursor II are shifted to lower wavenumber as compared to the band observed for the ligand (1594 cm−1). In the X-ray diffraction (XRD) patterns, the sharp and intense diffraction peaks observed (Figure 1) show successful

Figure 2. (a) TEM images and (b) their corresponding SAED patterns of TNC-I and TNC-II.

nanoclusters (TNC-I and TNC-II) can be seen. Figure 2Ib,IIb shows selected-area electron diffraction (SAED) patterns revealing the polycrystallinity of these materials. This was further confirmed by SEM and SEM−EDS elemental mapping. The SEM images show a spherical shaped morphology for asprepared Pd-based ternary nanoclusters (TNC-I and TNC-II) (Figure 3a,b). The elemental mapping analysis of as-prepared Pd-based ternary nanoclusters confirmed that palladium (Pd) 18664

DOI: 10.1021/acsomega.8b02924 ACS Omega 2018, 3, 18663−18672

ACS Omega

Article

excitation source to investigate the various phases present in the as-prepared Pd-based nanoclusters Figure 5. CdS shows

Figure 5. Raman spectra of as-prepared TNC-I (I) and TNC-II (II).

Raman vibrational bands at 263.7 and 565.6 cm−1 for TNC-I and 269.5 and 584.6 cm−1 for TNC-II, and the weak bands at 353.9 cm−1 for TNC-I and 354.7 cm−1 for TNC-II are assigned to the multiphonon scattering in CdS.38−40 The presence of TiO2 can be confirmed by the vibration peaks observed around 388−390, 484−499, and 611−613 cm−1 (for TNC-I and TNC-II).41−45 There is no intense Raman vibrational peak observed around 640 cm−1 (for TNC-I and TNC-II), which confirms the presence of Pd only. This is well supported by XRD results.44,45 The presence of Pd on CdS@TiO2 nanoclusters was confirmed by X-ray photoelectron spectroscopy (XPS) measurements. Figures 6 and 7 show the XPS spectra of asprepared nanoclusters TNC-I and TNC-II, which give the evidence of the presence of Cd, S, Ti, and O. It provides two states for cadmium (Cd 3d5/2 and Cd 3d3/2), two states for sulfur (S 2p3/2 and S 2p1/2), two states for titanium (Ti 2p3/2 and Ti 2p1/2), and one state for oxygen (O 1s). The binding energies located at 405 and 412 eV are attributed to Cd 3d5/2 and Cd 3d3/2 states of cadmium as shown in Figure 6 (TNCI). The peak located at 407 eV for Cd 3d5/2 and the peak located at 413 eV for Cd 3d3/2 are shown in Figure 7 (TNCII). These observations suggest that cadmium is in the Cd2+ chemical state.46−49 The binding energy of sulfur is 161 eV for S 2p3/2 and 163 eV for S 2p1/2 for TNC-I (Figure 6). For the material obtained from TNC-II, the binding energy for S 2p3/2 is 163 eV as shown in Figure 7. This suggests that sulfur is in the S2− chemical state.50,51 Similarly, the peaks obtained from TNC-I and TNC-II located around 459−460 eV for Ti 2p3/2 and 464−466 eV for Ti 2p1/2 (Figures 6 and 7) indicate that titanium is in the chemical state of Ti4+ and the oxygen shows binding energies of 530 and 532 eV as shown in Figure 6 and 533 eV as shown in Figure 7, corresponding to the oxygen O 1s peak. This is due to Ti−O in TiO2 and hydroxyl groups.52−57 Further, palladium was confirmed in the states of Pd 3d5/2 and Pd 3d3/2 with the binding energies of 336 eV (Pd 3d5/2) and 342 eV (Pd 3d3/2) as shown in Figure 6 (TNCI) and 338 eV (Pd 3d5/2) and 344 eV (Pd 3d3/2) as shown in Figure 7 (TNC-II). The peak positions were attributed to the Pd 3d level, which is well matched with the previous literature.58−60 The optical properties observed by UV−visible diffuse reflectance spectroscopy (UV-DRS) were found to be enhanced toward the visible region because of the addition of Pd, that is, Pd@CdS@TiO2, as shown in Figure 8. Their corresponding band gaps were found to be 2.09 and 2.10 eV for the materials obtained from TNC-I and TNC-II,

Figure 3. (a) SEM images and their elemental mapping of TNC-I. (b) SEM images and their elemental mapping of TNC-II.

is evenly distributed on the surface of CdS@TiO2. In addition, the EDS elemental mapping of as-prepared Pd-based ternary nanoclusters shows the even elemental distribution of Ti, O, S, Cd, and Pd as given in Figure 3a,b for TNC-I and TNC-II, respectively. Their corresponding EDS spectra are shown in Figure S6a,b. The SEM images of BNC-I and BNC-II are shown in Figure S7, which confirm the successful formation of binary nanocomposites. The formation of binary and Pd-based ternary nanoclusters was further studied by Fourier transform infrared (FTIR) analysis. The peaks observed at 602 and 609 cm−1 for BNC-I and BNC-II, respectively, as shown in Figure S8 and 599 and 627 cm−1 for TNC-I and TNC-II, respectively, as shown in Figure 4 confirm the stretching vibration bands of Cd−S.34,35 Along with that, the peaks observed in the vicinity of 400−800 cm−1 in both the nanocomposites correspond to the stretching vibration of Ti−O−Ti in TiO2.36,37 This was further confirmed by Raman analysis. Raman spectra of the as-prepared Pd-based ternary nanoclusters have been recorded with a 532 nm laser as the

Figure 4. FTIR spectra of as-prepared TNC-I (I) and TNC-II (II). 18665

DOI: 10.1021/acsomega.8b02924 ACS Omega 2018, 3, 18663−18672

ACS Omega

Article

Figure 6. XPS spectra of as-prepared TNC-I.

Figure 7. XPS spectra of as-prepared TNC-II.

Figure 8. UV-DRS spectra of as-prepared TNC-I (I) and TNC-II (II); the inset shows their corresponding band gaps calculated by Tauc’s plot.

respectively (inset Figure 8). They were calculated using Tauc’s plot.61,62 For comparison, similar band gap calculations

were carried for binary CdS@TiO2 nanocomposites. Their band gaps were found to be 2.30 and 2.28 eV for the materials 18666

DOI: 10.1021/acsomega.8b02924 ACS Omega 2018, 3, 18663−18672

ACS Omega

Article

Cr(VI).22,23 During the reduction process, successive color change from yellow to colorless is observed, which indicates successful reduction of Cr(VI) to Cr(III) (Figure 11A). This was further confirmed by the addition of excess of NaOH to the final solution. The color of solution changed to green after addition of excess NaOH because of the formation of hexahydroxochromate(III), which clearly gives the evidence for the presence of Cr(III).2 For comparison, reduction of Cr(VI) to Cr(III) using only HCOOH as a reducing agent (i.e., without the catalyst) (Figure S11) and also only binary CdS@TiO2 nanocomposites as a catalyst (Figure S12) was carried out. It was observed that even after 70 min of reaction time, no complete reduction of Cr(VI) takes place. The binary nanocomposites (BNC-I and BNC-II) show less efficiency for the reduction of Cr(VI) as compared to TNC-I and TNC-II nanoclusters because of the synergistic effect of Pd on the surface of TNCs. The plots of C/Co versus irradiation time for the catalytic reduction of Cr(VI) to Cr(III) in the presence as well as in the absence of the catalyst are shown in Figure 11B, and for comparison, reduction of Cr(VI) to Cr(III) was studied in the presence of the catalyst in the dark (Figure S13), where Co is the initial concentration and C is the final concentration at each time interval. The kinetic plots and rate constants calculated for the reduction of Cr(VI) to Cr(III) with binary and ternary nanocatalysts are of pseudo-first-order (Figure S14). It was observed that the rate constant k values of TNCs and BNCs are 0.433 (TNC-I), 0.516 (TNC-II), 0.015 (BNC-I), and 0.0218 (BNC-II). The higher k values demonstrate the enhancement of catalytic activity for the reduction of Cr(VI). TNC-I exhibits 28 times higher activity than BNC-I and TNC-II exhibits 24 times higher activity than BNC-II. From Figure S13, it can be seen that there is not much reduction in the dark. The catalytic reduction of Cr(VI) to Cr(III) is compared with the other catalysts reported in the literature (Table 1). It has been observed that the catalytic reduction efficiency of TNC-I and TNC-II nanoclusters reported in this work is remarkably better compared to those reported in the literature. The mechanism of catalytic activity of the materials can be explained as shown in Figure 12 and Scheme 1. Although TiO2 has been extensively used as a catalyst because it is less toxic, less expensive, and highly abundant and has excellent photochemical stability, it has a limitation of having a wide band gap due to which it absorbs UV radiation only.28 To be an effective catalyst, it is necessary that a material should absorb sunlight effectively and there should be sufficient electron−hole pair separation so that the electrons are available for reduction.

obtained from BNC-I and BNC-II, respectively. The absorption spectra and corresponding Tauc’s plots of binary nanocomposites are given in Figure S9. The photoluminescence (PL) spectra of binary and ternary nanocomposites are shown in Figure 9. The transfer and fate of

Figure 9. PL spectra of as-prepared TNC-I (a), TNC-II (b), BNC-I (c), and BNC-II (d).

photogenerated carriers are mainly investigated by PL spectra. The Pd-based ternary nanocomposites show lower intensities when compared with that of the binary nanocomposites. This phenomenon takes place because the Pd metal in ternary nanocomposites has higher Fermi energy level, which increases the electron-transfer reaction and leads to the prolonged electron−hole pair recombination.63,64 The catalytic efficiency of the as-prepared Pd-based ternary nanoclusters was evaluated at RT for the reduction of highly toxic Cr(VI) to Cr(III). It was monitored by UV−vis spectra with the absorption peak observed at 349 nm for Cr(VI) as shown in Figure 10. Yu et al.3 suggested that Pd provides large active sites and large surface for the enhancement of catalytic reduction of Cr(VI). It was observed that in the presence of Pd-based ternary nanoclusters, the absorption peaks of Cr(VI) decrease rapidly. Further, the peaks disappear completely, indicating the reduction of Cr(VI) into Cr(III) by HCOOH which was used as a reducing agent within 5 min for TNC-I and 7 min for TNC-II. It reveals that the as-prepared Pd-based ternary nanoclusters are more active and show greater catalytic reductive efficiency for Cr(VI). The recyclability of the catalysts TNC-I and TNC-II has also been performed. It is found that up to three cycles, the catalysts are stable and have good efficiency for the reduction of Cr(VI) (Figure S10). In the literature, it is reported that HCOOH gets adsorbed on the surface of the catalyst reducing itself into CO2 and H2. This plays a very important role in the catalytic reduction of

Figure 10. Catalytic reduction of Cr(VI) to Cr(III) using TNC-I and TNC-II at RT. 18667

DOI: 10.1021/acsomega.8b02924 ACS Omega 2018, 3, 18663−18672

ACS Omega

Article

Figure 11. (A) Photographs of Cr(VI) to Cr(III) reduction: (a) Cr(VI) solution, (b) Cr(III) solution, and (c) after addition of excess NaOH solution and (B) plots of C/Co vs irradiation time for the catalytic reduction of Cr(VI) to Cr(III): (a) TNC-I, (b) TNC-II, (c) BNC-I, (d) BNCII, and without the catalyst.

Table 1. Catalytic Efficiency of Various Catalysts Reported in the Literature along with the Present Work for the Reduction of Cr(VI) to Cr(III)a catalyst TiO2 (MT-1) AgCl@Ag CS-NCs Pd NPs@Pro-ESM ZnO BCN-BMO(Q) Ag/SnO2/NiO Fe-GCN Pd-NWWs (H2bpp)6{Fe[Mo6O12 (OH)3(HPO4)2H2PO4)2]2}2·11H2O TNC-I TNC-II

conditions [M] [M] [M] [M] [M] [M] [M] [M] [M] [M] [M]

= = = = = = = = = = =

0.2 mM (10 mg/L) [cat.] = 50 mL (1 g/L) 0.2 mM (10 mg/L) [cat.] = 20 mg 20 mM [cat.] = 15 mg 1.5 mM (75 mg/L) [cat.] = 2 g/L 0.2 mM (10 mg/L) [cat.] = 50 mg 0.2 mM (10 mg/L) [cat.] = 1g/L 0.4 mM (20 mg/L) [cat.] = 30 mg 0.8 mM [cat.] = 2 g/L 0.44 mM [cat.] = 10 mg 3 mM (150 mg/L) [cat.] = 5 mg 3 mM (150 mg/L) [cat.] = 5 mg

time (min)

k (min−1)

refs

30 8 26 60 20 60 120 15 180 5 7

0.079 0.125 0.133 0.044 0.147 0.023 0.021 0.282 0.0013 0.4339 0.5163

65 66 4 11 67 68 69 5 70 this work this work

a

[M] = chromium concentration, [cat.] = amount of the catalyst used.

electron−hole pair separation. Thus, the introduction of a noble metal (Pd) on the surface of BNCs shows the enhancement in the interfacial charge transfer as well as charge separation efficiency, followed by an increase in the catalytic reductive efficiency of Cr(VI) to Cr(III). The proposed design of TNCs that empower the efficient CdS → TiO2 → Pd pathway for the vectorial electron transfer is illustrated in Figure 12, which is also supported by PL studies.71−73 The catalytic reduction is further enhanced by scavenging holes present in the valence band of CdS using HCOOH as a scavenger (Scheme 1).74 This further hinders the charge carrier recombination rate, thereby increasing the catalytic performance. Thus, HCOOH acts not only as a reducing agent but also as a hole scavenger. Because of this, the materials used in this work show significantly improved catalytic activity.

Figure 12. Schematic possible mechanism of the catalytic reduction of Cr(VI) to Cr(III) over TNCs.

Scheme 1. Schematic Representation of HCOOH as a Hole Scavenger

3. CONCLUSIONS In conclusion, we report the synthesis and characterization of Pd-based Pd@CdS@TiO2 ternary nanoclusters by a solvothermal route using molecular precursors. The as-prepared TNCs show more catalytic efficiency to reduce highly toxic Cr(VI) to Cr(III) using HCOOH as a reducing agent at RT. Pd doped on the surface of BNCs activates the catalysts, hence it was observed that TNCs prepared from precursor I and precursor II show more efficiency than BNCs. From PL studies, it can be seen that TNCs show lower intensity as compared to BNCs, which results in slow electron−hole pair recombination and enhances the catalytic reduction of Cr(VI). HCOOH also acts as a scavenging agent for holes, thereby

Therefore, in order to facilitate the absorption of the sunlight and increase electron−hole pair separation, combination of a narrow band gap semiconductor such as CdS with a wide band gap semiconductor such as TiO2 is carried out. In this, TiO2 has more positive conduction band than that of the CdS conduction band. Therefore, the energetic electrons of the CdS conduction band transfer to the conduction band of TiO2. These electrons are then easily transported to the palladium surface and then subsequently used for reduction. The noble metal can act as a trap in the catalytic process to boost the 18668

DOI: 10.1021/acsomega.8b02924 ACS Omega 2018, 3, 18663−18672

ACS Omega

Article

Scheme 2. Schematic Representation of Synthesis of Pd-Based Ternary Pd@CdS@TiO2 Nanocomposites Using Precursor I (X = Cl) and Precursor II (X = I)

was characterized by elemental analysis, NMR, and FTIR (Yield: 2.286 g, 91.44%, mp 162 °C). Elemental analysis % found (cal.): C 29.37 (28.71), H 3.03 (2.67), N 11.04 (11.16), and Cd 15.93 (14.96). 1H NMR (δ in ppm): 2.3 (s, 3H, −CH3), 7.1−8.1 (m, 7H, −C6H5 + NH2), 11.4 (s, 1H, −NH−) (Figure S2a). 13C{1H} NMR (δ in ppm): 177.3 (>CS), 142.6 (>CN), 139.7, 131.0, 129.5, 127.4 (aromatic carbons), 20.7 (−CH3) (Figure S2b). IR: 3395, 3287 cm−1 (νNH2 assym and sym), 3188 cm−1 (νN−H), 1526 cm−1 (νCN), 955 cm−1 (νCS) (Figure S3b). Further, these molecular precursors (precursors I and II) were used for the preparation of binary CdS@TiO2 and ternary Pd@CdS@TiO2 nanocomposites. 4.3. One-Pot Synthesis of Nanocomposites. 4.3.1. Synthesis of Binary CdS@TiO2 Nanocomposites (BNC-I and BNC-II). Binary CdS@TiO2 nanocomposites using precursor I and precursor II were prepared by a solvothermal route. In a typical synthesis, Ti(OC3H7)4 (1 mL) and CdCl2(4MebenztsczH)2 (200 mg) (precursor I) were taken in a roundbottom flask containing 50 mL of EG. The mixture was sonicated for 30 min, and further it was allowed to stir and reflux for 2 h at 200 °C under nitrogen atmosphere. The reaction mixture was allowed to cool to RT. The yellowish product obtained was washed and centrifuged 4−5 times with methanol to remove the excess of EG. The final product obtained was dried under vacuum and characterized. Similarly, binary CdS@TiO2 (BNC-II) using CdI2(4MebenztsczH)2 (precursor II) was prepared and characterized. 4.3.2. Synthesis of Pd-Based Ternary Pd@CdS@TiO2 Nanocomposites (TNC-I and TNC-II). Pd-based Pd@CdS@ TiO2 (from precursor I) and Pd@CdS@TiO2 (from precursor II) ternary nanocomposites were prepared by a solvothermal route using molecular precursors as follows. In this synthesis, 1 mL of titanium isopropoxide and 200 mg of CdCl2(4-MebenztsczH)2 (precursor I) were taken in 50 mL of EG. The mixture was sonicated for 30 min. After sonication, the solution was allowed to stir and reflux. To this solution, 20 mg of PdCl2 was added, followed by the addition of 50 mg of NaBH4 under the refluxing condition. The final reaction mixture was refluxed for 2 h at 200 °C under nitrogen atmosphere. Then the reaction mixture was allowed to cool to RT. The blackish green product obtained was washed and centrifuged 4−5 times with methanol to remove the excess of EG and any other impurities present. The final product was dried under vacuum. Similarly, Pd-based Pd@CdS@TiO2 (TNC-II) was prepared using CdI2(4-MebenztsczH)2 (precursor II) and characterized (Scheme 2). 4.4. Catalytic Dichromate Reduction. The catalytic reduction of aqueous dichromate solution was further carried out using as-prepared binary CdS@TiO2 (BNC-I and BNC-

further enhancing the reduction of Cr(VI) to Cr(III). Thus, the environmental remediation of highly toxic Cr(VI) has been carried out with the help of environmental friendly, costeffective, and highly efficient catalysts.

4. EXPERIMENTAL DETAILS 4.1. Materials. All the solvents and metal salts used were of analytical grade and were used without further purification. Cadmium chloride (CdCl2, 95.0%), cadmium iodide (CdI2, 99.0%), potassium dichromate (K2Cr2O7, 99.99%), HCOOH (85.0%), palladium chloride (PdCl2, 99.0%), sodium borohydride (NaBH4, 98.0%), and ethylene glycol (EG, C2H6O2, 99.0%) were purchased from S.D. Fine Chemicals Limited. Ti(OC3H7)4 (98.0%) and 4-methylbenzaldehyde (C8H8O, 99.0%) were obtained from Sigma-Aldrich. Deionized water was used throughout the experiment. 4.2. Synthesis of Precursors. The precursors were prepared according to the previous study reported by our group.75 4.2.1. Preparation of CdCl2(4-MebenztsczH)2 Precursor I. A total of 1.696 g (4.387 mmol) of 4-MebenztsczH dissolved in 15 mL of dry methanol and 0.804 g (4.386 mmol) of CdCl2 dissolved in 20 mL of dry tetrahydrofuran (THF) were taken in separate round-bottom flasks. The former solution was added to the latter under inert atmosphere and allowed to stir for 48 h. A white colored product was obtained. It was separated by evaporating the solvent under vacuum. It was then washed with cyclohexane (3 × 5 mL), followed by nhexane (3 × 5 mL) to remove impurities. Finally, the free white solid was obtained and characterized by elemental analysis, NMR, and FTIR (Yield: 2.389 g, 95.56%, mp 188 °C). Elemental analysis % found (cal.): C 36.91 (37.93), H 3.90 (3.53), N 13.96 (14.74), S 12.94 (11.25), Cl 12.38 (12.44), and Cd 20.82 (19.72). 1H NMR (δ in ppm): 2.3 (s, 3H, −CH3), 7.2−8.0 (m, 7H, −C6H5 + NH2), 11.3 (s, 1H, −NH−) (Figure S1a). 13C{1H} NMR (δ in ppm): 177.4 (>CS), 142.6 (>CN), 139.9, 131.3, 129.2, 127.2 (aromatic carbons); 21.3 (−CH3) (Figure S1b). IR: 3433, 3259 cm−1 (νNH2 assym and sym), 3162 cm−1 (νNH), 1536 cm−1 (νCN), 946 cm−1 (νCS) (Figure S3a). 4.2.2. Preparation of CdI2(4-MebenztsczH)2 Precursor II. In a round-bottom flask, 1.027 g (2.656 mmol) of 4MebenztsczH was dissolved in 15 mL of dry methanol. In another round-bottom flask, 0.973 g (2.657 mmol) of CdI2 was dissolved in 20 mL of dry THF. These solutions were then mixed together and then allowed to stir for 48 h. A white product obtained was separated by evaporating the solvent under vacuum. It was then washed with cyclohexane (3 × 5 mL), followed by n-hexane (3 × 5 mL) to remove any impurities present. Finally, the free white solid was obtained. It 18669

DOI: 10.1021/acsomega.8b02924 ACS Omega 2018, 3, 18663−18672

ACS Omega II) and Pd-based ternary Pd@CdS@TiO2 nanocomposites (TNC-I and TNC-II) for the catalytic reduction of Cr(VI) to Cr(III) according to the reported method.76 Typically, 5 mg of as-prepared Pd-based nanocatalysts in the mixture of potassium dichromate (K 2 Cr 2 O 7 , 150 ppm, 15 mL), HCOOH (85%, 1.5 mL), and deionized water (24 mL) was taken in a 50 mL beaker, and the mixture was vigorously stirred using a magnetic stirrer. During the reaction, the aliquots of 3 mL of aqueous solutions were withdrawn at each predetermined time interval for checking the catalytic reduction efficiency of the catalysts for the reduction of Cr(VI) to Cr(III) by measuring the absorbance. A UV-2450 PC Shimadzu UV−vis spectrophotometer was used for this purpose. For comparison, the measurements were also carried out without the catalyst under similar experimental conditions. 4.5. Material Characterization. FTIR spectroscopy was recorded on a PerkinElmer FTIR spectrometer using KBr pellets in the range of 400−4000 cm−1. A Thermo Finnigan, Italy FLASH EA 1112 series analyzer was used for elemental analysis. The 1H and 13C{1H} NMR spectra were recorded in DMSO-d6 on a Bruker AVANCE 300 spectrophotometer. The internal standard, tetramethylsilane, was used for 1H and 13 C{1H} NMR spectra. The UV−vis spectra were recorded on a UV-2450 PC Shimadzu UV−visible spectrophotometer. Powder XRD was carried out on an XRD-7000 Shimadzu Xray diffractometer with Cu Kα radiation at λ = 0.154060 nm. The Raman spectra were recorded on a Kaiser Optical Systems Inc. (KOSI) laser Raman spectrometer. The morphologies and EDS were observed using JEOL JSM-7600 FEG-SEM with an operating voltage of 0.1−30 kV. TEM and SAED were recorded on PHILIPS, CM 200 with an operating voltage between 20 and 200 kV. The XPS recording was carried out on an X-ray photoelectron spectrometer (AXIS Supra) Kratos Analytical, UK (SHIMADZU group) using (Al Kα) 600 W Xray source, 1486.6 eV. The UV-DRS spectra were recorded on a UV-2450 PC Shimadzu UV−visible spectrophotometer using barium sulfate (BaSO4) as the standard. A PerkinElmer LS 55 fluorescence spectrometer was used for recording PL spectra.



ACKNOWLEDGMENTS



REFERENCES

The authors are thankful to the Department of Science and Technology (DST-PURSE and DST-EMRF), India, for providing financial support. The authors acknowledge the Department of Physics, Indian Institute of Technology, Bombay (IITB), for providing XPS analysis and Department of Earth Science (IITB) for providing Raman Facility.

(1) Li, Y.; Jin, Z.; Li, T.; Xiu, Z. One-Step Synthesis and Characterization of Core-Shell Fe@SiO2 Nanocomposite for Cr(VI) Reduction. Sci. Total Environ. 2012, 421−422, 260−266. (2) Celebi, M.; Yurderi, M.; Bulut, A.; Kaya, M.; Zahmakiran, M. Palladium Nanoparticles Supported on Amine-Functionalized SiO2 for the Catalytic Hexavalent Chromium Reduction. Appl. Catal., B 2016, 180, 53−64. (3) Li, H.-C.; Liu, W.-J.; Han, H.-X.; Yu, H.-Q. Hydrophilic Swellable Metal-Organic Framework Encapsulated Pd Nanoparticles as an Efficient Catalyst for Cr(VI) Reduction. J. Mater. Chem. A 2016, 4, 11680−11687. (4) Liang, M.; Su, R.; Qi, W.; Zhang, Y.; Huang, R.; Yu, Y.; Wang, L.; He, Z. Reduction of Hexavalent Chromium Using Recyclable Pt/ Pd Nanoparticles Immobilized on Procyanidin-Grafted Eggshell Membrane. Ind. Eng. Chem. Res. 2014, 53, 13635−13643. (5) Wei, L.-L.; Gu, R.; Lee, J.-M. Highly Efficient Reduction of Hexavalent Chromium on Amino-Functionalized Palladium Nanowires. Appl. Catal., B 2015, 176−177, 325−330. (6) Wang, N.; Zhu, L.; Deng, K.; She, Y.; Yu, Y.; Tang, H. Visible Light Photocatalytic Reduction of Cr(VI) on TiO2 In Situ Modified with Small Molecular Weight Organic Acids. Appl. Catal., B 2010, 95, 400−407. (7) Meichtry, J. M.; Brusa, M.; Mailhot, G.; Grela, M. A.; Litter, M. I. Heterogeneous Photocatalysis of Cr(VI) in the Presence of Citric Acid Over TiO2 Particles: Relevance of Cr(V)-Citrate Complexes. Appl. Catal., B 2007, 71, 101−107. (8) Cheng, Q.; Wang, C.; Doudrick, K.; Chan, C. K. Hexavalent chromium removal using metal oxide photocatalysts. Appl. Catal., B 2015, 176−177, 740−748. (9) Ku, Y.; Lin, C.-N.; Hou, W.-M. Characterization of Coupled NiO/TiO2 Photocatalyst for Photocatalytic Reduction of Cr(VI) in Aqueous Solution. J. Mol. Catal. A: Chem. 2011, 349, 20−27. (10) Naimi-Joubani, M.; Shirzad-Siboni, M.; Yang, J.-K.; Gholami, M.; Farzadkia, M. Photocatalytic Reduction of Hexavalent Chromium with Illuminated ZnO/TiO2 Composite. Ind. Eng. Chem. Res. 2015, 22, 317−323. (11) Qamar, M.; Gondal, M. A.; Yamani, Z. H. Laser-Induced Efficient Reduction of Cr(VI) Catalyzed by ZnO Nanoparticles. J. Hazard. Mater. 2011, 187, 258−263. (12) Lv, X.; Xue, X.; Jiang, G.; Wu, D.; Sheng, T.; Zhou, H.; Xu, X. Nanoscale Zero-Valent Iron (nZVI) Assembled on Magnetic Fe3O4/ graphene for Chromium (VI) Removal from Aqueous Solution. J. Colloid Interface Sci. 2014, 417, 51−59. (13) Liu, X.; Pan, L.; Lv, T.; Sun, Z.; Sun, C. Enhanced Photocatalytic Reduction of Cr(VI) by ZnO-TiO2-CNTs Composites Synthesized via Microwave-Assisted Reaction. J. Mol. Catal. A: Chem. 2012, 363−364, 417−422. (14) Pawar, R. C.; Lee, C. S. Sensitization of CdS Nanoparticles onto Reduced Graphene Oxide (RGO) Fabricated by Chemical Bath Deposition Method for Effective Removal of Cr(VI). Mater. Chem. Phys. 2013, 141, 686−693. (15) Yadav, M.; Xu, Q. Catalytic Chromium Reduction using Formic Acid and Metal Nanoparticles Immobilized in a MetalOrganic Framework. Chem. Commun. 2013, 49, 3327−3329. (16) Gong, K.; Wang, W.; Yan, J.; Han, Z. Highly Reduced Molybdophosphate as a Noble Metal-Free Catalyst for Reduction of Chromium using Formic Acid as a Reducing Agent. J. Mater. Chem. A 2015, 3, 6019−6027.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.8b02924. 1

H and 13C{1H} NMR and FTIR spectra of precursors I/II; XRD, FTIR, UV-DRS, and SEM of BNC-I/II; HRTEM and EDS spectra of TNC-I/II; chromium reduction (without the catalyst and using BNC-I/II); recyclability of TNC-I/II; C/Co versus time plot for TNC-I/II in the dark condition; and ln C/Co versus time graph of TNCs and BNCs (PDF)





Article

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected], [email protected]. Phone: +91-22-2654 3368. Fax: +91-22-2652 85 47 (S.S.G.). ORCID

Shivram S. Garje: 0000-0003-1017-5333 Notes

The authors declare no competing financial interest. 18670

DOI: 10.1021/acsomega.8b02924 ACS Omega 2018, 3, 18663−18672

ACS Omega

Article

(17) Omole, M. A.; K’Owino, I. O.; Sadik, O. A. Palladium Nanoparticles for Catalytic Reduction of Cr(VI) using Formic Acid. Appl. Catal., B 2007, 76, 158−167. (18) Vincent, T.; Guibal, E. Chitosan-Supported Palladium Catalyst. 1. Synthesis Procedure. Ind. Eng. Chem. Res. 2002, 41, 5158−5164. (19) Zhang, R.; Liu, H.; Wang, B.; Ling, L. Insights into the Preference of CO2 Formation from HCOOH Decomposition on Pd Surface: A Theoretical Study. J. Phys. Chem. C 2012, 116, 22266− 22280. (20) Mori, K.; Dojo, M.; Yamashita, H. Pd and Pd-Ag Nanoparticles within a Macroreticular Basic Resin: An Efficient Catalyst for Hydrogen Production from Formic Acid Decomposition. ACS Catal. 2013, 3, 1114−1119. (21) Dandapat, A.; Jana, D.; De, G. Pd Nanoparticles Supported Mesoporous γ-Al2O3 Film as Reusable Catalyst for Reduction of Toxic CrVI to CrIII in Aqueous Solution. Appl. Catal., A 2011, 396, 34−39. (22) Fu, G.-T.; Jiang, X.; Wu, R.; Wei, S.-H.; Sun, D.-M.; Tang, Y.W.; Lu, T.-H.; Chen, Y. Arginine-Assisted Synthesis and Catalytic Property of Single-Crystalline Palladium Tetrapods. ACS Appl. Mater. Interfaces 2014, 6, 22790−22795. (23) Borah, B. J.; Saikia, H.; Bharali, P. Reductive Conversion of Cr(VI) to Cr(III) over Bimetallic CuNi Nanocrystals at Room Temperature. New J. Chem. 2014, 38, 2748−2751. (24) Shao, F.-Q.; Feng, J.-J.; Lin, X.-X.; Jiang, L.-Y.; Wang, A.-J. Simple fabrication of AuPd@Pd core-shell nanocrystals for effective catalytic reduction of hexavalent chromium. Appl. Catal., B 2017, 208, 128−134. (25) Han, S.-H.; Bai, J.; Liu, H.-M.; Zeng, J.-H.; Jiang, J.-X.; Chen, Y.; Lee, J.-M. The One-Pot Fabrication of Hollow and Porous Pd-Cu Alloy Nanospheres and Their Remarkably Improved Catalytic Performance for the Hexavalent Chromium Reduction. ACS Appl. Mater. Interfaces 2016, 8, 30948−30955. (26) World Health Organization. Chromium (Environmental Health Criteria 61); International Programme on Chemical Safety: Geneva, Switzerland, 1990. (27) Wu, J.-H.; Shao, F.-Q.; Han, S.-Y.; Bai, S.; Feng, J.-J.; Li, Z.; Wang, A.-J. Shape-controlled synthesis of well-dispersed platinum nanocubes supported on graphitic carbon nitride as advanced visiblelight-driven catalyst for efficient photoreduction of hexavalent chromium. J. Colloid Interface Sci. 2019, 535, 41−49. (28) Acharya, R.; Naik, B.; Parida, K. Cr(VI) remediation from aqueous environment through modified-TiO2-mediated photocatalytic reduction. Beilstein J. Nanotechnol. 2018, 9, 1448−1470. (29) Pawar, A. S.; Mlowe, S.; Garje, S. S.; Akerman, M. P.; Revaprasadu, N. Zinc Thiosemicarbazone Complexes: Single Source Precursors for Alkylamine Capped ZnS Nanoparticles. Inorg. Chim. Acta 2017, 463, 7−13. (30) Mlondo, S. N.; Revaprasadu, N.; Christian, P.; Helliwell, M.; O’Brien, P. Cadmium Thiosemicarbazide Complexes as Precursors for the Synthesis of Nanodimensional Crystals of CdS. Polyhedron 2009, 28, 2097−2102. (31) Castegnaro, M. V.; Kilian, A. S.; Baibich, I. M.; Alves, M. C. M.; Morais, J. On the Reactivity of Carbon Supported Pd Nanoparticles during NO Reduction: Unraveling a Metal-Support Redox Interaction. Langmuir 2013, 29, 7125−7133. (32) Niu, S.; Guo, W.; Lin, T.-W.; Yu, W.; Wu, Y.; Ji, X.; Shao, L. Nanoscale Pd Supported on 3D Porous Carbon for Enhanced Selective Oxidation of Benzyl Alcohol. RSC Adv. 2017, 7, 25885− 25890. (33) Tan, X.; Wu, X.; Hu, Z.; Ma, D.; Shi, Z. Synthesis and Catalytic Activity of Palladium Supported on Heteroatom Doped Single-Wall Carbon Nanohorns. RSC Adv. 2017, 7, 29985−29991. (34) Arabzadeh, A.; Salimi, A. One Dimensional CdS Nanowire@ TiO2 Nanoparticles Core-Shell as High Performance Photocatalyst for Fast Degradation of Dye Pollutants under Visible and Sunlight Irradiation. J. Colloid Interface Sci. 2016, 479, 43−54. (35) Pawar, R. C.; Khare, V.; Lee, C. S. Hybrid Photocatalysts Using Graphitic Carbon Nitride/Cadmium Sulfide/Reduced Graphene

Oxide (g-C3N4/CdS/RGO) for Superior Photodegradation of Organic Pollutants under UV and Visible Light. Dalton Trans. 2014, 43, 12514−12527. (36) Wan, M.; Li, W.; Long, Y.; Tu, Y. Electrochemical Determination of Tryptophan Based on Si-doped Nano-TiO 2 Modified Glassy Carbon Electrode. Anal. Methods 2012, 4, 2860− 2865. (37) Das, K.; De, S. K. Optical Properties of the Type-II Core-Shell TiO2@CdS Nanorods for Photovoltaic Applications. J. Phys. Chem. C 2009, 113, 3494−3501. (38) Sasikala, R.; Shirole, A. R.; Sudarsan, V.; Girija, K. G.; Rao, R.; Sudakar, C.; Bharadwaj, S. R. Improved Photocatalytic Activity of Indium Doped Cadmium Sulfide Dispersed on Zirconia. J. Mater. Chem. 2011, 21, 16566−16573. (39) Nusimovici, M. A.; Birman, J. L. Lattice Dynamics of Wurtzite: CdS. Phys. Rev. 1967, 156, 925−938. (40) Fan, H. M.; Fan, X. F.; Ni, Z. H.; Shen, Z. X.; Feng, Y. P.; Zou, B. S. Orientation-Dependent Raman Spectroscopy of Single Wurtzite CdS Nanowires. J. Phys. Chem. C 2008, 112, 1865−1870. (41) Tompsett, G. A.; Bowmaker, G. A.; Cooney, R. P.; Metson, J. B.; Rodgers, K. A.; Seakins, J. M. The Raman Spectrum of Brookite, TiO2 (Pbca, Z = 8). J. Raman Spectrosc. 2005, 26, 57−62. (42) Wei, Y.; Jiao, J.; Zhao, Z.; Zhong, W.; Li, J.; Liu, J.; Jiang, G.; Duan, A. 3D Ordered Macroporous TiO2-Supported Pt@CdS CoreShell Nanoparticles: Design, Synthesis and Efficient Photocatalytic Conversion of CO2 with Water to Methane. J. Mater. Chem. A 2015, 3, 11074−11085. (43) Mani, A. D.; Subrahmanyam, C. One Pot Synthesis of CdS/ TiO2 Hetero-Nanostructures for Enhanced H2 Production from Water and Removal of Pollutants from Aqueous Streams. Mater. Res. Bull. 2016, 73, 377−384. (44) Kumar, A. P.; Kumar, B. P.; Kumar, A. B. V. K.; Huy, B. T.; Lee, Y.-I. Preparation of Palladium Nanoparticles on Alumina Surface by Chemical Co-precipitation Method and Catalytic Applications. Appl. Surf. Sci. 2013, 265, 500−509. (45) Baylet, A.; Marécot, P.; Duprez, D.; Castellazzi, P.; Groppi, G.; Forzatti, P. In Situ Raman and In Situ XRD Analysis of PdO Reduction and Pd0 Oxidation Supported on γ-Al2O3 Catalyst under Different Atmospheres. Phys. Chem. Chem. Phys. 2011, 13, 4607− 4613. (46) Yogamalar, N. R.; Sadhanandham, K.; Bose, A. C.; Jayavel, R. Band Alignment and Depletion Zone at ZnO/CdS and ZnO/CdSe Hetero-Structures for Temperature Independent Ammonia Vapour Sensing. Phys. Chem. Chem. Phys. 2016, 18, 32057−32071. (47) Li, W.; Li, M.; Xie, S.; Zhai, T.; Yu, M.; Liang, C.; Ouyang, X.; Lu, X.; Li, H.; Tong, Y. Improving the Photoelectrochemical and Photocatalytic Performance of CdO Nanorods with CdS Decoration. CrystEngComm 2013, 15, 4212−4216. (48) Xie, S.; Lu, X.; Zhai, T.; Gan, J.; Li, W.; Xu, M.; Yu, M.; Zhang, Y.-M.; Tong, Y. Controllable Synthesis of ZnxCd1‑xS@ZnO CoreShell Nanorods with Enhanced Photocatalytic Activity. Langmuir 2012, 28, 10558−10564. (49) Majumder, S.; Sankapal, B. R. Facile Fabrication of CdS/CdSe Core-Shell Nanowires Heterostructure towards Solar Cell Application. New J. Chem. 2017, 41, 5808−5817. (50) Ma, F.; Zhao, G.; Li, C.; Wang, T.; Wu, Y.; Lv, J.; Zhong, Y.; Hao, X. Fabrication of CdS/BNNSs Nanocomposites with Broadband Solar Absorption for Efficient Photocatalytic Hydrogen Evolution. CrystEngComm 2016, 18, 631−637. (51) Xia, J.; Lu, X.; Gao, W.; Jiao, J.; Feng, H.; Chen, L. Hydrothermal Growth of Sn4+-doped FeS2 Cubes on FTO Substrates and its Photoelectrochemical Properties. Electrochim. Acta 2011, 56, 6932−6939. (52) Weng, B.; Quan, Q.; Xu, Y.-J. Decorating Geometry- and SizeControlled sub-20 nm Pd Nanocubes onto 2D TiO2 Nanosheets for Simultaneous H2 Evolution and 1, 1-Diethoxyethane Production. J. Mater. Chem. A 2016, 4, 18366−18377. (53) Yang, G.; Yang, B.; Xiao, T.; Yan, Z. One-Step Solvothermal Synthesis of Hierarchically Porous Nanostructured CdS/TiO2 18671

DOI: 10.1021/acsomega.8b02924 ACS Omega 2018, 3, 18663−18672

ACS Omega

Article

Photoelectrocatalytic bactericidal Effects on Escherichia Coli. Biomaterials 2010, 31, 3317−3326. (73) Sun, L.; Wu, W.; Yang, S.; Zhou, J.; Hong, M.; Xiao, X.; Ren, F.; Jiang, C. Template and Silica Interlayer Tailorable Synthesis of Spindle-like Multilayer α-Fe2O3/Ag/SnO2 Ternary Hybrid Architectures and Their Enhanced Photocatalytic Activity. ACS Appl. Mater. Interfaces 2014, 6, 1113−1124. (74) Zheng, S.; Jiang, W.; Rashid, M.; Cai, Y.; Dionysiou, D.; O’Shea, K. Selective Reduction of Cr(VI) in Chromium, Copper and Arsenic (CCA) Mixed Waste Streams Using UV/TiO2 Photocatalysis. Molecules 2015, 20, 2622−2635. (75) Palve, A. M.; Garje, S. S. Preparation of Zinc Sulfide Nanocrystallites from Single-Molecule Precursors. J. Cryst. Growth 2011, 326, 157−162. (76) Huang, Y.; Ma, H.; Wang, S.; Shen, M.; Guo, R.; Cao, X.; Zhu, M.; Shi, X. Efficient Catalytic Reduction of Hexavalent Chromium Using Palladium Nanoparticle-Immobilized Electrospun Polymer Nanofibers. ACS Appl. Mater. Interfaces 2012, 4, 3054−3061.

Heterojunction with Higher Visible Light Photocatalytic Activity. Appl. Surf. Sci. 2013, 283, 402−410. (54) Zhou, P.; Le, Z.; Xie, Y.; Fang, J.; Xu, J. Studies on Facile Synthesis and Properties of Mesoporous CdS/TiO2 Composite for Photocatalysis Applications. J. Alloys Compd. 2017, 692, 170−177. (55) Mali, S. S.; Devan, R. S.; Ma, Y.-R.; Betty, C. A.; Bhosale, P. N.; Panmand, R. P.; Kale, B. B.; Jadkar, S. R.; Patil, P. S.; Kim, J.-H.; Hong, C. K. Effective Light Harvesting in CdS Nanoparticle Synthesized Rutile TiO2 Microspheres. Electrochim. Acta 2013, 90, 666−672. (56) Wu, L.; Yu, J. C.; Fu, X. Characterization and Photocatalytic Mechanism of Nanosized CdS Coupled TiO2 Nanocrystals under Visible Light Irradiation. J. Mol. Catal. A: Chem. 2006, 244, 25−32. (57) Yu, J. C.; Yu, J.; Tang, H. Y.; Zhang, L. Effect of Surface Microstructure on the Photoinduced Hydrophilicity of Porous TiO2 Thin Films. J. Mater. Chem. 2002, 12, 81−85. (58) Navas, M. P.; Soni, R. K. Bromide (Br‑) Ion-Mediated Synthesis of Anisotropic Palladium Nanocrystals by Laser Ablation. Appl. Surf. Sci. 2016, 390, 718−727. (59) Kundu, S.; Yi, S.-I.; Ma, L.; Chen, Y.; Dai, W.; Sinyukov, A. M.; Liang, H. Morphology Dependent Catalysis and Surface Enhanced Raman Scattering (SERS) Studies using Pd Nanostructures in DNA, CTAB and PVA Scaffolds. Dalton Trans. 2017, 46, 9678−9691. (60) Lou, Y.; Shantz, D. F. Palladium Dendron Encapsulated Nanoparticles Grown from MCM-41 and SBA-15. J. Mater. Chem. A 2017, 5, 14070−14078. (61) Pawar, A. S.; Garje, S. S.; Revaprasadu, N. Synthesis and Characterization of CdS Nanocystallites and OMWCNT-Supported Cadmium Sulfide Composite and Their Photocatalytic Activity under Visible Light Irradiation. Mater. Chem. Phys. 2016, 183, 366−374. (62) Abaker, M.; Umar, A.; Baskoutas, S.; Dar, G. N.; Zaidi, S. A.; Al-Sayari, S. A.; Al-Hajry, A.; Kim, S. H.; Hwang, S. W. Highly Sensitive Ammonia Chemical Sensor Based on α-Fe2O3 Nanoellipsoids. J. Phys. D: Appl. Phys. 2011, 44, 425401. (63) Chang, Y.-C.; Chen, D.-H. Catalytic Reduction of 4Nitrophenol by Magnetically Recoverable Au Nanocatalyst. J. Hazard. Mater. 2009, 165, 664−669. (64) Jostar, T. S.; Devadason, S.; Suthagar, J. Effect of CdS Layers on Opto-electrical Properties of Chemically Prepared ZnS/CdS/TiO2 Photoanodes. Mater. Sci. Semicond. Process. 2015, 34, 65−73. (65) Li, Y.; Bian, Y.; Qin, H.; Zhang, Y.; Bian, Z. Photocatalytic reduction behavior of hexavalent chromium on hydroxyl modified titanium dioxide. Appl. Catal., B 2017, 206, 293−299. (66) Han, S.-H.; Liu, H.-M.; Sun, C.-C.; Jin, P.-J.; Chen, Y. Photocatalytic performance of AgCl@Ag core−shell nanocubes for the hexavalent chromium reduction. J. Mater. Sci. 2018, 53, 12030− 12039. (67) Meng, Q.; Zhou, Y.; Chen, G.; Hu, Y.; Lv, C.; Qiang, L.; Xing, W. Integrating both homojunction and heterojunction in QDs selfdecorated Bi2MoO6/BCN composites to achieve an efficient photocatalyst for Cr(VI) reduction. Chem. Eng. Sci. 2018, 334, 334−343. (68) Rakibuddin, M.; Ananthakrishnan, R. Novel Ag Deposited Nano Coordination Polymers Derived Porous SnO2/NiO Heteronanostructure for Enhanced Photocatalytic Reduction of Cr(VI) under Visible Light. New J. Chem. 2016, 40, 3385−3394. (69) Wang, X.; Hong, M.; Zhang, F.; Zhuang, Z.; Yu, Y. Recyclable Nanoscale Zero Valent Iron Doped g-C3N4/MoS2 for Efficient Photocatalytic of RhB and Cr(VI) Driven by Visible Light. ACS Sustainable Chem. Eng. 2016, 4, 4055−4063. (70) Wang, X.; Wang, J.; Geng, Z.; Qian, Z.; Han, Z. Phosphomolybdate Assembly as a low-cost Catalyst for the Reduction of Toxic Cr(VI) in Aqueous Solution. Dalton Trans. 2017, 46, 7917− 7925. (71) Tada, H.; Mitsui, T.; Kiyonaga, T.; Akita, T.; Tanaka, K. AllSolid-State Z-Scheme in CdS-Au-TiO2 Three-Component Nanojuncton System. Nat. Mater. 2006, 5, 782−786. (72) Kang, Q.; Lu, Q. Z.; Liu, S. H.; Yang, L. X.; Wen, L. F.; Luo, S. L.; Cai, Q. Y. A Ternary Hybrid CdS/Pt-TiO2 Nanotube Structure for 18672

DOI: 10.1021/acsomega.8b02924 ACS Omega 2018, 3, 18663−18672