Abstract
Carbon nanotubes/akaganeite composite photocatalysts (CNTs/β-FeOOH) with different weight addition ratios of CNTs have been synthesized by a very simple hydrothermal method. X-ray diffraction, Fourier transform infrared spectrum, UV-vis diffuse reflectance spectra, scanning electron microscopy, and transmission electron microscope were used to characterize the composite photocatalysts. This study investigates the oxidative degradation of methyl orange, a probe contaminant, by a heterogeneous photo-Fenton-like process over the prepared composite photocatalysts. The CNTs/β-FeOOH composites exhibit remarkably enhanced photocatalytic activity toward the degradation of methyl orange in water under visible light irradiation as compared with bare β-FeOOH. The added CNTs have triple crucial roles in the promotion of photocatalytic activity of β-FeOOH. One is to trap electrons and effectively hinder recombination of the photoexcited carriers. Two is to act as a dispersing support to control the particle size of β-FeOOH in the CNTs/β-FeOOH composites. Three is to enhance the visible light absorption intensity. Finally, the catalytic mechanism of CNTs/β-FeOOH composite is proposed.
1 Introduction
In the practical application of Fenton or Fenton-like reaction, heterogeneous catalysts show a number of advantages over homogeneous catalysts due to the wider working pH range, easy recycling, and avoiding the generation of iron-containing sludge [1, 2]. Carbon nanotubes (CNTs) are a kind of sp2 carbon nanomaterials that exhibit unique properties such as high modulus, optical properties, electrical/thermal conductivity, and chemical stability [3, 4] and have lots of attractive applications in many important fields, including energy [5], electronics [6], biomedical area [7], and environment [8, 9]. In recent years, significant interest has been devoted to designing CNTs-semiconductor composite materials, aiming at enhancing their performance in heterogeneous photocatalysis [10–19].
Until now, there are only limited studies on akaganeite (β-FeOOH) as heterogeneous Fenton-like photocatalyst for water treatment. Zhao et al. [20] synthesized β-FeOOH coated resin and explored the catalytic activity for the degradation of 17β-estradiol under relatively weak UV irradiation. Yang et al. [21] researched surface acidity and reactivity of β-FeOOH/Al2O3 for pharmaceuticals degradation with ozone. In our previous work, the photocatalytic degradation of methyl orange over P25-supported β-FeOOH was studied under the ultraviolet and visible light irradiation [22]. For this study, we were motivated to fabricate and characterize the hybrid nanomaterials formed by nanostructured β-FeOOH onto single-walled CNTs. The photocatalytic activities of the hybrid nanomaterials (CNTs/β-FeOOH) were tested on the example of methyl orange (MO) degradation under visible light irradiation in the presence of H2O2. Enhancement of photocatalytic degradation of MO over CNTs/β-FeOOH composites was observed, and the photocatalytic oxidation mechanism in Fenton-like system was also discussed. To the best of our knowledge, this work may be the first report about utilizing single-walled CNTs supported β-FeOOH composite catalysts in the oxidation of MO with H2O2 under visible light irradiation.
2 Materials and methods
2.1 Catalyst preparation
All chemicals used in this work were of analytical grade, purchased from Shanghai Chemical Reagent Co., Ltd. (China), and used without any further purification.
Four CNTs/β-FeOOH materials, with the different content of single-walled CNTs, were obtained by impregnation-precipitation method. FeCl3·6H2O in the amount of 3.2435 g was added to a solution of 2.8800-g urea in 100-ml deionized water under permanent magnetic stirring for 30 min. Without adjusting the pH value, the required mass of single-walled CNTs was added to the above mixed solution. After additional stirring for 24 h, the obtained suspension was transferred into a Teflon-lined stainless steel autoclave (200 ml). The sealed autoclave was maintained at 90°C for 8 h and then cooled to room temperature naturally. The sample was collected by centrifugation, washed several times with absolute ethanol and deionized water, and finally dried at 60°C for 24 h. By applying this procedure, four samples with the single-walled CNTs weight of 0.010, 0.050, 0.10, and 0.20 g in the initial suspensions (denoted as 0.94%CNTs/β-FeOOH, 4.7%CNTs/β-FeOOH, 9.4%CNTs/β-FeOOH, and 18.8%CNTs/β-FeOOH) were prepared. The sample obtained by the same procedure but without addition of single-walled CNTs was denoted as β-FeOOH. For comparison, the photocatalyst denoted as 18.8%CNTs/β-FeOOH-mixing was also prepared by a simple mechanical mixing method. That is, β-FeOOH and CNTs were dispersed into 20 ml of anhydrous ethanol solution and ultrasonicated to make them mingle well. After that, this mixture was fully dried at 60°C in an oven for 24 h.
2.2 Photocatalytic degradation procedures
The photocatalytic degradation of MO was carried out in an XPA-7 photochemical reactor (Xujiang Electromechanical Plant, Nanjing, China). The temperature of reaction suspension was kept at 25±2°C by cooling water circulation. A 500-W xenon lamp was used as an irradiation source of visible light (VL). The initial pH values of suspensions were adjusted to 4.5 with dilute sulfuric acid solution and sodium hydroxide solution. The initial concentrations of MO and H2O2 were 80 mg l-1 and 0.3 g l-1, respectively, with a catalyst loading of 0.4 g l-1. All catalysts were dipped in MO solution and stirred in the dark for 60 min to establish adsorption/desorption equilibrium between the solution and the catalysts before irradiation. At given irradiation time intervals, a small quantity of the suspension was taken and centrifuged to separate the catalyst particles from the suspension. The concentration of MO was determined by using a UV-vis spectrophotometer (Beijing Ruili Corp., UV-9100) at 464 nm.
2.3 Catalyst characterization
The powder X-ray diffraction patterns (XRD) were recorded at a scanning rate of 4° min-1 in the 2θ range of 10–70° using a Bruker D8 Advance instrument with Cu-Kα radiation (λ=1.5406 Å) at room temperature. The morphologies and nanostructures of synthesized products were further observed using a Hitachi H-7650 transmission electron microscope (TEM) at the acceleration voltage of 80 kV. Fourier transform infrared (FTIR) spectrum measurements were performed on a Bruker Vector 22 FT-IR spectrophotometer at room temperature [23], with scanning from 4000 to 400 cm-1 using KBr pellets. The light reflectance property was studied with a Cary50 UV-vis-NIR spectrophotometer (Varian Australia PYT Ltd.), in which BaSO4 was employed as the internal reflectance standard. The photoluminescence (PL) spectra for solid samples were investigated on a Cary Eclipse Fluorescence spectrophotometer with an excitation wavelength of 360 nm.
3 Results and discussion
3.1 Structural and morphological characterization
Figure 1A depicts the XRD patterns of β-FeOOH and CNTs/β-FeOOH composites with different weight addition ratios of CNTs (0.94%, 4.7%, 9.4%, and 18.8%) to indicate the crystalline phase of akaganeite in the composites and the effects of CNTs on the crystalline of akaganeite. All the peaks produced by the β-FeOOH sample matched the diffraction of tetragonal akaganeite (JCPDS card No. 34-1266) very well with cell constants of a0=10.51 Å, b0=10.51 Å, and c0=3.033 Å, and no impurity peak can be detected. No typical diffraction peaks of CNTs are obviously observed in the CNTs/β-FeOOH composites. This is because the characteristic peak of CNTs (002) planes at 26.2° is probably overlapped with the (310) diffraction peak of akaganeite, which is consistent with the previous reports [16]. It is also clear to see that the diffraction peaks of akaganeite phase become weaker and broader when increasing the weight addition ratios of CNTs in the composites, which means that the akaganeite particle size reduces.

XRD patterns of the as-prepared β-FeOOH and CNT/β-FeOOH composites (A); characteristic parts of the FTIR spectra of CNTs and CNTs/β-FeOOH composites (B); TEM images of 0.94%CNTs/β-FeOOH (C); 4.7%CNTs/β-FeOOH (D); 9.4%CNTs/β-FeOOH (E); 18.8%CNTs/ β-FeOOH (F); UV-vis DRS of β-FeOOH and CNTs/β-FeOOH composites (G).
The as-prepared photocatalysts were analyzed by FTIR for further identification. As shown in Figure 1B, the absorption peaks at wavelength 842, 696, 649, and 492 cm-1 were assigned to the vibration modes of the FeO6 coordination octahedron [24]. The presence of the absorption peaks at 1639 cm-1 was assigned to O-H bending and aromatic C=C stretching (skeletal ring vibration) [15]. A band of wavenumber at 1402 cm-1 is most probably due to sorbed water and overlapping bands in region characteristic for carboxyl C-O groups [25]. Features at 1454 cm-1 attributed to CNTs vibrational modes (C-C stretching) are also apparent [25]. The FTIR spectrum of CNTs/β-FeOOH composites shows absorption peaks at 1656 and 1612 cm-1 corresponding to the stretching vibration of C=O from the carboxylic groups [26]. Moreover, the strong absorption bands observed in the range of 400–1000 cm-1 corresponded to Fe-O-Fe bonding, remarkably weakened with the increase of the weight addition ratios of CNTs. It suggests the presence of Fe-O-C bonds in the composites, indicating the chemical interaction between surface hydroxyl groups of akaganeite and functional groups of CNTs [13, 27].
Figure 1C–F display TEM images of the CNTs/β-FeOOH composites. As seen in these figures, the particle size of β-FeOOH decreases with the increase of the weight addition ratios of CNTs. This means that added CNTs have significant impact on the product particle size. During the synthesis of composites, CNTs can serve as heterogeneous nuclei for the growth of akaganeite, while the existence of CNTs can also inhibit the growth of akaganeite nanoparticles. It is in accordance with the XRD result. In particular, it is worth noting that the sample of 18.8%CNTs/β-FeOOH consists of long CNTs decorated with akaganeite nanoparticles on the surface (Figure 1F). The direct contact interface between akaganeite and CNTs is beneficial for the charge carriers transfer and separation, which might improve the photocatalytic performance of the 18.8%CNTs/β-FeOOH composite. Similar structure has been reported in a previous study [28]. However, the added amount of CNTs is much larger than that of the reference [28]. There are two possible reasons. One is that the concentration of iron ions in the reaction suspension is quite high. The other is that the diameter of single-walled CNT is too small compared with that of akaganeite. Therefore, the nanoparticles of akaganeite may grow and assemble on multiple parallel single-walled CNTs.
Figure 1G shows the ultraviolet-visible light diffuse reflectance spectra (UV-vis DRS) of β-FeOOH and CNTs/ β-FeOOH nanocomposites. Apparently, the visible light absorption increases with the increase of CNTs amount. The results demonstrate the significant influence of CNTs on the optical properties of CNTs/β-FeOOH composites. Similar to the case of TiO2-CNT or TiO2 chemically converted graphene composites, the phenomena in this study can be ascribed to the formation of Fe-O-C chemical bonding in the prepared composites [10, 13], which is also confirmed by the FTIR results.
3.2 Photocatalytic activity of CNTs/β-FeOOH nanocomposites
Figure 2 shows the photocatalytic activity of bare β-FeOOH and CNTs/β-FeOOH nanocomposites with different weight ratios of CNTs toward oxidation MO under visible light irradiation. It can be observed that the weight addition ratio of CNTs exhibits a significant influence on the photocatalytic activity of CNTs/β-FeOOH composites. Even with a small-amount addition of CNTs (0.94%), the photocatalytic activity of sample is noticeably increased. The degradation efficiency of MO follows the following order: 18.8%CNTs/β-FeOOH>9.4%CNTs/β-FeOOH>4.7%CNTs/ β-FeOOH>0.94%CNTs/β-FeOOH>bare β-FeOOH. The degradation efficiency of MO facilitated by 18.8%CNTs/ β-FeOOH was 88.9% after 120 min. Compared with previous studies (Table 1), the Fenton-like composite catalyst of 18.8%CNTs/β-FeOOH prepared in this work shows quite good VL catalytic performance.
![Figure 2: Dynamic curves of MO degradation over bare β-FeOOH and CNTs/β-FeOOH composites with different weight additions of CNTs under VL irradiation; (A) 18.8%CNTs/β-FeOOH+tert-butyl alcohol; (B) β-FeOOH; (C) 9.4%CNTs/β-FeOOH-mixing; (D) 18.8%CNTs/β-FeOOH without H2O2; (E) 0.94%CNTs/β-FeOOH; (F) 4.7%CNTs/β-FeOOH; (G) 9.4%CNTs/β-FeOOH; (H) 18.8%CNTs/β-FeOOH. Test conditions: pH=4.5; [H2O2]=0.3 g l-1; catalyst loading=0.4 g l-1.](/document/doi/10.1515/secm-2015-0212/asset/graphic/j_secm-2015-0212_fig_002.jpg)
Dynamic curves of MO degradation over bare β-FeOOH and CNTs/β-FeOOH composites with different weight additions of CNTs under VL irradiation; (A) 18.8%CNTs/β-FeOOH+tert-butyl alcohol; (B) β-FeOOH; (C) 9.4%CNTs/β-FeOOH-mixing; (D) 18.8%CNTs/β-FeOOH without H2O2; (E) 0.94%CNTs/β-FeOOH; (F) 4.7%CNTs/β-FeOOH; (G) 9.4%CNTs/β-FeOOH; (H) 18.8%CNTs/β-FeOOH. Test conditions: pH=4.5; [H2O2]=0.3 g l-1; catalyst loading=0.4 g l-1.
Previous studies conducted on the VL catalytic degradation of MO over different Fenton-like catalysts.
Photocatalysts | MO degradation efficiency | Testing conditions | Ref. |
---|---|---|---|
25%TiO2/β-FeOOH | 86.3% | [MO]=80 mg/l | [22] |
(after 120 min) | [photocatalyst]=400 mg/l | ||
V (H2O2, 30%)=0.05 ml | |||
pH=4.5 | |||
Light source: 500-W xenon lamp | |||
5%β-FeOOH/TiO2 | 97% | [MO]=80 mg/l | [29] |
(after 60 min) | [photocatalyst]=200 mg/l | ||
V (H2O2, 30%)=3 ml | |||
pH=5 | |||
Light source: 300-W halogen tungsten lamp | |||
TiO2/Hydroniumjarosite | 48.7% | [MO]=80 mg/l | [30] |
(after 60 min) | [photocatalyst]=200 mg/l | ||
V (H2O2, 30%)=0.05 ml | |||
pH=4.5 | |||
Light source: 500-W xenon lamp | |||
Bi2Fe4O9 | 84% | [MO]=20 mg/l | [31] |
(after 120 min) | [photocatalyst]=500 mg/l | ||
V (H2O2, 30%)≈0.01 ml | |||
pH: not mentioned in the article | |||
Light source: 500-W xenon lamp |
3.3 Possible mechanism involved in the oxidation of MO
Previous studies have demonstrated that ·OH radicals produced in H2O2-iron oxide systems showed a strong capacity to oxidize organics [32–34]. In order to identify whether ·OH was the key oxidant in the oxidation of MO catalyzed by CNTs/β-FeOOH composites, the photocatalytic degradation of MO was repeated with the addition of excess 2-propanol as an ·OH scavenger. Addition of tert-butyl alcohol (2% v/v) in the 18.8%CNTs/β-FeOOH/VL system inhibited the degradation of MO (Figure 2A), which confirms the critical role of hydroxyl radical for MO degradation. However, OH radicals were possibly derived from the photocatalytic process. Therefore, additional experiment in absence of H2O2 was carried out and the results are shown in Figure 2D. We can clearly see that MO was degraded by photocatalytic reaction, which exhibited the contribution of about 43% for the degradation of MO in the presence of H2O2. This means that both photocatalytic reaction and photo-Fenton-like reaction are responsible for the degradation of MO, which is shown in Figure 3. Based on previous studies [35–38], the degradation process of MO catalyzed by CNTs/β-FeOOH composites can be explained by the following mechanisms (“≡” signified the surface structure of catalysts):

The photocatalytic schematic diagram of CNTs/βFeOOH samples.
3.4 Analysis of improved catalytic performance of β-FeOOH
On the basis of the above characterizations and discussions, the significant enhancement of photoactivity after the hybridization of β-FeOOH with CNTs can be attributed to the following three factors. Firstly, the photogenerated electrons from β-FeOOH under irradiation can be accepted by CNTs [14, 19], thereby decreasing the recombination probability of the photoexcited electron-hole pairs. To better understand the importance of interfacial interaction between β-FeOOH and CNTs on the degradation efficiently of MO, the sample of 18.8%CNTs/β-FeOOH-mixing was prepared. As seen in Figure 2C, 18.8%CNTs/β-FeOOH-mixing shows much worse photoactivity than 18.8%CNTs/β-FeOOH toward the oxidation of MO under identical reaction conditions. This result suggests that the interfacial interaction at the direct contact between β-FeOOH and CNTs can efficiently contribute to the oxidation of MO, whereas this effect cannot be obtained over 18.8%CNTs/β-FeOOH-mixing with poor interfacial interaction. Actually, CNTs have 1D structure of carbon-based nanocylinders, where the electrons can move freely without any scattering from atoms or defects [27]. Furthermore, 18.8%CNTs/ β-FeOOH composite shows decreased PL intensity than that of bare β-FeOOH (displayed in Figure 4A), suggesting the effectively reduced charge recombination. Secondly, the presence of CNTs in the composite materials is able to enhance the visible light absorption intensity as reflected by the above UV-vis DRS, which is beneficial to both photocatalytic and photo-Fenton-like reaction. Thirdly, CNTs act as a dispersing template or support to decrease the particle size of β-FeOOH in the synthesis process as confirmed by the above TEM images. It is well known that the small-sized catalyst particles have larger specific area, which ensures efficient light absorption and provides more photoreactive sites for degradation reactions [14].

Photoluminescence spectra of CNTs, 18.8%CNTs/β-FeOOH and β-FeOOH (A); repeated use of 18.8%CNTs/β-FeOOH in the photocatalytic oxidation of MO under visible light irradiation (B).
3.5 Stability and reusability of 18.8%CNTs/ β-FeOOH composite
The stability and reusability of 18.8%%CNTs/β-FeOOH catalyst were evaluated by four successive recycling tests for the oxidation of MO under visible light irradiation, as displayed in Figure 4B. After each cycle, the 18.8%%CNTs/β-FeOOH catalyst was directly centrifuged and entered into the next cycle without any other post treatment. After four cycles of repeated use, 83.8% percentage of MO can still be degraded, which was basically close to the level of the first trial. This indicates that the reuse of the composite catalyst is cost-effective because the catalyst does not have to be replaced over a relatively long period.
4 Conclusions
To sum up, a series of CNTs/β-FeOOH composites with different added ratios of CNTs have been synthesized by a simple hydrothermal method. It is found that CNTs/ β-FeOOH composites exhibit remarkably enhanced visible light photocatalytic activity than bare β-FeOOH toward the oxidation of MO in the aqueous phase. The enhanced visible light catalytic activity can be attributed to the formation of CNTs/β-FeOOH hybrid structure, which plays an important role in inhibiting the recombination of photo-excited electrons and holes. In addition, the increase in visible light absorption intensity and the decrease of β-FeOOH particle size are conducive to both photocatalytic and photo-Fenton-like reaction for the degradation of MO. It would be of great promise for the industrial application of this composite catalyst for the treatment of wastewater containing MO.
Funding source: National Natural Science Foundation of China
Award Identifier / Grant number: 41371476
Award Identifier / Grant number: 21477054
Funding statement: This study was financially supported by the National Natural Science Foundation of China (41371476, 21477054).
Acknowledgments:
This study was financially supported by the National Natural Science Foundation of China (41371476, 21477054).
References
[1] Xu ZH, Liang JR, Zhou LX. J. Alloys Compd. 2013, 546, 112–118.10.1016/j.jallcom.2012.08.087Search in Google Scholar
[2] Huang CP, Huang YH. Appl. Catal. A Gen. 2008, 346, 140–148.10.1016/j.apcata.2008.05.017Search in Google Scholar
[3] Xu BY, Ju Y, Cui YB, Song GB. Mater. Sci. Eng. C 2015, 51, 182–188.10.1016/j.msec.2015.02.052Search in Google Scholar PubMed
[4] Ma X, Yang ST, Tang H, Liu YF, Wang HF. J. Colloid Interface Sci. 2015, 448, 347–355.10.1016/j.jcis.2015.02.042Search in Google Scholar PubMed
[5] De Volder MF, Tawfick SH, Baughman RH, Hart AJ. Science 2013, 339, 535–539.10.1126/science.1222453Search in Google Scholar PubMed
[6] Tamura H, Martinazzo R, Ruckenbauer M, Burghardt I. J. Chem. Phys. 2012, 137, 22A540.10.1063/1.4751486Search in Google Scholar PubMed
[7] Cao A, Liu Z, Chu S, Wu M, Ye Z, Cai Z, Chang Y, Wang S, Gong Q, Liu Y. Adv. Mater. 2010, 22, 103–106.10.1002/adma.200901920Search in Google Scholar PubMed
[8] Liu Z, Tabakman S, Welsher K, Dai H. Nano Res. 2009, 2, 85–120.10.1007/s12274-009-9009-8Search in Google Scholar PubMed PubMed Central
[9] Yang ST, Chen S, Chang Y, Cao A, Liu Y, Wang H. J. Colloid Interface Sci. 2011, 359, 24–29.10.1016/j.jcis.2011.02.064Search in Google Scholar PubMed
[10] Yu Y, Yu JC, Yu JG, Kwok YC, Che YK, Zhao JC, Ding L, Ge WK, Wong PK. Appl. Catal. A Gen. 2005, 289, 186–196.10.1016/j.apcata.2005.04.057Search in Google Scholar
[11] Ma LL, Sun HZ, Zhang YG, Lin YL, Li JL, Wang E.K, Yu Y, Tan M, Wang JB. Nanotechnol. 2008, 19, 115709.10.1088/0957-4484/19/11/115709Search in Google Scholar PubMed
[12] Yao Y, Li G, Ciston S, Lueptow RM, Gray KA. Environ. Sci. Technol. 2008, 42, 4952–4957.10.1021/es800191nSearch in Google Scholar PubMed
[13] Zhang H, Lv X, Li Y, Wang Y, Li J. ACS Nano 2010, 4, 380–386.10.1021/nn901221kSearch in Google Scholar PubMed
[14] Xu YJ, Zhuang Y, Fu X. J. Phys. Chem.C 2010, 114, 2669–2676.10.1021/jp909855pSearch in Google Scholar
[15] Nguyen-Phan TD, Pham VH, Shin EW, Pham HD, Kim S, Chung JS, Kim EJ, Hur SH. Chem. Eng. J. 2011, 170, 226–232.10.1016/j.cej.2011.03.060Search in Google Scholar
[16] Peng T, Zeng P, Ke D, Liu X, Zhang X. Energy Fuels 2011, 25, 2203–2210.10.1021/ef200369zSearch in Google Scholar
[17] Zhang Y, Zhang N, Tang ZR, Xu YJ. ACS Nano 2012, 6, 9777–9789.10.1021/nn304154sSearch in Google Scholar PubMed
[18] Chen Z, Zhang N, Xu YJ. Cryst. Eng. Comm. 2013, 15, 3022–3030.10.1039/c3ce27021aSearch in Google Scholar
[19] Weng B, Liu SQ, Zhang N, Tang ZR, Xu YJ. J. Catal. 2014, 309, 146–155.10.1016/j.jcat.2013.09.013Search in Google Scholar
[20] Zhao YP, Hu JY, Chen HB. J. Photochem. Photobiol. A Chem. 2010, 212, 94–100.10.1016/j.jphotochem.2010.04.001Search in Google Scholar
[21] Yang L, Hu C, Nie YL, Qu JH. Appl. Catal. B Environ. 2010, 97, 340–346.10.1016/j.apcatb.2010.04.014Search in Google Scholar
[22] Xu ZH, Zhang M, Wu JY, Liang JR, Zhou LX, Lǚ B. Water Sci. Technol. 2013, 68, 2178–2185.10.2166/wst.2013.475Search in Google Scholar PubMed
[23] Xu H, Alur S, Wang YQ, Cheng AJ, Kang K, Sharma Y, Park M, Ahyi C, Williams J, Gu CK, Hanser A, Paskova T, Preble EA, Evans KR, Zhou Y. J. Electron. Mater. 2010, 39, 2237–2242.10.1007/s11664-010-1304-3Search in Google Scholar
[24] Ristic M, Music S, Orehovec Z. J. Mol. Struct. 2005, 744, 295–300.10.1016/j.molstruc.2004.10.051Search in Google Scholar
[25] Misra A, Tyagi PK, Singh MK, Misra DS. Diamond Relat. Mater. 2006, 15, 385–388.10.1016/j.diamond.2005.08.013Search in Google Scholar
[26] Singh S, Singh BR, Khan W, Naqvi AH. J. Mol. Struct. 2014, 1056–1057, 194–201.10.1016/j.molstruc.2013.10.020Search in Google Scholar
[27] Williams G, Seger B, Kamat PV. ACS Nano 2008, 2, 1487–1491.10.1021/nn800251fSearch in Google Scholar PubMed
[28] Wang ZY, Luan DY, Madhavi S, Hu Y, Lou XW. Energy Environ. Sci. 2012, 5, 5252–5256.10.1039/C1EE02831FSearch in Google Scholar
[29] Chowdhury M, Ntiribinyange M, Nyamayaro K, Fester V. Mater. Res. Bull. 2015, 68, 133–141.10.1016/j.materresbull.2015.03.044Search in Google Scholar
[30] Xu ZH, Fang D, Shi WC, Xu JY, Lu AM, Wang KB, Zhou LX. Environ. Eng. Sci. 2015, 32, 497–504.10.1089/ees.2014.0404Search in Google Scholar
[31] Zhang H, Tong T, Cao WH, Chen JG, Jin DR, Cheng JR. J. Sol-Gel Sci. Technol. 2015, 75, 481–485.10.1007/s10971-015-3751-zSearch in Google Scholar
[32] Zhao YP, Hu JY, Jin W. Environ. Sci. Technol. 2008, 42, 5277–5284.10.1021/es703253qSearch in Google Scholar PubMed
[33] Han SK, Hwang TM, Yoon YJ, Kang JW. Chemosphere 2011, 84, 1095–1101.10.1016/j.chemosphere.2011.04.051Search in Google Scholar PubMed
[34] Zhao YP, Hu JY. Appl. Catal. B Environ. 2008, 78, 250–258.10.1016/j.apcatb.2007.09.026Search in Google Scholar
[35] Lin SS, Gurol MD. Environ. Sci. Technol. 1998, 32, 1417–1423.10.1021/es970648kSearch in Google Scholar
[36] Kwan WP, Voelker BM. Environ. Sci. Technol. 2002, 36, 1467–1476.10.1021/es011109pSearch in Google Scholar PubMed
[37] Yang L, Ma QL, Cai YG, Huang YM. Appl. Surf. Sci. 2014, 292, 297–300.10.1016/j.apsusc.2013.11.134Search in Google Scholar
[38] Gu CK, Shannon C. J. Mol. Catal. A: Chem. 2007, 262, 185–189.10.1016/j.molcata.2006.08.029Search in Google Scholar
©2018 Walter de Gruyter GmbH, Berlin/Boston
This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.
Articles in the same Issue
- Frontmatter
- Review
- A model on the curved shapes of unsymmetric laminates including tool-part interaction
- Original articles
- Enhanced catalytic performance of β-FeOOH by coupling with single-walled carbon nanotubes in a visible-light-Fenton-like process
- Investigation of the microstructure and properties of W75-Cu/W55-Cu brazed joint with Cu-Mn-Co filler metal
- In situ polymerization approach to poly(arylene ether nitrile)-functionalized multiwalled carbon nanotube composite films: thermal, mechanical, dielectric, and electrical properties
- Fabrication and characterization properties of polypropylene/polycarbonate/clay nanocomposites prepared with twin-screw extruder
- Torsional vibration of functionally graded carbon nanotube reinforced conical shells
- A comparative study performance of cationic organic montmorillonite prepared by different methods
- Design and analysis with different substrate materials of a new metamaterial for satellite applications
- An investigation on dry sliding wear behaviour of pressure infiltrated AA1050-XMg/B4C composites
- Investigation of mechanical performances of composite bolted joints with local reinforcements
- Effect of alkali treatment on the flexural properties of a Luffa cylindrica-reinforced epoxy composite
- Thermal expansion, electrical conductivity and hardness of Mn3Zn0.5Sn0.5N/Al composites
- Effect of modified layered double hydroxide on the flammability and mechanical properties of polypropylene
- A unified formulation for free vibration of functionally graded plates
- Friction-stir welding of aluminum alloy with an iron-based metal as reinforcing material
- Hybridization effect of coir fiber on physico-mechanical properties of polyethylene-banana/coir fiber hybrid composites
- Micromechanical properties of unidirectional composites filled with single and clustered shaped fibers
- Structure and microwave absorbing properties of carbon-filled ultra-high molecular weight polyethylene
- Investigation and optimization of the electro-discharge machining parameters of 2024 aluminum alloy and Al/7.5% Al2O3 particulate-reinforced metal matrix composite
- Structural behavior of load-bearing sandwich wall panels with GFRP skin and a foam-web core
- Synthesis, thermal and magnetic behavior of iron oxide-polymer nanocomposites
- High-temperature damping capacity of fly ash cenosphere/AZ91D Mg alloy composites
- Investigation of penetration into woven fabric specimens impregnated with shear thickening fluid
Articles in the same Issue
- Frontmatter
- Review
- A model on the curved shapes of unsymmetric laminates including tool-part interaction
- Original articles
- Enhanced catalytic performance of β-FeOOH by coupling with single-walled carbon nanotubes in a visible-light-Fenton-like process
- Investigation of the microstructure and properties of W75-Cu/W55-Cu brazed joint with Cu-Mn-Co filler metal
- In situ polymerization approach to poly(arylene ether nitrile)-functionalized multiwalled carbon nanotube composite films: thermal, mechanical, dielectric, and electrical properties
- Fabrication and characterization properties of polypropylene/polycarbonate/clay nanocomposites prepared with twin-screw extruder
- Torsional vibration of functionally graded carbon nanotube reinforced conical shells
- A comparative study performance of cationic organic montmorillonite prepared by different methods
- Design and analysis with different substrate materials of a new metamaterial for satellite applications
- An investigation on dry sliding wear behaviour of pressure infiltrated AA1050-XMg/B4C composites
- Investigation of mechanical performances of composite bolted joints with local reinforcements
- Effect of alkali treatment on the flexural properties of a Luffa cylindrica-reinforced epoxy composite
- Thermal expansion, electrical conductivity and hardness of Mn3Zn0.5Sn0.5N/Al composites
- Effect of modified layered double hydroxide on the flammability and mechanical properties of polypropylene
- A unified formulation for free vibration of functionally graded plates
- Friction-stir welding of aluminum alloy with an iron-based metal as reinforcing material
- Hybridization effect of coir fiber on physico-mechanical properties of polyethylene-banana/coir fiber hybrid composites
- Micromechanical properties of unidirectional composites filled with single and clustered shaped fibers
- Structure and microwave absorbing properties of carbon-filled ultra-high molecular weight polyethylene
- Investigation and optimization of the electro-discharge machining parameters of 2024 aluminum alloy and Al/7.5% Al2O3 particulate-reinforced metal matrix composite
- Structural behavior of load-bearing sandwich wall panels with GFRP skin and a foam-web core
- Synthesis, thermal and magnetic behavior of iron oxide-polymer nanocomposites
- High-temperature damping capacity of fly ash cenosphere/AZ91D Mg alloy composites
- Investigation of penetration into woven fabric specimens impregnated with shear thickening fluid