Single-step sensitization of reduced graphene oxide sheets and CdS nanoparticles on ZnO nanorods as visible-light photocatalysts

https://doi.org/10.1016/j.apcatb.2013.06.022Get rights and content

Highlights

  • Facile fabrication of graphene and CdS sensitized ZnO nanorods.

  • Effective photodegradation of MB under visible light irradiation.

  • Excellent separation of photogenerated electron-hole pairs.

Abstract

Novel composite photocatalysts composed of reduced graphene oxide (RGO) sheets, CdS nanoparticles (CNPs), and ZnO nanorods were developed using a simple low-temperature chemical approach. The RGO sheets and CNPs were sensitized via single-step chemical bath deposition at 70 °C. Because of the incorporation of RGO into the CdS and ZnO structures, the photodegradation performance toward methylene blue (MB) degradation under visible light illumination was significantly improved compared with that of just CdS and ZnO. The highest kinetic rate constant (k = 0.028 min−1) using 0.20 wt% of added RGO was four times higher than that without graphene. This composite degraded MB within 40 min. Such improved performance resulted from effective separation of electron-hole pairs by the RGO sheets. Photoluminescence quenching, increased optical absorbance, and high specific surface area (23.08 m2 g−1) also contributed to the high degradation efficiency. An energy level diagram was constructed to explain how the RGO sheets minimized the recombination rate. The developed RGO–CdS–ZnO composites have a potential application in water purification devices.

Graphical abstract

Schematic of electron transportation from CdS nanoparticles to ZnO naorods and graphene sheets for MB photodegradation.

  1. Download : Download full-size image

Introduction

Heterogeneous photocatalysts play crucial roles in advanced water purification devices because of their broad applicabilities and better performances compared to homogeneous catalysts [1], [2]. Among the photocatalysts, ZnO has been shown to be promising because of its high electron mobility at room temperature (155 cm2 V−1 s−1), rich family of nanostructures, non-toxicity, low cost, and a bandgap that is similar to that of TiO2 [3], [4], [5], [6]. Despite their good performances and stabilities, ZnO-based photocatalysts have limited applications because of a wide bandgap energy (3.37 eV) and high recombination rate. Hence, various research groups have tried to improve its photocatalytic activity using different approaches such as element doping, sensitization with a visible bandgap semiconductor, and use of gold (Au) nanoparticles [7], [8], [9], [10]. Cho et al. developed visible (VIS) and ultraviolet (UV) photocatalysts based on a ZnSe-sensitized ZnO composite [11]. Kundu et al. explored the high absorptivity of ZnO nanorods sensitized with CdS particles and showed that the particle density coverage was important for increasing photocatalytic activity [12]. Yan et al. synthesized a ZnO/TiO2 core-brush structure and demonstrated that an interaction effect can reduce the recombination rate [13]. Han et al. compared the photocatalytic performance of ZnO nanorods and flowerlike structures and observed that the latter exhibited superior activity because of high surface areas and direct electron transport through branches [14]. Lai et al. reported that the larger oxygen vacancies in flowerlike ZnO led to a higher reaction rate and catalytic activity toward rhodamine B photodegradation [15]. Eskizeybek et al. reported an improvement in photodegradation efficiency in sunlight with ZnO nanostructures and polyaniline composites [16]. Bizarro studied Al-doped ZnO nanostructures and found that photocatalytic activity was affected by surface morphology [17]. Udawatte et al. synthesized Au/ZnO nanoparticle composites to minimize the photoelectron recombination rate [18]. However, these combinations suffered from low catalytic performance, low visible light absorptivity and stability, and high cost.

Recently, low bandgap semiconductor sensitization has been reported as the best option spanning the UV as well as the VIS portions of the solar spectrum [19], [20]. A low bandgap semiconductor can potentially use multiple electron-hole pair generation per incident photon to achieve higher photocatalytic activity [21]. Various researchers have proposed visible-bandgap semiconductors for effective sensitization, such as CdSe, InP, CdTe, PbS, and CdS [22], [23], [24], [25], [26]. Among these sensitizers, CdS is highly promising because of its reasonable bandgap (2.42 eV), which offers new opportunities for light harvesting. Recently, our group reported CdS-sensitized ZnO nanorods synthesized using a chemical bath deposition method and their use in a solar cell application [27]. Wang et al. improved the stability and catalytic activity by synthesizing a CdS–ZnO core–shell structure using a modified hydrothermal method [28]. Li et al. used a simple two-step process to prepare CdS–ZnO heterostructure photocatalysts having improved performance because of low recombination rates [29]. Although these had satisfactory photodegradation performance, it was not sufficient for commercial photocatalytic devices. Improving the photocatalytic efficiency is essential by avoiding recombination losses and promoting fast electron transportation.

With the discovery of RGO, a wide range of potential applications was expected because of its remarkable properties including high electron mobility at room temperature, large theoretical specific surface area (2630 m2 g−1), excellent thermal conductivity (3000–5000 Wm−1 K−1), good optical transparency (97.7%), and high Young's modulus (∼1 TPa) [30]. RGO has been combined with various semiconductors including ZnO, CdS, TiO2, Fe3O4, SnO2, Cu2O, and WO3, and its use has been explored in supercapacitors, solar cells, gas sensors, batteries, and photocatalysts [31], [32], [33], [34]. It plays a crucial role in photocatalysis for the effective separation of photogenerated electron-hole pairs because of its electron-capturing ability [35]. RGO may also be effective for the photodegradation of water pollutants, and methylene blue (MB) dye can be used as a test pollutant. We investigated the incorporation of RGO into CNPs and ZnO nanorods for the effective photodegradation of MB dye in visible light. A unique low-temperature, water-based method for the single-step sensitization of RGO and CdS on ZnO nanorods is described in this paper.

We measured the photocatalytic performance in visible light of composites composed of RGO, CdS, and ZnO. RGO sheets and CNPs were coated onto the surface of ZnO nanorods at 70 °C. We varied the weight percentage of RGO to study its effect on the photocatalytic activity of the CNPs and ZnO nanorods. We correlated the optoelectronic properties of the composites to their photocatalytic activities. A schematic diagram describing the electron transport through the RGO–CdS–ZnO composites was constructed based on these studies.

Section snippets

Preparation of ZnO nanorods

ZnO nanorods were grown using an aqueous chemical process. A uniform ZnO buffer layer was deposited onto ultrasonically cleaned glass and fluorine-doped tin oxide (FTO)-coated glass substrates. The seed solution was prepared in absolute ethanol from 0.05 M zinc acetate (Zn(CH3COO)2·2H2O, 99.5%) and 0.05 M diethanolamine (HN(CH2CH2OH)2, DEA, 99.5%). The clean substrate was dip-coated for 10 s in the seed solution and then dried overnight at room temperature. The dried films were annealed at 400 °C

XRD

Structural features of synthesized RGO powders and RGO–CdS–ZnO composite films with varying amounts of graphene were analyzed by XRD (Fig. 1). The ZnO nanorods film with and without RGO–CdS sensitization had similar diffraction patterns that corresponded to a hexagonal wurtzite crystal structure (JCPDS #01-079-2205). This indicated that the sensitization of graphene and CNPs did not affect the orientation and structure of the ZnO nanorods. The XRD pattern of RGO showed two peaks for the (0 0 2)

Conclusions

Composites made with RGO–CdS-sensitized ZnO nanorods have been successfully synthesized using a single-step chemical method at low temperature. FE-SEM and TEM analyses showed that the CNPs and ZnO nanorods were uniformly distributed across the surface of the RGO sheets. XPS spectra confirmed the formation of RGO–CdS–ZnO composites. The quenching of the visible emission peak in the room-temperature PL spectra indicated effective separation of photogenerated electron-hole pairs. The RGO played a

Acknowledgments

This research was supported by the Basic Science Research Program of the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (2011-0029862), by a Human Resources Development of the Korean Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Ministry of Knowledge Economy, Republic of Korea (No. 20124030200130) and by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MEST) (No.

References (54)

  • J.H. Sun et al.

    Journal of Hazardous Materials

    (2009)
  • M.J. Height et al.

    Applied Catalysis B: Environmental

    (2006)
  • R.C. Pawar et al.

    Materials Letters

    (2011)
  • M. Bizarro et al.

    Catalysis Today

    (2011)
  • S. Chakrabarti et al.

    Journal of Hazardous Materials

    (2008)
  • R. Qiu et al.

    Journal of Hazardous Materials

    (2008)
  • Z. Han et al.

    Journal of Hazardous Materials

    (2012)
  • Y. Lai et al.

    Applied Catalysis B: Environmental

    (2011)
  • V. Eskizeybek et al.

    Applied Catalysis B: Environmental

    (2012)
  • M. Bizarro

    Applied Catalysis B: Environmental

    (2010)
  • C. Li et al.

    Applied Catalysis A: General

    (2011)
  • J.S. Jang et al.

    Journal of Catalysis

    (2008)
  • A.J. Nozik

    Physica E: Low-dimensional Systems and Nanostructures

    (2002)
  • S.A. Vanalakar et al.

    Materials Letters

    (2011)
  • B. Li et al.

    Journal of Physics and Chemistry of Solids

    (2011)
  • S. Stankovich et al.

    Carbon

    (2007)
  • S.A. Vanalakar et al.

    Electrochimica Acta

    (2011)
  • A. Benayad et al.

    Chemical Physics Letters

    (2009)
  • R.C. Pawar et al.

    Sensors and Actuators B: Chemical

    (2010)
  • R.C. Pawar et al.

    Scripta Materialia

    (2013)
  • G. Zhu et al.

    Journal of Electroanalytical Chemistry

    (2011)
  • A. Fujishimaa et al.

    International Journal of Hydrogen Energy

    (2007)
  • W.Y. Teoh et al.

    The Journal of Physical Chemistry Letters

    (2012)
  • G. Kenanakisa et al.

    Catalysis Today

    (2010)
  • Q. Wang et al.

    Environmental Science Technology

    (2009)
  • S. Cho et al.

    Nanoscale

    (2012)
  • P. Kundu et al.

    Journal of Materials Chemistry

    (2011)
  • Cited by (198)

    View all citing articles on Scopus
    View full text