Anda di halaman 1dari 6

Journal of Physics and Chemistry of Solids 72 (2011) 15481553

Contents lists available at SciVerse ScienceDirect

Journal of Physics and Chemistry of Solids


journal homepage: www.elsevier.com/locate/jpcs

Synthesis of ZnO nanospheres with uniform nanopores by a hydrothermal process


Yongzhe Zhang a,b, Jooyoung Chung a, Juneyoung Lee a, Jihyun Myoung a, Sangwoo Lim a,n
a b

Department of Chemical and Biomolecular Engineering, Yonsei University, 134 Shinchon-dong, Seodaemoon-gu, Seoul 120-749, Korea Department of Physics, Jinan University, Guangzhou 510632, China

a r t i c l e i n f o
Article history: Received 26 October 2010 Received in revised form 22 June 2011 Accepted 25 September 2011 Available online 1 October 2011 Keywords: A. Nanostructures B. Crystal growth C. Electron microscopy C. X-ray Diffraction D. Surface properties

a b s t r a c t
ZnO nanospheres were successfully synthesized by a hydrothermal process (S1 sample) and a wetchemical method (S2 sample). Following synthesis, calcination treatment at 450 1C was performed for the sample prepared by the wet-chemical method (S3 sample). All of the samples possessed a regular spherical shape. A polycrystalline wurtzite structure was conrmed in the S1 and S3 samples by X-ray diffraction and selected area electron diffraction, whereas a mixture of ZnO nanoparticles and amorphous materials was observed in the sample S2. The surface area and pore structure of the samples were investigated by nitrogen adsorptiondesorption measurements. Uniform nanopores with a diameter of 4.07 nm were present in the S1 sample while a broad pore size distribution was obtained for the S2 and S3 samples. The highest surface area was obtained for the S1 sample and a possible formation mechanism was studied. & 2011 Elsevier Ltd. All rights reserved.

1. Introduction Nanostructured ZnO has attracted signicant attention due to its unique properties in light-emitting devices, photodetectors, solar cells, photocatalysts, and sensors [1]. In particular, nanosphere ZnO consisting of numerous ZnO nanoparticles appears to be a promising candidate in photochemical and electrochemical devices such as dye-sensitized solar cells (DSSCs) and biosensors [27]. Its high surface area, abundant surface chemical groups, and good crystal structure are three important factors for these applications. In the cases of DSSC and biosensor applications, a special molecule such as a dye or enzyme is usually introduced to attach onto the surface of ZnO. Thus, the higher surface area leads to a greater amount of absorbed molecules, which may lead to a higher photo-electricity conversion efciency (Z) or sensitivity. Moreover, dye or enzyme molecules can attach to the ZnO surface by chemical adsorption [8] or electrostatic forces [9], respectively. Both of the absorption mechanisms may be strongly inuenced by surface functional groups such as OH, CH2 and CH. Therefore, abundant surface chemical groups on the surface of ZnO are also important. Otherwise, the pore structure in the ZnO nanospheres inuences the surface area and absorption of molecules to some

Corresponding author. Tel.: 82 2 21235754; fax: 82 23126401. E-mail address: swlim@yonsei.ac.kr (S. Lim).

extent. An appropriate pore size distribution would lead to the absorption of a large amount of molecules [10]. Indeed, a lot of studies have been conducted regarding the synthesis of ZnO nanospheres [1117]. ZnO nanoparticles have been easily synthesized from the zinc salts such as zinc nitrate, zinc acetate, and zinc chloride with hydrolysis [1117]. However, very limited study has been performed to understand how the surface area and pore structure are decided. Therefore, it is worth focusing on studying the formation mechanism of ZnO pore structures for their device application. In addition, since a good crystal structure is necessary for quick electron transfer in these devices, heat treatment is usually adopted to achieve a good crystal structure [11,14,16]. During the annealing treatment, ZnO nanoparticles in the nanosphere fuse together and the surface area of the ZnO nanospheres signicantly decreases. The binding groups on the ZnO surface can also be destroyed. Therefore, in general, there is a trade-off between a good crystal structure and high surface area with abundant surface chemical groups. To avoid heat treatment, low temperature preparation of ZnO nanospheres with a good crystal structure is necessary, which can be applicable to the fabrication of exible devices. Hydrothermal process is a method to synthesize single crystals, which depends on the solubility of minerals in hot solution under high pressure. Compared to other methods, the hydrothermal method has inherent advantages such as low cost, low temperature, easy synthesis, and controllable growth by adjusting solvent, temperature, and catalyst. So far, various materials with a good crystal structure have been synthesized by this method

0022-3697/$ - see front matter & 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.jpcs.2011.09.016

Y. Zhang et al. / Journal of Physics and Chemistry of Solids 72 (2011) 15481553

1549

at low temperatures [1822]. In particular, when the hydrothermal synthesis was applied, ZnO nanoparticles whose sizes are less than 20 nm were formed below 200 1C without any further heat treatment [2326]. Therefore, the hydrothermal method is suitable for preparing ZnO nanospheres and it would solve the tradeoff between crystal structure and surface area. In addition, one of the main characteristic of those synthesizing conditions is to apply NaOH with the zinc acetate or zinc nitrate for the adjustment of pH. Residual sodium salt may have an effect on the conversion efciency of DSSC [27], thus elimination of sodium from the precursor is necessary. In this work, ZnO nanospheres were synthesized by a hydrothermal process using diethylene glycol without NaOH. For comparison, ZnO nanospheres were also prepared by a normal wet-chemical method. The morphology, structure, and surface properties of the samples were investigated. The surface area and pore structure were studied and a possible mechanism is presented.

2. Experimental details All of the chemicals used in this study were analytic grade reagents (Sigma-Aldrich) without further purication. The experimental procedure was adopted as follows. First, 0.004 mol of zinc acetate dihydrate (Zn(CH3COO)2 2H2O) was dissolved in 40 ml of diethylene glycol. The solution was then transferred into a Teonlined stainless steel autoclave (45 mL). The autoclave was sealed and placed into an oven. The temperature was increased from room temperature to 160 1C in 20 min and was maintained for 5 h. After the reaction was completed, the autoclave was naturally cooled down to room temperature. The white products were isolated by centrifugation at a speed of 5000 rpm for 30 min, cleaned by three cycles of centrifugation/washing/redispersion in alcohol, and oven-dried at 120 1C for 24 h. The nal product was designated as S1. As a comparison, ZnO nanospheres were also synthesized by a normal wet-chemical method in which all of the experiment details were the same as mentioned above except for the use of a beaker instead of the autoclave. The nal product was designated as S2. Then, calcination was performed on the S2 sample at 450 1C for 2 h and this product was designated as S3. The structural properties of the lms were studied by X-ray diffraction (XRD, Bruker D8 DISCOVER (CuKa)). Transmission electron micrographs (TEM), high resolution electron micrographs (HRTEM), and selected area electron diffraction (SAED) patterns were obtained using a JEM-3010 transmission electron microscope. The morphologies of the samples were investigated by scanning electron microscopy (SEM) with the same Pt sputtering process (Hitachi, S-4200). Nitrogen adsorption and desorption at 77 K (Model ASAP 2020, Micromeritics Corp., Norcross, GA) was carried out to measure the Brunauer Emmett Teller (BET) surface area and nanopores in the ZnO nanospheres.
Fig. 1. SEM images of the (a) S1, (b) S3, and (c) S3 samples.

3. Results and discussion The morphologies of the samples were investigated by SEM, as shown in Fig. 1. All of the nanospheres possessed a regular spherical shape. Although no calcination was applied, a clear SEM picture was captured from the S1 sample as shown in Fig. 1(a), indicating that limited organic and amorphous materials remained. This may have arisen from the special high pressure environment in the autoclave. For the S1 sample, the diameter ranged from several tens to several hundreds of nanometers. For the S2 sample, the appearance of the nanosphere surface could not be clearly observed, as shown in Fig. 1(b). The blurred SEM picture obtained from the S2 sample may be due to the presence of organic materials and amorphous Zn(OH)2 in the nanosphere as one can see from the following reactions of ZnO nanospheres synthesis. ZnCH3 COO2 xH2 O-ZnOH x CH3 COO 2x xCH3 COOH ZnOH x CH3 COO 2x -ZnO x1H2 O 2xCH3 COOH 1 2

After heat treatment (S3 sample), ZnO nanoparticles fused together and mesopores ranging from several nanometers to several tens of nanometers were formed.

1550

Y. Zhang et al. / Journal of Physics and Chemistry of Solids 72 (2011) 15481553

Fig. 2. XRD patterns of ZnO nanosphere samples and FTO substrate.

The structural properties of the ZnO nanospheres were studied by XRD as shown in Fig. 2. All the samples possessed a typical wurtzite structure. The average grain size of ZnO was estimated by the Scherrer formula D 0:89l=B cos y 3

where D, l, y, and B are the mean grain size, the X-ray wavelength of 0.154056 nm, the Bragg diffraction angle and the full width at half maximum (FWHM) of the (100) and (101) peaks, respectively. Therefore, the average obtained grain sizes of the S1, S2, and S3 samples are 10.77, 10.74, and 15.93 nm, respectively. The grain size of the S2 sample increased from 10.74 to 15.93 nm after heat treatment, which may be due to the fusing of ZnO nanoparticles or decomposing of Zn(OH)2 in the nanospheres as follows: ZnOH2 -ZnO H2 O 4

However, a signicant change in the grain size was not observed between S1 and S2 samples. Therefore, it is suggested that grain size is mainly increased by further heat treatment after ZnO nucleation. To obtain more structural information, TEM, HRTEM, and SAED measurements were performed, as shown in Fig. 3. For the S1 sample hydrothermally synthesized in the autoclave without any further heat treatment, a clear crystal lattice and diffraction spots were obtained, which implies a good crystal structure. Moreover, it is noted that a number of nanopores about 5 nm in size (the white spots in Fig. 3(a)) were found in the S1 nanospheres. A clear crystal structure was observed from the S1 sample, as shown in the SAED pattern in the inset of Fig. 3(b). For the S2 sample, an individual ZnO nanosphere with a diameter of about 1 mm was captured in Fig. 3(c). Although a ZnO nanoparticle size of 10.74 nm was obtained from the XRD results, no clear particles boundaries or a crystal lattice were observed in the S2 sample as shown in Fig. 3(d), which probably attributed to the presence of organic materials and amorphous Zn(OH)2. More direct evidence of this nding was obtained from the SAED results. Obvious rings accompanied with several diffraction spots indicate that the nanospheres consist of ZnO nanoparticles and amorphous materials. After annealing at 450 1C, a good crystal structure was obtained from the S3 sample, as shown in Fig. 3(f) and the inset. Because surface area and pore structure are important factors to apply this material for various applications such as DSSC and biosensors, BET measurements were performed. The nitrogen

adsorptiondesorption isotherms of the samples are shown in Fig. 4(a). The three samples show obviously different isotherms indicating that different pore structures were formed. For the S1 sample, only one small hysteresis appears in the relative pressure range of 0.450.8. The increase in slope at p/p0 0.45 corresponds to capillary condensation, which is a typical trend of mesoporous materials with uniform pore systems. Therefore, it is suggested that a uniform pore size was achieved, which corresponds to typical characteristics of mesoporous materials formation. This is conrmed by the pore size distribution curve, where an exactly uniform pore size distribution centered at 4.07 nm was found, as shown in Fig. 4(b). For the S2 sample, the isotherm shows obviously different characteristics: a type IV (IUPAC classication) isotherm with some type II contribution, a feather of mesoporous material [28]. Two hysteresis loops were found in the relative pressure range of 0.451.0, indicating bimodal pore size distributions in different pore size regions. The rst hysteresis appears at a low relative pressure of 0.45, followed by a nearly linear increase up to p/p0 0.85 in both the adsorption and desorption branches. A hysteresis at a high relative pressure between 0.85 and 1.0 belongs to the type H3 loop, which is associated with slitlike pores. The bimodal pore size distribution is ascribed to two different pores: ne intra-aggregated pores formed between intra-agglometrated primary crystallites and large inter-aggregated pores produced by inter-aggregated secondary particles [29]. For the S2 nanospheres, there were many ZnO nanoparticles nearly 10.74 nm in size with amorphous structures. Some complexes consisted of ZnO nanoparticles and an amorphous-like structure may exist in the nanospheres. The intra-aggregated pores may be formed between the ZnO nanoparticles and the inter-aggregated pores could be produced between complexes. This bimodal pore-size distribution is further conrmed by its corresponding pore size distribution shown in Fig. 4(b). The distribution shows that the S2 nanospheres contain uniform 3.67 nm sized mesopores and larger sized pores of about 20 nm. On the other hand, for the S3 sample, one hysteresis appears in the high relative pressure range of 0.61.0. This hysteresis loop is of type H2, indicating a broad pore distribution in the mesopore range. Its corresponding pore size distribution is shown in Fig. 4(b). Compared to the S1 sample, the pore size in the S3 sample is larger and the pore size distribution is broader. A possible mechanism is presented by the schematic illustration shown in Fig. 5. The S1 nanospheres consist of numerous uniform ZnO nanoparticles. Thus, a uniform pore size distribution was obtained. The BET surface areas of the S1, S2, and S3 samples were 36.0, 28.2, and 16.4 m2/g, respectively. The highest surface area of the S1 sample is due to its uniform pore size. A large number of initial nuclei synthesized in the nucleation stage may result in a large number of nanoparticles with small size and narrow distribution [30]. The change of Gibbs free energy of a spherical nucleus in the nucleation stage, DG, can be expressed as

DG Dmv Dms 4pr3 DGv =3 4pr2 g


and

DGv kT ln1 s=O

where Dmv is the change of volume free energy, Dms is the change of surface free energy, r is the radius of spherical nucleus, DGv is the change of Gibbs free energy per unit volume of the solid phase, g is the surface energy per unit area, k is the Boltzmann constant, T is the temperature, s is the supersaturation, and O is the atomic volume. Therefore, it is conjectured that the increase in the pressure, which was applied for the synthesis of S1 sample, induced modication of the surface energy or supersaturation. In particular, because the solubility or the supersaturation is not strongly affected by the pressure in the liquid phase, the change

Y. Zhang et al. / Journal of Physics and Chemistry of Solids 72 (2011) 15481553

1551

Fig. 3. TEM, HRTEM images, and SAED patterns (inset) of the (a) and (b) S1, (c) and (d) S2, and (e) and (f) S3 samples.

in the volume free energy, Dmv, may be constant under the process conditions for the synthesis of S1 and S2 samples. However, because the surface energy per unit area, g, is dened as the energy required to create a unit area of new surface [30], it is proportional to the surface area [31]. Therefore, it is suggested that a higher surface area of the S1 sample produced at a higher pressure provides higher surface energy of the nuclei. Assuming that change of volume free energy is almost constant and only the change of surface free energy increases, the critical size for the nucleation, rn, which is given as 2g/DGv, will be increased. Finally, it is suggested that a uniform pore size distribution with a higher surface area obtained from the S1 sample is due to the increased change of surface energy. Despite of current

conclusion, further study is needed to completely understand this mechanism.

4. Conclusion ZnO nanospheres were successfully synthesized by a wetchemical method and a hydrothermal process. Compared to the samples prepared by the normal wet-chemical method, ZnO nanospheres synthesized by the hydrothermal process could achieve a good crystal structure without annealing and maintained abundant surface chemical groups. Moreover, numerous nanopores with an exactly uniform size of about 4.07 nm were found in the ZnO

1552

Y. Zhang et al. / Journal of Physics and Chemistry of Solids 72 (2011) 15481553

National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (2009-0093823, 2010-0010573).

References
[1] Z.L. Wang, Zinc oxide nanostructures: growth, properties and applications, J. Phys.: Condens. Matter 16 (2004) R829R858. [2] T.P. Chou, Q.F. Zhang, G.E. Fryxell, G.Z. Cao, Hierarchically structured ZnO lm for dye-sensitized solar cells with enhanced energy conversion efciency, Adv. Mater. 19 (2007) 25882592. [3] Q.F. Zhang, T.P. Chou, B. Russo, S.A. Jenekhe, G.Z. Cao, Aggregation of ZnO nanocrystallites for high conversion efciency in dye-sensitized solar cells, Angew. Chem. Int. Ed. 47 (2008) 24022406. [4] Q.F. Zhang, T.P. Chou, B. Russo, S.A. Jenekhe, G.Z. Cao, Light scattering enhanced energy conversion efciency in hierarchically-structure ZnO dye sensitized solar cells, Adv. Funct. Mater. 18 (2008) 16541660. [5] Q.F. Zhang, C.S. Dandeneau, S. Candelaria, D.W. Liu, B.B. Garcia, X.Y. Zhou, Y.H. Jeong, G.Z. Cao, Effects of lithium ions on dye-sensitized ZnO aggregate solar cells, Chem. Mater. 22 (2010) 24272433. [6] K. Park, Q.F. Zhang, B.B. Garcia, X.Y. Zhou, Y.H. Jeong, G.Z. Cao, Effect of an ultrathin TiO2 layer coated on submicrometer-sized ZnO nanocrystallite aggregates by atomic layer deposition on the performance of dye-sensitized solar cells, Adv. Mater 22 (2010) 23292332. [7] M. Ahmad, C.F. Pan, L. Gan, Z. Nawaz, J. Zhu, Highly sensitive amperometric cholesterol biosensor based on Pt-incorporated fullerene-like ZnO nanospheres, J. Phys. Chem. C 114 (2010) 243250. [8] P. Falaras, Synergetic effect of carboxylic acid functional groups and fractal surface characteristics for efcient dye sensitization of titanium oxide, Sol. Energy Mater. Sol. Cells 53 (1998) 163175. [9] A. Wei, X.W. Suna, J.X. Wang, Y. Lei, X.P. Cai, C.M. Li, Z.L. Dong, W. Huang, Enzymatic glucose biosensor based on ZnO nanorod array grown by hydrothermal decomposition, Appl. Phys. Lett. 89 (2006) 123902. [10] S.J. Bao, C.M. Li, J.F. Zang, X.Q. Cui, Y. Qiao, J. Guo, New nanostructured TiO2 for direct electrochemistry and glucose sensor applications, Adv. Funct. Mater. 18 (2008) 591599. [11] A. Eftekhari, F. Molaei, H. Arami, Flower-like bundles of ZnO nanosheets as an intermediate between hollow nanosphere and nanoparticles, Mater. Sci. Eng. A 437 (2006) 446450. [12] Z.L.S. Seow, A. SWWong, V. Thavasi, R. Jose, S. Ramakrishna, G.W. Ho, Controlled synthesis and application of ZnO nanoparticles, nanorods and nanospheres in dye-sensitized solar cells, Nanotechnology 20 (2009) 045604. [13] D. Pradhan, K.T. Leung, Template-free single-step electrochemical synthesis of ZnO hollow nanospheres: self-assembly of hollow nanospheres from nanoparticles, J. Mater. Chem. 19 (2009) 49024905. [14] Y. Rajeswari, A. Senthamizhan, S. Ramasamy, V. Ajayan, A. Katsuhiko, B.A. Chandra, An investigation on co-precipitation derived ZnO nanospheres, J. Nanosci. Nanotechnol. 9 (2009) 59665972. [15] Shalaka C. Navale, S.W. Gosavi, I.S. Mulla, Controlled synthesis of ZnO from nanospheres to micro-rods and its gas sensing studies, Talanta 75 (2008) 13151319. [16] Z.W. Tao, X.B. Yu, X.Y. Fei, J. Liu, G.Q. Yang, Y.H. Zhao, S.P. Yang, L.Z. Yang, Synthesis and photoluminescence of Cl-doped ZnO nanospheres, Opt. Mater. 31 (2008) 15. [17] A. Ishizumi, Y. Takahashi, A. Yamamoto, Y. Kanemitsu, Fabrication and optical properties of Eu3 -doped ZnO nanospheres and nanorods, Mater. Sci. Eng. B 146 (2008) 212215. [18] J.M. Wang, P.S. Lee, J. Ma, One-pot synthesis of hierarchically assembled tungsten oxide (hydrates) nano/microstructures by a crystal-seed-assisted hydrothermal process, Cryst. Growth Des 9 (2009) 22932299. [19] X.C. Song, Y. Zhao, Y.F. Zheng, Synthesis of MnO2 nanostructures with sea urchin shapes by a sodium dodecyl sulfate-assisted hydrothermal process, Cryst. Growth Des. 7 (2007) 159162. [20] W.X. Zhang, Z.H. Hui, Y.W. Cheng, L. Zhang, Y. Xie, Y.T. Qian, A hydrothermal method for low-temperature growth of nanocrystalline pyrite nickel diselenide, J. Cryst. Growth 209 (2000) 213216. [21] Y.G. Wang, G. Xu, L.L. Yang, Z.H. Ren, X. Wei, W.J. Weng, P.Y. Du, G. Shen, G.R. Han, Formation of single-crystal SrTiO3 dendritic nanostructures via a simple hydrothermal method, J. Cryst. Growth 311 (2009) 25192523. [22] J.F. Liu, X.L. Li, Y.D. Li, Synthesis and characterization of nanocrystalline niobates, J. Cryst. Growth 311 (2009) 47594762. [23] B. Baruwati, D.K. Kumar, S.V. Manorama, Hydrothermal synthesis of highly crystalline ZnO nanoparticles: a competitive sensor for LPG and EtOH, Sensors Actuators B 119 (2006) 676682. [24] P.M. Aneesh, K.A. Vanaja, M.K. Jayaraj, Synthesis of ZnO nanoparticles by hydrothermal method, Proc. SPIE 6639 (2007) 66390J-166390J-9. [25] P.M. Aneesh, M.K. Jayaraj, Red luminescence from hydrothermally synthesized Eu-doped ZnO nanoparticles under visible excitation, Bull. Mater. Sci. 33 (2010) 227231. [26] S.K. Mishra, R.K. Srivastava, S.G. Prakash, R.S. Yadav, A.C. Panday, Photoconductivity, dark-conductivity and photoluminescence study of hydrothermally synthesized ZnO nanoparticles, in: Proceedings of the 2009

Fig. 4. (a) Nitrogen adsorptiondesorption isotherms and (b) BJH pore-size distributions of ZnO nanospheres.

Fig. 5. Schematic illustration of the pore structures of the S1 and S2 samples.

nanospheres prepared by the hydrothermal method and the largest BET surface area was obtained. The pore structures of all the samples were studied in detail and a possible mechanism was presented in this work. The results of this study are important for applying ZnO nanospheres in the photochemical and electrochemical devices especially in the dye-sensitized solar cells.

Acknowledgments This work was supported by the Priority Research Centers Program and Basic Science Research Program through the

Y. Zhang et al. / Journal of Physics and Chemistry of Solids 72 (2011) 15481553

1553

International Conference on Emerging Trends in Electronic and Photonic Devices & Systems, 2009, pp. 461464. [27] J.-C. Tinguely, R. Solarska, A. Braun, T. Graule, Low-temperature roll-to-roll coating procedure of dye-sensitized solar cell photoelectrodes on exible polymer-based substrates, Semicond. Sci. Technol. 26 (2011) 045007. [28] M. Kruk, M. Jaroniec, Gas adsorption characterization of ordered organic inorganic nanocomposite materials, Chem. Mater. 13 (2001) 31693183.

[29] J.G. Yu, W. Liu, H.G. Yu, A one-pot approach to hierarchically nanoporous titania hollow microspheres with high photocatalytic activity, Cryst. Growth Des. 8 (2008) 930934. [30] G. Cao, Nanostrutures and Nanomaterials: Synthesis, properties, and Applications, Imperial College Press, London, UK, 2004. [31] H.D. Shih, F. Jona, D.W. Jepsen, P.M. Marcus, Atomic underlayer formation during the reaction of Ti{0001} with nitrogen, Surf. Sci. 60 (1976) 445465.

Anda mungkin juga menyukai