CuS Nanoballs

Embed Size (px)

Citation preview

  • 8/13/2019 CuS Nanoballs

    1/6

    A facile solution chemical route to self-assembly of CuS ball-flowers andtheir application as an efficient photocatalyst

    Zhiguo Cheng, Shaozhen Wang, Qian Wang and Baoyou Geng*

    Received 22nd July 2009, Accepted 11th September 2009

    First published as an Advance Article on the web 30th September 2009

    DOI: 10.1039/b914902c

    Ball-flower shaped CuS structures have been synthesized by using mixed copper chloride and thiourea

    in a simple hydrothermal process employing poly(vinylpyrrolidone) (PVP) as the surfactant. The

    morphological investigations by field emission scanning electron microscope (FE-SEM) reveal that the

    ball-flower shaped nanostructures are monodispersed in large quantities. The ball-flower shaped

    morphologies are strongly dependent on the different ratios of copper chloride to thiourea, the reaction

    temperature and reaction time. The possible growth mechanism of the formation of ball-flower shaped

    CuS products is discussed in detail. In addition, the photocatalytic activity of ball-flower shaped CuS

    architectures has been tested by the degradation of rhodamine B (RhB) under UV light irradiation,

    showing that the as-prepared ball-flower shaped CuS structures exhibit high photocatalytic activity for

    the degradation of RhB.

    1. Introduction

    The architectural controlled synthesis of nanostructured semi-

    conductors has received much research attention, due to their

    outstanding physical and chemical properties and potential

    applications in numerous fields. As an important semiconductor

    material, copper sulfide has found many applications due to its

    optical properties,1,2 chemical-sensing capability,3 ideal charac-

    teristics for solar energy absorption,4 thermoelectric cooling

    properties,5 fast-ion conduction at high temperature,6,7 etc.

    Recently, various morphologies of copper sulfide have been

    reported, such as plate-like,8 rod-like,9 tube-like structures,10

    hollow spheres,11,12

    flower-like,13

    snowflake-like, urchin-likepatterns,14 etc. Many techniques have been established to prepare

    copper sulfide in various forms, including hydrothermal or sol-

    vothermal methods,15,16 microwave irradiation,17 chemical vapor

    deposition,18 in situ template-controlled method.19 Obtaining

    new materials via a relative simple route and developing the

    morphology-controlled synthesis methodologies are a goal and

    evoke great interest in materials chemistry.20 In solution

    approaches to copper sulfide, thiourea is a commonly used sulfur

    source, for instance, Zhu and co-workers21 have synthesized

    hollow spherical copper sulfide in water solution containing

    Cu(CH3COO)2, thiourea and 2-HP-b-CD. Qian et al. prepared

    CuS crystals with various morphologies using shuttle-like CuO

    and thiourea as raw reagents at low temperature.22

    Wanget al.proposed a sonochemical approach to single crystalline CuS

    nanoplates.23 However, up to now, there have been no reports on

    the synthesis of monodispersive ball-flower superstructures

    constructed with CuS nanosheets, and the controlled synthesis of

    CuS hierarchical nanostructures with different morphologies has

    not been performed.

    In recent years, considerable attention has been given to the

    environmental problem involving organic pollutants in water.

    Photocatalysis is a promising technology for the treatment of

    contaminants, especially for the removal of organic compounds.

    Many investigations have been reported on utilizing metal oxide

    nanomaterials as photocatalysts to decompose or destroy the

    organic pollutants in water.24 However, there is little research on

    the photocatalytic property of CuS superstructures with a high

    specific surface area, though other properties of CuS micro-

    structure have been widely investigated.

    In this work, we synthesized 3D hierarchical CuS ball-flowers

    by a newly designed hydrothermal self-assembly process. The

    simplest synthetic route to 3D nanostructures is probably a self-assembly in which ordered aggregates are formed in a sponta-

    neous process.25 Thiourea was also used as a sulfur source.

    However, now, an interesting but challenging question is raised,

    namely, what role does thethiourea play in theformation of metal

    sulfide nanostructures? To learn about the thiourea-basedprocess

    in detail, we investigated the simplest aqueous system containing

    only CuCl2 and thiourea. As expected, we found that thiourea

    reduces Cu(II) to Cu(I) as it combines to form a complex. The

    detailed shape evolution process from the intermediate complex

    to the final product was clearly shown, and the mechanism of

    formation of the hierarchical structure was studied. The photo-

    catalytic properties of the as-obtained 3D hierarchical architec-

    tures were investigated under UV light irradiation. The resultsshow that the synthesized CuS structures exhibit the superiority

    of photocatalytic performance and have exhibited much better

    photocatalytic properties for the photodegradation of RhB than

    that of other photocatalysts, such as TiO2nanopowders.

    2. Experimental

    2.1. Chemicals and synthesis

    All reagents were of analytical grade and were used without

    further purification. In a typical experiment, an appropriate

    College of Chemistry and Materials Science, Anhui Key Laboratory ofFunctional Molecular Solids, Anhui Laboratory of Molecular-BasedMaterials, Anhui Normal University, Wuhu, 241000, P. R. China.E-mail: [email protected]; Fax: +86-553-3869303; Tel: +86-553-3869303

    144 | CrystEngComm, 2010, 12, 144149 This journal is The Royal Society of Chemistry 2010

    PAPER www.rsc.org/crystengcomm | CrystEngComm

    View Article Online / Journal Homepage / Table of Contents for this issue

    http://pubs.rsc.org/en/journals/journal/CE?issueid=CE012001http://pubs.rsc.org/en/journals/journal/CEhttp://dx.doi.org/10.1039/b914902c
  • 8/13/2019 CuS Nanoballs

    2/6

    amount of copper chloride was dissolved in distilled water

    (10 mL) and formed a blue solution of a certain concentration

    under constant stirring for 10 min. Then, appropriate amounts of

    thiourea (Tu) and PVP were dissolved in distilled water (15 mL)

    and formed a colorless solution. Then the above two solutions

    were mixed slowly, a white suspension was formed in a few

    minutes, which was then transferred into a 30 mL Teflon-lined

    autoclave and maintained at 120180 C for 624 h, then cooled

    to room temperature naturally. The black precipitate was filteredoff, washed with distilled water and absolute ethanol several

    times, and then dried in vacuum at 60 C for 4 h. For compar-

    ison, the bulk CuS powders were prepared by a liquid state

    reaction. Simply, appropriate amounts of CuCl2and Na2S were

    added into the beaker under stirring for 15 min and as a result,

    black products were obtained.

    2.2. Characterization

    X-Ray powder diffraction (XRD) was carried out on an

    XRD-6000 (Japan) X-ray diffractometer with Cu Ka radiation

    (l 1.54060 A) at a scanning rate of 0.05 s1. Scanning electron

    microscopy (SEM) micrographs were taken using a HitachiS-4800 Scanning electron microscope coupled with energy

    dispersive X-ray spectroscopy (EDS).

    2.3. Photocatalytic property

    The photocatalytic activity experiments of the obtained CuS for

    the decomposition of RhB in air were performed at ambient

    temperature. A cylindrical Pyrex flask (capacityca.100 mL) was

    used as the photoreactor vessel. CuS microstructues as catalyst

    (10 mg) was added to an aqueous RhB solution (2.0 105 M,

    50 mL) and was magnetically stirred in the dark for 30 min to

    ensure establishing an adsorptiondesorption equilibrium. The

    UV light was generated from a 300 W high-pressure mercurylamp. As a comparison, the photocatalytic activity of commer-

    cial TiO2 powders (Degussa P25, Degussa Co., the surface area

    ca. 45 m2 g1) was also tested in the same experimental condi-

    tions. UV-vis absorption spectra were recorded at different

    intervals to monitor the reaction using Solid Spec-3700 UV/vis

    spectrophotometer.

    3. Results and discussion

    3.1. Structure and morphology

    The typical XRD patterns as shown in Fig. 1 reveal the phase and

    purity of the as-obtained CuS ball-flower structures. The

    diffraction peaks are indexed to the data of the hexagonal phaseCuS (JCPDS Card File No. 060464). Compared with the

    standard diffraction data of the hexagonal phase CuS, we can

    find that all peaks shift to the lower angle slightly. The reasons

    may be ascribed to the following reasons: on the one hand,

    during sample preparation, the surface of the sample was above

    the sample platform or the zero point of the instrument was not

    accurate. On the other hand, the crystallite size was a non-

    uniform distribution after a hydrothermal treatment, which may

    also lead the peaks shift to the lower angle.

    The morphology of the prepared samples was further investi-

    gated with FE-SEM, and their typical images are displayed in

    Fig. 2. Fig. 2a displays a representative overview of the ball-

    flower CuS architectures (the concentration ratio of thiourea to

    Cu2+ (4:2), 24 h, 120 C), which shows that the as-obtained

    products are composed of large-scale ball-flower architectures

    with diameters of 1.82.4 mm. Fig. 2b and c show high-magnifi-

    cation SEM images of CuS ball-flower architectures from

    a different angle of view, which vividly demonstrate that the ball-

    flower CuS structures are built of two-dimensional nanoplates.

    The nanoplates are well-ordered and oriented to form ball-flower

    architectures. Fig. 2d shows the EDS pattern of the obtained

    ball-flower products, which clearly reveals that the obtained

    products mainly consisted of Cu and S elements, the weak O

    signal comes from the oxidation of air. The atomic ratio of Cu to

    S is 1.08:1, which closes to the stoichiometry of CuS.

    3.2. Influencing factors

    3.2.1. Concentration ratio of Cu2+ to thiourea. In this work,

    we found that CuS ball-flower structures could be selectively

    obtained by adjusting the concentration ratio of CuCl2$2H2O to

    thiourea and hydrothermal temperature. To determine the effect

    Fig. 1 XRD pattern of CuS hierarchical ball-flower structures prepared

    at 120 C for 24 h.

    Fig. 2 SEM images of (a), (b), (c) different magnifications of the

    as-obtained CuS ball-flower structures. (d) EDS spectrum of the obtained

    products.

    This journal is The Royal Society of Chemistry 2010 CrystEngComm, 2010, 12, 144149 | 145

    View Article Online

    http://dx.doi.org/10.1039/b914902c
  • 8/13/2019 CuS Nanoballs

    3/6

    of the concentration ratio of CuCl2$2H2O to thiourea on the

    formation of the products with different shapes, the images of the

    products obtained from the solutions with a concentration ratio

    of CuCl2$2H2O to thiourea of 1/1.5, 1/2.0, 2/3.0, 2/4.0 and

    hydrothermal heat treatment at 120 C for 24 h are shown in

    Fig. 3. It is clearly seen that the morphology changed greatly with

    the change of the concentration ratio of CuCl2$2H2O to thio-

    urea. Fig. 3a and 3b display the images of the product obtained at

    a concentration ratio of CuCl2$

    2H2O to thiourea of 1/1.5. It canbe seen that trepang-like hierarchical architectures were ach-

    ieved. With further increasing the thiourea concentration to

    2.0 M, a large number of dandelion-like nanostructures

    composed of lots of ultrathin nanosheets were observed (Fig. 3c).

    When the concentration ratio of CuCl2$2H2O to thiourea was

    2.0/3.0, microsphere-like architectures were obtained, as seen

    from Fig. 3e. The as-synthesized product entirely takes on the

    ball-flower morphology when the thiourea concentration is

    4.0 M (Fig. 3g).

    From the above results, it can be deduced that both thiourea

    and copper ion concentrations influence the CuS morphologies

    critically. When thiourea was added to the copper(II) salt solu-

    tion, the color of the solution changed from typically blue to

    green, indicating the thioureacopper(II) complex is formed.

    Further addition of thiourea or just leaving the green solution

    standing for a longer time will cause the solution to become

    colorless, indicative of conversion into thioureacopper (I)

    complexes.26 During the experiment, when the concentration of

    Tu was 1.5 or 2.0, the green color of the thioureacopper(II)

    reaction solution was persistently kept until the solution was

    transferred to an autoclave. However when the concentration of

    Tu was further increased to 3.0 and 4.0, the reaction solutionrapidly became colorless, implying that excess thiourea signifi-

    cantly speeds up the redox process. On the basis of the above

    analysis, formation of different coordination precursors is

    a crucial factor determining the crystal growth process of the

    copper sulfide crystals.

    3.2.2. Reaction temperature.To investigate the effect of the

    reaction temperature on the formation of CuS ball-flower

    structures, a series of comparative experiments was carried out

    through similar processes. It is found that the reaction temper-

    ature has a significant influence on the morphology of the

    as-synthesized CuS products. When the reaction temperature is

    below 120 C, not only the yield of the products is low, but alsothe morphology of the sample is irregular. When the reaction

    temperature is elevated to 120 C, the product is composed of

    ball-flower structures, and no other structures are observed any

    more (sample S4, Fig. 4a). At a higher reaction temperature of

    140 C, (Fig. 4b), the CuS ball-flower structures (sample S5,

    Fig. 4b) became thick and rigid. No obvious changes were

    observed in the morphology of CuS synthesized from the higher

    reaction temperatures 160 (Fig. 4c), however, the surface of the

    CuS hierarchical ball-flower structures became compact. When

    the reaction temperature increased to 180 C, the sheets assem-

    bled in the ball-flower grew thicker than those appearing at 160C and the CuS ball-flower structures assembled by the

    compactly packed flakes were observed (Fig. 4d). Therefore, wecan find that the exterior morphology of the final products did

    not change obviously with the increase of the reaction tempera-

    ture; but the sheets composed ball-flowers thickened with the

    Fig. 3 FE-SEM images of the CuS nanostructures obtained by

    a hydrothermal treatment for 24 h with thiourea concentration of (a), (b)

    1.5 M; (c), (d) 2.0 M; (e), (f) 3.0 M; (g), (h) 4.0 M at 120 C.

    Fig. 4 FE-SEM images of the CuS nanostructures obtained by

    a hydrothermal treatment for 24 h at (a) 120 C; (b) 140 C; (c) 160 C; (d)

    180 C with a copper ion concentration of 0.1 M.

    146 | CrystEngComm, 2010, 12, 144149 This journal is The Royal Society of Chemistry 2010

    View Article Online

    http://dx.doi.org/10.1039/b914902c
  • 8/13/2019 CuS Nanoballs

    4/6

    reaction temperature. It might be related to the reaction rate

    increased at high temperature but could not alter the growth

    process in the present reaction system. More in-depth studies are

    necessary to further understand their growth process.

    The detailed experimental parameters together with the

    morphological properties of the corresponding products (deno-

    ted as S1S7) are listed in Table 1. Typical SEM images of the

    samples are shown in Fig. 2a, 2c, 2e and 2g.

    3.3. Formation of the CuS structures

    In order to understand the evolution process of the ball-flower

    structures of CuS material, we carried out time-dependent

    experiments during which samples were collected at different

    time intervals from the reaction mixture.

    As shown in Fig. 5, the CuS self-assembled into hierarchical

    ball-flower structures consisting of sheets. In the initial stage

    (6 h), only a small amount of the product was obtained. As

    shown in Fig. 5a, the morphology of the as-produced samples at

    the early stage consisted of very thin flakes with sizes of about

    1.21.5 mm, and a small quantity of aggregation. With further

    increasing the reaction time to 10 h, these sheets aligned withclearly oriented layers, pointing toward a common center, as

    displayed in Fig. 5b. With the increase of the hydrothermal time

    to 14 h, no single sheets remained. All of the sheets assembled

    into nest-like structures (Fig. 5c). Finally, when the reaction was

    further prolonged to 18 h, the cores in the center of the spheres

    dissolved completely, resulting in the formation of nest-like

    structures with hollow cores, as demonstrated in Fig. 5d. As the

    reaction proceeded, the size of the pores among the flakes

    decreased obviously. At the elongated reaction time of 24 h, the

    morphology of the final product is shown in Fig. 2a. It is clear to

    see that much more sheets were assembled into hierarchical

    structures, and the typical hierarchical ball-flower structures

    formed. From this point, the size and morphology of the productremained the same even at longer reaction times. In addition, the

    PVP molecules play an important role in preventing the spheres

    from agglomerating as many polymer molecules do in the

    formation of other spherical species.27,28 On the other hand, PVP

    in this study acts as a directing agent to promote the preferential

    3D growth of CuS ball-flower. We also found that other

    surfactants, such as CTAB and SDBS etc. did not reach the

    equivalent effect as the PVP did. So we can say that the PVP

    molecules may be unique for the formation of the ball-flower

    structures in the present reaction system.

    The XRD patterns of the products prepared at different

    reaction times are shown in Fig. 6. As shown in Fig. 6, the

    products prepared at different reaction times have similar XRDpatterns, except for relative peak intensity levels which were due

    to the random orientation. All diffraction peaks can be indexed

    as hexagonal CuS structure with calculated lattice constants of

    about a 3.792 A and c 16.344 A. These calculated lattice

    constant values are consistent with the reported data for CuS

    (JCPDS Card no. 060464).

    On the basis of the above analysis, we proposed a reasonable

    mechanism for the formation of CuS ball-flower structures. The

    whole evolution process is illustrated in Fig. 7. In this formation

    process, time was the most important controlling factor. Such

    a process is consistent with previous reports of a so-called two-

    stage growth process, which involves a fast nucleation of amor-

    phous primary particles followed by a slow aggregation andcrystallization of primary particles.2932 In our experiment,

    thiourea first coordinated with CuCl2 to produce thiourea

    copper (II) complexes, which precipitated to become the nuclei

    and quickly grew into the primary particles. In the following

    secondary growth stage, the primary particles aggregated into

    sheets which became the base of the nest-like structure. The

    sheets continued to grow by combining with the remaining

    Table 1 Morphologies and structures of CuS samples

    Sample T/C T/hThe molar ratio ofCuCl2$2H2O to thiourea Morphologhy

    S1 120 24 1/1.5 TreapangS2 120 24 1/2.0 DandelionS3 120 24 2/3.0 MicrosphereS4 120 24 2/4.0 Ball-flowerS5 140 24 2/4.0 Ball-flower

    S6 160 24 2/4.0 Ball-flowerS7 180 24 2/4.0 Ball-flower

    Fig. 5 SEM images of the CuS nanostructures obtained by hydro-

    thermal treatment at 120 C for (a) 6 h; (b) 10 h; (c)14 h; (d) 18 h with

    thiourea concentration of 4.0 M.

    Fig. 6 The XRD pattern of products at different reaction times: 6 h (a),

    10 h (b), 14 h (c) and 18 h.

    This journal is The Royal Society of Chemistry 2010 CrystEngComm, 2010, 12, 144149 | 147

    View Article Online

    http://dx.doi.org/10.1039/b914902c
  • 8/13/2019 CuS Nanoballs

    5/6

    primary particles, finally forming the ball-flower structure. Many

    forces cause the sheets to self-assemble into the ball-flower

    morphology, such as electrostatic and dipolar fields associated

    with the aggregate, hydrophobic interactions, hydrogen bonds,

    crystal-face attraction, and van der Waals forces.33,34 For

    example, Zhong35 described the evolution mechanism of

    a flower-like iron oxide precursor in the presence of CTAB,

    which followed Ostwald ripening kinetics. Further work is

    underway to investigate the details of the self-assembly growth

    mechanism.

    3.4. Photocatalytic properties of CuS ball-flower

    To demonstrate the potential application of as-synthesized

    hierarchical ball-flower CuS microstructures in the degradation

    of organic contaminants, we have investigated their photo-

    catalytic activities by choosing the photocatalytic degradation of

    RhB as a model reaction. Fig. 8a shows the optical absorption

    spectra of RhB aqueous solution (initial concentration: 2.0

    105 M, 50 mL) with 10 mg of the as-prepared CuS powders afterexposure to ultraviolet light (UV) for different durations. The

    main absorption peak locates at 553 nm, which corresponds to

    the RhB molecules, decreases rapidly with extension of the

    exposure time, and completely disappears after about 60 min.

    Further exposure leads to no absorption peak in the whole

    spectrum, which indicates the total decomposition of RhB. It is

    clear that the intense pink color of the starting solution gradually

    disappears with increasing exposure time to the UV light.

    Further experiments were carried out to compare the catalytic

    activity of the as-prepared CuS architectures, bulk CuS powders

    and commercial Degussa P25 TiO2 powders. Fig. 8b shows the

    curves of the concentration of residual RhB with irradiation

    time. Without any catalyst, only a slow decrease in the concen-tration of RhB was detected under UV irradiation. The catalytic

    activity of the CuS architectures in the dark was also performed,

    which showed that only a little decrease in the concentration of

    RhB was detected in the dark, which confirmed the concentra-

    tion decrease of RhB solution is mainly due to photodegradation

    of the products. The addition of catalysts leads to obvious

    degradation of RhB, and the photocatalytic activity depends on

    the morphology. For the CuS with a hierarchical micro-

    architecture, however, the activity is much higher than that of

    bulk CuS powders (line c) and commercial Degussa P25 TiO2powders (line d). The RhB solution is decolorized completely by

    using the CuS ball-flowers after UV irradiation for 5060 min

    (line e).

    It is generally accepted that the catalytic process is mainly

    related to the adsorption and desorption of molecules on the

    surface of the catalyst. The high specific surface area can provide

    more reactive adsorption/desorption sites for photocatalytic

    Fig. 7 Formation mechanism of hierarchical ball-flower structures of

    CuS.

    Fig. 8 (a) Time-dependent absorption spectrum of a solution of RhB

    solution (2.0 105 M, 50 mL) in the presence of architectural CuS

    (10 mg) under exposure to UV light. (b) The RhB normalization

    concentration (from the optical absorbance measurements at 550 nm) in

    the solution with different catalysts (10 mg) vs. the exposure time.

    Starting RhB concentration C0: 2.0 105 M. Line a: without any

    catalyst under UV light. Line b: the architectural CuS in the dark. Line c:

    the bulk CuS powders under UV light; Line d: Degussa P25 TiO2powders under UV light; Line e: the architectural CuS under UV light.

    Fig. 9 BET measurement of CuS architectures.

    148 | CrystEngComm, 2010, 12, 144149 This journal is The Royal Society of Chemistry 2010

    View Article Online

    http://dx.doi.org/10.1039/b914902c
  • 8/13/2019 CuS Nanoballs

    6/6

    reactions.36 In order to evaluate the surface area of the obtained

    CuS architectures, full nitrogen sorption isotherms were

    measured (Fig. 9). The specific surface area was thus evaluated to

    be 78.6 m2 g1 from data points in this pressure range by the BET

    equation. The result showed that the obtained CuS architectures

    possess a larger specific surface area than that of Degussa

    P25 powders (45 m2 g1).

    Therefore, the photocatalytic superiority of the CuS archi-

    tectures may be attributed to their special structural features. Thehierarchically microstructured CuS can provide a greater surface

    area than the reference samples Degussa P25 powders, which is

    obviously beneficial for the enhancement of photocatalytic

    performance. In addition, good dispersing and uniformity also

    can provide a large active surface area.37,38

    4. Conclusions

    In summary, nearly monodispersed CuS ball-flower structures

    have been successfully synthesized via a facile hydrothermal

    process at low-temperature. The morphologies of the architec-

    tures can be selectively produced by adjusting the concentration

    ratio of CuCl2$

    2H2O to thiourea and hydrothermal temperature.The possible growth mechanism is also proposed, and this

    method may be extends to the preparation of other metal chal-

    cogenide semiconductors. The photocatalytic properties of the

    hierarchical CuS architectures was explored; the results revealed

    that the obtained products exhibited excellent photocatalytic

    activity for degradation of RhB under exposure to UV light

    irradiation.

    Acknowledgements

    This work was supported by the National Natural Science

    Foundation of China (20671003, 20971003), the Key Project of

    Chinese Ministry of Education (209060), the EducationDepartment of Anhui Province (2006KJ006TD) and the

    Program for Innovative Research Team in Anhui Normal

    University.

    References

    1 K. Takase, M. Koyano, T. Shimizu, K. Makihara, Y. Takahashi,Y. Takano and K. Sekizawa,Solid State Commun., 2002, 123, 531.

    2 E. J. Silvester, F. Grieser, B. A. Sexton and T. W. Healy, Langmuir,1991, 7, 2917.

    3 A. Etkus, A. Galdikas, A. Mironas, I. Simkiene, I. Ancutiene,V. Janickis, S. Kaciulis, G. Mattogno and G. M. Ingo, Thin SolidFilms, 2001,391, 275.

    4 S. T. Lakshmikvmar and A. C. Rastogi,Sol. Energy Mater. Sol. Cells,1994, 32, 7.

    5 E. Ramli, T. B. Rauchfuss and C. L. Stern, J. Am. Chem. Soc., 1990,112, 4043.

    6 M. T. Nair and P. K. Nair, Semicond. Sci. Technol., 1989, 4, 191.7 J. C. Folmer and F. Jellinek, J. Less-Common Met., 1980, 76, 153.8 J. Zhang and Z. K. Zhang, Mater. Lett., 2008, 62, 2279.9 G. Mao, W. Dong and D. G. Kurth, Nano Lett., 2004, 4, 249.

    10 J. Y. Gong, S. H. Yu, H. S. Qian, L. B. Luo and X. M. Liu, Chem.Mater., 2006,18, 2012.

    11 H. L. Zhu, X. Ji, D. R. Yang, Y. J. Ji and H. Zhang, MicroporousMesoporous Mater., 2005, 80, 153.

    12 X. L. Yu, C. B. Cao, H. S. Zhu, Q. S. Li, C. L. Liu and Q. H. Gong,Adv. Funct. Mater., 2007, 17, 1397.

    13 S. Gorai, D. Ganguli and S. Chaudhuri, Cryst. Growth Des., 2005,5,875.

    14 L. Y. Zhu, Y. Xie, X. W. Zheng, X. Liu and G. E. Zhou, J. Cryst.Growth, 2004, 260, 494.

    15 J. Zou, J. X. Zhang, B. H. Zhang, P. T. Zhao and K. X. Huang,Mater. Lett., 2007, 61, 5029.

    16 S. Gorai, D. Ganguli and S. Chaudhuri, Mater. Res. Bull., 2007, 42,345.

    17 X. H. Liao, N. Y. Chen, S. Xu, S. B. Yang and J. J. Zhu, J. Cryst.Growth, 2003, 252, 593.

    18 S. K. Haram, A. R. Mahadeshwar and S. G. Dixit, J. Phys. Chem.,1996, 100, 5868.

    19 J. Lu, Y. Zhao, N. Chen and Y. Xie, Chem. Lett., 2003,32, 30.

    20 X. F. Duan and C. M. Lieber, Adv. Mater., 2000, 12, 298.21 J. Z. Xu, S. Xu, J. Geng, G. X. Li and J. J. Zhu, Ultrason. Sonochem.,

    2006, 13, 451.22 K. B. Tang, D. Chen, Y. F. Liu, G. Z. Shen, H. G. Zheng and

    Y. T. Qian,J. Cryst. Growth, 2004,263, 232.23 H. L. Xu, W. Z. Wang and W. Zhu, Mater. Lett., 2006, 60, 2203.24 X. F. Song and L. Gao, J. Phys. Chem. C, 2008,112, 15299.25 G. M. Whitesides and M. Boncheva, Proc. Natl. Acad. Sci. U. S. A.,

    2002, 99, 4769.26 M. Q. Ai, P. E. Yue, D. O. Huang, Q. L. Han and Y. S. Cheng, Cryst.

    Growth Des., 2005,5, 855.27 X. H. Li, D. H. Zhang and J. S. Chen, J. Am. Chem. Soc., 2006,128,

    8382.28 U. Jeong, Y. Wang, M. Ibisate and Y. Xia, Adv. Funct. Mater., 2005,

    15, 1907.29 C. Burda, X. B. Chen, R. Narayanan and M. A. El-Sayed, Chem.

    Rev., 2005, 105, 1025.30 Y. Cheng, Y. S. Wang, Y. H. Zheng and Y. Qin, J. Phys. Chem. B,2005, 109, 11548.

    31 R. L. Penn, J. Phys. Chem. B, 2004,108, 12707.32 J. Park, V. Privman and E. Matijevic, J. Phys. Chem. B, 2001, 105,

    11630.33 H. Colfen and M. Antonietti, Angew. Chem., Int. Ed., 2005,44, 5576.34 H. Colfen and S. Mann,Angew. Chem., Int. Ed., 2003, 42, 2350.35 L. S. Zhong, J. S. Hu, H. P. Liang, A. M. Cao, W. G. Song and

    L. J. Wan,Adv. Mater., 2006, 18, 2426.36 L. H. Zhang, H. Q. Yang, J. Yu, F. H. Shao, L. Li, F. H. Zhang and

    H. Zhao, J. Phys. Chem. C, 2009,113, 5434.37 C. H. Ye, Y. Bando, G. Z. Shen and D. Golberg, J. Phys. Chem. B,

    2006, 110, 15146.38 J. S. Hu, L. L. Ren, Y. G. Guo, H. P. Liang, A. M. Cao, L. J. Wan and

    C. L. Bai, Angew. Chem., Int. Ed., 2005, 44, 1269.

    This journal is The Royal Society of Chemistry 2010 CrystEngComm, 2010, 12, 144149 | 149

    View Article Online

    http://dx.doi.org/10.1039/b914902c