You are on page 1of 9

22758

J. Phys. Chem. B 2005, 109, 22758-22766

Novel Multilamellar Mesostructured Molybdenum Oxide Nanofibers and Nanobelts: Synthesis and Characterization
Rui-Qi Song, An-Wu Xu,*,, Bin Deng, and Yue-Ping Fang
School of Chemistry and Chemical Engineering, Sun Yat-Sen UniVersity, Guangzhou 510275, China, and Max Planck Institute of Colloids and Interfaces, Department of Colloid Chemistry, MPI Research Campus Golm, D-14424, Potsdam, Germany ReceiVed: June 21, 2005; In Final Form: September 30, 2005

One-dimensional molybdenum oxide nanostructures with layered mesostructures were prepared directly from commercial bulk MoO3 crystals by a surfactant-templated hydrothermal process. X-ray diffraction, scanning electron microscopy, transmission electron microscopy, infrared spectra, and thermal analyses have been used to characterize the obtained molybdenum oxide nanomaterials. By use of cetyltrimethylammonium bromide as the structure-directing template, novel molybdenum oxide nanofibers with triple interlayer distances of 2.84, 2.66, and 2.46 nm have been obtained. The nanofibers have diameters of 20-100 nm and length up to 20 m. The growth of multilamellar molybdenum oxide nanofibers can be interpreted by the combination of surfactant/inorganic self-assembly process and host/guest intercalation chemistry. On the basis of the X-ray diffraction and infrared results, a possible arrangement of surfactant in the interlayer space of molybdenum oxide by bilayer micelles with different tilt angles has been proposed. In addition, the thermal stability of surfactant has been improved by intercalation. Moreover, molybdenum oxide nanobelts with two kinds of interlayered structures were also produced in the presence of n-alkylamines (n ) 12, 14, 16, and 18) following a similar method, these nanobelts show length up to more than 10 m, width ranging between 200 and 600 m, and width-to-thickness ratios of about 3-12. A linear relationship is observed between the interlayer distance and the number of carbon atoms in n-alkyl chains.

Introduction One-dimensional (1D) nanostructured inorganic materials exhibit properties that differ from their corresponding bulk materials due to the reduced size and the large surface-to-volume ratios.1 Over the past decades, there has been considerable effort devoted to fabricate 1D inorganic nanomaterials with different morphologies, such as nanowires, nanobelts, and nanotubes, to optimize and utilize chemical and physical properties of these materials.2 Layered inorganic oxides have attracted continuing interest with regard to their many applications resulting from the sorption properties and, therefore, for sorbents, production of catalysts, ion exchangers, and electrochemical materials.3 Particularly, these materials have recently been investigated to build nanomaterials, in cases where 1D hollow nanostructures have been fabricated by a high temperature-driven rolling process from 2D sheets of artificial lamellar structures.4 Molybdenum trioxide (MoO3), an important layered oxide, is especially attractive to be examined on the nanoscale, due to its layered crystal structure in the bulk and the multifaced functional properties. It shows excellent catalytic activity in selective oxidation reactions, with applications not only in petroleum refining and chemical production but also on pollution control industries.5 A wide range of application of MoO3 and its derivatives can also be found in secondary batteries, electronic display devices, sensors, recording systems, and lubricants.6-9 For these applications, the efficient or sensitive
* To whom correspondence should be addressed. E-mail: An-Wu.Xu@ mpikg.mpg.de. Phone: 49-331-5679505. Fax: 49-331-5679502. Sun Yat-Sen University. Max Planck Institute of Colloids and Interfaces.

parameters have been found to be strongly dependent on the surface area or interfacial properties of the MoO3 materials. Owing to the large aspect and surface-to-volume ratio, a significantly large surface area could be expected when the size range of a material is downscaled to 1-100 nm in one dimension, as compared to its bulk counterpart. A variety of techniques have been developed to control the architectures and morphological patterns of MoO3 materials. Examples include electrochemical process,10 carbon nanotubes template,11 acidification of ammonium heptamolybdate tetrahydrate or molybdic acid,12,13 intercalation of neutral primary amine into molybdic acid,14 homogeneous solution reaction of sodium molybdate and perchloric acid,15 and infrared irradiation.16 Successful synthesis of 1D MoO3 nanomaterials can be realized in these strategies; however, the study on the synthesis of MoO3 nanofibers and nanobelts with multilamellar mesostructures based on surfactant/ inorganic self-assembly process is still limited. Our group has published extensively on the fabrication of 1D nanostructures by solid-liquid-solid (SLS) growth from oxide bulk crystals under hydrothermal treatment.17-19 Rare earth hydroxide nanotubes,17 lithium manganese oxide nanobelts,18 and porous magnesium hydroxide nanoplates19 have been synthesized by this method that dose offer advantages over other synthetic approaches. One of the advantages is the low cost and availability of raw materials, which favors scaled-up industrial manufacturing. Another advantage is the simplicity of the hydrothermal process, which proceeds at a mild temperature in aqueous solution under a sealed environment. We present here a novel approach to assemble commercial bulk MoO3 into triple-layered MoO3 nanofibers in aqueous solution by using cetyltrimethylammonium bromide (CTAB) as the

10.1021/jp0533325 CCC: $30.25 2005 American Chemical Society Published on Web 11/05/2005

Lamellar Molybdenum Oxide 1D Nanostructures

J. Phys. Chem. B, Vol. 109, No. 48, 2005 22759

Figure 1. SEM image (a) and XRD pattern (b) of the starting material (commercial bulk MoO3 crystals).

Figure 2. SEM images of the products synthesized in the presence of CTAB by hydrothermal treatment. CTAB concentration: 1.1 10-2 M (a), 1.5 10-2 M (b), 7.6 10-2 M (c), and 1.5 10-1 M (d).

structure-directing template. The approach of hydrothermal synthesis, in combination with the structure-directing properties of organic components, could be exploited in the isolation of metastable inorganic-organic composites that maintain the structural elements of the synthetic precursors. Furthermore, MoO3 nanobelts can also be formed together with two kinds of interlayers by the same method using aliphatic primary amines as the template. A possible arrangement of surfactant in the interlayer space of molybdenum oxide by bilayer micelles with different tilt angles has been proposed. Experimental Section Preparation. In a typical synthesis, 0.8 g of MoO3 powders (Shanghai Chinese Colloid Chemical Factory, 99.5% purity) was

mixed with CTAB (Alfa, 99.9% purity) at the required molar ratio in 18 mL of distilled water under ambient stirring for 2 h. Then the mixture was transferred into a stainless Teflon-lined 25-mL capacity autoclave. The autoclave was sealed and maintained at 180 C for 4 days and then air cooled to room temperature. The resulting MoO3 nanofibers were collected and washed with ethanol and then dried at 80 C in air overnight. MoO3 nanobelts were prepared by the same method in the presence of primary monoamines with different alkyl chain length as the structure-directing agents. Characterization. The low-angle and wide-angle X-ray powder diffraction (XRD) diagrams of all samples were carried out on a D/Max-IIIA X-ray diffractometer using Cu KR radiation ( ) 1.5418 ), with the operation power maintained

22760 J. Phys. Chem. B, Vol. 109, No. 48, 2005

Song et al.

Figure 3. Low-angle (a) and wide-angle (b) XRD patterns of the product synthesized in the presence of 7.6 10-2 M CTAB. (c) Low-angle and wide-angle (the inset) XRD pattern of the product prepared at CTAB concentration of 1.5 10-2 M.

at 3 kW. Scanning electron microscopy (SEM) images were obtained with a JEOL JSM-6330F operated at a beam energy of 10.0 kV. For SEM measurements, the samples were coated with platinum for higher magnification. Transmission electron microscopy (TEM), high-resolution TEM (HRTEM), and selectedarea electron diffraction (SAED) were performed on a JEOL JEM-2010 operated at acceleration voltage of 200 kV. Sample grids were prepared by sonicating dry samples in ethanol for 15 min and depositing one drop of the suspension onto a carbon foil supported on a copper grid for TEM investigations. Fourier transformation infrared (FT-IR) spectra were recorded on pellets of the samples mixed with KBr on a Bruker EQUINOX FT-IR spectrometer in the range of 500-4000 cm-1 at a resolution of 2 cm-1. A NETZSCH TG-209 unit was used to carry out the thermogravimetric and differential thermal analyses (TGA and DTA) under nitrogen atmosphere with a heat rate of 5 Cmin-1. Results and Discussion The SEM image and XRD patterns of the starting material are shown in parts a and b of Figure l, respectively. The starting material has irregular large particles with micrometer sizes and shows bulk orthorhombic MoO3 polycrystalline, which matches well the standard JCPDS data (JCPDS 76-1003). Figure 2 shows the morphology of the samples obtained by hydrothermal treatment of bulk MoO3 crystals in the presence of CTAB, measured by SEM. From this it can be clearly seen that CTAB concentration has a significant effect on the growth of MoO3. The morphology changed dramatically with CTAB concentration variation. As shown in parts b and c of Figure 2,

MoO3 nanofibers were obtained at CTAB concentrations ranging from 1.5 to 7.6 10-2 M, whereas excess CTAB in aqueous solution led to exclusively entangled microsized fibers. CTAB could be responsible for the formation of the as-synthesized nanofibers due to the fact that no fibers can be observed in the absence of CTAB. The reaction product consists almost exclusively of nanofibers when CTAB concentration was increased to 7.6 10-2 M. It is also found that the fibers become uniform in shape and warp slightly only in their ends at higher concentration of CTAB, accompanied by the decrease in the average diameter. The maximum length of the nanofibers obtained with CTAB concentration of 1.5 10-2 M is up to 20 m, and diameters range from several tenths to ca. 100 nm. XRD patterns of the nanofibers formed by the hydrothermal process with different CTAB concentrations are shown in Figure 3. Parts a and b of Figure 3 are the low-angle XRD patterns of the obtained products. It is interesting to note that triple interlayered mesostructures in MoO3 nanofibers can be observed, which can be identified by the appearance of three groups of 00l reflections. The peak (I) with the lowest intensity at 2 ) 3.05 reflects one type of interlayer distance with d value of 2.84 nm. The peak (II) with the intermediate intensity at 2 ) 3.27 and the peak (III) with the highest intensity at 2 ) 3.56 correspond to the other two types of interlayer distances with d values of 2.66 and 2.46 nm, respectively. The wide-angle XRD pattern of the sample is shown in Figure 3b and the inset in Figure 3c, with the reflections a little weaker than the 00l reflections in the corresponding low-angle XRD patterns. All diffraction peaks can be readily indexed to a pure orthorhombic

Lamellar Molybdenum Oxide 1D Nanostructures

J. Phys. Chem. B, Vol. 109, No. 48, 2005 22761

Figure 5. FTIR spectra of reagent crystalline CTAB (a), MoO3 as purchased (b), and the product prepared in the presence of 1.5 10-2 M CTAB (c).

Figure 4. Typical TEM image (a), SAED pattern of an individual nanofiber (b), and HRTEM image (c) of the obtained molybdenum oxide nanofibers at CTAB concentration of 7.6 10-2 M. (d) HRTEM image of lamellar MoO3 nanofibers obtained at CTAB concentration of 1.5 10-2 M.

phase (space group Pbnm62) with calculated lattice constants R ) 3.965 , b ) 13.857 , and c ) 3.694 , in agreement with the reported data (JCPDS 76-1003). It reveals that the crystal phase of the MoO3 layers is the same as that of the starting material. TEM was also employed to provide further insight into structure and morphology of the as synthesized MoO3 nanofibers obtained in the presence of CTAB. Figure 4 shows the representative results of TEM characterization of the as-prepared nanofibers, which confirms that the nanofibers are triple interlayered mesostructures. The selected ultralong nanofiber shown in Figure 4c has a diameter of approximately 30 nm and a respect ratio of 32. The walls appear as altering fringes of dark and bright contrast. The dark contrasts of narrow fringes represent the molybdenum oxide layers, between which the CTA+ cations are embedded. The layers of oxide in the selected nanofiber are separated by three characteristic distances of ca. 2.9, 2.7, and 2.4 nm, respectively, as indicated in Figure 4c, which agree well with the results calculated from the d values of the 00l reflections of the low angle XRD pattern. Figure 4b shows the corresponding SAED pattern taken from a single nanofiber shown in Figure 4a. The SAED pattern supports the well-crystallized structure inside the MoO3 fiber walls. Moreover, the lamellar structure is also confirmed by the multiple fringes observed in the image from the HRTEM investigation on the sample obtained at CTAB concentration of 1.5 10-2

M, as shown in Figure 4d. It is also found that most layers within the fibers are indeed unequally separated from each other. The repeat varying interlayer spacings between corresponding planes in the fiber wall structure give rise to strong reflections at the low-angle XRD pattern. There has long been much interest in the use of MoO3 because of the ability of this material to act as host for intercalation chemistry, in which guest molecules are believed to insert into the host lattice by electron transfer. In general, when smaller cations, such as alkalis, silver, and ammonium, are present in the reaction system, tunnel structures or three-dimensional lattices are formed.20 For the intermediate size cations such as tetramethylammonium, layered structures as well as Kegginlike cluster complexes dominate. For the larger cations, such as C12H25N+(CH3)3, flakelike layered microcrystals are found. In contrast, the combination of hydrothermal treatment with the use of CTAB here permits to promote the formation of MoO3 nanofibers with high aspect ratios. On the other hand, the assynthesized fibers have unique wall structures consisting of MoO3 layers that are separated by triple interlayer spacings. In the cationic surfactant-inorganic layered solid intercalating systems, the cationic headgroup is believed to be tethered to the internal surface of the galleries of the layered inorganic hosts by Coulomb forces.21 The surfactant molecules can adopt a variety of structures in the galleries, such as monolayer, bilayer, and paraffin-type monolayer, depending on the grafting density and the structure of the surfactant methylene tail. In addition, the methylene chain can behave as the trans or gauche conformer, which differ from each other distinctively in their vibrational modes.22 FT-IR spectroscopy has been used as a sensitive tool to probe the structure and organization of alkyl chain assemblies. The room-temperature FT-IR spectra in the 500-4000-cm-1 region of solid CTAB, lamellar MoO3 nanofibers prepared with CTAB concentration of 1.5 10-2 M and bulk MoO3 crystal are shown in Figure 5. On the basis of the literature values, the molybdenum oxide lattice vibrations appear at 553, 858, and 994 cm-1, respectively. The IR contour in Figure 5c indicates the signals of three bands at 550, 853, and 996 cm-1, which can be assigned to the lattice vibration mode of MoO3. Because of the presence of CTAB in the synthetic process, it is reasonable to expect the appearance of the relevant CTAB bands, especially together with XRD results. It has been reported that the antisymmetric and symmetric deformation modes pertaining to C16H33(CH3)3N+ of the CTAB headgroup appear at around 1390

22762 J. Phys. Chem. B, Vol. 109, No. 48, 2005

Song et al.

Figure 6. (a) Cartoon illustrating the formation of CTAB-mediated assembly of multilamellar structured MoO3 nanofibers. (b) Structure model of MoO3 sheets as seen along the [010] direction.

Figure 7. TGA and DTA curves collected from the pure CTAB (a) and the as-synthesized MoO3 nanofibers in the presence of 7.6 10-2 M CTAB (b). In all graphs, dashed lines correspond to the DTA curves and solid lines correspond to the TG curves.

and 1490 cm-l.22 In fact, from Figure 5 it can be seen that the bands shift to lower frequency as layered MoO3 nanofibers were formed. Two intense bands appear at 2846 and 2920 cm-1 in the IR spectrum of the multilamellar MoO3 nanofibers, which can be attributed to the antisymmetric and symmetric C-H stretching modes of the methylene groups, respectively. These frequencies are almost identical to those for the corresponding bands of solid CTAB, indicating that trans conformers dominate in the methylene chains.23 The absence of splitting of the 720 cm-1 band also supports the large distance between chains of trans conformers, leading to weaker lateral interchain interactions than those of the gauche conformers. Additional bands in the regions of 3250-3750 cm-1 can be also observed in all samples sets, indicating the presence of hydration of the cationic headgroup of the surfactant. The presence of CTA+ cations is most likely responsible for the anisotropic growth and the appearance of the three distinct interlayer spacings. The c-axis interlayer spacings of the nanosized fiberlike MoO3 intercalate, as calculated from the 00l reflections in the XRD patterns in Figure 3a, are 2.84, 2.66, and 2.46 nm, respectively. This corresponds to intercalations with lattice expansions of 2.15, 1.97, and 1.77 nm, compared to those of MoO3 bulk crystal.24 By comparison with the

expansions in d spacing with the length of the fully stretched CTA+ cations, 2.2 nm,25 it is concluded that the intercalated CTA+ cations adopt bilayer structures with different tilt angles between the MoO3 layers. Figure 6a shows the possible arrangements of the intercalated CTA+ cation that can account for the observed occurrence of three different expansions in interlayer spacings. Three panels as shown in Figure 6a have three tilted bilayers inserted without interdigitation, in which two all-trans methylene chains sandwich between the internal surfaces of the opposing oxide layers. The value of the tilt angles is determined from the X-ray lattice expansions assuming that the chains are all trans and that there is no interdigitation in the intercalated CTA+ cations bilayer. It is well known that surfactant molecules in aqueous solution above a critical micelle concentration self-aggregate to form micelles.26 And in our synthetic system, the bilayer micelles formed by CTAB serve as template and direct the formation of the MoO3 nanofibers. As shown in Figure 1b, the precursor is indexed to the orthorhombic phase of MoO3. By consideration of the structural particularity of MoO3, a selective adsorption mechanism is employed to interpret the formation of MoO3 nanofibers assisted by CTAB. As shown in Figure 6b, the orthorhombic structured MoO3 has a distinct layered structure,

Lamellar Molybdenum Oxide 1D Nanostructures

J. Phys. Chem. B, Vol. 109, No. 48, 2005 22763

Figure 8. SEM images of the MoO3 nanobelts with lamellar structures prepared by hydrothermal treatment in the presence of dodecylamine (a), tetradecylamine (b), hexadecylamine (c), and octadecylamine (d).

in which one monolayer is one layer of MoO6 octahedrons with corner sharing along the [100] direction and edge sharing along the [001] direction. And two MoO3 monolayers have the structure in which one of the slabs orient perpendicular to the [010] direction.27 Charge sensitivity analysis and periodic boundary density functional calculation have been developed to identify reactivity trends and chemisorption activity of MoO3 surfaces. It is proposed that the (010) MoO3 surfaces exhibit stronger preference for adsorption and catalytic activity.28 During the 1D anisotropic growth, the oxide bulk crystals dissolved and recrystallized under hydrothermal conditions, further renucleated at the selected surfactant-inorganic interfaces. CTAB micelles would play capping effect on the (010) MoO3 surfaces owing to the stronger activity and higher selectivity of the surfaces for the adsorption of CTAB. This inhibits crystal growth along the [010] direction. On the other hand, the bonding situations for MoO6 along the [100] direction is different from that along the [001] direction.29 Two Mo-O bonds will be formed if the octahedron grow along the [001] direction, while only one Mo-O bond forms in case of the growth along the [100] direction. From the point of energy view, the greater energy released in the growth along the [001] direction is favorable for the anisotropic growth along the [001] direction. Hence, the formation of the nanofibers is preferred. And the surfactant micelles are suggested to behave as glue that holds nanocrystals together with different lamellar spacings by steric and electrostatic interaction. As to different morphologies subject to variation in CTAB concentration, the main reason can be attributed to the competition between formation and stabilization of surfactant micelle structures, which is controlled by the counterion condensation. Prior to the addition of oxide precursor,

Figure 9. Low-angle XRD pattern of the nanobelts prepared with equivalent molar ratio of MoO3 and dodecylamine under hydrothermal treatment at 180 C for 7 days. Inset: wide-angle XRD pattern of the corresponding nanobelts.

the surfactant molecules are in a dynamic equilibrium between micelles and single molecules. Upon the addition of oxide, the number density of nanofibers increases with the strong growth of micelles of the surfactant. A bilayer of CTA+ should make the MoO3 layers positively charged. Thus, the growth into 1D nanostructures could be significantly restricted due to the higher condensation of Br- on the micelle surfaces with the increment in the size of surfactant micelle, as the solution becomes more concentrated.

22764 J. Phys. Chem. B, Vol. 109, No. 48, 2005

Song et al. the oxide layers, which is effective in improving the thermal stability of CTAB. In addition to the CTAB-directing 1D MoO3 multilamellar nanofibers described in this report, assembly of bulk MoO3 crystals into 1D nanobelts with lamellar mesostructures was also realized by using neutral long-chain primary amine (CnH2+1NH2, n ) 12, 14, 16, and 18) as the structure-directing template instead of CTAB. Following the same scenario in the fabrication of MoO3 nanofibers, layered MoO3 nanobelts were obtained by hydrothermal treatment of commercial bulk MoO3 crystals in the presence of monoamines. SEM images corresponding to the samples prepared by using dodecylamine, tetracylamine, hexacylamine, and octacylamine are shown in Figure 8, respectively. It shows that nanobelts were formed from bulk MoO3 crystals by hydrothermal treatment at 180 C for 7 days. It is found that the quantity of alkylamines is an important factor in determining the morphology of the final products. Welldefined 1D nanobelts can be grown at equivalent molar ratio of MoO3 to amines. As shown in Figure 8, some nanobelts were scrolled to form interesting morphologies such as water pipe for fire protection when dodecylamine or tetracylamine was used as the template. The length of the obtained nanobelts is usually several tenths of micrometers, and their widths are typically 200-600 nm, significantly wider and longer than those of MoO3 nanofibers obtained in the presence of CTAB. As a result, widthto-thickness ratios of the nanobelts are usually in the range of ca. 3-12. These lamellar mesostructures were formed by the intercalation of surfactant into MoO3 layers, and lamellar sheets were further rolled into the nanobelts under the certain conditions. It should be noted that the low-angle XRD patterns for the products obtained by the use of n-alkylamines have two groups of 00l reflections, as similarly observed for the case in the presence of CTAB. Taking the intercalation of dodecylamine in MoO3 as an example, Figure 9 shows the XRD patterns of the corresponding products with morphology as shown in Figure 8a. After assembly of MoO3 directed by dodecylamine, the reflections of MoO3 bate and a new phase appears with two

Figure 10. A linear relationship between interlayer distance and the number of carbon atoms in alkylamines. Interlayer distances were determined from the d value of the two groups of 00l reflections in the XRD patterns of the MoO3 nanobelts obtained by using alkylamines as the template.

Thermal stability of the as-synthesized nanofibers was also investigated to further comprehend the formation process of the nanofibers. Figure 7 shows the TGA and DTA curves of pure CTAB and the as-synthesized MoO3 nanofibers. As expected, the TGA-DTA of pure solid CTAB exhibits a single mass loss, a distinct exothermic peak accompanied by 100% mass loss occurred at temperatures in the range of 200-270 C. On the other hand, the TGA curve of obtained MoO3 nanofibers can be divided into three stages. In the first stage, no other components but lattice H2O molecules were released before 300 C with a mass loss of 1.27%. Two sequential exothermic peaks between 300 and 400 C with a general mass loss of 24.0% could be attributed to the CTA+ decomposition and the elimination of bromide species. The fact that the synthesized MoO3 nanofibers is thermally stable until 300 C, suggesting the affinity between the CTA+ cations and the bonding sites of

Figure 11. TEM images of the beltlike MoO3 with lamellar structures prepared by the hydrothermal method using different neutral amines as the template at 180 C for 7 days with 1:1 molar ratio of MoO3 to amines: (a) dodecylamine, (b) tetradecylamine, (c) hexadecylamine.

Lamellar Molybdenum Oxide 1D Nanostructures interlayer d spacings of 2.88 and 3.04 nm, resulting from the concurrent adsorption and insertion of dodecylamine in the MoO3 sheets during the hydrothermal process. As shown in Figure 9, the two distinct interlayer distances correspond to two different orientations of the bilayers formed by dodecylamines between the adjacent oxide layers. The inset shows less intense peaks at smaller d values, corresponding to the structure within the layers. The diffusing reflections can be indexed to orthorhombic phase MoO3. Figure 10 shows the variations in the distance between the layers, caused by varying template with different alkyl chains. It indicates that the interlayer distance increases proportional to alkyl length of the n-alkylamine. The 00l peaks shift slightly to lower angles, reflecting an increase in the lamellar spacing with increasing the number of carbon atoms in the n-alkylamine templates chosen for study here. Although the reflections of type 00l move to larger d values with the increment of the alkyl chain length, the reflections in wide-angle XRD patterns of the corresponding products are independent of the alkylamine templates. The result suggests that template has no effect on the wall structure within the layers. TEM images of the beltlike products obtained using neutral amines with alkyl chains of different lengths are presented in Figure 11. It is evident from the images that the lamellar structures could be formed across long distance independent of the alkyl length of the amines. The warped shape of the nanobelts might be attributed to mechanical stress caused by increased interlayer expansion due to the insertion of nalkylamines between two adjacent oxide layers, as compared to the case for CTAB-templated assembly of nanofibers. The reason might lie in that neutral long-chain primary amines have lower steric effects than CTAB with respect to few branched groups. Conclusions Our present study shows that 1D MoO3 nanofibers and nanobelts with multilamellar mesostructures can be synthesized from commercial bulky MoO3 crystals by a surfactant-templated hydrothermal process. The combination of hydrothermal treatment with the use of surfactant promotes the anisotropic growth via SLS transformation. CTAB can be successfully inserted into MoO3 layers with three tilt angles to form a novel intercalation structure with three interlayer distances in aqueous solution, and these lamellar sheets are further grown into 1D nanostructures. The as-synthesized fiber-shaped oxide with CTA+ cations intercalated in improves the thermal stability of the surfactant. The choice of surfactants and their concentration greatly influence the morphology and structure of the resultant products. MoO3 nanobelts with bilamellar mesostructures were also prepared by the same method employing neutral amine with long alkyl chain as intercalating template. Well-developed nanobelts with two interlayer distances can be obtained with equivalent molar ratio of Mo/alkylamines after duration in aqueous solution. The interlayer distances can be easily controlled by using monoamines with different length of alkyl chains. The present work provides much inspiration toward the formation of a target structure with desired morphologies, arrangements, and even improved functional properties. The flexible strategy reported herein is being expanded to prepare other 1D nanostructured materials. For example, starting from bulk Co2O3 crystals with the use of alkylamines, we have prepared cobalt oxide nanowires. Tungsten oxide nanosheets have also been obtained from bulk WO3 crystals in the presence

J. Phys. Chem. B, Vol. 109, No. 48, 2005 22765 of alkylamines. Surfactant-assisted hydrothermal synthesis of 1D binary oxidic nanomaterials from the corresponding commercial bulk powders is also in progress. Acknowledgment. We are grateful to the National Natural Science Foundation of China (20371053) and the Natural Science Foundation of Guangdong Province (031574, 04300770) for financial support. A.-W. Xu thanks the Alexander von Humboldt Foundation for granting a research fellowship. This work was also supported by the Young Teacher Starting Project of Sun Yat-Sen University (for R.-Q. Song). References and Notes
(1) (a) Alivisatos, A. P. Science 1996, 271, 933. (b) Heath, J. R.; Shiang, J. J. Chem. Soc. ReV. 1998, 27, 65. (c) El-Sayed, M. A. Acc. Chem. Res. 2001, 34, 257. (d) Huang, M. H.; Mao, S.; Feick, H.; Yan, H. Q.; Wu, Y. Y.; Kind, H.; Weber, E.; Russo, R.; Yang, P. D. Science 2001, 292, 1897. (e) Law, M.; Kind H.; Messer. B.; Kim, F.; Yang, P. D. Angew. Chem., Int. Ed. 2002, 41, 2405. (f) Xiong, Y. J.; Xie, Y.; Li, Z. Q.; Li, X. X.; Gao, S. M. Chem.-Eur. J. 2004, 10, 654. (g) Luo, Y.; Lee, S. K.; Hofmeister, H.; Steinhart, M.; Gosele, U. Nano Lett. 2004, 4, 143. (h) Sun, Z. Y.; Liu, Z. M.; Han, B. X.; Wang, Y.; Du, J. M.; Xie, Z. L.; Han G. J. AdV. Mater. 2005, 17, 928. (i) Hu, J.; Odom, W.; Lieber, C. M. Acc. Chem. Res. 1999, 32, 435. (2) (a) Tenne, R.; Margulis, L.; Genut, M.; Hodes, G. Nature 1992, 360, 444. (b) Trentler, T. J.; Hickman, K. M.; Geol, S. C.; Viano, A. M.; Gibbons, P. C.; Buhro, W. E. Science 1995, 270, 179l. (c) Dai, H. J.; Wong, E. W.; Lu, Y. Z.; Fan, S. S.; Lieber, C. M. Nature 1995, 375, 769. (d) Krumeich, F.; Muhr, H. J.; Niederberger, M.; Bieri, F.; Schnuder, B.; Nesper, R. J. Am. Chem. Soc. 1999, 121, 8324. (e) Pan, Z. W.; Dai, Z. R.; Wang, Z. L. Science 2001, 291, 1947. (f) Fang, Y. P.; Xu, A. W.; Song, R. Q.; Zhang, H. X.; You, L. P.; Yu, J. C.; Liu, H. Q., J. Am. Chem. Soc. 2003, 125, 16025. (g) Zhan, B. Z.; White, M. A.; Sham, T. K.; Pincock, J. A.; Doucet, R. J.; Rao, K. V. R.; Robertson, K. N.; Cameron, T. S. d. J. Am. Chem. Soc. 2003, 125, 2195. (h) Zhang, L. Z.; Yu, J. M.; Xu, A. W. Li, Q.; Kwong, K. W.; Wu, L. Chem. Commun. 2003, 2910. (i) Fang, Y. P.; Xu, A. W.; Dong, W. F. Small 2005, 1, 967. (j) Patzke, G. P.; Krumeich, f.; Nesper, R. Angew. Chem., Int. Ed. 2002, 41, 2446. (3) (a) Schaak, R. E.; Afzal, D.; Schottenfeld, J. A.; Mallouk, T. E. Chem. Mater. 2002, 14, 442. (b) Salah, M. B.; Vilminot, S.; Mhiri, S.; Kurmoo, M. Eur. J. Inorg. Chem. 2004, 2272. (c) Ookuba, K. O.; Hayashi, H. Langmuir 1993, 9, 1418. (d) Constantino, V. R.; Pinnavvania, T. J. Inorg. Chem. 1995, 34, 883. (e) Xu, A. W.; Cai, Y. P.; Zhang, H. X.; Zhang, L. Z.; Yu, J. C. Angew. Chem., Int. Ed. 2002, 41, 3844. (f) Hagrman, P. J.; Laduca, R. L., Jr.; Koo, H. J.; Rarig, R., Jr.; Haushalter, R. C.; Whangbo, M.-H.; Zubieta, J. Inorg. Chem. 2000, 39, 4311. (g) Kamath, P. V.; Dixit, M.; Indira, L.; Shukla, A. K.; Kumar, V. G.; Munichandraiah, N. J. Electrochem. Soc. 1994, 141, 1487. (4) (a) Niederberger, M.; Muhr, H.-J.; Krumeich, F.; Bieri, F.; Gu nther, D.; Nesper, R. Chem. Mater. 2000, 12, 1995. (b) Chen, X.; Sun, X. M.; Li, Y. D. Inorg. Chem. 2002, 41, 4524. (5) (a) Liu, H. F.; Liu, R. S.; Liew, K. Y.; Johnson, R. E.; Lunsford, J. H. J. Am. Chem. Soc. 1984, 106, 4117. (b) Baiker, A.; Gasser, D. J. Phys. Chem. 1986, 149, 119. (c) Zhang, W.; Oyama, S. T. J. Phys. Chem. 1996, 100, 10759. (d) Gu nther, S.; Marsi, M.; Kolmakov, A.; Kiskinova, M.; Noeske, M.; Taglauer, E.; Mestl, G.; Schubert, U. A.; Knozinger, H. J. Phys. Chem. B 1997, 101, 10004. (e) Silva, I. F.; Klimkiewicz, M.; Eser, S.; Energy Fuels 1998, 12, 554. (f) Baiker, A.; Dollemeier, P.; Reller, A. J. Catal. 1987, 103, 394. (g) Ressler, J.; Wienold, T.; Jentoft, R. E. J. Catal. 2002, 210, 67. (6) Kerr, T. A.; Leroux, F.; Nazar, L. F.Chem. Mater. 1998, 10, 2588. (7) (a) Yao, J. N.; Yang, Y. A.; Loo, B. H. J. Phys. Chem. B 1998, 102, 1856. (b) Yang, Y. A.; Cao, Y. W.; Loo, B. H.; Yao, J. N.; J. Phys. Chem. B 1998, 102, 9392. (c) Ferreira, F. F.; Souza Cruza, T. G.; Fantini M. C. A.; Tabacniks, M. H.; de Castro S. C.; Morais, J.; de Siervo, A.; Landers, R.; Gorenstein, A. Solid State Ionics 2000, 136-137, 357. (8) Hosono, K.; Matsubara, I.; Murayama, N.; Woosuck, S.; Izu, N. Chem. Mater. 2005, 17, 349. (9) Wang, J.; Rose, K. C.; Lieber, C. M. J Phys. Chem. B 1999, 103, 8405. (10) McEvoy, T. M.; Stevenson, K. J. J. Mater. Res. 2004, 19, 429. (11) Satishkumar, B. C.; Govindaraj, A.; Nath, M.; Rao, C. N. R. J. Mater. Chem. 2000, 10, 2115. (12) Lou, X. W.; Zeng, H. C. Chem. Mater. 2002, 14, 4781. (13) Patzke, G. R.; Michailovski, A.; Krumeich, F.; Nesper, R.; Grunwaldt, J. D.; Baiker, A. Chem. Mater. 2004, 16, 1126. (14) Niederberger, M.; Krumeich, F.; Muhr, H.-J.; Mu ller, M.; Nesper, R. J. Mater. Chem. 2001, 11, 1941. (15) Li, X. L.; Liu, J. F.; Li, Y. D. Appl. Phys. Lett. 2002, 81, 4832.

22766 J. Phys. Chem. B, Vol. 109, No. 48, 2005


(16) Li, Y. B.; Bando, Y. S. Chem. Phys. Lett. 2002, 364, 484. (17) (a) Xu, A. W.; Fang, Y. P.; You, L. P.; Liu, H. Q. J. Am. Chem. Soc. 2003, 125, 1494. (b) Fang, Y. P.; Xu, A. W.; You, L. P.; Song, R. Q.; Yu, J. C.; Zhang, H. X.; Li, Q.; Liu, H. Q. AdV. Func. Mater. 2003, 13, 955. (18) Zhang, L. Z.; Yu, J. C.; Xu, A. W.; Li, Q.; Kwong, K. W.; Wu, L. Chem. Commun. 2003, 2910. (19) Yu, J. C.; Xu, A. W.; Zhang, L. Z.; Song, R. Q.; Wu, L. J. Phys. Chem. B 2004, 108, 64. (20) Whittingham, M. S.; Guo, J. D.; Chen, R. J.; Chirayil, T.; Janauer, G.; Zavalij, P. Solid State Ionics 1995, 75, 257. (21) Dugaev, V. K. Phys. Stat. Sol. B 2000, 219, 31. (22) (a)Venkataraman, N. V.; Vasudevan, S. J. Phys. Chem. B 2000, 104, 11179. (b) Venkataraman, N. V.; Vasudevan, S. J. Phys. Chem. B 2002, 106, 7766. (23) (a) Lagaly, G. Angew. Chem., Int. Ed. Engl. 1976, 15, 575. (b) Snyder, R. G.; Strauss, H. L.; Elliger, C. A. J. Phys. Chem. 1982, 86, 5145.

Song et al.
(c) MacPhail, R. A.; Strauss, H. L.; Synder, R. G.; Elliger, C. A. J. Phys. Chem. 1984, 88, 334. (d) Vaia, R. A.; Teukolsky, R. K.; Giannelis, E. P. Chem. Mater. 1994, 6, 1017. (e) Li, H. Y.; Tripp, C. P. Langmuir 2002, 18, 9441. (24) Bissessur, R.; DeGroot, D. C.; Schindler, J. L.; Kannewurf, C. R.; Kanatzidis, M. G. J. Chem. Soc., Chem. Commun. 1993, 687. (25) (a) Venkataraman, N. V.; Vasudevan, S. Proc. Ind. Acad. Sci. 2001, 113, 539. (b) Weidemaier, K.; Taverniner, H. L.; Fayer, M. D. J. Phys. Chem. B 1997, 101, 9352. (26) Mohamed, B.; Barney, L. B.; Raoul, Z. J. Phys. Chem. B 2003, 107, 13432. (27) Kihlborg, L. AdVances in chemistry: The crystal chemistry of molybdenum oxides; American Chemical Society: Washington, 1963. (28) (a) Nalewajski, R. F.; Michalak, A. J. Phys. Chem. A 1998, 102, 636. (b) Yin, X. L.; Han, H. M.; Miyamoto, A. J. Mol. Model 2001, 7, 207. (29) Zeng, H. C. Inorg. Chem. 1998, 37, 1967.

You might also like