Вы находитесь на странице: 1из 11

J Nanopart Res (2010) 12:28832893 DOI 10.

1007/s11051-010-9879-4

RESEARCH PAPER

Morphological control of Ni/NiO core/shell nanoparticles and production of hollow NiO nanostructures
Nitin Chopra Leslie Claypoole Leonidas G. Bachas

Received: 18 September 2009 / Accepted: 9 February 2010 / Published online: 27 February 2010 Springer Science+Business Media B.V. 2010

Abstract Chemical synthesis coupled with a microwave irradiation process allowed for the control of size (640 nm), shape, and shell thickness of Ni/NiO core/shell nanoparticles. In this unique synthetic route, the size of Ni nanoparticles (NiNPs) was strongly inuenced by the nickel salt-to-stabilizer ratio and the amount of the stabilizer. Interestingly, it was observed that the shape of the nanoparticles was altered by varying the reaction time, where longer reaction times resulted in annealing effects and rupture of the stabilizer micelle leading to distinct shapes of Ni/NiO core/shell nanostructures. Product cooling rate was another important parameter identied in this study that not only affected the shape, but also the crystal structure of the core/shell nanoparticles. In addition, a simple and cost-effective method of microwave irradiation of NiNPs led to the formation of distinctly shaped hollow NiO nanoparticles.
L. Claypoole2007 NSF-REU Fellow. L. G. Bachas (&) Department of Chemistry, University of Kentucky, Lexington, KY 40506, USA e-mail: bachas@uky.edu N. Chopra Department of Metallurgical and Materials Engineering, Center for Materials for Information Technology (MINT), University of Alabama, Tuscaloosa, AL 35487, USA L. Claypoole Fairmont State University, Fairmont, WV, USA

These high surface area core/shell nanoparticles with well-controlled morphologies are important and can lead to signicant advancement in the design of improved fuel cells, electrochromic display devices, and catalysis systems. Keywords Core/shell nanoparticles Nickel Nickel oxide Morphology Hollow nanostructures Crystal defects Nanocomposites Nanomanufacturing

Introduction Developing simple, economical, and high-yield methods for the production of nanoparticles is critical for their application in devices and catalytic systems (Service 2004; Bratlie et al. 2007; Caruso et al. 1998; Shenhar et al. 2005; Hall et al. 2000; Palchik et al. 1999a). In spite of successes in manufacturing various nanostructures, including core/shell nanoparticles (Fendler 1998; Rao et al. 2004; Sako et al. 1994; Meneses et al. 2007), morphological control of metal/ metal oxide core/shell nanoparticles remains a challenge. The physical properties of core/shell nanoparticles are strongly inuenced by their shape and size as well as the surface chemistry of their shells (Lu et al. 2007; Lee et al. 2006). Most synthetic approaches for metal/metal oxide core/shell nanoparticles rely on reduction and precipitation of the metal ion followed by oxidation in air. Because of the high surface

123

2884

J Nanopart Res (2010) 12:28832893

tension of metals, such approaches result predominantly in a spherical-shaped nanoparticle core, with the shell forming over this core. However, under certain stringent synthetic conditions that favor anisotropic crystal growth, formation of cubic, tetrahedral, and triangular nanoparticles (not in a core/ shell conguration) has been observed (Dumestre et al. 2004; Ahmadi et al. 1996; Chen and Caroll 2002). Control of anisotropic growth in crystals in a manner that results in core/shell nanoparticles with well-controlled morphology remains a challenge. Nanostructured nickel and its oxides are of special interest because of their electronic and magnetic properties and their surface chemistry (Lee et al. 2006; Pinarbasi et al. 2000). NiO is a stable wide direct band-gap material (3.56 eV) and exhibits ptype semiconducting behavior (Sato et al. 1993). Nanosized NiO can be superparamagnetic and superanti-ferromagnetic (Khadar et al. 2003; Makhlouf et al. 1997). Such nanostructures have potential applications in catalysis, rechargeable batteries, and fuel cells (Park et al. 2005a, b; Poizot et al. 2000). For all these applications, it is critical that the nanostructures have maximum available surface area. The synthesis of NiO nanostructures, including Ni/NiO core/shell nanoparticles, employing a variety of precursor nickel salts and synthetic routes, has been previously reported (Lee et al. 2006; Hou et al. 2005; Cordente et al. 2001; Parada and Moran 2006; He et al. 2005; Li et al. 2006a, b; Ghosh et al. 2006; Xiang et al. 2002; Jana et al. 2004; Li et al. 2004). Most of these methods are tedious and energyintensive because of the high temperatures required for the growth of the nanostructures. There are four major synthetic issues that are yet to be addressed for Ni/NiO core/shell nanoparticles, namely, (1) availability of a facile and reproducible production method to give large yields of nanoparticles with uniform size distribution, (2) ability to control the shape and size of the core and the shell, (3) a faster dry oxidation method, surpassing the limitations of the much slower and energy-consuming air/thermal oxidation, to result in an oxide shell with wellcontrolled thickness, and (4) understanding the effects of various growth parameters on the morphology of the core/shell nanoparticles. In regards to the last issue, wet-chemical synthetic routes are challenging when it comes to controlling the shape of nanoparticles. Numerous reports demonstrate the use

of organic surfactants to control the shape of semiconducting nanoparticles (He et al. 2005). To the authors knowledge, shape control of Ni/NiO core/shell nanoparticles remains elusive. A previous attempt to synthesize different shapes of Ni nanoparticles (NiNPs) with NiO shell resulted predominantly in near-spherical and elongated shapes (Nayak et al. 2006). In this study, we address all four issues mentioned above through the development of a new synthetic methodology that couples chemical synthesis and microwave irradiation to successfully achieve control of the size (640 nm) and, most importantly, the shape and shell thickness of Ni/NiO core/shell nanoparticles. Furthermore, we report for the rst time a one-step extended microwave irradiation of NiNPs to form hollow NiO nanostructures. These results offer a robust technique to tailor the morphology of Ni/NiO core/shell nanoparticles and also aid in understanding the mechanism involved.

Experimental details Materials Ni(II) acetate tetrahydrate (98%), oleylamine (70%), and N,N-dimethylformamide (99.8%) were obtained from Aldrich (Milwaukee, WI). Tri-n-octylphosphine oxide (TOPO, 98%) and tri-n-octylphosphine (TOP, 90%) were obtained from Alfa Aesar (Ward Hill, MA). Ethanol (200 proof) and acetone were obtained from Decon Labs (King of Prussia, PA) and Fisher Scientic (Pittsburgh, PA), respectively. n-Hexane (85%) was obtained from Mallinckrodt (Hazelwood, MO). All the above chemicals were used without further purication. Transmission electron microscopy (TEM) was carried out on a JEOL 2010 F HRTEM (Japan). Ultrasonication was performed in a Branson 2210 Ultrasonicator (Danbury, CT), and the centrifuge separation was done in an Avanti J-25I centrifuge (Fullerton, CA). Microwave processing was performed in a Sharp Carousel domestic microwave oven (Model #: R209BK, 950 W, 120 V, Single Phase, 60 Hz, AC only). Samples were stored in a vacuum oven. XRD data were taken with a Rigaku D-Max III X-Ray Diffraction Unit (The Woodlands, TX).

123

J Nanopart Res (2010) 12:28832893

2885

Ni/NiO core/shell nanoparticles synthesis Various concentrations of reactants and reaction times were utilized to synthesize the Ni/NiO core/ shell nanoparticles. Nickel(II) acetate tetrahydrate (8 9 10-4, 4 9 10-3, and 20 9 10-3 mol) and 712 mL (2.1 9 10-23.6 9 10-2 mol) of oleylamine were added to a 100 mL, two-neck round bottom ask, purged with Ar gas and covered with a rubber septum. Unless stated, all reactions were carried out in an inert (Ar) atmosphere and with 7 mL oleylamine. The ask was heated in an oil bath, with constant stirring, between 80 and 90 C for *30 min. TOPO (0, 1.3 9 10-4, 7.8 9 10-3 mol) and 1 mL (2.2 9 10-3 mol) of TOP were added to the ask. The solution turned from green to dark green. The ask was slowly heated to temperatures between 245 and 255 C and held there for 30 480 min. The solution turned black after reaching a temperature of approximately 230 C. The color change to black signaled the formation of NiNPs. The reaction mixture was allowed to cool down to room temperature. Thereafter, the precipitate was rigorously washed with ethanol and acetone and briey ultrasonicated. The precipitate was vacuumdried overnight, collected, and immediately stored in a desiccator purged with Ar. The nanoparticles were placed in a microwave oven (Sharp Carousel domestic microwave oven Model R209BK) with microwave power set at high for a duration of 30180 min to allow the formation of an oxide shell. It was observed that these nanoparticles, dispersed in water, showed magnetic character (Fig. 1). Nanoparticles were characterized using TEM at the Microscopy Facility (University of Kentucky). The sample was prepared by dispersing a drop of solution (in acetone) onto a Cu-Lacey carbon TEM grid obtained from EM Sciences (Hateld, PA), which was allowed to dry in a vacuum oven (23 inch Hg) at room temperature overnight. In regard to the surface species on NiNPs, previous studies (Sharma et al. 2008; Davar et al. 2009) on other nanoparticle syntheses using the same stabilizers have shown that TOPO or oleylamine easily bind to the nanoparticle surface as evidenced by Fourier transform infrared spectroscopy. TOPO, TOP, and oleylamine have certain afnity for the NiNP surface, and no other stabilizer was utilized in our study. Thus, these stabilizer molecules are present on the

Fig. 1 Highly magnetic Ni/NiO core/shell nanoparticles dispersed in water. Nanoparticles tend to attach to the tweezer tip (inset)

NiNPs. It is expected that these surface species may be eliminated under three circumstances: (a) the second step of the synthesis where TOPO and TOP are injected and the reaction is carried out above ash points of either of these stabilizers, (b) rigorous postsynthesis washing of the nanoparticles, and (c) microwave oxidation of the nanoparticles.

Results and discussion In this work, we employ a chemical precipitation route to synthesize NiNPs, where thermal decomposition of nickel acetateoleylamine complex occurs in the presence of stabilizers (TOPO and TOP) at *245255 C. The nickel acetateoleylamine complex was formed by heating nickel acetate with a xed volume of oleylamine at 80 C for 3040 min. Unless stated, all the reactions were carried out in an inert (Ar) atmosphere. The overall synthetic procedure for Ni/NiO core/shell nanoparticles is shown in Scheme 1. Stabilizers (TOPO and TOP) were used to control the growth of nanoparticles and coat the nanoparticles to prevent them from further oxidation and aggregation. These nanoparticles were easily dispersed in ethanol or hexane to form a homogeneous colloidal solution, which is usually stable for several days. NiNPs can be oxidized in air to form the oxide shell outside but that process would have taken several days and is uncontrolled (Lee et al. 2006). To enhance the oxidation process and to control the shell

123

2886 Scheme 1 Chemical reaction route to synthesis of Ni/NiO coreshell nanoparticles

J Nanopart Res (2010) 12:28832893

i) 80 C, 30 min

(C2H3O2)2Ni. 4H2O + C18H37N


ii) 245 C TOPO, TOP
o

P PP P P Ni P P P P P PP

Washing Microwave

Ni
NiO shell

thickness, microwave processing was performed on NiNPs resulting in magnetic Ni/NiO core/shell nanoparticles (Fig. 1). Systematic studies were performed to control the size and shape of Ni/NiO core/ shell nanoparticles. The effect of precursor nickel salt concentration, stabilizer (TOPO) concentration, reaction time, product cooling rate, and microwave duration were thoroughly investigated. Effect of precursor nickel salt and stabilizer amounts on the nanoparticle size We observed that the concentration of the precursor nickel salt (nickel acetate) inuenced the size of the NiNPs. For a xed amount of TOPO (7.8 9 10-3 mol), changing the amount of nickel salt (ratio of nickel salt/TOPO from 0.09 to 2.6 mol/mol) resulted in an increase in the size of the nanoparticles (from 7.4 1.0 to 37.3 6.4 nm, Fig. 2). On the other hand, increasing the amount of stabilizer only (TOPO, 07.8 9 10-3 mol) resulted in size reduction of NiNPs (from 302 47 to 26.5 5.0 nm, Fig. 3). Apart from being an important size-controlling parameter, TOPO acts as a passivating agent preventing oxidation of NiNPs. This stabilizer forms a micelle around individual Ni nuclei, preventing coalescence of the latter while facilitating crystal growth (Yin and Alivisatos 2005). It is the density or packing of the micelle that is affected by the concentration of the stabilizer (Hou et al. 2005). A higher concentration of TOPO leads to a tightly bound micelle on the nickel nuclei as TOPO can form

complex with nickel. This is the reason for the impeded crystal growth and the reduction in the size of NiNPs with increasing TOPO concentration. The fact that TOPO can also aid in controlling the growth rate of nanoparticles (Hou et al. 2005) suggests that the stabilizer concentration becomes an important parameter for controlling the size of NiNPs. It was also observed that TOPO, when present at the concentrations tested, did not affect the shape of the nanoparticles. Thus, contrary to a proposed growth mechanism for metal phosphide nanorods (Park et al. 2005a, b). TOPO, TOP, and any remaining oleylamine do not appear to be selective to a particular crystal facet in the growing NiNPs. This is further evidenced by the spherical shape of the NiNPs even in the absence of TOPO and TOP (Fig. 3B). Further, atmospheric exposure of nanoparticles led to the formation of a thin oxide shell (\5 nm), resulting in Ni/NiO core/shell nanoparticles. Effect of reaction duration on the nanoparticle shape In order to understand the effect of reaction duration on the morphology of the NiNPs, xed amounts of nickel acetate (4 9 10-3 mol) and oleylamine (7 mL) were reacted at 80 C for 3040 min. Subsequently, TOPO (3.8 9 10-3 mol) and TOP (1 mL) were injected and the temperature increased to *245255 C. This reaction was carried out for different durations and the produced nanoparticles were characterized to observe the changes in

123

J Nanopart Res (2010) 12:28832893 Fig. 2 Transmission electron micrographs (TEM) and size (diameter, dA) distribution of nanoparticles prepared using different nickel acetate/TOPO (mol/mol) ratios; A mole ratio 0.09 (dA = 7.4 1.0 nm), B mole ratio 0.51 (dA = 21.9 3.1 nm), and C mole ratio 2.57 (dA = 37.3 6.4 nm). Note: Nanoparticles lose their 2D superlattice structure as the size of the nanoparticles increases (evident in A, but not in B)
80 70 60

2887

A
Counts

50 40 30 20 10 0 5 10 15 20 25 30 35 40 45 50 55 60 65

Diameter (nm)

50 40

Counts

30 20 10 0

20 nm

10 15 20 25 30 35 40 45 50 55 60 65

Diameter (nm)
50 40

Counts

30 20 10 0 5 10 15 20 25 30 35 40 45 50 55 60 65

Diameter (nm)

morphology. To overcome the challenge of controlling the shapes of the nanoparticles in a wet-chemical synthetic route, we undertook a shape evolution study to identify important parameters that can be used to tailor the morphology of Ni/NiO core/shell nanoparticles. We observed that variations in the reaction time alone, after the injection of TOPO and TOP, did not have a signicant effect on the dimensions (18 20 nm) of the Ni/NiO core/shell nanostructures. However, our studies show that after 1-h reaction time (Fig. 4A), some particles developed kinks or corners. After 2-h reaction time (Fig. 4B), some rhomboid-shaped structures were observed that further changed into squares/rectangles beyond 48 h reaction duration (Fig. 3C, D). Thus, shorter reaction

times resulted in spherical nanoparticles, while longer reaction times introduced corners, producing triangular, rhomboid, hexagonal, and square shapes. Slow growth rates at shorter reaction times ensure that each growing particle is allowed to achieve equilibrium manifested by a spherical shape. Longer reaction times favor unidirectional growth of one facet resulting in more kinked or elongated shapes. As stated, the stabilizers used in the synthesis were not selective to any crystal facet of the growing NiNP and did not affect the shape of the nanoparticles. This unidirectional growth is due to annealing at longer reaction times, damaging the stabilizer micelle around the growing crystal. The location of the rupture of the micelle allows crystal growth on that

123

2888

J Nanopart Res (2010) 12:28832893

80 70

80 70 60

Counts

60 50

Variable 3 g TOPO, 21.90 +/- 3.082 nm 0.05 a TOPO, 26.48 +/- 4.994 nm g 0 g TOPO, 302.4 +/- 47.36 nm

Effect of cooling rate on the crystal structure of the nanoparticle It has been reported that the morphology of nanoparticles depends, among other factors, on the cooling rate of the product; faster cooling rates generate strains and dislocations affecting the defect distribution in the nal structure (Shen et al. 1997; Kim and Kwon 2003). These defects act as trapping and scattering centers determining the electrical and optical properties of the nanomaterials. We observed that when the NiNP solution was quenched on ice immediately after synthesis, the nanoparticles acquired oval shapes and developed defects (Fig. 5) as compared to the spherical nanoparticles observed under slower cooling (Fig. 2). The crystal structure of the nanoparticles can be affected at higher temperatures ([200 C) because of atom diffusion or phase transitions (Puntes et al. 2001). One of the effects can be the creation of defects in the nanoparticle crystal structure. Keeping this in mind, fast cooling of nanoparticles (on ice) could cause sudden cessation of such processes resulting in lattice and thermal strains. This can be a reason that led to crystal defects and the oval shape of the NiNPs (Fig. 5B). Such an observation for NiNPs has not been reported earlier. However, a molecular dynamics (MD) study indicates that cooling rates can affect the crystal structure of NiNPs (Gafner et al. 2004). In addition, similar cooling effects on morphology have been observed for Mg nanostructures, where the shape changes as a function of the cooling rates (Kang et al. 2007). Microwave processing of the NiNPs and the formation of NiO shell In order to achieve full control over the morphology of Ni/NiO core/shell nanoparticles, it is not only important to tailor the morphology of the Ni core, but also to control the thickness of the oxide shell. The latter is an important parameter in determining the properties of core/shell nanostructures. Previous studies have demonstrated that air oxidation of NiNPs results in NiO shells (Lee et al. 2006; Park et al. 2005a, b; Song et al. 2008; Karmhag et al. 2001). However, it is extremely important to develop a synthetic method that will allow significant reduction in the oxide growth temperatures and

50 40 30 20 10 0

Counts

40 30 20 10 0 60 120

16

20

24

28

32

36

40

Diameter (nm)

c
180 240 300 360 420 480

Diameter (nm)

1 m

Fig. 3 Size (diameter) distribution of Ni/NiO core/shell nanoparticles corresponding to different concentrations of phosphine-based stabilizing agent (TOPO); A size distribution for three different TOPO amounts (a: 7.8 9 10-3 mol, dA (average diameter): 21.9 3.1 nm; b: 1.3 9 10-4 mol, dA: 26.5 5.0 nm; c: 0 mol, dA: 302 47 nm). Inset Nanoparticle size distributions for a and b. B TEM image of the Ni/NiO core/shell nanoparticles synthesized without TOPO. Note: As the amount of TOPO was varied, the amount of nickel acetate (4.0 9 10-3 mol), oleylamine (7 mL), and TOP (1 mL) were kept constant

facet resulting in a new shape or the formation of corners in the nanoparticles. This micelle is most likely composed of oleylamine, TOPO, and TOP. Of these, oleylamine, which is a weak stabilizer and has a ashpoint of 170 C, is eliminated at the reaction temperature (245255 C), whereas TOPO is more thermally stable (ash point: 252 C). The reaction temperature being near the ashpoint of TOPO causes it to partially vaporize with time. This leaves gaps on the micelle around the growing nanoparticle leading to corners. The proposed mechanism is schematically illustrated in Fig. 4E. Loss of surfactant capping layers during the annealing process leads to nanoparticle aggregation and faster oxidation upon air exposure. Thus, as demonstrated in this study, the reaction time is an important parameter for controlling the shape of the core and, after oxidation, the shell of Ni/NiO core/shell nanoparticles.

123

J Nanopart Res (2010) 12:28832893 Fig. 4 TEM image of Ni/ NiO core/shell nanostructures synthesized using different reaction times: A formation of corners and edges on the nanoparticles after 1-h reaction time; B hexagonal and rhomboid-shaped nanoparticles start forming after 2 h of reaction; C formation of square and rectangular nanoparticle after 4-h reaction time; D complete transformation of spherical nanoparticles to different shaped nanoparticles (mainly square and rectangular) after 8-h reaction time; and E mechanism of shape change as a function of reaction duration. Note: An oxide shell was formed on these nanoparticles due to their exposure to air. The amount of nickel acetate used was 4 9 10-3 mol with 7 mL of oleylamine. After 30-min reaction, 3.8 9 10-3 mol of TOPO was injected with 1 mL of TOP and the reaction was carried out for different durations

2889

After 6 h reaction time

At 0.5 h reaction

At 1 h reaction time Stabilizer molecule

At 2 h reaction time

After 6 h reaction time

After 6 h reaction time

shorter reaction times. This will further prevent grain growth and aggregation/sintering of the nanoparticles. In the air oxidation process, as the oxide barrier lm grows, it becomes too thick for ionic diffusion to occur and thus limits the oxide formation after some time (Karmhag et al. 2001). Still, all these methods do not facilitate the formation of oxide shells with well-controlled thickness and

morphology. In this regard, microwave-based processing is of great importance because of its high penetration depth and energy efciency. This process reduces the nanoparticle diffusion and surface migration rates, minimizes the thermal gradients, shortens the processing times, and is a less energyintensive and an environmentally friendly method (Whiteeld and Davidson 2000; Shi and Hwang

123

2890 Fig. 5 Ni/NiO core/shell nanoparticles produced by ice quenching of the nanoparticle solution immediately after the reaction completion. A High resolution TEM image of a core/shell nanoparticle indicating the emergence of various defects (T twin boundaries, SF stacking faults, D dislocations, and M missing atoms). B Formation of oval-shaped nanoparticles on ice quenching

J Nanopart Res (2010) 12:28832893

2003). Because of these advantages, we performed microwave processing of NiNPs using a domestic microwave as a cost-efcient NiO shell formation process. This process resulted in uniform nanoparticle sizes, with well-dened morphology, and with negligible sintering. This led to an accelerated and controlled growth of the NiO shell, contrary to the several days required to oxidize the Ni core in air (Lee et al. 2006). Varying the microwave duration (60180 min) changes the NiO shell thickness from 2 to 8 nm. Nanoparticles were characterized by XRD before and after microwave processing (Fig. 6). Reections of cubic NiO (JCPDS card no. 47-1049) were conrmed by the presence of peaks corresponding to (111) and (220) planes. Similarly, face-centered cubic-Ni peaks (JCPDS card no. 4-0850) were assigned to (111), (200), and (220) planes. Peaks for NiO were observed after 1 h of microwave duration. While the peak intensity was saturated after 3 h, a large population of hollow NiO nanostructures was observed with thicker shells (Fig. 6C). An attempt was made to characterize these nanoparticles using elemental analysis methods. However, a small fraction of NiNPs was not completely oxidized in the sample set, which resulted in a mixed XRD signal of Ni and NiO. We also believe that, for some hollow NiO nanoparticles, there is a very thin layer of Ni that is not detected by high resolution TEM but still present in the inner surface of the hollow NiO core.

The production of such hollow nanostructures using a simple and economic microwave irradiation process is an important extension of core/shell nanoparticle technology, where the core can be utilized as a sacricial template. The uniqueness of this process is that it overcomes difculties encountered in previous hollow nanoparticle syntheses (Caruso et al. 1998; Meneses et al. 2007; Jung et al. 2007), namely use of wet chemistry routes, rupturing of hollow nanoparticles, formation of a polycrystalline shell, and tedious template removal methodologies. In addition, the microwave process allows for a nely controlled oxidation of the Ni core that results in control over the shell thickness of the formed oxide. The mechanism of formation of NiO shell is still unknown, but could be attributed to two simultaneous processes: (1) sparking at the surface of NiNPs and (2) enhanced rate of diffusion of oxygen and nickel across the growing NiO barrier lm resulting in the formation of the oxide shell. It is proposed here that these processes are accompanied by the formation of highly reactive oxygen species that can easily penetrate the initially formed oxide shell (Palchik et al. 1999b). Subsequently, the generated oxygen species react with Ni resulting in a NiO shell. It is important to note that the core/shell nanoparticles have the same diameter as the starting NiNPs. This also indicates that the oxide shell is formed simultaneously as the core nanoparticle is consumed.

123

J Nanopart Res (2010) 12:28832893

2891

Ni (111)

Intensity (Arbitrary Units)

Ni (200) Ni 220

32

40

50

60

70

80

90

2 (degrees)

Ni (111)

Intensity (Arbitrary units)

Ni 200 NiO (111) NiO (220)

Ni 220

31

40

50

60

70

80

2 (degrees)
Fig. 6 XRD peaks indicating the presence of Ni in the nanoparticles A before microwave processing as well as NiO B after 3 h of microwave processing. C A large population of differently shaped hollow NiO nanoparticles is produced beyond 3 h of microwave processing

Conclusion In conclusion, Ni/NiO core/shell nanoparticles were successfully prepared using a chemical synthesis method coupled with microwave processing that is both cost-effective and scalable. We demonstrated that an increase in the nickel salt/TOPO ratio or a decrease in TOPO alone led to an increase in the nanoparticle size. More importantly, varying the reaction time assisted in tailoring the shape of the

core/shell nanoparticles. The shape evolution is attributed to the rupturing of the stabilizer micelles surrounding the NiNPs due to prolonged annealing of the nanoparticle solution. Further, microwave processing was used to enhance the formation of an oxide shell with controlled thickness. Prolonged microwave oxidation ultimately led to hollow NiO nanostructures. It is important to note that the microwave process did not cause any sintering or changes in the morphology and monodispersity of the

123

2892

J Nanopart Res (2010) 12:28832893 He Y, Li X, Swihart MT (2005) Laser-driven aerosol synthesis of nickel nanoparticles. Chem Mater 17:10171026. doi: 10.1021/cm048128t Hou Y, Kondoh H, Ohta T, Gao S (2005) Size-controlled synthesis of nickel nanoparticles. Appl Surf Sci 241:218 222. doi:10.1016/j.apsusc.2004.09.045 Jana NR, Chen Y, Peng X (2004) Size- and shape-controlled magnetic (Cr, Mn, Fe, Co, Ni) oxide nanocrystals via a simple and general approach. Chem Mater 16:39313935. doi:10.1021/cm049221k Jung KY, Lee JH, Koo HY, Kang YC, Park SB (2007) Preparation of solid nickel nanoparticles by large-scale spray pyrolysis of Ni(NO3)2 6H2O precursor: effect of temperature and nickel acetate on the particle morphology. Mater Sci Eng B 137:1019. doi:10.1016/j.mseb.2006.09.025 Kang DH, Park SS, Oh YS, Kim NJ (2007) Effect of nanoparticles on the creep resistance of MgSn based alloys. Mater Sci Eng A 449451:318321. doi:10.1016/j.msea. 2006.02.332 Karmhag R, Niklasson GA, Nygren M (2001) Oxidation kinetics of nickel nanoparticles. J Appl Phys 89:3012 3017. doi:10.1063/1.1325002 Khadar MA, Biju V, Inoue A (2003) Effect of nite size on the magnetization behavior of nanostructured nickel oxide. Mater Res Bull 38:13411349. doi:10.1016/S00255408(03)00139-9 Kim IG, Kwon JW (2003) Reduction of grown-in defects by vacancy-assisted oxygen precipitation in high density dynamic random access memory. Appl Phys Lett 83: 48634865. doi:10.1063/1.1632536 Lee IS, Lee N, Park J, Kim BH, Yi YW, Kim T, Kim TK, Lee IH, Paik SR, Hyeon T (2006) Ni/NiO core/shell nanoparticles for selective binding and magnetic separation of histidine-tagged proteins. J Am Chem Soc 128:10658 10659. doi:10.1021/ja063177n Li J, Qin Y, Kou X, Huang J (2004) The microstructure and magnetic properties of Ni nanoplatelets. Nanotechnology 15:982986. doi:10.1088/0957-4484/15/8/020 Li Y, Cai M, Rogers J, Xu Y, Shen W (2006a) Glycerolmediated synthesis of Ni and Ni/NiO core-shell nanoparticles. Mater Lett 60:750753. doi:10.1016/j.matlet. 2005.10.005 Li X, Zhang X, Li Z, Qian Y (2006b) Synthesis and characteristics of NiO nanoparticles by thermal decomposition of nickel dimethylglyoximate rods. Solid State Commun 137:581584. doi:10.1016/j.ssc.2006.01.031 Lu AH, Salabas EL, Schuth F (2007) Magnetic nanoparticles: synthesis, protection, functionalization, and application. Angew Chem Int Ed 46:12221244. doi:10.1002/anie. 200602866 Makhlouf SA, Parker FT, Spada FE, Berkowitz AE (1997) Magnetic anomalies in NiO nanoparticles. J Appl Phys 81:55615563. doi:10.1063/1.364661 Meneses CT, Flores WH, Garcia F, Sasaki JMA (2007) A simple route to the synthesis of high-quality NiO nanoparticles. J Nanopart Res 9:501505. doi:10.1007/s11051006-9109-2 Nayak BB, Vitta S, Nigam AK, Bahadur D (2006) Ni and Ni nickel oxide nanoparticles with different shapes and a coreshell structure. Thin Solid Films 505:109112. doi:10.1016/j.tsf.2005.10.018

Ni/NiO core/shell nanoparticles. The ability to manipulate the morphology of Ni/NiO core/shell nanoparticles is critical to the development of future devices based on these nanoparticles.
Acknowledgments This work was funded by NSF (NSFREU program) and the Army Research Laboratory under Cooperative Agreement No. W911NF-04-2-0023. The views and conclusions contained in this document are those of the authors and should not be interpreted as representing the ofcial policies, either expressed or implied, of the Army Research Laboratory or the U.S. Government. The U.S. Government is authorized to reproduce and distribute reprints for Government purposes notwithstanding any copyright notation hereon.

References
Ahmadi TS, Wang ZL, Green TC, Henglein A, El-Sayed MA (1996) Shape-controlled synthesis of colloidal platinum nanoparticles. Science 272:19241925. doi:10.1126/ science.272.5270.1924 Bratlie KM, Lee H, Komvopoulos K, Yang P, Somorjai GA (2007) Platinum nanoparticle shape effects on benzene hydrogenation selectivity. Nano Lett 7:30973101. doi: 10.1021/nl0716000 Caruso F, Caruso RA, Mohwald H (1998) Nanoengineering of inorganic and hybrid hollow spheres by colloidal templating. Science 282:11111114. doi:10.1126/science. 282.5391.1111 Chen S, Caroll DL (2002) Synthesis and characterization of truncated triangular silver nanoplates. Nano Lett 2:1003 1007. doi:10.1021/nl025674h Cordente N, Respaud M, Senocq F, Casanove M-J, Amiens C, Chaudret B (2001) Synthesis and magnetic properties of nickel nanorods. Nano Lett 1:565568. doi:10.1021/ nl0100522 Davar F, Fereshteh Z, Salavati-Niasari M (2009) Nanoparticles Ni and NiO: Synthesis, characterization and magnetic properties. J Alloys Compd 476:797801. doi:10.1016/ j.jallcom.2008.09.121 Dumestre F, Chaudret B, Amiens C, Renaud P, Fejes P (2004) Superlattices of iron nanocubes synthesized from Fe [N(SiMe3)2]2. Science 303:821823. doi:10.1126/science. 1092641 Fendler JH (1998) Nanoparticles and nanostructured lms. Wiley-VCH, Weinheim Gafner YY, Gafner SL, Entel P (2004) Formation of an icosahedral structure during crystallization of nickel nanoclusters. Phys Solid State 46:13271330. doi:10.1134/ 1.1778460 Ghosh M, Biswas K, Sundaresan A, Rao CNR (2006) MnO and NiO nanoparticles: synthesis and magnetic properties. J Mater Chem 16:106111. doi:10.1039/b511920k Hall SR, Davis SA, Mann S (2000) Cocondensation of organosilica hybrid shells on nanoparticle templates: a direct synthetic route to functionalized core-shell colloids. Langmuir 16:14541456. doi:10.1021/la9909143

123

J Nanopart Res (2010) 12:28832893 Palchik O, Kataby G, Mastai Y, Gedanken A (1999a) New method for nanofabrication of structures analogous to core-shell vesicles. Adv Mater 11:12891292 Palchik O, Avivi S, Pinkert D, Gedanken A (1999b) Preparation and characterization of Ni/NiO composite using microwave irradiation and sonication. Nanostruct Mater 11:415420. doi:10.1016/S0965-9773(99)00321-9 Parada C, Moran E (2006) Microwave-assisted synthesis and magnetic study of nanosized Ni/NiO materials. Chem Mater 18:27192725. doi:10.1021/cm0511365 Park J, Kang E, Son SU, Park HM, Lee MK, Kim J, Kim KW, Noh HJ, Park JH, Bae CJ, Park JG, Hyeon T (2005a) Monodisperse nanoparticles of Ni and NiO: synthesis, characterization, self-assembled superlattices, and catalytic applications in the suzuki coupling reaction. Adv Mater 17:429434. doi:10.1002/adma.200400611 Park J, Koo B, Yoon KY, Hwang Y, Kang M, Park JG, Hyeon T (2005b) Generalized synthesis of metal phosphide nanorods via thermal decomposition of continuously delivered metal-phosphine complexes using a syringe pump. J Am Chem Soc 127:84338440. doi:10.1021/ ja0427496 Pinarbasi M, Metin S, Gill H, Parker M, Gurney B, Carey M, Tsang C (2000) Antiparallel pinned NiO spin valve sensor for GMR head application. J. Appl. Phys. 87:57145719. doi:10.1063/1.372499 Poizot P, Laruelle S, Grugeon S, Dupont L, Tarascon JM (2000) Nano-sized transition-metal oxides as negativeelectrode materials for lithium-ion batteries. Nature 407:496499. doi:10.1038/35035045 Puntes VF, Krishnan KM, Alivisatos AP (2001) Colloidal nanocrystal shape and size control: the case of cobalt. Science 291:21152117. doi:10.1126/science.1057553 Rao CNR, Muller A, Cheetham AK (2004) The chemistry of nanomaterials: synthesis, properties, and applications, vols 1, 3. Wiley-VCH, Weinheim Sako S, Ohshima K, Sakai M, Bandow S (1997) Magnetic property of NiO ultrane particles with a small Ni core. J Vac Sci Technol B 15:13381342

2893 Sato H, Minami T, Takata S, Yamada T (1993) Transparent conducting p-type NiO thin lms prepared by magnetron sputtering. Thin Solid Films 236:2731. doi:10.1016/ 0040-6090(93)90636-4 Service RF (2004) Nanotechnology grows up. Science 304:17321734. doi:10.1126/science.304.5678.1732 Sharma SN, Sharma H, Singh G, Shivaprasad SM (2008) Studies of interaction of amines with TOPO/TOP capped CdSe quantum dots: role of crystallite size and oxidation potential. Mater Chem Phys 110:471480. doi:10.1016/ j.matchemphys.2008.02.038 Shen B, Zhang XY, Yang K, Chen P, Zhang R, Shi Y, Zheng YD, Sekiguchi T, Sumino K (1997) Gettering of Fe impurities by bulk stacking faults in Czochralski-grown silicon. Appl Phys Lett 70:18761878. doi:10.1063/ 1.118718 Shenhar R, Norsten TB, Rotello VM (2005) Polymer-mediated nanoparticle assembly: structural control and applications. Adv Mater 17:657669. doi:10.1002/adma.200401291 Shi S, Hwang JY (2003) Microwave-assisted wet chemical synthesis: advantages, signicance, and steps to industrialization. J. Miner Mater Character Eng 2:101110 Song PX, Wen DS, Guo ZX, Korakianitis T (2008) Oxidation investigation of nickel nanoparticles. Phys Chem Chem Phys 10:50575065. doi:10.1039/b800672e Whiteeld PS, Davidson JJ (2000) Microwave synthesis of Li1.025Mn1.975O4 and Li1?xMn2-xO4-yFy (x = 0.05, 0.15; y = 0.05, 0.1). J Electrochem Soc 147:44764484 Xiang L, Deng XY, Jin Y (2002) Experimental study on synthesis of NiO nano-particles. Scripta Mater 47:219224. doi:10.1016/S1359-6462(02)00108-2 Yin Y, Alivisatos AP (2005) Colloidal nanocrystal synthesis and the organicinorganic interface. Nature 437:664670. doi:10.1038/nature04165

123

Вам также может понравиться