Synthesis of Co-Doped CdS Nanocrystals by Direct Thermolysis of ...

3 downloads 0 Views 8MB Size Report
Jun 7, 2015 - longjiang Postdoctoral Fund (no. LBH-Z14012), Youths. Science Foundation of Heilongjiang Province of China. (QC2011C010), and the ...
Hindawi Publishing Corporation Journal of Nanomaterials Volume 2015, Article ID 109734, 11 pages http://dx.doi.org/10.1155/2015/109734

Research Article Synthesis of Co-Doped CdS Nanocrystals by Direct Thermolysis of Cadmium and Cobalt Thiolate Clusters Jianing Zhao, Xiaoli Li, and Zhiguo Li School of Material Science and Engineering, Northeast Forestry University, No. 26 Hexing Road, Harbin 150040, China Correspondence should be addressed to Zhiguo Li; [email protected] Received 29 March 2015; Revised 5 June 2015; Accepted 7 June 2015 Academic Editor: Ruibing Wang Copyright © 2015 Jianing Zhao et al. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. Co-doped CdS (Co:CdS) nanocrystals with controllable morphology (quantum dots and nanorods) were easily synthesized by direct thermolysis of (Me4 N)2 [Co4 (SC6 H5 )10 ] and (Me4 N)4 [S4 Cd10 (SPh)16 ] under different precursor concentration, in virtue of the ions exchange of molecular clusters. The Co:CdS quantum dots were produced under low precursor concentration, and the Co:CdS nanorods could be obtained under higher precursor concentration. The Co-doping effect on the structure, growth process, and property of CdS nanocrystals was also investigated. The results indicated that the Co-doping was favorable for the formation of the nanorod structures for a short reaction time. In addition, the Co-doping in the CdS lattice resulted in the ferromagnetic property of the Co:CdS quantum dots at room temperature. Moreover, compared with the CdS quantum dots, the Co:CdS quantum dots exhibited obvious quantum confinement effect and photoluminescence emission with slightly red-shift.

1. Introduction The doping of magnetic ions into semiconductor nanoparticles potentially gives a new class of materials, diluted magnetic semiconductors (DMSs) [1]. In II–VI materials, the magnetic coupling for Mn and Co ions in bulk semiconductors has been observed [2, 3]. DMSs are currently receiving much attention due to their potential applications in magnetic, electronic [4], and biological labeling fields [5], especially in the emerging field of spin-based electronics, or “spintronics.” The objective of spintronics is to control electron spins in order to transmit information and provide new functionality to semiconductor devices, for example, semiconducting alloys [6], semiconducting film [7], and nonmagnetic semiconductor [8, 9]. The potential applications and the possibility of novel magneto-optical phenomena in these materials have motivated the exploration about the synthesis techniques of magnetic ions doped nanocrystals (NCs). CdS is a popular II–VI semiconductor material and can be easily synthesized, which makes it an effective test vehicle in training as well as in the demonstration of innovative technologies [10]. Furthermore, CdS has a large band gap

(2.42 eV) and can form a doped semiconductor by adding a third element. Much effort of magnetic ions doping has focused on CdS NCs such as Mn [11, 12]. Mn can easily replace Cd because the radius of Mn ion is close to that of Cd ion. In addition, the incorporation of Mn in CdS crystal lattice slightly disturbs CdS crystal structure. In previous reports [11, 13], Mn2+ -doped CdS NCs have been synthesized by colloidal route based on the simultaneous precipitation of CdS and MnS or by a coprecipitation reaction in reverse micelles using Na2 S as the sulfur source. However, the results show that only a small fraction of the initial Mn2+ added can be incorporated into the crystal lattice, while a large proportion of the Mn2+ remain on the surface or form MnS precipitates. Consequently, in the case of nanostructured materials, the small size and large surface area-to-volume ratio present a significant challenge to impurity doping. Co2+ -doped II–VI systems could exhibit stronger coupling than that of Mn-doped NCs, due to the increased Co d-orbital mixing with the valence band and conduction band of the II–VI host. It is possible for the Co-doped CdS NCs to induce novel magneto-optical phenomena which could open new possibilities for spintronics studies and applications. Recently, room-temperature ferromagnetism

2 has been reported for some Co2+ -doped NCs, such as TiO2 [14, 15] and ZnO [16]. For Co2+ -doped NCs, doping occurs primarily at the surface or near-surface sites in these materials by simple coprecipitation [17] and standard inverted micelle coprecipitation methods [18]. Radovanovic and Gamelin [18] have shown that normal soft chemical processing does not render cobalt substitution in a CdS matrix and that the dopants are diverted to the surface. It is attributed to the large mismatch between the radius of Co and Cd ions, which is responsible for exclusion of Co ions during CdS NCs growth. They have proposed an interesting core/shell strategy to solve this problem. Unfortunately, no magnetic properties are examined. Bogle and coworkers [19] synthesized CdS:Co NCs via a route of pulsed high energy electron induced synthesis, and the incorporation of Co is realized in the CdS NCs under stoichiometric proportion in liquid medium condition. Nevertheless, the method needs special instrument and is less controllable especially for the NCs morphology and optical properties. The difficulty of controlling doping is to develop convenient methods for the preparation of highquality Co-doped CdS NCs. To obtain magnetic ions internal doped NCs, the chemical synthesis of molecular clusters is another possible strategy, and this synthesis method has ability to exchange metal and chalcogenide atoms among each other [20]. The character is explored during the synthesis of Co ions doped CdS [21] and Mn ions doped ZnSe NCs [22]. Inorganic clusters such as (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] are a class of materials that exist as discrete units related to a fragment of a bulk lattice [23, 24]. As the metal-chalcogen bond has already existed in the (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ], thermolysis of the cluster provided a preformed template for the construction of nanocrystals (which has been synthesized in our previous work [25]). Application of molecular-cluster-based precursors for NCs growth offered an alternative synthetic methodology for material preparation with novel properties. In this work, Co-doped CdS NCs with different morphologies (quantum dots and nanorods) were synthesized in virtue of the ions exchange of molecular clusters. By direct thermal decomposition of two clusters, (Me4 N)2 [Co4 (SC6 H5 )10 ] and (Me4 N)4 [S4 Cd10 (SPh)16 ], in hexadecylamine (HDA), uniform CdS:Co quantum dots were obtained under low precursor concentration and CdS:Co nanorods were formed by increasing the cluster concentration. The morphology and structure of the Codoped CdS NCs were thoroughly characterized by TEM, HRTEM, and XRD. The luminescent and ferromagnetic properties of the nanocrystals were further investigated by ultraviolet-visible (UV-vis) absorption spectroscopy, photoluminescence (PL) spectroscopy, and physical properties measurement system (PPMS).

Journal of Nanomaterials research were of analytical grade. (Me4 N)4[S4 Cd10 (SC6 H5 )16 ] and (Me4 N)2 [Co4 (SC6 H5 )10 ] were prepared according to the reported methods [23, 26]. 2.2. Synthesis of Co-Doped CdS Quantum Dots. In the typical synthesis of the CdS:Co quantum dots, the cadmium thiolate cluster (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] (1.2 g) and the cobalt thiolate cluster (Me4 N)2 [Co4 (SC6 H5 )10 ] (0.2 g and 0.5 g) were added in 80 g HDA at 80∘ C under an atmosphere of nitrogen. The temperature of the reaction mixture was raised to 220∘ C and the growth of the CdS quantum dots was monitored by recording the optical absorption spectra. At the desired crystal size, the crystalline growth was subsequently terminated by cooling the reaction mixture to approximately 60∘ C, followed by precipitation from HDA via the addition of anhydrous methanol to produce precipitated samples. The purification of the CdS:Co quantum dots was carried out by suspension in toluene followed by reprecipitation in methanol and isolation by centrifugation. After purification, the dried samples are about 10.8 and 12.4 g/L (with 0.2 g and 0.5 g cobalt thiolate cluster, resp.). To efficiently eliminate those Co ions which may be physically adsorbed on the surface of NCs, the as-prepared samples were stripped with pyridine to perform the ligand exchange [27]. Particle size for the Co:CdS samples was determined by UV-vis absorption and TEM measurements. As a reference, CdS NCs were prepared by the similar method without addition of cobalt thiolate cluster. 2.3. Synthesis of Co-Doped CdS Nanorods. Synthesis of Co-doped CdS nanorods was performed by using the parallel methodology by thermolysis of 1.2 g (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] and of 0.2 g or 0.5 g (Me4 N)2 [Co4 (SC6 H5 )10 ] in different contents of HDA. The following process and purification of CdS:Co nanorods are the same as those of quantum dots. 2.4. Characterization. TEM and HRTEM were performed on a Tecnai S-Twin electron microscope operating at 200 kV. ICP was performed on an Optima 5300 DV (Perkin Elmer Inc.). Samples were dissolved in a 2% HNO3 solution, and the Cd and Co concentrations were measured against known Cd and Co standards (High Purity Standards). XRD patterns of the nanocrystals were recorded with a Rigaku D/max-𝛾B diffractometer equipped with a rotating anode and a Cu K𝛼 source (𝜆 = 0.154056 nm). UV-vis absorption spectra were recorded at room temperature on a UV-2550 spectrophotometer. Photoluminescence (PL) spectra were recorded at room temperature on a FP-6500 spectrophotometer. Magnetic measurements were carried out using the physical properties measurement system (PPMS) of Quantum Design with a magnetic field up to 5 T.

2. Materials and Methods 2.1. Materials. Cd(NO3 )2 ⋅4H2 O, Co(NO3 )2 ⋅6H2 O, triethylamine, acetonitrile, sulfur powder, tetramethylammonium chloride, and hexadecylamine (HDA) were purchased from Aldrich Chemical. Pyridine, methanol, and toluene used in

3. Results and Discussion 3.1. Synthesis of Co-Doped CdS Quantum Dots. The thiophenolate complex of (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] is used as the precursor, and metal ion exchange reactions in the clusters

Journal of Nanomaterials

3

Average diameter = 3.8 ± 0.4 nm

40

Frequency (%)

30

20

10

0 2 (a)

3

4 Particle size (nm)

5

6

(b)

(c)

Figure 1: Cd0.924 Co0.076 S quantum dots synthesized by thermolysis of 1.2 g (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] and 0.2 g (Me4 N)2 [Co4 (SC6 H5 )10 ] in HDA: (a) TEM image, (b) corresponding size distribution, (c) HRTEM image.

have been revealed [20, 28, 29]. For example, metal exchange has been shown to occur in [S4 M10 (SPh)16 ]4− (M=Cd, Zn) for Zn2+ /Cd2+ by NMR analysis [28]. Løver et al. [29] have also investigated the metal ion exchange process using electrospray mass spectrometry in the thiophenolate-capped clusters, such as (Me4 N)4 [E4 Cd10 (SC6 H5 )16 ] (E=S, Se), and generated mixed-metal clusters of cadmium and zinc. Even at room temperature, such metal ion substitution process could be easy to achieve because the ion radius of Zn2+ and Cd2+ is close. However, between the radii of Co and Cd ions large mismatch existed, resulting in the Co ions doping occurring primarily at the surface or near-surface sites in NCs [17, 18]. The cobalt thiophenolate cluster (Me4 N)2 [Co4 (SC6 H5 )10 ] is used as the Co precursor. The complexation of Co with thiophenolate could improve the reactivity of Co2+ , which has been shown to undergo rapid metal ion exchange within similar clusters such as Fe2+ /Co2+ , Co2+ /Zn2+ , and Co2+ /Cd2+ exchange that occurred in [M4 (SPh)10 ]2− (M=Fe,

Co, Zn, Cd) complex [20, 29]. For the formation of Codoped CdS nanocrystals, the supposition was performed by direct thermolysis of (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] and (Me4 N)2 [Co4 (SC6 H5 )10 ] in HDA. Figures 1 and 2 show the TEM image, histogram of size distribution, and HRTEM image of the prepared CdS NCs doped with different Co concentration (the Co concentration was determined by ICP in the next section). The Co-doped CdS (CdS:Co) quantum dots are well-defined and nearly spherical nanoparticles, and the quantum dots are prepared by thermolysis of 1.2 g (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] and 0.2 or 0.5 g (Me4 N)2 [Co4 (SC6 H5 )10] at 80 g HAD. The size distribution of the samples is demonstrated in Figures 1(b) and 2(b). The average particle size is 3.8 nm ± 10% for Cd0.924 Co0.076 S NCs and 4.1 nm ± 12% for Cd0.815 Co0.185 S NCs, respectively. Figure 1(c) shows a typical HRTEM image of Cd0.924 Co0.076 S NCs, and the products

4

Journal of Nanomaterials

Average diameter = 4.1 ± 0.5 nm

Frequency (%)

30

20

10

0 2 (a)

3

4 5 Particle size (nm)

6

7

(b)

(c)

Figure 2: Cd0.815 Co0.185 S quantum dots synthesized by thermolysis of 1.2 g (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] and 0.5 g (Me4 N)2 [Co4 (SC6 H5 )10 ] in HDA: (a) TEM image, (b) corresponding size distribution, (c) HRTEM image.

are nearly spherical particles with slight prolate deviations. The space between adjacent planes is 0.336 nm, which is consistent with the interplanar distance of the (002) planes of hexagonal CdS. Figure 2(c) demonstrates the HRTEM image of Cd0.815 Co0.185 S NCs along with (002) plane. The particles are also nearly spherical with slight prolate deviations. Figure 3 shows the room-temperature UV-vis absorption and photoluminescence (inset) spectra of the CdS:Co quantum dots. The excitonic absorption peaks are well-defined and rather sharp indicative of narrow size distribution of the NCs. The excitonic absorption peak is associated with the lowest optical transition and provides a simple way to determine the band gap of NCs. The maximum absorption peaks as shown in Figure 3 are 418 nm (for CdS), 415 nm (for Cd0.924 Co0.076 S), and 413 nm (for Cd0.815 Co0.185 S), corresponding to the band gap of 2.97, 2.99, and 3.00 eV, respectively. The band-gap energy of the synthesized NCs is higher than that of CdS bulk

crystal (2.42 eV) [30], indicating quantum confinement of the CdS:Co quantum dots. The similar absorption spectra of the NCs demonstrate that the incorporation of Co ions in the CdS NCs lattice does not significantly change the absorption edge of the NCs. The average particle size could be estimated from the first excitonic absorption peak [31], which is all about 4.0 nm for the Co ions doped samples. This is in agreement with the value of TEM results. In the PL spectra as shown in Figure 3, the emission band is red-shifted compared with the absorption band, which is frequently observed for CdS NCs [32–34]. Bulk CdS is reported to have a broad emission with the emission maximum in the 500∼700 nm region of the luminescence spectrum, which is attributed to the recombination from surface defects (predominantly sulfur vacancies) [35]. For the NCs, the photoluminescence arises from the radiative recombination of electrons and holes at defect levels brought about by surface trap, and red-shifted

(c) (b) (a)

(b)

300

400 500 Wavelength (nm)

600

Figure 3: UV-vis absorption and PL spectra (inset) of the (a) CdS, (b) Cd0.924 Co0.076 S, and (c) Cd0.815 Co0.185 S quantum dots.

20

30

(c)

40

112

103

650

110

500 550 600 Wavelength (nm)

100 002 101

450

(a)

102

(a)

Intensity (a.u.)

(b) (c)

5

PL intensity

Absorbance (a.u.)

Journal of Nanomaterials

50

60

70

2𝜃 (∘ )

Figure 5: X-ray diffraction (XRD) pattern of (a) CdS, (b) Cd0.924 Co0.076 S, and (c) Cd0.815 Co0.185 S quantum dots. Table 1: Comparison of 𝑑-spacing values of the as-synthesized samples with standard JCPDS CdS: (a) CdS, (b) Cd0.924 Co0.076 S, and (c) Cd0.815 Co0.185 S quantum dots. 𝑑/nm (a) 0.356 0.334 — 0.247 0.206 0.189 0.176

𝑑/nm (b) 0.356 0.332 0.316 0.246 0.205 0.189 0.176

𝑑/nm (c) 0.352 0.330 — 0.245 0.204 0.188 0.175

𝑑/nm (lit) 0.359 0.336 0.316 0.245 0.207 0.190 0.176

hkl (lit) 100 002 101 102 110 103 112

Note: (a) CdS, (b) Cd0.924 Co0.076 S, (c) Cd0.815 Co0.185 S, and (lit) standard JCPDS value of CdS.

Figure 4: TEM image of the as-synthesized CdS nanocrystals.

photoluminescence is observed. No deep trap emission is observed in the long wave region, indicating that HDA is an effective passivating agent. The particles of CdS and CdS:Co NCs were prepared using the same process, and the size of the CdS:Co NCs was consistent with that of the CdS NCs (as shown in Figure 4, 4.2 nm for CdS quantum dots). Consequently, the absorption peaks of CdS:Co NCs are consistent with that of CdS NCs. Compared with the PL spectrum of CdS NCs, the spectra of Co ions doped NCs are slightly red-shifted in the similar size regime. The observed shift in PL spectra implies that perhaps the initial Co-doping induces defects due to self-contraction. 3.2. Proof of Co Internal Doping in CdS Nanocrystals. Inductively coupled plasma (ICP) is used to measure the Co ions concentration of the prepared NCs. To further confirm that the Co ions concentration measured reflects substitution of internal Cd sites, the method of pyridine stripping is adopted. This method has previously been used to remove

surface-bound contamination in doped NCs [36]. The Co ions concentration after pyridine stripping yields doping concentrations of 𝑥 = 0.076, 0.185 moles of cobalt in Cd1−𝑥 Co𝑥 S and 0.082, 0.196 moles of cobalt prior to ligand exchange, respectively. The differences between Co ions concentrations in as-grown and pyridine-capped NCs are about 7% and 6% for the two samples, indicating that pyridine stripping can remove surface-adsorbed Co ions. The Co ions were internally doped in the NCs based on the ICP measured results. Figure 5 shows the XRD patterns of the synthesized CdS and Co-doped CdS NCs with similar method. The XRD data are given in Table 1. The diffraction patterns show broad peaks typical of particles in the nanosize. The XRD patterns are consistent with the hexagonal phase of CdS. The (110), (103), and (112) planes of wurtzite NCs are clearly distinguishable in the patterns (JCPDS file number 41-1049). From the CdS XRD pattern (Figure 5(a)), a strong (002) peak is observed in contrast to other peaks, while the (102) peak is low and broad. For the Co-doped CdS NCs, the (002) peaks are much stronger than that of CdS NCs as shown in Figures 5(b) and 5(c).

6

Journal of Nanomaterials characterized in the previous sections. Further investigation is in progress for the magnetism of CdS:Co NCs.

0.10 300 K

0.00

0.01

M (emu/g)

M (emu/g)

0.05

−0.05

−0.10 −50000

0.00

−0.01 −400 −300 −200 −100

−25000

0

0 100 H (Oe)

25000

200

300

400

50000

H (Oe)

Figure 6: Hysteresis loop of the Cd0.815 Co0.185 S quantum dots; the inset figure shows the enlarged hysteresis loop.

The stacking faults along the (002) plane are attributed to low synthesis temperature, and the result is similar to the CdS NCs synthesized by Murray et al. [30]. Meantime, this defect feature is very prominent in bulk II–VI materials [37]. All the detectable peaks of the three samples are almost identical to the positions, but the measured lattice constant is different. The average lattice constant (𝑎) of the prepared NCs was estimated by Cohen’s method [38] and a lattice contraction was found after introduction of Co ions. For ˚ for Cd0.815 Co0.185 S NCs and 4.096 for example, 𝑎 is 4.085 A ˚ for CdS NCs. As the Co Cd0.924 Co0.076 S compared to 4.115 A ˚ is smaller than Cd ion radius (0.97 A), ˚ ion radius (0.74 A) the substitution of Cd by Co ions may cause some stress and strain effects to shrink the lattice structure. From the variation of the NCs lattice constant, the Co ions could be distributed inside the NCs rather than on the surface. The magnetic properties of the prepared Cd0.815 Co0.185 S NCs were measured on physical properties measurement system (PPMS) of Quantum Design. Figure 6 shows the field dependence of magnetic moment (M-H) curve of the sample measured at 300 K. There is a hysteresis loop because of ferromagnetism, and the coercive field and remnant magnetization are 84 Oe and 3.681 × 10−3 emu/g, respectively. The similar phenomena have been reported for Co ions doped ZnO nanocrystals [39, 40], but the origin of roomtemperature ferromagnetism in the transition metal doped NCs is still controversial due to the possibility of phase separation and uncertainty of magnetic interactions. In this work, any detectable traces of secondary phases or clusters cannot be observed from the HRTEM and XRD. The exhibition of room-temperature ferromagnetism perhaps relates to the NCs defects. According to the bound magnetic polaron theory [41], the spins of the magnetic transition metal cations are aligned by the spins of carriers localized on lattice defects. Such defects, as demonstrated by XRD, could provide the necessary source of carriers required to mediate ferromagnetic ordering. Therefore, the ferromagnetism observed in CdS:Co NCs is associated with the substitutional Co-dopants

3.3. Influence of Co-Doping on Growth Process of Quantum Dots. In order to investigate the influence of Co-doping on growth process of quantum dots, the growth process of Cd0.924 Co0.076 S quantum dots was monitored by UVvis spectrum. All the samples for absorption experiments were used as prepared, without any further purification. Figure 7(a) shows the absorption spectra of Cd0.924 Co0.076 S quantum dots obtained at different reaction temperatures. No obvious absorption peak is observed at temperature below 120∘ C. When the reaction temperature is raised to 120∘ C, the absorption peak appears, implying that nuclei are formed massively. With the temperature rising from 120∘ C to 190∘ C, the absorption peaks are slowly red-shifted from 370 nm to 384 nm. When the reaction temperature is increased above 190∘ C, the absorption peaks are redshifted largely from 384 nm to 424 nm. The large shift of the absorption peaks indicates that the quantum dots grow rapidly as the temperature exceeds 190∘ C [25]. Figure 7(b) demonstrates the change of absorption peaks with reaction temperature. The absorption peak increases slightly from 100∘ C to 150∘ C, but the absorption peak increases rapidly when the temperature is raised above 190∘ C. Figure 8 shows the absorption spectra of Cd0.924 Co0.076 S quantum dots at 220∘ C with different reaction times. The absorption peak increases with prolonging time at 220∘ C. Based on the above results, the growth process of quantum dots could be divided into two stages, including nucleation and rapid growth. During the nucleation stage (below 190∘ C), the quantum dots grow slowly and the nucleation is dominant. As the temperature is increased above 190∘ C, a rapid growth stage of quantum dots occurred. The growth process of Co-doped quantum dots is similar to that of CdS quantum dots [25]. This similarity reveals that the Codoping in the quantum dots does not change the growth process of NCs. It can be reasonably concluded that the Cd/Co ion exchange between clusters is rapid under lower temperature and is accomplished before the formation of quantum dots. The rapid exchange ability of Cd/Co ion leads to the formation and growth of Co-doping NCs consistent with that of CdS NCs. Thus, the way of preparing Co-doping NCs by inorganic clusters does not disturb the growth process of NCs and can be easy to achieve. 3.4. Synthesis of Co-Doped CdS Nanorods. Based on our previous works [25, 42] and other research results [43], the reaction conditions especially precursor concentration significantly affect the morphology of NCs. The effect of precursor concentration on NCs morphology with Co-dopant is explored by reducing HDA dosage. Figures 9 and 10 show the sample TEM images prepared by thermal decomposition of 1.2 g (Me4 N)4[S4 Cd10 (SC6 H5)16 ] and 0.2 g or 0.5 g (Me4 N)2 [Co4 (SC6 H5 )10 ] in 25 g HDA at 220∘ C for 1 h. After pyridine stripping, the Co ion concentrations measured by ICP are 0.079 and 0.180, respectively. Compared

Journal of Nanomaterials

7

450 Wavelength (nm)

Absorbance (a.u.)

220∘ C

190∘ C 150∘ C 140∘ C 120∘ C

400

110∘ C 100∘ C 300

400 500 Wavelength (nm)

350

600

120 160 Reaction temperature (∘ C)

(a)

200

(b)

Figure 7: UV-vis absorption spectra of Cd0.924 Co0.076 S quantum dots at different reaction temperatures (a) and the evolution of the first absorption peak of Cd0.924 Co0.076 S quantum dots with reaction temperature (b).

Wavelength (nm)

Absorbance (a.u.)

440

7h 5h

420

400

3h 1h 0.5 h 0h 300

400 500 Wavelength (nm)

600

(a)

380

0

2

4 Reaction time (h)

6

8

(b)

Figure 8: UV-vis absorption spectra of Cd0.924 Co0.076 S quantum dots taken at 220∘ C for reacting different time (a) and the evolution of the first absorption peak of Cd0.924 Co0.076 S quantum dots with reaction time (b).

with the results of CdS:Co quantum dots, the Co-doping concentration in CdS NCs is identical under the same content of (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] and (Me4 N)2 [Co4 (SC6 H5 )10 ]. The morphology of the NCs changes obviously with HDA dosage reducing. The products are straight nanorods by thermolysis of 1.2 g (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] and 0.2 g (Me4 N)2 [Co4 (SC6 H5 )10 ] in 25 g HDA (as shown in Figure 9). The nanorods sizes are an average diameter of 3.9 nm and length of 42 nm, and no aggregation can be observed. Furthermore, the interplanar distance along the growth axis of the nanorods is about 0.336 nm (the HRTEM image), which is consistent with the interplanar distance of the (002) plane in the wurtzite structure of CdS [44]. Thus, the results confirm that the nanorods are lengthened along the axis. For the NCs obtained by thermolysis of 1.2 g (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ]

and 0.5 g (Me4 N)2 [Co4 (SC6 H5 )10 ] in 25 g HDA, the products are the mixture of straight and branched nanorods (small quantity) as shown in Figure 10. The branched nanorods with small quantity are formed by increasing (Me4 N)2 [Co4 (SC6 H5 )10 ] usage to 0.5 g under other same conditions. The precursor concentration significantly influences the morphology of the CdS:Co NCs, and the morphology of NCs changes from quantum dots to nanorods with precursor concentration increase. In order to further prove such results, cluster of 1.2 g (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] and 0.5 g (Me4 N)2 [Co4 (SC6 H5 )10 ] is thermolysized in different contents of HDA. It has been shown in Figure 1 that the products obtained are quantum dots by thermal decomposition of the same clusters in 80 g HDA. Decreasing HDA mass to 40 g,

8

Journal of Nanomaterials

(a)

(b) Average length = 42 nm ± 11%

Average diameter = 3.9 nm ± 9.4%

30

Frequency (%)

Frequency (%)

40

20

20

10

0

0

2

4 Diameter (nm)

6

8

0 30

35

40 Length (nm)

(c)

45

50

(d)

Figure 9: The sample synthesized by thermolysis of 1.2 g (Me4 N)4 [S4 Cd10 (SPh)16 ] and 0.2 g (Me4 N)2 [Co4 (SC6 H5 )10 ] in 25 g HDA: (a) TEM image, (b) HRTEM image, (c) diameter, and (d) length distribution histogram of the nanorods.

II

I

III (a)

(b)

Figure 10: The sample synthesized by thermolysis of 1.2 g (Me4 N)4 [S4 Cd10 (SPh)16 ] and 0.5 g (Me4 N)2 [Co4 (SC6 H5 )10 ] in 25 g HDA: (a) TEM image and (b) enlarged image of (a).

Journal of Nanomaterials

9

(a)

(b)

Figure 11: The sample synthesized by thermolysis of 1.2 g (Me4 N)4 [S4 Cd10 (SPh)16 ] and 0.2 g (Me4 N)2 [Co4 (SC6 H5 )10 ] in (a) 40 g HDA and (b) 35 g HDA. Table 2: Comparison of the synthesis conditions for CdS and CdS:Co nanorods.

Cd cluster (g) 1.0 1.0

(a)

CdS Reaction HDA (g) time (h) 40 3 30 3

CdS:Co Product

Cd cluster (g)

(a)

Co cluster (g)

(b)

HDA (g)

Reaction time (h)

Product

QDs QDs + NRs

1.2

0.5

40

1

NRs

1.2 1.2 1.2

25 25 25

1 3 5

QDs QDs + NRs NRs

1.2

0.2

25

1

NRs

3.0

20

5

Branched NRs

1.2

0.5

25

1

Branched NRs

Note: (a) (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] and (b) (Me4 N)2 [Co4 (SC6 H5 )10 ]. QDs: quantum dots and NRs: nanorods.

straight nanorods are formed as shown in Figure 11(a). By use of 35 g HDA, straight nanorods with a small quantity of branched nanorods are obtained as shown in Figure 11(b). Based on the evolution of NCs morphology with precursor concentration, it can be concluded that the nanorods could be formed by aggregation of nanoparticles. The nanoparticles are formed first in the process and then aggregate together along the (002) plane of CdS wurtzite crystal lattice to form nanorods [44]. Thus, the formation of Co-doping CdS nanorods follows oriented attachment mechanism [45]. The diameters of the nanoparticles and nanorods are almost identical at the same reaction conditions. The appearance of small quantity of nanoparticles in Figure 9(b) also approves the mechanism. The reaction time has to be 5 h or even longer for the formation of CdS nanorods without Co-doping by thermolysis of (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] [42]. Compared with the fabrication of Co-doping and pure CdS nanorods, the addition of (Me4 N)2 [Co4 (SC6 H5 )10 ] cluster means Codoping in CdS NCs and is favorable for the formation of nanorods, even the reaction time is shortened. It has been shown that the nanorods are obtained at 220∘ C for just reacting 1 h by addition of 0.2 g (Me4 N)2 [Co4 (SC6 H5 )10 ]. Table 2 shows the products obtained at different conditions for comparison of doping and undoping nanorods.

The products are quantum dots only by thermal decomposition 1.0 g (Me4 N)4 [S4 Cd10 (SC6 H5 )16 ] in 40 g HDA for 3 h, while the nanorods are formed by addition of 0.5 g (Me4 N)2 [Co4 (SC6 H5 )10 ], and the reaction time is just 1 h. The above results demonstrate that the dopant of Co ions in the CdS is favorable for the formation of nanorods. It has been shown [46] that magnetic particles such as Fe3 O4 are easy to be aggregated as one-dimensional chain structure in organic solvent. For magnetic Fe3 O4 nanoparticles behaving like magnetic dipoles, the effect of interparticle forces plays principle role in ordering of the nanoparticles in onedimensional structures. According to the phenomenon, the Co-doping CdS nanoparticles could be considered as magnetic dipoles in HDA and maybe there are similar magnetic interparticle interactions. The interaction between Co-doped CdS particles is favorable for the oriented aggregation of particles, further promoting the formation of the nanorods.

4. Conclusions In summary, Co-doped CdS NCs were successfully synthesized by direct thermolysis of (Me4 N)2 [Co4 (SC6 H5 )10 ] and (Me4 N)4 [S4 Cd10 (SPh)16 ] molecular clusters in HDA.

10 Uniform Co:CdS quantum dots were obtained under low precursor concentration, and Co:CdS nanorods were formed by increasing the precursor concentration. It was revealed that the formation of the Co:CdS nanorods followed an oriented attachment mechanism. The Co-doping facilitated the morphology evolution of resulting product to nanorod structures, even in shorter reaction time. Additionally, the multifunctionality of the Co:CdS NCs including luminescence and ferromagnetism at room temperature was achieved due to the Co-doping. These Co:CdS NCs with controllable morphology were considered to have potential applications in spintronic devices and biological labeling fields, and this feasible synthesis method could be anticipated to be employed in industry manufacturing of Co:CdS NCs.

Conflict of Interests The authors declare that there is no conflict of interests regarding the publication of this paper.

Journal of Nanomaterials

[9]

[10]

[11]

[12]

[13]

[14]

Acknowledgments [15]

This work was supported by the Fundamental Research Funds for the Central Universities (2572014DB03), Heilongjiang Postdoctoral Fund (no. LBH-Z14012), Youths Science Foundation of Heilongjiang Province of China (QC2011C010), and the Science and Technology Research Foundation of Heilongjiang Province Department of Education (12513005). The authors thank Professor Yinong Wang of Dalian University of Technology for his help with TEM and HRTEM measurements and analysis.

References [1] T. Dietl, “A ten-year perspective on dilute magnetic semiconductors and oxides,” Nature Materials, vol. 9, no. 12, pp. 965–974, 2010. [2] M. J. Seong, H. Alawadhi, I. Miotkowski, A. K. Ramdas, and S. Miotkowska, “Raman electron paramagnetic resonance in Zn1−𝑥 Co𝑥 Te and Cd1−𝑥 Co𝑥 Te,” Physical Review B, vol. 63, no. 12, Article ID 125208, 7 pages, 2001. [3] J. M. Vila-Fungueiri˜no, B. Rivas-Murias, and F. Rivadulla, “Strong interfacial magnetic coupling in epitaxial bilayers of LaCoO3 /LaMnO3 prepared by chemical solution deposition,” Thin Solid Films, vol. 553, no. 28, pp. 81–84, 2014. [4] A. Alsaad, “Structural, electronic and magnetic properties of Fe, Co, Mn-doped GaN and ZnO diluted magnetic semiconductors,” Physica B: Condensed Matter, vol. 440, pp. 1–9, 2014. [5] S. Santra, H. Yang, P. H. Holloway, J. T. Stanley, and R. A. Mericle, “Synthesis of water-dispersible fluorescent, radioopaque, and paramagnetic CdS:Mn/ZnS quantum dots: a multifunctional probe for bioimaging,” Journal of the American Chemical Society, vol. 127, no. 6, pp. 1656–1657, 2005. [6] H. Ohno, D. Chiba, F. Matsukura et al., “Electric-field control of ferromagnetism,” Nature, vol. 408, no. 6815, pp. 944–946, 2000. [7] H. Ohno, “Making nonmagnetic semiconductors ferromagnetic,” Science, vol. 281, no. 5379, pp. 951–956, 1998. [8] Y. Ohno, D. K. Young, B. Beschoten, F. Matsukura, H. Ohno, and D. D. Awschalom, “Electrical spin injection in a ferromagnetic

[16]

[17]

[18]

[19]

[20]

[21]

[22]

[23]

semiconductor heterostructure,” Nature, vol. 402, no. 6763, pp. 790–792, 1999. R. Flederling, M. Kelm, G. Reuscher et al., “Injection and detection of a spin-polarized current in a light-emitting diode,” Nature, vol. 402, no. 6763, pp. 787–790, 1999. D. K. Dwivedi, D. Shankar, and M. Dubey, “Synthesis, structural and optical characterization of CdS nanoparticles,” Journal of Ovonic Research, vol. 6, no. 1, pp. 57–62, 2010. Q. Pang, B. C. Guo, C. L. Yang et al., “Cd1-xMnxS quantum dots: new synthesis and characterization,” Journal of Crystal Growth, vol. 269, no. 2–4, pp. 213–217, 2004. S. Salimian and S. Farjami Shayesteh, “Structural, optical and magnetic properties of Mn-doped CdS diluted magnetic semiconductor nanoparticles,” Journal of Superconductivity and Novel Magnetism, vol. 25, no. 6, pp. 2009–2014, 2012. A. K. Gupta and R. Kripal, “EPR and photoluminescence properties of Mn2+ doped CdS nanoparticles synthesized via coprecipitation method,” Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy, vol. 96, pp. 626–631, 2012. B. Santara, B. Pal, and P. K. Giri, “Signature of strong ferromagnetism and optical properties of Co doped TiO2 nanoparticles,” Journal of Applied Physics, vol. 110, no. 11, Article ID 114322, 2011. V. R. Singh, K. Ishigami, V. K. Verma et al., “Ferromagnetism of cobalt-doped anatase TiO2 studied by bulk- and surfacesensitive soft x-ray magnetic circular dichroism,” Applied Physics Letters, vol. 100, no. 24, Article ID 242404, 2012. P. Varshney, G. Srinet, R. Kumar et al., “Room temperature ferromagnetism in sol-gel prepared Co-doped ZnO,” Materials Science in Semiconductor Processing, vol. 15, no. 3, pp. 314–318, 2012. V. Ladizhansky and S. J. Vega, “Doping of CdS nanoparticles by Co2+ ions studied by NMR,” The Journal of Physical Chemistry, vol. 104, no. 22, pp. 5237–5241, 2000. P. V. Radovanovic and D. R. Gamelin, “Electronic absorption spectroscopy of cobalt ions in diluted magnetic semiconductor quantum dots: demonstration of an isocrystalline core/shell synthetic method,” Journal of the American Chemical Society, vol. 123, no. 49, pp. 12207–12214, 2001. K. A. Bogle, S. Ghosh, S. D. Dhole et al., “Co:CdS diluted magnetic semiconductor nanoparticles: radiation synthesis, dopant-defect complex formation, and unexpected magnetism,” Chemistry of Materials, vol. 20, no. 2, pp. 440–446, 2008. K. S. Hagen, D. W. Stephan, and R. H. Holm, “Metal ion exchange reactions in cage molecules: the systems [M4nM’n(SC6H5)10]2- (M, M’ = Fe(II), Co(II), Zn(II), Cd(II)) with adamantane-like stereochemistry and the structure of [Fe4(SC6H5)10]2-,” Inorganic Chemistry, vol. 21, no. 11, pp. 3928–3936, 1982. G. Giribabu, G. Murali, D. A. Reddy, C. Liu, and R. P. Vijayalakshmi, “Structural, optical and magnetic properties of Co doped CdS nanoparticles,” Journal of Alloys and Compounds, vol. 581, no. 25, pp. 363–368, 2013. R. Zeng, M. Rutherford, R. Xie, B. Zou, and X. Peng, “Synthesis of highly emissive Mn-Doped ZnSe nanocrystals without pyrophoric reagents,” Chemistry of Materials, vol. 22, no. 6, pp. 2107–2113, 2010. I. G. Dance, A. Choy, and M. L. Scudder, “Syntheses, properties, and molecular and crystal structures of (Me4 N)4 [E4 M10(SPh)16 ] (E = S, Se; M = Zn, Cd): molecular supertetrahedral fragments of the cubic metal chalcogenide lattice,” Journal of the American Chemical Society, vol. 106, no. 21, pp. 6285–6295, 1984.

Journal of Nanomaterials [24] T. Vossmeyer, G. Reck, B. Schulz, L. Katsikas, and H. Weller, “Double-layer superlattice structure built up of Cd32 S14 (SCH2 CH(OH)CH3 )36 .cntdot.4H2 O clusters,” Journal of the American Chemical Society, vol. 117, no. 51, pp. 12881– 12882, 1995. [25] Z. Li, W. Cai, and J. Sui, “Large-scale preparation of CdS quantum dots by direct thermolysis of a single-source precursor,” Nanotechnology, vol. 19, no. 3, Article ID 035602, 2008. [26] I. G. Dance, “Synthesis, crystal structure, and properties of the hexa(𝜇-benzenethiolato)tetra(benzenethiolatocobaltate(II)) dianion, the prototype cobalt(II)-thiolate molecular cluster,” Journal of the American Chemical Society, vol. 101, no. 21, pp. 6264–6273, 1979. [27] T. H. Ji, W.-B. Jian, and J. Y. Fang, “The first synthesis of Pb1-xMnxSe nanocrystals,” Journal of the American Chemical Society, vol. 125, no. 28, pp. 8448–8449, 2003. [28] I. G. Dance, “Applications of Cd N.M.R to polycadmium complexes. III. effects of Zn/Cd and of Se/S substitution in the tetra-adamantanoid cage [S4 Cd10 (Sph)16 ]4− in relation to heterometal metallothioneins,” Australian Journal of Chemistry, vol. 38, no. 12, pp. 1745–1755, 1985. [29] T. Løver, W. Henderson, G. A. Bowmaker, J. M. Seakins, and R. P. Cooney, “Electrospray mass spectrometry of thiophenolatecapped clusters of CdS, CdSe, and ZnS and of cadmium and zinc thiophenolate complexes: observation of fragmentation and metal, chalcogenide, and ligand exchange processes,” Inorganic Chemistry, vol. 36, no. 17, pp. 3711–3723, 1997. [30] C. B. Murray, D. J. Norris, M. G. Bawendi, and C. B. Murray, “Synthesis and characterization of nearly monodisperse CdE (E=sulfur, selenium, tellurium) semiconductor nanocrystallites,” Journal of the American Chemical Society, vol. 115, pp. 8706–8715, 1993. [31] W. W. Yu, L. Qu, W. Guo, and X. Peng, “Experimental determination of the extinction coefficient of CdTe, CdSe, and CdS nanocrystals,” Chemistry of Materials, vol. 15, pp. 2854–2860, 2003. [32] J. Cizeron and M. P. Pileni, “Solid solution of Cd𝑦 Zn1−𝑦 S nanosized particles: photophysical properties,” The Journal of Physical Chemistry B, vol. 101, no. 44, pp. 8887–8891, 1997. [33] G. Counio, T. Gacoin, and J. P. Boilot, “Synthesis and photoluminescence of Cd1−𝑥 Mn𝑥 S (x ≤ 5) nanocrystals,” The Journal of Physical Chemistry B, vol. 102, no. 27, pp. 5257–5260, 1998. [34] D. Kim, M. Miyamoto, and M. Nakayama, “Photoluminescence properties and energy transfer processes from excitons to Mn2+ ions in Mn2+ -doped CdS quantum dots prepared by a reversemicelle method,” Journal of Applied Physics, vol. 100, no. 9, Article ID 094313, 2006. [35] N. Chestnoy, T. D. Harris, R. Hull, and L. E. Brus, “Luminescence and photophysics of CdS semiconductor clusters: the nature of the emitting electronic state,” The Journal of Physical Chemistry, vol. 90, no. 15, pp. 3393–3399, 1986. [36] F. V. Mikulec, M. Kuno, M. Bennati, D. A. Hall, R. G. Griffin, and M. G. Bawendi, “Organometallic synthesis and spectroscopic characterization of manganese-doped CdSe nanocrystals,” Journal of the American Chemical Society, vol. 122, no. 11, pp. 2532– 2540, 2000. [37] M. Aven and J. S. Prener, Physics and Chemistry of II-VI Compounds, North-Holland, Amsterdam, The Netherlands, 1967. [38] B. D. Cullity and S. R. Stock, Elements of X-Ray Diffraction, Prentice Hall, New Jersey, NJ, USA, 2nd edition, 1978. [39] B. D. Yuhas, S. Fakra, M. A. Marcus, and P. Yang, “Probing the local coordination environment for transition metal dopants in

11

[40]

[41]

[42]

[43]

[44]

[45]

[46]

zinc oxide nanowires,” Nano Letters, vol. 7, no. 4, pp. 905–909, 2007. D. A. Schwartz, N. S. Norberg, Q. P. Nguyen, J. M. Parker, and D. R. Gamelin, “Magnetic quantum dots: synthesis, spectroscopy, and magnetism of Co2+ - and Ni2+ -doped ZnO nanocrystals,” Journal of the American Chemical Society, vol. 125, no. 43, pp. 13205–13218, 2003. J. M. D. Coey, M. Venkatesan, and C. B. Fitzgerald, “Donor impurity band exchange in dilute ferromagnetic oxides,” Nature Materials, vol. 4, no. 2, pp. 173–179, 2005. W. Cai, Z. Li, and J. Sui, “A facile single-source route to CdS nanorods,” Nanotechnology, vol. 19, no. 46, Article ID 465606, 2008. S. G. Thoma, A. Sanchez, P. P. Provencio, B. L. Abrams, and J. P. Wilcoxon, “Synthesis, optical properties, and growth mechanism of blue-emitting CdSe nanorods,” Journal of the American Chemical Society, vol. 127, no. 20, pp. 7611–7614, 2005. Z. Li, J. Sui, X. Li, and W. Cai, “Oriented attachment growth of quantum-sized CdS nanorods by direct thermolysis of singlesource precursor,” Langmuir, vol. 27, no. 6, pp. 2258–2264, 2011. Y. Cao, P. Hu, and D. Jia, “Phase- and shape-controlled hydrothermal synthesis of CdS nanoparticles, and oriented attachment growth of its hierarchical architectures,” Applied Surface Science, vol. 265, pp. 771–777, 2013. A. Sergey, A. S. Trifonov, M. Gicrsig et al., “One-and twodimensional arrays of magnetic nanoparticles by the LangmuirBiodgett technique,” Advanced Materials, vol. 11, no. 5, pp. 388– 392, 1999.

Journal of

Nanotechnology Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

International Journal of

International Journal of

Corrosion Hindawi Publishing Corporation http://www.hindawi.com

Polymer Science Volume 2014

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Smart Materials Research Hindawi Publishing Corporation http://www.hindawi.com

Journal of

Composites Volume 2014

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Journal of

Metallurgy

BioMed Research International Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Nanomaterials

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Submit your manuscripts at http://www.hindawi.com Journal of

Materials Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Journal of

Nanoparticles Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Nanomaterials Journal of

Advances in

Materials Science and Engineering Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Journal of

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Journal of

Nanoscience Hindawi Publishing Corporation http://www.hindawi.com

Scientifica

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Journal of

Coatings Volume 2014

Hindawi Publishing Corporation http://www.hindawi.com

Crystallography Volume 2014

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

The Scientific World Journal Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Journal of

Journal of

Textiles

Ceramics Hindawi Publishing Corporation http://www.hindawi.com

International Journal of

Biomaterials

Volume 2014

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014