J. Phys. Chem. C 2007, 111, 18087-18098

18087

Aqueous Phase Surfactant Selective Shape Controlled Synthesis of Lead Sulfide Nanocrystals Mandeep Singh Bakshi,*,†,‡,§ Pankaj Thakur,§,| Shweta Sachar,§ Gurpreet Kaur,§ Tarlok Singh Banipal,| Fred Possmayer,†,⊥ and Nils O. Petersen*,‡,# Department of Obstetrics and Gynecology, Department of Biochemistry, and Department of Chemistry, UniVersity of Western Ontario, 339 Windermere Road, London, Ontario, Canada N6A 5A5, National Institute for Nanotechnology, Edmonton, Alberta, Canada, and Department of Chemistry and Department of Applied Chemistry, Guru Nanak DeV UniVersity, Amritsar 143005, Punjab, India ReceiVed: July 13, 2007; In Final Form: July 22, 2007

Aqueous phase synthesis at 80 °C was carried out to synthesize lead sulfide (PbS) nanocrystals (NC) and microcrystals (MC) by using cationic twin-tail surfactants (TTS) such as 12-0-12, 10-2-10, 12-2-12, and 14-2-14 as capping agents in the concentration range from 0.1 -to 2 mM. The effect of hydrophobicity on the shape and size of PbS NC was evaluated by choosing DTAB as a reference surfactant for all TTS. TEM micrographs of PbS MC synthesized in the presence of DTAB indicated the formation of star-shaped MC with sizes between 3 -and 5 µm. An increase in the hydrophobicity, by introducing another tail in the basic structure of DTAB to make 12-0-12, significantly controlled the shape and size and lead to the formation of well-defined nanocubes and spheres 50-100 nm in size. Similarly, the effect of the hydrocarbon tail length on the shape controlled synthesis of PbS NC was systematically evaluated. Pyrene fluorescence measurements were used to determine the variation in the degree of hydrophobicity with respect to both chemical structure as well as concentration of TTS. It was concluded that a stronger hydrophobic character and higher concentration produced well-defined geometries of PbS NC. No significant concentration effect within a range of 0.1-2 mM DTAB and 10-2-10 was observed on the morphology of PbS NC probably due to a much weaker hydrophobicity of these surfactants. An attempt was made to present all TEM results in a schematic phase diagram. This phase diagram provided the best correlation between the shape and the size of PbS NC and the surfactant parameters (i.e., hydrophobicity and concentration effects). Apart from this, shape dependence UV-vis absorbance was also noted and discussed in context with an overall preview of all shapes of PbS NC/MC obtained. The shape controlled synthesis of PbS NC was obtained due to the preferential adsorption of TTS on the {111} crystal planes that directed the overall growth predominantly at the {100} planes. FTIR measurements were used to evaluate the adsorption of TTS on the PbS surface. A large shift in the stretching vibrations of TTS head functional groups suggested their orientation toward the PbS surface. This was further supported by the high-resolution XPS spectra of C 1s and N 1s of adsorbed TTS on the PbS surface. An effective interfacial adsorption of TTS on the surface of PbS NC driven by a stronger hydrophobic character is the key to achieve controlled PbS NC growth at the nanoscale.

Introduction Semiconductor nanocrystals have been extensively studied recently due to their vital applications in numerous technological areas such as microelectronics, electrooptics, nonlinear optics, light energy conversion, photocatalysis, photoelectrochemistry, and biological fluorescence labeling.1 Such applications are attributed to their quantum confinement effects, optical, electrical, and chemical properties, which are drastically different from the bulk.2 To synthesize novel architectures based on semiconductor nanocrystals is one of the challenging issues in material science.3 Among conventional semiconductors, lead sulfide (PbS) is particularly important because of its * Corresponding author. E-mails: [email protected]; Nils.Petersen@ nrc-cnrc.gc.ca. † Department of Obstetrics and Gynecology, University of Western Ontario. ‡ Department of Chemistry, University of Western Ontario. § Department of Chemistry, Guru Nanak Dev University. | Department of Applied Chemistry, Guru Nanak Dev University. ⊥ Department of Biochemistry, University of Western Ontario. # National Institute for Nanotechnology.

narrow band gap of 0.41 eV in the bulk form, a large excitation Bohr radius of 18 nm, and diverse morphologies.4 Most prominent among them include spheres,4a cubes,4b rodlike shapes,4a tubes,4c wires,4d truncated octahedrons,4b and dendritic4e, star-shaped4f, and flower-shaped structures.4g Different synthetic routes such as colloidal solutions,5a reverse micelles,5b solvothermal analysis,5c microwave analysis,5d and self-assembly4e have been employed for the fabrication of PbS nanoparticles. Recently, Zhao and Qi6a and Zhou et al.6b have reported the synthesis of various morphologies of PbS NC by using cetyltrimethylammonium bromide (CTAB), sodium dodecyl sulfate (SDS), their mixtures, and poly(vinylpyrrolidone) (PVP). They have mentioned that a precise control of surfactant concentration governs the overall morphology of PbS NC. According to their results, the surfactant molecules prefer to adsorb at the {111} facets due the presence of a greater charge density in comparison to {100}. This absorbtion promotes the growth at {100} facets of fcc geometry and results in the formation of various morphologies. Even the reaction time also

10.1021/jp075477c CCC: $37.00 © 2007 American Chemical Society Published on Web 11/20/2007

18088 J. Phys. Chem. C, Vol. 111, No. 49, 2007 SCHEME 1: Structure Formulas of Various TTS

significantly influences the morphology of PbS NC due to Ostwald ripening or surface diffusion effects. While surfactant-assisted synthesis of semiconductor NP has attracted much attention due to its soft-template effect, reproducibility, and aqueous phase reaction, little is known about the influence of surfactant hydrophobicity on the shape controlled aspects. We have attempted to address this issue by choosing a series of twin-tail cationic surfactants (TTS) (Scheme 1) and then using them as capping agents in the synthesis of PbS NC. TTS mentioned in Scheme 1 are more hydrophobic in comparison to their basic monomeric homologous (e.g., dodecyltrimethylammonium bromide, DTAB) and, hence, are considered to be even better shape directing agents due to their stronger amphiphilic nature. The capping ability of a shape directing ionic surfactant is directly related to its surface adsorption behavior, which in return is further related to its hydrophobicity as well as polarity. In the present study, both properties were systematically varied in a series of TTS to understand this effect. It was observed that hydrophobicity is the main contributing factor for a shape directing agent such as a surfactant to control the morphology of PbS NC. Experimental Procedures Materials. DTAB (99%) and sodium dodecyl sulfate (SDS, 99%) were received from Lancaster and Acros, respectively. Didodecyldimethylammonium bromide (12-0-12), dimethylenebis(decyldimethylammonium bromide) (10-2-10), dimethylenebis(dodecyldimethylammonium bromide) (12-2-12), and dimethylenebis(tetradecyldimethylammonium bromide) (14-214) were synthesized as reported in the literature.7 All surfactants were used after repeated crystallization from ethanol. Lead

Bakshi et al. acetate (99.9%), Pluronic P103 ((EO)17(PO)60(EO)17), thioacetamide (TAA, 98%), and acetic acid (99.5%) were purchased from Aldrich. Water was used after purification through double distillation. Preparation of PbS Nanoparticles. In a typical procedure,6 30 mL of distilled water was taken in a round-bottomed glass flask. Under constant stirring, 5 mL of aqueous TTS solution was added to the flask. This was followed by the addition of 4 mL of 1 M aqueous acetic acid. After this, 2 mL of 0.5 M aqueous lead acetate and 2 mL of aqueous 0.5 M thioacetamide were introduced into the solution under constant stirring at room temperature. After mixing all the components, the reaction mixture was kept in an oil bath at a temperature of 80 °C for 48 h under static conditions. This led to the formation of a black colloidal solution, indicating the formation of PbS NC. The colloidal suspension of PbS was collected and washed twice with water followed by several washings with methanol. Table 1 lists various TTS with different concentrations used in this study to synthesize PbS NC. Methods. UV-vis spectra of aqueous solutions were taken by a UV spectrophotometer (PerkinElmer Lambda 25) in the wavelength range of 200-900 nm. The shape and size of PbS NC were characterized by transmission electron microscopy (TEM). The samples were prepared by mounting a drop of a solution on a carbon coated Cu grid and allowing it to dry in air. The samples were observed with the help of a Philips CM10 transmission electron microscope operating at 100 kV. The X-ray diffraction (XRD) patterns were characterized with graphite monochromatized Cu- KR irradiation. The hydrophobicity of TTS in the aqueous reaction solution before heating at 80 °C for 48 h was studied with the help of pyrene fluorescence using a Hitachi fluorescence spectrophotometer F2500. The ratio of the intensities of I1 to I3 vibronic bands of the pyrene emission spectrum gives the polarity of the medium in which it is solubilized. The emission spectra were recorded, employing an excitation wavelength of 334 nm, and the intensities I1 and I3 were measured at wavelengths corresponding to the first and third vibronic bands located at ca. 373 and 384 nm. The surface chemical composition of PbS NC was confirmed with the help of X-ray photoelectron spectroscopy (XPS) measurements. A portion of the PbS solution in MeOH was placed onto a clean silicon wafer and then was put into the introduction chamber of the XPS instrument. The solvent was then pumped away under high vacuum. The sample was analyzed by using a Kratos Axis Ultra X-ray photoelectron spectrometer. XPS can detect all elements except hydrogen and helium and can probe the surface of the sample to a depth of 7-10 nm. Survey scan analyses were carried out with an analysis area of 300 µm × 700 µm. Infrared absorption measurements were recorded with a FTIR spectrometer (Shimadzu) in the range of 4000-400 cm-1 in the form of KBr pellets. Each spectrum was measured in transmission mode with 256 scans and 4 cm-1 resolution. Results Various TTS along with their cmc values used as capping agents in the present study are listed in Table 1. Mainly, a fixed range of concentrations of each TTS from 1 to 2 mM was used in the synthesis of PbS NC, although in some cases (e.g., 142-14), even less than 1 mM amounts were also used to understand the proper capping mechanism. DTAB (single-tail cationic surfactant) and P103 (a block copolymer and a nonionic surfactant) were used for the sake of comparison. The hydrophobicity of the present cationic surfactants was systematically

Shape Controlled Synthesis of PbS Nanocrystals

J. Phys. Chem. C, Vol. 111, No. 49, 2007 18089

TABLE 1: Various TTS with Different Concentrations Used as Shape Directing Agents for Synthesis of PbS Nanoparticlesa TTS/cmc DTABb /15.8 12-0-12 /0.2

concentration (mM), shape (in parentheses), and size 0.1 (star shape) ≈3 µm 0.1 (star shape) ≈2 µm (core-shell) 114 ( 15 nm 0.1 (star shape) ≈4 µm

0.5 (star shape) ≈3 µm 0.5 (cubes) 64.7 ( 5.4 nm

14-2-14 /0.16

0.01 (star shape) ≈3 µm 0.2 (cubes) 120 ( 18 nm

10-2-10 /6.5 P103b /0.026

0.1 (star shape) ≈3 µm 0.05 (star shape) ≈3 µm

0.02 (plate-like) 1.5 µm (cubes, 19%) 227 ( 39.2 nm (hexagonal, 81%) 277 ( 19 nm 0.5 (star shape) ≈3 µm

12-2-12 /0.84

0.5 (tree shape) variable sizes

1.0 (star shape) ≈5 µm 1.0 (cubes) 56 ( 3.8 nm (rods) 2.04 ( 0.11 1.0 (spheres, 62%) 62 ( 6.1 nm (rectangles, 38%) 55 ( 3.0 nm 0.05 (cubes) 144 ( 18 nm (truncated cubes) 103 ( 17.2 nm 1.0 (star shape) ≈3 µm

2.0 (star shape) ≈5 µm 2.0 (truncated cubes) 55.1 ( 7.1 nm 2.0 (hexagonal) 60 ( 4.9 nm 0.1 (cubes) 99 ( 14 nm 2.0 (star shape) ≈3 µm 2.0 (star shape) ≈3 µm

a Different shapes have been listed in parentheses along with the sizes. cmc is the critical micelle concentration (mM). b Listed for the sake of comparison.

Figure 1. TEM micrographs of PbS MC in the presence of DTAB. (a) [DTAB] ) 0.1 mM; (b) [DTAB] ) 1 mM; (c) magnified view of single PbS star-shaped MC with an approximate size of 5 µm; and (d) XRD pattern of PbS MC in the presence of DTAB ([DTAB] ) 5 mM).

increased by choosing DTAB as a reference surfactant. The first column of Table 1 and Scheme 1 lists various TTS and critical micelle concentration (cmc) values in pure water at 25 °C.8 TEM Measurements. All results of TEM measurements are summarized in Table 1. Figure 1 shows a typical fern-like starshaped PbS MC synthesized in the presence of 0.1 mM (Figure 1a) and 1 mM (Figure 1b) DTAB with an average size distribution close to 5 µm (Figure 1c). Similar morphologies in the presence of 0.5 and 2 mM DTAB have been shown in Supporting Information Figures S1a and S1b, respectively. We do not observe any appreciable change in the morphology of PbS MC with an increase in DTAB concentration from 0.1 to 2 mM. The XRD patterns of PbS MC prepared with 5 mM

DTAB are shown in Figure 1d. All peaks are very much prominent and refer to the cubic rock salt structure of a crystalline PbS. The intensity of the (200) peak is much higher than that of the (111) peak, which suggests a higher growth rate on the {100} facets in comparison to the {111} facets. DTAB is a strong ionic surfactant and is expected to adsorb electrostatically at the {111} facets of fcc geometry of PbS (Pb+/ S-) due to the presence of a greater charge driven by high atomic density.6 This would leave low atomic density {100} facets poorly capped or uncapped with surfactant molecules and remain available for further nucleation. This would consequently result in a preferential growth at the {100} facets in comparison to the {111} facets. On the contrary, in the case of a noble metal

18090 J. Phys. Chem. C, Vol. 111, No. 49, 2007

Bakshi et al.

Figure 2. TEM micrographs of PbS NC in the presence of different amounts of 12-0-12. Panel a shows core-shell-type arrangement, the left inset shows the thickness of shell ∼20 nm, and the right inset shows the amorphous nature of both core and shell at [12-0-12] ) 0.1 mM. Panel b shows fine monodispersed cubic crystalline NC at [12-0-12] ) 0.5 mM. Panel c shows spherical NC along with nanorods shown by block arrows at [12-0-12] ) 1 mM. Panel d shows truncated cubic NC at [12-0-12] ) 2 mM.

NP such as Au and Ag, surfactant molecules (CTAB and SDS) prefer to adsorb at the {100} facets of fcc geometry9 in comparison to the {111} facets. This preference practically comes from a relatively low atomic density at the {100} facets with appropriate atomic spacing suitable for the size of surfactant head groups to interact. Thus, PbS being a semiconductor possesses insufficient electron density at the {100} facets to attract surfactant molecules in comparison to that of Au or Ag. This fundamental difference of surfactant preferential adsorption at the {111} facets in the case of the PbS semiconductor from that at the {100} facets of Au or Ag tells us why mostly cubic or star-shaped geometries are obtained in the former case in comparison to nanorods/nanowires in the latter case.9 Interestingly, if another tail is introduced into a DTAB molecule to make it 12-0-12 (Scheme 1), the shape and size of the PbS MC dramatically changed. Figure 2 shows the TEM images at different concentrations of 12-0-12. At 0.1 mM, the sample still contains a few aggregates of predominantly starshaped MC with an average size close to 2 µm (Figure S2) but mostly consists of several small spherical NC (114 ( 15 nm, Figure S3a) with a clear core-shell-type arrangement (Figure 2a). The shell is approximately 20 nm thick (left inset of Figure 2a) and cannot be made up of a surfactant bilayer with a 0.1 mM amount. The identical selected area diffraction pattern (right inset of Figure 2a) for both core and shell indicates that they are amorphous without any crystalline nature. Recently, a similar behavior was also demonstrated by cadmium sulfide NC from our unpublished work and can be attributed to the presence of

an insufficient amount of surfactant (i.e., 0.1 mM). An increase in the amount at 0.5 mM (Figure 2b) produced fine monodisperse crystalline (see selected area diffraction pattern, left inset, XRD pattern shown in Figure S3XRD) cubic NC with an average size distribution of 64.7 ( 5.4 nm (Figure S3b). As the concentration of 12-0-12 is increased to 1 mM, all cubic NC were changed predominantly into spheres (Figure 2c) with an average size distribution of 56.4 ( 3.8 nm (Figure S3csphere). This sample also shows a few nanorods (shown by block arrows in Figure 2c) with an average aspect ratio of 2.04 ( 0.11 (Figure S3crod). A further increase in the concentration to 2 mM produced clear truncated cubic NC (Figure 2d) with an average size distribution still close to 55.1 ( 7.1 nm (Figure S3d). When a spacer of two methylene groups is introduced in the head group of 12-0-12, it is converted into a Gemini surfactant (i.e., 12-2-12). Use of 12-2-12 as a capping agent produces starshaped structures with an average size distribution close to 5 µm at 0.1 mM 12-2-12 (Figure S4a) and 0.2 mM 12-2-12 (Figure S4b). At 0.5 mM 12-2-12 (Figure 3a), the star-shaped morphologies were converted into typical fern-like structures, where one can find dark spots as small polyhedral NC embedded in the branches of a fern (shown by arrows in Figure 3b). Figure 3c shows XRD patterns indicating growth at the {100} facets. An increase in the concentration (12-2-12 ) 1 mM) produces a combination of spherical (62 ( 6.1 nm, Figure S4c) and rectangular (55 ( 3.0 nm, Figure S4d) NC with approximately 62% of the former and 38% of the latter. At 2 mM 12-2-12, a well-defined monodisperse hexagonal geometry with an average

Shape Controlled Synthesis of PbS Nanocrystals

J. Phys. Chem. C, Vol. 111, No. 49, 2007 18091

Figure 3. TEM micrographs of PbS NC in the presence of 12-2-12. Panel a shows fern-like MC at [12-0-12] ) 0.1 mM. Panel b shows small polyhedral NC embedded in the branches of a fern shown by block arrows. (c) XRD patterns showing a growth at {100} facet. Panel d shows spherical and cubic NC at [12-2-12] ) 1 mM. Panel e shows fine monoodisperse rectangular NC at [12-0-12] ) 2 mM.

size distribution of 60 ( 4.9 nm (Figure S4e) is obtained. It shows how a systematic increase in the amount of 12-2-12 from 0.1 to 2 mM transfers the star-shaped MC into monodisperse hexagonal NC. Consequently, if we increase the hydrophobicity further as in the case of 14-2-14 (Figure 4), again different morphologies (i.e., star-shaped and well-defined cubic NC) are observed but at much lower concentrations than those observed in the case of 12-2-12. At 0.01 mM 14-2-14 (Figure S5a), starshaped NC are obtained, while at 0.02 mM 14-2-14 (Figure S5b), the number of star NC decreases, and plate-like geometries appear (indicated by block arrows) that eventually seem to transform into predominantly large cubic NC with an average size distribution of 144 ( 17.8 nm at 0.05 mM 14-2-14 (Figure S5c). A further increase in the concentration (i.e., 0.1 and 0.2 mM) produces well-defined monodisperse cubic NC (Figure 4a,b, respectively) with average size distributions of 99 ( 14.1 nm (Figure S5d) and 120 ( 17.7 nm (Figure S5e), respectively. The shape transformation occurs when the concentration of 142-14 increases to 0.5 mM (Figure 4c). It generates a combination of large cubes (19%) and hexagonal (81%) NC with average size distributions of 227 ( 39.2 nm (Figure S5f) and 277 (

19.0 nm (Figure S5g), respectively. When a similar PbS synthesis was carried out in the presence of 10-2-10, which is less hydrophobic in comparison to 12-2-12 and 14-2-14 due to smaller twin-tails, only star-shaped NC were obtained at all 102-10 concentrations with appreciable anisotropic growth (Figure S6). No appreciable change was observed with an increase in the concentration from 0.1 to 2 mM just like that observed in the case of DTAB. The only noted difference was the predominant presence of hexadentate star-shaped MC (Figure S6e) with an average size of 2.75 µm in comparison to tetradentate (Figure 1c) obtained in the presence of DTAB. To check the surfactant specific shape controlled behavior, we performed identical reactions in mixed surfactant systems by selecting another oppositely charged surfactant component (SDS). The PbS NC/MC thus synthesized in the presence of 12-0-12+SDS ([12-0-12] ) 10 mM + [SDS] ) 10 mM) and 14-2-14+SDS ([14-2-14] ) 0.5 mM + [SDS] ) 0.5 mM) are shown in Figure 5. It should be noted that the addition of SDS in the case of 12-0-12 leads to the formation of star-shaped MC (Figure 5a) in comparison to the presence of well-defined spherical NC (see Figure 2). A similar addition in the case of

18092 J. Phys. Chem. C, Vol. 111, No. 49, 2007

Bakshi et al.

Figure 4. TEM micrographs of PbS NC in the presence of 14-2-14. Panel a shows cubic NC at [14-2-14] ) 0.1 mM. Panel b also shows cubic NC and its XRD pattern at [14-2-14] ) 0.2 mM. Panel c shows a combination of large cubic and hexagonal NC at [14-2-14] ) 0.5 mM.

Figure 5. TEM micrographs of PbS NC in the presence of different combinations of mixed surfactants. Panel a shows star-shaped MC in the presence of 12-0-12+SDS ([12-0-12] ) 10 mM + [SDS] ) 10 mM). Panel b shows large polyhedral and thread-like NC in the presence of 14-2-14+SDS ([14-2-14] ) 0.5 mM + [SDS] ) 0.5 mM).

14-2-14 eliminates the presence of well-defined PbS nanocubes (see Figure 4) and produces several micrometer-sized rod-shaped structures along with polyhedral geometries with an average size distribution of 259 ( 48 nm (Figure 5b). Thus, the induction of a negatively charged anionic surfactant such as SDS alters the capping ability of 12-0-12 and 14-2-14 by undergoing strong electrostatic interactions10 with them, and this consequently changes the morphology of MC. Apart from this, if P103, a nonionic surfactant, is used as a capping agent instead of TTS, then no ordered morphology is obtained except large PbS flaketype structures (Figure S7). Therefore, the absence of any

ordered morphology of PbS MC in the presence of mixed surfactant systems (i.e., 12-0-12+SDS and 14-2-14+SDS) as well as P103 ultimately suggests a high degree of TTS selectivity in controlling the overall shape of PbS NC. Reaction Time Dependent TEM Measurements. All TEM images shown in the previous section were taken for the samples after 48 h of reaction time. To check the effect of reaction time on the growth of PbS NC, samples were drawn out from the reaction mixture at different time intervals. Figure 6 shows various TEM micrographs of PbS NC synthesized in the presence of 1 mM 12-2-12 at different intervals of time. After

Shape Controlled Synthesis of PbS Nanocrystals

J. Phys. Chem. C, Vol. 111, No. 49, 2007 18093

Figure 6. TEM micrographs at different intervals of time with upper frames showing PbS NC in the presence of 12-2-12 ) 1 mM.

30 min after the start of reaction, mainly roughly spherical NC with rugged surfaces arranged in a typical pearl-necklace model were obtained. This situation essentially remains the same even after 1 h of reaction. After 2 h, shape transformation sets on with the appearance of cubic geometries (indicated by block arrows) but still with rugged surfaces, whereas after 4 h of reaction time, the shape of PbS NC becomes even more clear, which ultimately (after 48 h) takes a final clear shape of nanocubes and spheres (Figure 3d). As far as the size of the NC is concerned, there is no appreciable increase up to 4 h (∼30-40 nm), but it becomes almost double in 48 h (∼55-65 nm). This demonstrates that a longer reaction time provides a suitable reaction condition for annealing13 of the amorphous material to acquire well-defined crystalline geometries. The role of surfactant adsorption with respect to time is still unclear, but it seems that the annealing effect is the shape determining factor rather than any change in the surfactant adsorption since the free energy of interfacial adsorption should remain the same in a homogeneous thermodynamically stable system. An overview of TEM images indicates that the hydrophobicity of TTS appears to be the main shape directing factor. Figure 7 shows the variation of the pyrene I1/I3 intensity ratio at 25 °C of a pyrene emission spectrum in the presence of aqueous reaction surfactant solution. A value close to 1.7 suggests an aqueous polar environment, while a value close to 1.0 indicates a complete hydrophobic environment in which pyrene is solubilized.11 Although all reactions were carried out at 80 °C,

Figure 7. Variation of pyrene I1/I3 intensity ratio of the pyrene emission spectrum at 25 °C in a reaction mixture at different surfactant concentrations.

the variation of hydrophobicity for the present surfactants remains the same (not shown). A comparison among these values for all surfactants further demonstrates that DTAB with a maximum higher value produces the least hydrophobic environment, while 12-0-12 with a minimum value generates the maximum hydrophobic environment. Furthermore, there is a significant change in the I1/I3 value from 0.1 to 2 mM in the case of DTAB and 10-2-10, while it remains almost close to

18094 J. Phys. Chem. C, Vol. 111, No. 49, 2007

Figure 8. Schematic phase diagram showing different morphologies of PbS NC/MC with respect to hydrophobicity as well as concentration of each TTS (see text for details).

1.2 for 12-2-12 and 14-2-14. The I1/I3 value shows a regular decrease with the amount of 12-0-12 and remains close to 1. Keeping the hydrophobicity of the present surfactants in consideration, we have found that the shape controlling ability of each surfactant is directly related to this property.12 To authenticate this hypothesis, we summarized all TEM results in a schematic phase diagram (Figure 8) generated by the hydrophobicity parameter (I1/I3) (we used the same order as shown in Figure 7) versus the concentration of each TTS used in the present study. DTAB due to its least hydrophobicity (Figure 7) proves to be the weakest shape directing agent and allows maximum anisotropic growth (Figure 1). Note the tetradentate PbS MC, which is predominantly obtained at all concentrations. 10-2-10 is a little more hydrophobic than DTAB and mostly produces hexadentate star-shaped MC at all concentrations (Figure S6). 12-2-12 is significantly hydrophobic (Figure 7) as compared to DTAB and 10-2-10, and thus, starshaped structures are limited to 0.5 mM. Exceeding this amount shows an effective capping ability that drastically reduces the size of MC to a nanoscale range with a combination of nanocubes that ultimately acquires fine monodisperse hexagonal geometries at 2 mM. 14-2-14 with an even greater hydrophobicity than 12-2-12 (Figure 7) shows the presence of star-shaped and plate-like geometries only below 0.05 mM. From 0.05 to 0.2 mM, it leads to the formation of fairly monodisperse nanocubes that attain a much larger size along with the presence of hexagonal NC as the concentration reaches 0.5 mM. 12-012, being the maximum hydrophobic among all TTS, shows the presence of spherical core-shell-type PbS NC even at 0.1 mM, which take the form of monodisperse nanocubes at 0.5 mM, spherical NC at 1 mM, and truncated nanocubes at 2 mM. Thus, Figure 8 shows a strong dependence of both shape and size of PbS NC on the degree of hydrophobicty of the capping TTS, which is provided by two hydrocarbon chains (12-0-12) instead of one (DTAB) as well as by a longer hydrophobic tail (14-2-14) (Table 1). Apart from this, an increase in the amount of a TTS also proves to be significantly effective in controlling the anisotropic growth and results in the formation of monodisperse NC (Table 1). UV Measurements. The characteristic UV-vis absorbance properties of colloidal PbS NC synthesized in the presence of different TTS are shown in Figure 9. The measurements were performed after diluting at least 10 times the as-prepared solution in each case. Four different kinds of curves were observed for various PbS NC. The type A curve as in the case of DTAB

Bakshi et al.

Figure 9. UV-vis absorption spectra of aqueous colloidal PbS NC/ MC solution showing four types of curves corresponding to different morphologies of PbS NC/MC in the presence of (A) DTAB, (B) 120-12, (C) 12-0-12+SDS (1:1), and (D) 14-2-14.

shows a broad absorbance without any clear maximum. It is particularly due to the position dependent quantum size effects of multifaceted fern-type large PbS MC including scattering effects (Figure 1). The type B curve for 12-0-12 shows a small hump close to 340 nm (shown by an arrow) for small-sized nanocubes (Figure 2b) and is very much in line with the results reported in the literature for similar shape and size of PbS NC.6 This curve takes the shape of curve C when 12-0-12 is mixed with SDS in a 1:1 mol ratio, where a small hump is still visible around 350 nm (shown by an arrow) and a broad absorbance over the rest of the wavelength region might be due to the scattering of various fern-type morphologies observed for this sample (Figure 5a). On the other hand, the presence of fine PbS nanocubes in the case of 14-2-14 (Figure 4b) with almost double the size than those obtained for 12-0-12 (Figure 2a) shows a clear maximum around 750 nm (curve D in Figure 9). It takes the form similar to curve C when SDS is mixed with 14-2-14 due to the formation of star-shaped NC (not shown). The size dependent UV-vis absorbance for PbS NC has already been mentioned by Zhao and Qi.6 The primary cause for such a behavior arises from the large Bohr radius of PbS, which is close to 18 nm and demonstrates a contrasting difference from other semiconductor materials with a much smaller Bohr radius. This is why the latter materials do not show any significant shape and size dependent optical properties. The four different curves shown in Figure 9 for different kinds of PbS NC/MC are very much reproducible as far as their shape and size are concerned. Hence, one can quantify the UV-vis results of even semiconductor materials such as PbS with a large Bohr radius on the basis of their shape and size. FTIR Measurements. To quantify the surface adsorption of TTS on PbS NC, FTIR spectral studies of PbS NC synthesized in the presence of various TTS were performed. Figure 10 shows representative examples of the FTIR spectra of pure 12-0-12, 14-2-14, and that of PbS NC synthesized in the presence of both surfactants. The results for various surfactants are compared in Table 2. Table 2 shows that the symmetric and antisymmetric CH2 stretching vibrations (νsym(C-H) and νasym(C-H))14 of pure DTAB and in the presence of PbS NC ([DTAB] ) 5 mM) shift toward higher frequencies. Likewise, the scissoring mode of vibrations (δs(C-H)) shows a slight shift toward lower frequencies. There is not much shift in the νasym(N+-CH3) peak for DTAB in the presence of PbS NC, suggesting weak interactions of the alkylammonium head group with the PbS surface.

Shape Controlled Synthesis of PbS Nanocrystals

Figure 10. FTIR spectra of pure 12-0-12, 14-2-14, and that of PbS NC prepared in the presence of these surfactants.

However, the absence of (ν(C-N+)) bands at 935 and 912 cm-1 and the appearance of new bands at 1193, 1087, and 860 cm-1 in the presence of PbS NC clearly indicates that head groups of DTAB molecules are directed toward the PbS surface. Other weak peaks at 860 and 1165 cm-1 correspond to PbS.15,16 The νsym(C-H) and νasym(C-H) vibrations of 12-0-12 ([120-12] ) 5 mM) in the presence of PbS NC also shift to higher frequencies. The shift in the νasym(C-H) is much higher than that for DTAB. A higher gauche/trans conformer ratio of the methylene chain of 12-0-12 in the presence of PbS NC could be the cause of this shift.17,18 The scissoring modes of vibration have an almost similar behavior to that of DTAB. But, a significant shift in νasym(N+-CH3) toward higher frequencies suggests much stronger interactions of the 12-0-12 head group with the PbS surface in comparison to that of DTAB. This is further evident from the appearance of new bands of symmetric C-N+ stretching modes at 1193, 1135, and 860 cm-1 in the presence of PbS NC. Almost a similar behavior is observed in

J. Phys. Chem. C, Vol. 111, No. 49, 2007 18095 the case of 10-2-10, 12-2-12, and 14-2-14 (each 1 mM) with PbS NC as far as symmetric and antisymmetric CH2 stretching vibrations are concerned, but relatively much weaker shifts in the ν(C-N+) vibration modes of these surfactants in the presence of PbS indicate relatively weaker interactions with the PbS NC surface in comparison to that of 12-0-12. The data reported in Table 2 do not show any significant difference among the behavior of vibrational frequencies of twin-tails of different TTS (i.e., νsym(C-H) and νasym(C-H)), but one can observe a clear difference among the number of symmetric C-N+ stretching (ν(C-N+)) modes that provide useful information regarding the strength of head group interactions with the PbS surface. The number of ν(C-N+) modes is more in the case of DTAB and less in the case of 12-0-12, which demonstrates that some of the 12-0-12 head group stretching modes are absent, most probably due to much stronger interactions with the PbS surface. Although both DTAB and 12-0-12 possess the same tetralkylammonium head group, the much stronger hydrophobicity of the latter might be responsible for stronger interactions and consequently lead to this difference. In the case of dimeric gemini TTS with a spacer (i.e., 10-2-10, 12-2-12, and 14-2-14), the shift in the frequencies is not so significant (e.g., 1082 to 1083 cm-1 in the case of 10-2-10, 1166 to 1160 cm-1 in 12-2-12, and 1050 to 1060 cm-1 in 14-2-14), which shows their relatively weaker head group interactions with the PbS surface in comparison to that of 12-0-12. XPS Measurements. The surface composition of TTS capped PbS NC has been further studied by XPS measurements. Figures S8-S10 show the XPS spectra of PbS NC synthesized in the presence of 0.01, 0.02, and 0.2 mM 14-2-14, respectively. The high-resolution images for Pb 4f, S 2p, C 1s, and N 1s for PbS capped with 0.01 mM 14-2-14 are shown in Figure 11. The corresponding binding energies and area occupied by various species are listed in Table 3. Figure 11a shows strong emission peaks of Pb 4f7/2 and Pb 4f5/2 at 137.11 and 141.97 eV, while similar weak emission peaks are located at 138.30 and 143.16 eV, respectively.19 The former stronger peaks of Pb 4f are due to crystalline PbS, while the corresponding latter ones may be attributed to PbO formation.19d Figure 11b shows clear emission peaks of S 2p1/2 and S 2p3/2 at 161.51 and 160.33 eV, respectively. Figure 11c shows strong emission due to C 1s. Several species of C 1s from different functional groups constitute this strong emission. The percentage area occupied by each species indicates that the maximum emission (82%) comes from C-C and C-H functional groups, which constitute the hydrophobic tails of 14-2-14 (Table 3). On the other hand, weak emission due to N 1s (Figure 11d) is also evident from the dimeric ammonium head groups of 14-2-14. Thus, XPS results clearly indicate the presence of 14-2-14 on the surface of PbS as a capping agent. Similar behavior is shown by PbS NC with 0.02 and 0.2 mM 14-2-14(Table 3). Discussion All results from different studies indicate that TTS are efficient capping agents and hence effectively control the shape and size of PbS NC. Hydrophobicity is considered to be the main shape controlling factor as is evident from the TEM micrographs. However, FTIR and XPS studies as well as reported works6 indicate that the ionic surfactants interact with the {111} facets of PbS fcc geometry through electrostatic interactions and that the surfactant head group is oriented toward the PbS surface. If this is the situation, then how does hydrophobicity play a part in controlling the shape of PbS NC? We know that a stronger hydrophobic surfactant has a lesser

18096 J. Phys. Chem. C, Vol. 111, No. 49, 2007

Bakshi et al.

Figure 11. High-resolution XPS spectra for Pb 4f7/2 (a), S 2p (b), C 1s (c), and N 1s (d) for PbS NC capped with 14-2-14 ([14-2-14] ) 0.01 mM).

TABLE 2: Mode Assignments of Various TTS in the Absence and Presence of PbS NCa peak assignment νsym (C-H) νasym (C-H) δs (C-H) νasym (N+-CH3) ν (C-N+)

a

DTAB

DTAB-PbS

12-012

12-0-12-PbS

10-2-10

10-2-10-PbS

12-2-12

12-2-12-PbS

14-2-14

14-2-14-PbS

2849 2918 1473 1464 1487 1064 1045 1038 964 935 912

2883 2921 1463 1452 1484 1193 1087 961

2853 2921 1475 1468 1487 975 926

2883 2947 1473 1456 1508

2854 2914 1463

2883 2953 1466

2848 2921 1473

2848 2916 1470

1479 1082 1050 989

1508 1083 951

1489 1166 1062 980 937 903

1507 1161 1042 970

2850 2914 1473 1452 1492 1050 1023 984

2881 2947 1462 1456 1506 1060 951

ν, stretching; sym, symmetric; asym, antisymmetric; and δs, methylene scissoring.

solubility in the aqueous phase in comparison to a weaker hydrophobic surfactant. Thus, the former will have higher interfacial adsorption in comparison to the latter. Hence, a stronger hydrophobicity acts as a driving force for a better surface adsorption of surfactant molecules.20 This will push a greater number of strongly hydrophobic surfactant molecules from aqueous bulk at the water-PbS NC interface where they

would undergo preferential adsorption at the {111} planes with a greater charge density. Even a greater amount of a strongly hydrophobic surfactant will have a greater interfacial adsorption and would effectively control the shape evolution during the crystal growth. This is what we have seen in Figure 8 especially for 12-0-12, 12-2-12, and 14-2-14. A shape transformation of cubic geometry to a hexagonal/sphere (Figures 2-4) is expected

Shape Controlled Synthesis of PbS Nanocrystals

J. Phys. Chem. C, Vol. 111, No. 49, 2007 18097

TABLE 3: Binding Energies (eV) (I) and Area (%) (II) of Some Species from the Surface Composition of PbS NC in the Presence of Different Amounts of 14-2-14 as the Capping Agent Pb 4f5/2

Pb 4f7/2

S 2p1/2

S 2p3/2

C 1s

14-2-14

(I)

(II)

(I)

(II)

(I)

(II)

(I)

(II)

(I)

(II)

0.01 mM

141.97 143.16 142.36 143.32 142.07 143.18

31.5 11.2 38.5 4.2 33.5 9.3

137.11 138.30 137.50 138.46 137.21 138.32

42.2 15.0 51.5 5.7 44.8 12.5

161.51

33.3

160.33

66.7

285.00

82.3

161.90

33.3

160.72

66.7

285.00

81.7

161.64

33.3

160.46

66.7

285.00

80.5

0.02 mM 0.2 mM

to be driven by a nucleation shift from the {100} facets to the {110} facets. In the case of fcc noble metals such as Au and Ag, the ionic surfactants (such as CTAB) prefer to adsorb at the {100} or {110} facets due to larger interatomic distances among the surface atoms in comparison to that on the {111} facets, and thus, the growth is directed at the {111} facets.21 But, in the case of semiconductor nanomaterials such as PbS, due to the preferential adsorption of TTS on the {111} facets, the growth is directed at the {100} or {110} planes. It appears that a change in the interfacial environment with the increase in TTS concentration might shift the growth from the {100} planes to the {110} planes because the next preferential planes for TTS adsorption could be from the {111} to {100} facets. Regarding the concentration effects, although a large change in the degree of hydrophobicity of DTAB and 10-2-10 with respect to their amounts is observed (Figure 7), this has brought little change in the morphologies of PbS MC (Figure 8). Apart from this, it is also quite difficult to quantify the size of PbS MC over the range of concentrations of DTAB and 10-2-10 because of their highly anisotropic structures in both cases (Figure 1 and Figure S6, respectively). This means that both parameters(i.e., the hydrophobicity as well as the concentration) are unable to control the shape and size of PbS MC when DTAB and 10-2-10 are used within the said concentration range. A complete absence of well-defined PbS NC (e.g., nanocubes or spheres) even at 2 mM DTAB and 10-2-10 might indicate an insufficient degree of hydrophobicity that might require an effective interfacial adsorption to govern the shape and size. It is to be noted that cmc values of both DTAB and 10-2-10 at 25 °C are much higher than the concentration range selected here. Therefore, the pre-micellar concentration used for these surfactants might be insufficient to produce the required interfacial adsorption. We have preferred not to exceed the concentration range used here to compare the capping behavior among all present surfactants. On the contrary, there is a little change in the hydrophobicity of 12-2-12, 14-2-14, and 12-0-12 over the concentrations used (Figure 7), but an increase in their amounts from 0.1 to 2 mM shows a drastic effect on the shapes of PbS NC (Table 1). Thus, the concentration becomes the predominant shape controlling factor only for TTS with greater hydrophobicity. Conclusion PbS NC were synthesized in the presence of a series of TTS by varying their hydrophobicity and concentration in the range of 0.1-2 mM in an aqueous phase at 80 °C. It was observed that the hydrophobicity is the main contributing parameter for the shape controlled synthesis of PbS NC of ordered geometries. DTAB with the least hydrophobicity within the selected concentration range did not produce any well-defined MC with sizes in the nanometer range. The same is true of 10-2-10 with an almost same degree of hydrophobicity to that of DTAB. On the contrary, 12-0-12, 12-2-12, and 14-2-14 showed a significant

influence on the morphology of PbS NC. At low concentrations, all produced only large star-shaped MC, the shape of which slowly transformed into well-defined geometries such as spheres, cubes, and hexagonals as the concentration increased. The shape controlled effect is attributed to the effective adsorption of TTS on the {111} fcc crystal planes of PbS. This was mainly driven by the stronger amphiphilic character of the stronger hydrophobic TTS such as 12-0-12, 12-2-12, and 142-14 than DTAB and 10-2-10, which directs the subsequent crystal growth at the {100} or {110} planes. All results were summarized in a schematic phase diagram based on hydrophobicity as well as concentration parameters where a clear dependence of shape and size on these parameters was observed. It has been concluded that a stronger hydrophobic character of a capping surfactant is the key to achieve the shape controlled synthesis of PbS semiconductor nanomaterials at the nanoscale. To generalize this idea, we are working on other semiconductor materials such CdS and CuS with same TTS. Acknowledgment. This study was supported by Grants MOP 66406 and FRN 15462 from the Canadian Institutes of Health Research. Supporting Information Available: This material is available free of charge via the Internet at http://pubs.acs.org. References and Notes (1) (a) Honma, I.; Hirakawa, S.; Yamada, K.; Bae, J. M. Solid State Ionics 1999, 118, 29. (b) Phely Bobin, T. S.; Muisener, R. J.; Koberstein, J. T.; Papadimitrakopoulos, F. Synth. Methods 2001, 116, 439. (c) Curri, M. L.; Comparelli, R.; Cozzoli, P. D.; Mascolo, G.; Agostiano, A. Mater. Sci. Eng. C 2003, 23, 285. (d) Bruchez, M., Jr.; Moronne, M.; Gin, P.; Weiss, S.; Alivisators, A. P. Science (Washington, DC, U.S.) 1998, 281, 2013. (e) Chan, W. C.; Nie, S. M. Science (Washington, DC, U.S.) 1998, 281, 2016. (f) Mattoussi, H.; Mauro, J. M.; Goldman, E. R.; Anderson, G. P.; Sundar, V. C.; Mikulec, F. V.; Bawendi, M. G. J. Am. Chem. Soc. 2000, 447, 12142. (g) Bakueva, L.; Musikhin, S.; Hines, M. A.; Chang, T.-W. F.; Tzolov, M.; Scholes, G. D.; Sargent, E. H. Appl. Phys. Lett. 2003, 82, 2895. (h) Bakueva, L.; Konstantatos, G.; Levina, L.; Musikhin, S.; Sargent, E. H. Appl. Phys. Lett. 2004, 84, 3459. (2) (a) Sun, S.; Murray, C. B.; Weller, D.; Folks, L.; Moser, A. Science (Washington, DC, U.S.) 2000, 287, 1989. (b) Hyeon, T. Chem. Commun. 2003, 927. (c) Nirmal, M.; Brus, L. E. Acc. Chem. Res. 1999, 32, 407. (d) Jacobs, K.; Zaziski, D.; Scher, E. C.; Herhold, A. B.; Alivisatos A. P. Science (Washington, DC, U.S.) 2001, 293, 1803. (e) Michler, P.; Imamoglu, A.; Mason, M. D.; Carson, P. J.; Strouse, G. F.; Buratto, S. K. Nature (London, U.K.) 2000, 406, 968. (f) Kim, S.-W.; Son, S. U.; Lee, S. S.; Hyeon, T.; Chung, Y. K. Chem. Commun. 2001, 2212, 477. (3) Peng, X. Chem.sEur. J. 2002, 8, 335. (4) (a) Wang, S.; Yang, S. Langmuir 2000, 16, 389. (b) Lee, S. M.; Jun, Y. W.; Cho, S. N.; Cheon, J. W. J. Am. Chem. Soc. 2002, 124, 11244. (c) Leontidis, E.; Orphanou, M.; Kyprianidou-Leondidou, T.; Krumeich, F.; Caseri, W. Nano Lett. 2003, 3, 569. (d) Dai, H.; Wong, E. W.; Lu, Y. Z.; Fan, S.; Lieber, C. M. Nature (London, U.K.) 1995, 375, 769. (e) Kuang, D.; Xu, A.; Fang, Y.; Liu, H.; Frommen, C.; Fenske, D. AdV. Mater. 2003, 15, 1747. (f) Ma, Y.; Qi, L.; Ma, J.; Cheng, H. J. Cryst. Growth Des. 2004, 4, 351. (g) Ni, Y.; Liu, H.; Wang, F.; Liang, Y.; Hong, J.; Ma, X.; Xu, Z. J. Cryst. Growth Des. 2004, 4, 759. (5) (a) Rossetti, R.; Hull, R.; Gibson, J. M.; Brus, L. E. J. Chem. Phys. 1985, 83, 1406. (b) Khiew, P. S.; Radiman, S.; Huang, N. M.; Ahmad, M. S. J. Cryst. Growth. 2003, 254, 235. (c) Wang, D.; Yu, D.; Shao, M.; Liu,

18098 J. Phys. Chem. C, Vol. 111, No. 49, 2007 X.; Yu, W.; Qian, Y. Cryst. Growth Des. 2003, 257, 384. (d) Ni, Y.; Liu, H.; Wang, F.; Liang, Y.; Hong, J.; Ma, X.; Xu, Z. Cryst. Res. Technol. 2004, 39, 200. (6) (a) Zhao, N.; Qi, L. AdV. Mater. 2006, 18, 359. (b) Zhou, G.; Lu, M.; Xiu, Z.; Wang, S.; Zhang, H.; Zhou, Y.; Wang, S. J. Phys. Chem. 2006, 110, 6543. (7) (a) Bai, G.; Wang, J.; Yan, H.; Li, Z.; Thomas, R. K. J. Phys. Chem. B 2001, 105, 3105. (b) Zana, R.; Benrraou, M.; Rueff, R. Langmuir 1991, 7, 1072. (c) Wettig, S. D.; Verrall, R. E. J. Colloid Interface Sci. 2001, 244, 377. (8) (a) Menger, F. M.; Littau, C. A. J. Am. Chem. Soc. 1991, 113, 1451. (b) Zana, R. AdV. Colloid Interface Sci. 2002, 97, 205. (c) Bakshi, M. S.; Singh, J.; Kaur, G. J. Colloid Interface Sci. 2004, 285, 403. (d) Bakshi, M. S.; Sachar, S. Colloid Polym. Sci. 2005, 283, 671. (e) Bakshi, M. S.; Singh, J.; Kaur, G. J. Photochem. Photobiol. 2005, 173, 202. (f) Bakshi, M. S.; Kaur, G. J. Colloid Interface Sci. 2005, 289, 551. (g) Kanaebal, A.; Oda, R.; Mendes, E.; Candau, S. J. Langmuir 2000, 16, 2489. (9) (a) Ni, C.; Hassan, P. A.; Kaler, E. W. Langmuir 2005, 21, 3334. (b) Kuo, C. H.; Huang, M. H. Langmuir 2005, 21, 2012. (c) Kuo, C. H.; Chiang, T. F.; Chen, L. J.; Huang, M. H. Langmuir 2004, 20, 7820. (d) Johnson, C. J.; Dujardin, E.; Davis, S. A.; Murphy, C. J.; Mann, S. J. Mater. Chem. 2002, 12, 1765. (e) Bakshi, M. S.; Kaur, G.; Thakur, P.; Banipal, T. S.; Possmayer, F.; Petersen, N. O. J. Phys. Chem. C 2007, 111, 5932. (f) Bakshi, M. S.; Possmayer, F.; Petersen, N. O. Chem. Mater. 2007, 19, 1257. (g). Bakshi, M. S.; Kaura, A.; Bhandari, P.; Kaur, G.; Torigoe, K.; Esumi, K. J. Nanosci. Nanotechnol. 2006, 6, 1405. (h) Bakshi, M. S.; Sharma, P.; Banipal, T. S.; Kaur, G.; Torigoe, K.; Petersen, N. O; Possmayer, F. J. Nanosci. Nanotechnol. 2007, 7, 916. (10) (a) Shang, Y. Z.; Xu, Y.; Liu, H. L.; Hu, Y. J. Dispersion Sci. Technol. 2006, 27, 105. (b) Wang, B. Q.; Shang, Y. Z.; Liu, H. L.; Hu, Y. Fluid Phase Equilib. 2005, 228-229, 109. (c) Cheon, H. Y.; Jeong, N. H.; Kim, H. U. Bull. Korean Chem. Soc. 2005, 26, 107. (d) Kumar, A.; Alami, E.; Holmberg, K.; Seredyuk, V.; Menger, F. M. Colloids Surf., A 2003, 228, 197. (e) Alargova, R. G.; Kochijashky, I. I.; Sierra, M. L.; Kwetkat, K.; Zana, R. J. Colloid Interface Sci. 2001, 235, 119. (11) (a) Bakshi, M. S.; Kaur, N.; Mahajan, R. K. J. Photochem. Photobiol., A 2006, 183, 146. (b) Bakshi, M. S.; Kaur, G.; Yoshimura, T.; Esumi, K. Colloids Surf., A 2006, 281, 163. (c) Bakshi, M. S.; Singh, K.; Kaur, G.; Yoshimura, T.; Esumi, K. Colloids Surf., A 2006, 278, 129. (d) Bakshi, M. S.; Singh, K.; Singh, J. J. Colloid Interface Sci. 2006, 297, 284. (e) Bakshi, M. S.; Singh, J.; Kaur, G. Chem. Phys. Lipids 2005, 138, 81. (f) Bakshi, M. S.; Kaur, G. J. Colloid Interface Sci. 2005, 289, 551. (g) Bakshi, M. S.; Singh, K. J. Colloid Interface Sci. 2005, 287, 288. (h) Ray,

Bakshi et al. G. B.; Chakraborty, I.; Moulik, S. P. J. Colloid Interface Sci. 2006, 294, 248. (i) Haque, M. E.; Das, A. R.; Moulik, S. P. J. Phys. Chem. 1995, 99, 14032. (j) Nakano, T. Y.; Sugihara, G.; Nakashima, T.; Yu, S. C. Langmuir 2002, 18, 8777. (12) (a) Eguchi, M.; Du, Y. Z.; Ogawa, Y.; Okada, T.; Yumoto, N.; Kodaka, M. Int. J. Pharm. 2006, 311, 215. (b) Singh, B.; Florence, A. T. Int. J. Pharm. 2005, 298, 348. (c) Jiang, Z. L.; Pan, H. C.; Yuan, W. E. Chem. Res. Chin. UniV. 2004, 20, 523. (13) Espiau de Lamaestre, R.; Majimel, J.; Jomard, F.; Bernas, H. J. Phys. Chem. B 2005, 109, 19148. (b) de Lamaestre, R. E.; Bernas, H.; Ricolleau, C.; Majimel, J. Nucl. Instr. Methods Phys. Res., Sect. B 2006, 242, 214. (14) Sun, Y.; Hao, E.; Zhang, X.; Yang, B.; Shen, J.; Chi, L.; Fuchs, H. Langmuir 1997, 13, 5168. (15) Parvathy, N. N.; Pajonk, G. M.; Rao, A. V. J. Cryst. Growth 1997, 179, 249. (16) Cao, H.; Wang, G.; Zhang, S.; Zhang, X. Nanotechnology 2006, 17, 3280. (17) Weers, J. G.; Scheuing, D. R. Micellar shape to rod transitions. In Transform Infrared Spectroscopy in Colloid and Interface Science; ACS Symposium Series 447; American Chemical Society: Washington, DC, 1991. (18) Scheuing, D. R. Fourier transform infrared spectroscopy. In Transform Infrared Spectroscopy in Colloid and Interface Science; ACS Symposium Series 447; American Chemical Society: Washington, DC, 1991. (19) (a) Wang, S. Q.; Pan, A. L.; Yin, H. Y.; He, Y. P.; Lei, Y.; Xu, Z. D.; Zou, B. S. Mater. Lett. 2006, 60, 1242. (b) Lobo, A.; Moller, T.; Nagel, M.; Borchert, H.; Hickey, S. G.; Weller, H. J. Phys. Chem. B 2005, 109, 17422. (c) Chen, L. J.; Zhang, S. M.; Wu, Z. S.; Zhang, Z. J.; Dang, H. X. Mater. Lett. 2005, 59, 3119. (d) Gurin, V. S.; Kasparov, K. N.; Tyavlovskaya, E. A. Colloids Surf., A 1998, 139, 1. (20) (a) Zhou, Q.; Wu, Y. F.; Rosen, M. J. Langmuir 2003, 19, 7955. (b) Windsor, R.; Neivandt, D. J.; Davies, P. B. Langmuir 2002, 18, 2199. (c) Miller, R.; Fainerman, V. B.; Makievski, A. V.; Kragel, J.; Grigoriev, D. O.; Kazakov, V. N.; Sinyachenko, O. V. AdV. Colloid Interface Sci. 2000, 86, 39. (21) (a) Xiang, Y.; Wu, X.; Liu, D.; Jiang, X.; Chu, W.; Li, Z.; Ma, Y.; Zhou, W.; Xie, S. Nano Lett. 2006, 6, 2290. (b) Yu, Y. Y.; Chang, S. S.; Lee, C. L.; Wang, C. R. J. Phys. Chem. B 1997, 101, 6661. (c) Zhang, S.-H.; Jiang, Z.-Y.; Xie, Z.-X.; Xu, X.; Huang, R.-B.; Zheng, L.-S. J. Phys. Chem. B 2005, 109, 9416.

Aqueous Phase Surfactant Selective Shape Controlled ...

fluorescence using a Hitachi fluorescence spectrophotometer ..... we performed identical reactions in mixed surfactant systems .... shaped structures are limited to 0.5 mM. .... The data reported in Table 2 do not show any significant difference.

1MB Sizes 0 Downloads 233 Views

Recommend Documents

Synthesis of sulfur nanoparticles in aqueous surfactant ...
Dec 6, 2009 - After equilibration, the sample was soni- cated in a bath for 2 min and particle size was measured by DLS method immediately. CMC values of ...

Surfactant Selective Synthesis of Gold Nanowires by ...
HPS shift to lower frequencies indicating a less mobile ... DPPC+SDS+Au NP (IV) (see details in the text). TABLE 1: Mode .... Internet at http://pubs.acs.org.

Surfactant Selective Synthesis of Gold Nanowires by Using a DPPC ...
Fred Possmayer,†,‡ and Nils O. Petersen‡,§,#. Department of Obstetrics and Gynaecology, Department of Biochemistry, and Department of Chemistry,. UniVersity of Western Ontario, 339 Windermere Road, London, ON, Canada N6A 5A5, National Institut

Solution Phase Interactions Controlled Ordered ...
as DNA, chitosan, phospholipids, and BSA by using seed-mediated ... arranged in a typical pearl-necklace type arrangement except in the presence of BSA.

Aqueous-Phase Room-Temperature Synthesis of Gold ...
This step was repeated at least five to ten times to remove the maximum ..... Shalaev, V. M., Moskovits, M., Eds.; American Chemical Society: Washington, DC ...

Modeling Current-Mode-Controlled Three-Phase ...
Pursuing rapid simulation speed and requiring reflections on more circuit details are always contradictory in computer simulations, especially for those of ..... verification using a virtual prototyping system,” IEEE APEC 2005,. Austin, Texas, USA,

Continuous source of phase-controlled entangled two ... - umexpert
Jul 11, 2007 - C. H. Raymond Ooi*. Department of Physics, Korea Advanced Institute of Science and Technology, Daejeon, 305-701 Korea;. Frick Laboratory, Department of Chemistry, Princeton University, New Jersey 08544, USA; and Institute for Quantum S

Continuous source of phase-controlled entangled two ...
Jul 11, 2007 - This provides the possi- bility of coherently controlling the degree of entanglement in ... First, we discuss the physics of a two-photon emission laser using the ... Color online a Double Raman atom in a doubly resonant optical ...

Phase transitions for controlled Markov chains on ...
Schrödinger operators, J. Math-for-Ind., 4B (2012), pp. 105–108. [12] F. den Hollander, Large deviations, Fields Institute Monographs 14, AMS, Prov- idence, RI, 2000. [13] A. Hordijk, Dynamic Programming and Markov Potential Theory, Math. Centre. Tra

Continuous source of phase-controlled entangled two ... - umexpert
Jul 11, 2007 - J2. − i c p. Z*. 2 − Ic − Ip. ,. C5. J4 p c. Z bc 1 − i ac. + i 2 − Ic − Ip. ,. C6 where. Z. Ic − 2 ac bc − 2Ic + Ic − Ip. 2 + iIp ac + bc . C7. APPENDIX D: ...

Nonideal Mixing of SeTe in Aqueous Micellar Phase for ...
Nov 3, 2009 - concentration, large plate-like morphologies were obtained at xSe) 1, while the morphologies at .... Atmospheric Ultrathin Window (UTW) and the data were ..... because they were too small and appreciably water-soluble.

Nonideal Mixing of SeTe in Aqueous Micellar Phase for ...
Nov 3, 2009 - in versatile photocells and solar cells.1,2 Te also exhibits photoconductivity to ..... greater surface energy over rest of the crystal planes. The active ... on the surface of a NC shown in Figure 6b become the active sites for further

Attending to color and shape: The special role of location in selective ...
of location in selective visual processing. YEHOSHUA TSAL and NILLI LAVIE. Tel Aviv University, Tel Aviv, Israel. Subjects were presented with circular arrays of letters and were instructed to report first a given target (or targets) and then any oth

Controlled Breaking of Phase Symmetry in a ”Which ...
detector, which consists of an interferometer with quantum dots (QDs) in each of its .... describe electron motion around the interferometer in opposite directions.

Selective Memoization
Jan 16, 2002 - the programmer to express programs that reveal their true data dependences ...... the exploration is data-sensitive, only certain aspects of the ...

Regularized Active Shape Model for Shape Alignment
locating a landmark. ai equals to 0 implies that classifier can perfectly predict shape's boundary; while ai equals to 1 means classifier fails to ..... asm/index.html.

ALE 17. Phase Changes and Phase Diagrams
The temperature decreases as more and more of the liquid is converted into a gas. ... Use the data below to calculate the total heat in Joules needed to convert ...

Recent Advances in Surfactant EOR.pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Recent Advances in Surfactant EOR.pdf. Recent Advances in Surfactant EOR.pdf. Open. Extract. Open with. Sign

USA_V_CAMERON selective prosecution.pdf
UNITED STATES DISTRICT COURT. DISTRICT OF MAINE. UNITED STATES OF AMERICA ). ) v. ) CR-09-24-B-W. ) JAMES CAMERON ). AMENDED1 ORDER ...

Speed controlled machine tool
Sep 10, 2007 - 408/238; 483/30. See application ?le for complete search history. ... drive the electric motor by the rotation of the spindle, a tool holding part for ...... 160, monitor information concerning the rotational state of the generator 70

Electrolysis of Aqueous Solutions Lab.pdf
Electrolysis of Aqueous Solutions Lab.pdf. Electrolysis of Aqueous Solutions Lab.pdf. Open. Extract. Open with. Sign In. Main menu. Displaying Electrolysis of ...

Learning Selective Sum-Product Networks
Signal Processing and Speech Communication Lab, Graz University of Technology ... els are easy to learn and compete well with state of the art. 1. Introduction.